Você está na página 1de 728

BASEMENT TECTONICS 8

Proceedings of the International Conferences on Basement Tectonics


VOLUME 2
The titles published in this series are listed at the end a/this volume.
BASEMENT TECTONICS 8
Characterization and Comparison
of Ancient and Mesozoic
Continental Margins
Proceedings of the
Eighth International Conference on Basement Tectonics,
held in Butte, Montana, USA, August 1988
Edited by
MERVIN J. BARTHOLOMEW
Earth Sciences ami Resources Institute,
University of South Carolina, Columbia, SC, USA
DONALD W. HYNDMAN
Department ofGeology,
University of Montana, Missoula, MT, USA
DA VID W. MOGK
Department of Earth Sciences,
Montana State University, Bozeman, MT, USA
and
ROBERT MASON
Department of Geological Sciences,
Queen's University, Kingston, Ontario, Canada
SPRINGER SCIENCE+BUSINESS MEDIA, B.V.
Library of Congress Cataloging-in-Publication Data
International Conference on Basement Tectonics (8th
Mont. )
1988 Butte,
Basement tectonics 8 characterization and comparison of ancient
and Mesozoic continental margins proceedings of the Eighth
International Conference on Basement Tectonics, Butte, Montana, USA,
August 8-12, 1988 I edited by Mervin J. Bartholomew ... [et al.l.
p. cm. -- (Proceedings of the International Conferences on
Basement Tectonics ; v. 2)
Inc 1 udes index.
ISBN 978-94-010-4703-6
DOI 10.1007/978-94-011-1614-5
1. Geology, Structural--Congresses. 2. Continental margins-
-Congresses. 3. Geology,Stratigraphic--Meso2oic--Congresses.
4. Geology, Stratigraphic--Precambrian--Congresses. 5. Geology,
Stratigraphic--Paleozoic--Congresses. 1. Bartholomew, Mervin J.
II. Title. III. Title, Basement tectonics eight. IV. Series.
QE601.I6 1988
551.8--dc20 92-41537
ISBN 978-94-010-4703-6
Printed on acid-free paper
AII Rights Reserved
1992 Springer Science+Business Media Dordrecht
Originally published by Kluwer Academic Publishers in 1992
Softcover reprint of the hardcover 1 st edition 1992
No part of the material protected by this copyright notice may be reproduced or
utilized in any form or by any means, electronic or mechanical,
including photocopying, recording or by any information storage and
retrieval system, without written permission from the copyright owner.
ISBN 978-94-011-1614-5 (eBook)
CONFERENCE COMMITTEES
CHAIR
Mervin J. Bartholomew
Earth Sciences & Resources Institute, University of South Carolina" Columbia, SC 29208 USA
PROGRAM COMMITTEE
Donald W. Hyndman
Department of Geology, University of Montana, Missoula, MT 59812, USA
David W. Mogk
Department of Earth Sciences, Montana State University, Bozeman, MT 59717, USA
Robert Mason
Department of Geological Sciences, Queen's University, Kingston, Ontario, K7L 3N6, Canada
FIELD TRIP COMMITTEE
Sharon E. Lewis
Environmental Restoration Department, Westinghouse Savannah River Company,
P.O. Box 616, Aiken, SC 29802, USA
Richard B. Berg
Montana Bureau of Mines & Geology, Montana Tech, Butte, MT 59701, USA
Bill Bonnichsen
Tdaho Geological Survey, University of Idaho, Moscow, ID 83843, USA
J. Michael O'Neill
U.S. Geological Survey, Denver Federal Center, Denver, CO 80225, USA
TECHNICAL EDITING COMMITTEE
Mervin J. Bartholomew
Earth Sciences & Resources Institute, University of South Carolina, Columbia, SC 29208, USA
Constance M. Prynne
Earth Sciences & Resources Institute, University of South Carolina, Columbia, SC 29208, USA
William B. Macgregor
Humanities & Social Sciences Division, Montana Tech, Butte, MT 59701, USA
Brenda C. Sholes
Montana Bureau of Mines & Geology, Montana Tech, Butte, MT 59701, USA
v
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION, INC.
675 South 400 East, Salt Lake City, Utah 84111, USA
TRUSTEES
(1988 - 1990)
CHIEF TRUSTEE / CHAIRMAN OF THE BOARD OF TRUSTEES
John J. Gallagher, Jr.
ARCO Oil and Gas Company
2300 West Plano Parkway
Plano, TX 75075, USA
DEPUTY CHIEF TRUSTEE
Patrick J. Barosh
35 Potter Street
Concord, MA 01742, USA
ASSOCIATE TRUSTEE / TREASURER
Douglas F. Black
U.S. Geological Survey, National Center
Reston, VA 22092, USA
DEPUTY TRUSTEE
M. James Aldrich
Los Alamos National Laboratory, MSD-462
Los Alamos, NM 87545, USA
ASSISTANT TRUSTEE / SECRETARY
John James Prucha
Department of Geology, Syracuse University
Syracuse, NY 13244, USA
vi
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
FOREWORD
The 8th International Conference on Basement Tectonics was held in Butte, Montana, August 8-12,1988.
Historically, basement tectonics conferences have focused on such topics as reactivation of faults, the influence
of basement faults on metallogeny and hyrocarbon accumulation, and the use of geophysical and remote sensing
techniques to interpret subsurface and surface geology. The 8th Conference diverged from past conferences in
that a unifying theme was selected. Because ancient major terrane or cratonic boundaries are often postulated
to be fault zones which are subsequently reactivated, the conference was organized to examine all aspects of
ancient continental margins and terrane boundaries and to compare younger (Mesozoic) ones, about which more
is known, with older (Paleozoic and Precambrian) ones. Moreover, because the 8th Conference was held in the
northwestern United States, a greater emphasis was placed on the Mesozoic margin of western North America
and the North American shield. The seven oral sessions and four poster sessions all dealt with aspects of the
conference theme: characterization and comparison of ancient continental margins.
The organizers extend their thanks to those individuals who graciously consented to serve as moderators for
the oral sessions: John M. Bartley, Mark S. Gettings, M. Charles Gilbert, John M. Guilbert, Donald W.
Hyndman, William P. Leeman, Robert Mason, and A. Krishna Sinha. The program with abstracts volume was
prepared by S. E. Lewis and M. J. Bartholomew.
Ninety percent of the submitted manuscripts were accepted for publication in this proceedings volume after
peer review by at least two reviewers and appropriate revision. These forty-eight papers and three extended
abstracts make up the first six chapters of this volume. Some papers deal with more than one aspect of ancient
continental margins, so the editors placed each paper in the chapter which appeared to be most relevant.
Abstracts for the other thirty-six conference presentations are presented in Chapter 7. The time, effort, and
patience of the reviewers listed overleaf is greatly appreciated, particularily those many reviewers who
consented to review two papers.
The 8th Conference was preceded by a field trip across Precambrian terranes in Montana and Wyoming.
The post-meeting field trip crossed from the Mesozoic cratonic area into accreted terranes in Idaho and Oregon.
The eleven field guides for these excursions and eleven papers on related geology were published for the 8th
Conference by the Montana Bureau of Mines and Geology (Lewis and Berg, 1988).*
Financial support for the 8th Conference and for the plate margins and field guide volume was primarily
provided through the Office of Basic Energy Sciences, U.S. Department of Energy Grant number DE-FG02-
88ER13962 (to Bartholomew). ARCO Oil and Gas Company provided funds for preparation of the circular used
to advertise the 8th Conference. Financial support for publication of this proceedings volume was primarily
provided through the Office of Basic Energy Sciences, U.S. Department of Energy Grant number DE-FG02-
91ER14201 (Bartholomew), and through the Earth Sciences & Resources Institute, University of South
Carolina (Bartholomew). Additional financial support was obtained through grants from the Research and
Scholarly Activities Committee, Montana Tech (Bartholomew), the Office of Research Administration,
University of Montana (Hyndman), the Department of Geology, University of Montana (Hyndman), the
Department of Earth Sciences, Montana State University (Mogk), and the Department of Geological Sciences,
Queen's University (Mason). All contributions are greatly appreciated.
Assistance with page layout, drafting, and photography was provided by John W. Jones and Ric Pantonial
in Columbia, South Carolina, and by Mayrose E. Tompkins, Roger Holmes, Robert Dal Porto, Kristine L. Shifty,
Dana M. Penrose, and Laurie M. Boyle in Butte, Montana.
*LEWIS, S. E., and R. B. BERG (eds.), 1988, Precambrian and Mesozoic Plate Margins: Montana, Idaho and Wyoming,
with Field Guides for the 8th International Conference on Basement Tectonics (August 1988): Montana Bureau of
Mines and Geology, Special Publication 96, 195 p.
vii
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
REVIEWERS OF MANUSCRIPTS
The Editors extend their thanks to the following individuals who graciously consented to serve as reviewers
for manuscripts which were considered for publication in this proceedings volume:
M. James Aldrich
Elaine A. Aliberti
DavidAlt
Fred Barker
Richard B. Berg
Steven C. Bergman
Andy R. Bobyarchick
Ronald A. Burwash
Dugald M. Carmichael
Alan Clark
John K. Costain
Stewart S. Farrar
Michael F. FolIo
John J. Gallagher, Jr.
M. Charles Gilbert
Arthur Goldstein
John W. Goodge
John M. Guilbert
Vicki L. Hansen
Norman L. Hatch, Jr.
Robert D. Hatcher, Jr.
Herwart H. Helmstaedt
Darrell J. Henry
Robert J. Hooper
Scott S. Hughes
Jonathan Husch
Harold L. James
David Jones
Stephen A. Kish
Byron Kulander
David R. Lageson
Hans G. Ave Lallemant
IanM. Lange
Sharon E. Lewis
Michael T. Lukert
Calvin F. Miller
viii
Gautam Mitra
Alan P. Morris
Sharon Mosher
LeRoy A. Odom
Jim Oliver
J. Michael O'Neill
K. Howard Polsen
Karen W. Porter
Robert P. Raeside
Michael R. Rampino
Douglas W. Rankin
Nicholas Rast
Brady P. Rhodes
Francois Robert
Edwin S. Robinson
Christopher J. Schmidt
James G. Schmitt
Klaus J. Schulz
James W. Sears
Donald T. Secor
Mark A. Sholes
Carol Simpson
A. Krishna Sinha
WalterS. Snyder
J. Alexander Speer
Edgar W. Spencer
William A. Thomas
Richard P. Tollo
James F. Tull
Alexis V olborth
Neil A. Wells
Donald U. Wise
Charles Wood
Joseph L. Wooden
Lee A. Woodward
Gerald A. Zieg
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
TABLE OF CONTENTS
FOREWORD ..................................................................... . vii
CHAPTER 1: CONCEPTS, TECHNIQUES, AND AMBIGUITIES ....................................... 1
Evolution of convergent plates
W. B. Hamilton
Ancient collisional continental margins in the Canadian Shield: Geophysical signatures and
3
derived crustal transects ........................................................................... 5
M.D. Thomas
Tectonothermics of modern and ancient continental margins
M. J.Drury
Geochronological studies offault-related rocks
P. D. Fullagar
The importance of subduction-related margin-parallel shear zones along transpressional
27
37
convergent plate margins ......................................................................... 51
V. L. Hansen
Kinematic indicators in shear zones
C. K.Mawer
Zone of weakness concept: A review and evaluation
J. J. Prucha
Petrogenetic evaluation oftrace element discrimination diagrams
F. a.Dudas
Metallogeny at Precambrian and Mesozoic continental margins
J. M. Guilbert
IX
67
83
93
129
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
CHAPTER 2: CORDILLERAN MESOZOIC MARGIN .............................................. 143
Post-Archean crustal evolution, terrane accretion, and metamorphism of the Western
Cordillera, with emphasis on northern and central California ....................................... 145
W. G. Ernst
Mesozoic and Cenozoic intrusions and batholiths of the circum-Pacific region as analogues
of pre-Phanerozoic batholiths: A summary ........................................................ 169
D.A.Brew
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada, California,
and implications for transform offset of the SierranIKlamath convergent margins ..................... 179
Y. Dilek and E. M. Moores
Problems concerning collisional vs noncollisional deformation at continental-margin
orogens, with an example from the Mesozoic Cordilleran orogen ..................................... 197
S. H. Edelman
Plutonism across the Tujunga-North American terrane boundary: A middle to upper
crustal view of two juxtaposed magmatic arcs ...................................................... 205
J. L. Anderson, A. P. Barth, E. D. Young, E. E. Bender,
M. J. Davis, D. L. Farber, E. M. Hayes, and K. A. Johnson
Geophysical investigations of the cratonic margin in the Pacific Northwest (USA)
R. L. Thiessen, K. R. Johnson, and G. B. Mohl
Sedimentation and basin evolution along the Mesozoic margin of western North America
W. L. Bilodeau
Thick-skinned overstep tectonics in the Jurassic Winnemucca fold-and-thrust belt,
231
241
north-central Nevada, U.S.A.: Evidence from the Sonoma Range ................................... 249
S. D. Stahl
CHAPTER 3: PRECAMBRIAN MARGINS ........................................................ 263
Generation of granitoids at Archaean continental margins in southern Africa
D. R. Hunter
The northern Wyoming Province: Contrasts in Archean crustal evolution
D. W. Mogk, P. A Mueller, J. L. Wooden, and D. R.Bowes
Evidence for the amalgamation of Archean oceanic and continental blocks to form
265
283
the Beartooth Plateau ........................................................................... 299
J. K. Meen
x
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Precambrian geology and ductile normal faulting in the southwest corner of the
Beartooth Uplift, Montana ....................................................................... 313
E. A. Erslev
Geochemistry and origin of amphibolite and ultramafic rocks, Branham Lakes area,
Tobacco Root Mountains, southwestern Montana ................................................... 323
M. L. Cummings and W. R. McCulloch
Multiple reactivation of a collisional boundary: An example from southwestern Montana
B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
Long-lived basement weak zones and their role in extensional magmatism in the Ogilvie
341
Mountains, Yukon Territory ..................................................................... 359
C. F. Roots and R. 1. Thompson
Geotectonics of the cratonic margin from paleomagnetism of the Middle Proterozoic
Aldridge (Prichard) Formation and Moyie Sills of British Columbia and Montana ...................... 373
D. T. A. Symons and E. A. Timmins
Impact origin oflarge intracratonic basins, the stationary Proterozoic crust, and
the transition to modern plate tectonics ............................................................ 385
J. W. Sears and D. Alt
CHAPTER 4: APPALACHIAN MARGINS ......................................................... 393
Late Precambrian tectonism - The opening of the Iapetus Ocean
N. Rast
Transition from alluvial to deep-water sedimentation in the Lower Lynchburg Group
395
(Upper Proterozoic), Virginia ..................................................................... 407
F. Wehr
The Robertson River Igneous Suite (Blue Ridge Province, Virginia) - Late Proterozoic,
anorogenic (A-type) granitoids of unique petrochemical affinity ...................................... 425
R. P. Tollo and S. Arav
Structural characterization ofthe Late Proterozoic (post-Grenville) continental
margin of the Laurentian craton .................................................................. 443
M. J. Bartholomew
Effect of the Transylvania fracture zone on evolution of the western margin of the
Central Appalachian basin ....................................................................... 469
S. 1. Root
xi
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Late Paleozoic destruction of the western proto-Atlantic margin in the southern
Appalachians ................................................................................... 481
S.A. Kish
The Pine Mountain window of Alabama: Basement-cover evolution in the southernmost
exposed Appalachians ........................................................................... 491
M. G. Steltenpohl
Tectonic implications of the brittle fracture history of the Permian Narragansett Pier
Granite, Rhode Island ........................................................................... 503
M. J. Hozik
Comparison of sedimentary petrologic sequences produced in Mesozoic extensional
and convergent tectonic settings in North America ................................................. 527
P. T.Ryberg
Structure of the Mesozoic Marietta-Tryon graben, South Carolina and adjacent
North Carolina ................................................................................. 539
J. M. Garihan and W. A. Ranson
CHAPTER 5: PROBLEMS AT MARGINS IN EUROPE AND ASIA ................................... 557
Strain pattern in the Purbeck-Isle of Wight Monocline: A case study offolding due
to dip-slip fault in the basement .................................................................. 559
M.S.Ameen
Characteristics of the circum-Tyrrhenian Hercynian massifs and their role in the
Alpine-Apennine chains ......................................................................... 579
N. Minzoni
Structure and tectonic evolution of the western continental margin ofIndia: Evidence
from subsidence studies for a 25-20 Ma plate reorganization in the Indian Ocean . . . . . . . . . . . . . . . . . . . . . . . 583
A. Agrawal and J. J. W. Rogers
Tectonics and regional unconformities of eastern India basins
J. J. Gallagher, Jr
Basement-cover relationships of the Peninsular Gneiss with the greenstone belts of
591
Karnataka, India ............................................................................... 599
A. G. U garkar
xii
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
CHAPTER 6: METALLOGENY AND CONTINENTAL MARGINS .................................... 601
Earthquake faulting, induced fluid flow, and fault-hosted gold-quartz mineralization
R. H. Sibson
Transpressive tectonics and the Archean gold deposits of Superior Province,
603
Canadian Shield ................................................................................ 615
K. H. Poulsen, F. Robert, and K. D. Card
Structural controls during formation and deformation of Archean lode-gold deposits
in the Canadian Shield .......................................................................... 625
R. Mason and H. H. Helmstaedt
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera
W. J. McMillan and A. Panteleyev
Tectonic setting of Au-Ag deposits hosted by Proterozoic strata along the Lewis and
633
Clark Line, west-central Montana ................................................................ 653
L. A. Woodward
Geotectonic setting of western Pacific gold deposits
R. H. Sillitoe
Felsic magmatism and hydrothermal gold deposits: A tectonic perspective
R. Mason
Two-stage basement fault-block deformation in the development of the
665
679
Witwatersrand goldfields, South Africa ............................................................ 689
R. Myers, I. G. Stanistreet, and T. S. McCarthy
Archean crustal lead in the Helena Embayment of the Belt Basin, Montana
R. E. Zartman
Tectonic development of base-metal and barite-vein deposits associated with the
699
early Mesozoic basins of eastern North America .................................................... 711
G. R. Robinson, Jr.
CHAPTER 7: ABSTRACTS OF OTHER CONFERENCE PRESENTATIONS .......................... 727
Features of volcanic rocks and tectonic settings
A. T. Anderson, Jr.
The role of water in the formation of granulite and amphibolite facies rocks
L. M. Angeloni
xiii
728
728
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Comparison of the late Precambrian-early Paleozoic and Mesozoic continental margins
of the northeast United States .................................................................... 728
P. J. Barosh
Tectonic setting of Tertiary hydrothermal systems in the Great Basin region
B. R. Berger
Precise geochronology and isotope systematics as tools in tectonic and petrologic analysis
M. E. Bickford
Regional structural setting of sediment-hosted gold deposits in Nevada
H. F. Bonham, Jr.
Precambrian continental margins
K. Burke
The western Canada convergent margin: Structure and tectonic history
R.M. Clowes
Two types off oreland basin in Sichuan microcraton margin, China
H. Dong, H. Lu, X. Deng, and B. Wu
The formation and emplacement oflarge thrust sheets on the eastern margin of
729
729
730
730
730
731
North America ................................................................................. 731
M. A. Evans
Sedimentology and organic geochemistry of the Jurassic Newark Basin, northern
New Jersey ..................................................................................... 731
M. S. Fedosh, J. P. Smoot, and R. K. Kotra
Tectonic controls on gold mineralization in the southern Appalachian Piedmont
P. G. Feiss
Activation of the central crystalline root zone of the Himalaya
L. N. Gupta and N. Chaudhri
Recognition of ancient continental margins in the Canadian Shield
H. H. Helmstaedt
U sing paleomagnetism and aeromagnetism to delineate ancient continental margins
J. W. Hillhouse
The knowledge offaults and lithostratigraphy for a reevaluation of the Abitibi and
732
732
732
733
Pontiac subprovinces ............................................................................ 733
M.Hocq
Tectonics of gold mineralization in the Precambrian of South India
D. Jayakumar
xiv
734
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Metallotectonic analysis of the Canadian Appalachians
J.D. Keppie
From passive to active to transform: Structural history of the European-Apulian margins
734
during the Alpine cycle .......................................................................... 734
B.Lammerer
Tampere Schist Belt: Structural style within a convergent plate margin in the Early
Proterozoic of southern Finland ................................................................... 735
M. Nironen
Archean volcanic-sedimentary (greenstone) belts
R. W. Ojakangas
Use of granitic type in tectonic analysis: Mid-Paleozoic magmatism, eastern Massachusetts
M. L. PaigeandR. Hon
Thermal history of the Bakersville metagabbro and metadiabase dikes in the Blue
735
736
Ridge, northwestern North Carolina and adjacent Tennessee ........................................ 736
L. D. Rainey, S. A. Goldberg, and J. R. Butler
The Midland fish-bed: An Early Jurassic transgressive-regressive lacustrine sequence
in the Culpeper Basin, Virginia ................................................................... 736
S. c. Roberts and P. J. W. Gore
Varieties of Proterozoic orogeny
J. J. W. Rogers
Metamorphism and tectonics
J. W. Sears
The Red Sea line - A Late Proterozoic transcurrent fault
A. E. Shimron
The geological definition ofseismogenic continental crust
R. H. Sibson
Late Precambrian and Paleozoic accretionary history of the central and southern
737
737
737
738
Appalachians: Characterization through magmatic and deformational events ........................ 738
A. K. Sinha, R. L. Badger, E. Hund, J. P. Hogan, and R. E. Guy
Crustal extensional processes in Cordilleran metamorphic core complexes
A. W. Snoke
Fluids in the crust: Use of stable isotopes to distinguish magmatic/hydrothermal
739
events in fragments of oceanic and arc crust ........................................................ 739
D. S. Stakes
xv
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Large-scale hydrothermal systems and gold deposition along major fault zones in collisional
margins: Characteristics of the Mother Lode, California, and similar systems in B.C.,and
their comparison to other types of geothermal systems .............................................. 740
B. E. Taylor
Structural analysis of the central Columbia Plateau utilizing radar, Landsat, digital
topography, and magnetic data bases ............................................................. 740
R. L. Thiessen, J. R. Eliason, C. W. Brougher, K. R. Johnson,
G. B. Mohl, M. Foley, and D. E. Beaver
Late Cretaceous tectonics of the Blythe-Quartzite region, SE California and SW Arizona:
Deformation within an oblique-slip orogen ........................................................ 740
R. M. Tosdal
Mesoscopic analysis of brittle deformation zones and fault rocks: Mesozoic extensional
reactivation of the late Paleozoic Hylas fault zone, Virginia ......................................... 741
R. Venkatakrishnan and A. Watkins
Pb isotopic characteristics of Mesozoic intrusive magmatism along the craton margin
in the western USA ............................................................................. 741
J. L. Wooden, J. S. Stacey, A. C. Robinson, R. W. Kistler,
R. M. Tosdal, and M. J. Whitehouse
AUTHOR INDEX ............................................................................... 743
xvi
CHAPTER 1
CONCEPTS, TECHNIQUES,
AND AMBIGUITIES
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Evolution of convergent plates
WARREN B. HAMILTON
U.S. Geological Survey, Denver, Colorado 80225, USA
(received and accepted September 8,1988)
EXTENDED ABSTRACT
Plate boundaries vary greatly along strike at
anyone time, change rapidly, and evolve complexly,
so histories of most arcs are highly variable along
trend. Subducting plates sink more steeply than
they dip, and overriding plates advance over them
as hinges roll back. Magmatic arcs are products of
sinking slabs and approximately track lOO-km
depth contours on slabs. Oceanic arcs migrate
forward above sinking slabs and lengthen with
time, and new oceanic crust forms by sea-floor
spreading behind oceanic arcs or is plated out in
lumpy form by fast-migrating magmatic arcs.
Subduction occurs beneath only one side of an
internally rigid plate at a time.
Collisions progress along trend and can be
between two active arcs, intervening lithosphere
sinking beneath both, or between active and passive
margins. Arcs, aggregated by offshore collisions,
collide in turn with continents. Collisions produce
great crustal shortening, but plates overriding
subducting oceanic lithosphere commonly are ex-
tended. Extensional thinning can exceed magmatic
crustal thickening, as in modern North Island, New
Zealand, or be subordinate to it, as in much of the
Andes. Most collisions are followed by subduction
reversals or jumps as new subduction systems break
through beneath the aggregates; new trenches
commonly are by-products of collisions. A strip of
back-arc basin crust often remains attached to a
collided arc after subduction reversal and becomes
the ophiolitic basement of a fore-arc basin. Sedi-
mentation in trenches is mostly longitudinal so
clastic sediments incorporated into accretionary
wedges can come from distant sources rather than
from adjacent arcs. High-pressure metamorphism
occurs beneath overriding plates, not in front of
them.
Exposures of deep arc crust are dominated by
rocks equilibrated at or near the temperature of
relatively dry mafic melts. Above the Mohorovicic
discontinuity, as exposed in an obliquely eroded
mature island arc in Kohistan, northern Pakistan,
and in the continental Ivrea Zone crustal slab in
northwest Italy, are variably layered, complexly
injected, and mostly plagioclase-bearing igneous
and meta-igneous rocks. Beneath the Moho in both
sections are ultramafic and plagioclase-free residual
and magmatic rocks. The Moho is a self-perpetua-
ting density filter for rising melts and is the shallow
limit of crystallization of voluminous ultramafic
rocks and the deep limit of stability of plagioclase in
rocks of most compositions. Melts variously of
mantle, secondary-crustal, and hybrid origins crys-
tallize in the lower crust commonly in middle parts
of the granulite facies - orthopyroxene + plagio-
clase or sanidine, with or without clinopyroxene.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) <edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 3-4. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
3
4 W. B. Hamilton
Deep-crustal ratios of pressure to temperature are
not low enough to permit olivine + plagioclase to
coexist. Lower-crust igneous rocks typically are
retrograded under falling-temperature garnet +
clinopyroxene granulite-facies conditions. Deforma-
tion accompanying such retrogression is more com-
monly driven by rising and spreading magmas than
by tectonic shortening.
Magmatic arcs evolve as their crust is inflated
into geanticlines by intrusive masses and as early or
pre-existing rocks are partly recycled into young
magmas. Many middle-crust igneous rocks (as, two-
mica granite in continental arcs, trondhjemite in
oceanic ones) crystallize from hydrous, low-tempera-
ture melts whose water comes partly from sub-
ducted slabs but probably mostly from breakdown of
wallrock minerals by heat introduced by melts
rising from the lower crust; water saturation forces
crystallization within the middle crust. Magmas
that traverse the middle crust without becoming
hydrated can rise to the upper crust and form
intrusive or volcanic rocks from relatively hot, dry
melts. The common presence of supracrustal rocks
in exposures of the middle crust and the upper part
of the lower crust (where they have been de-
granitized by partial melting) may reflect primarily
depression by emplacement of magma above them
rather than by compressive downbuckling or im-
brication.
Many petrologic discriminants for environment
off ormation of magmatic rocks have been developed
from volcanic rocks of obvious end-member asso-
ciations and are inapplicable to many volcanic
assemblages as well as to middle- and deep-crustal
rocks. Many modern volcanic suites show hybrid
arc and rift petrologic characteristics, and con-
tinental rift rocks can have arc characteristics. Arc
rocks often are strongly bimodal within scales of
time and space such as 1 million years and 5000
square km. Submarine island-arc rocks are vari-
ably enriched in sodium and depleted in calcium by
hydrothermal reaction with water and differ erra-
tically from the subaerial volcanic rocks on which
popular discriminants are based.
Tectonic and magmatic histories of arc com-
plexes are difficult to decipher even in immature
stages when broad tracts of ocean intervene, as in
the modern western Pacific region. Most orogenic
belts now parts of the continents record the end
results of convergent-plate processes; wherein
oceanic plates have disappeared by subduction and
all island-arc and continental-crustal components
are squashed together and variably overprinted by
younger tectonism and magmatism. Problems of
interpretation are formidable.
These and related matters are in part sum-
marized in the three papers listed below, which also
provide extensive references.
REFERENCES
HAMILTON, W. B., 1988, Plate tectonics and island arcs:
Geological Society of America, Bulletin, v. 100, p. 1503-
1527.
__ , 1988, Tectonic setting and variations with depth of
some Cretaceous and Cenozoic structural and magmatic
systems of the western United States; pp. 1-40 in W. G.
Ernst (ed.), Metamorphism and Crustal Evolution of
the Western United States (Rubey Volume 7): Prentice-
Hall Inc., Englewood Cliffs, New Jersey, 1153 p.
__ , 1989, Convergent-plate tectonics viewed from the
Indonesian region; pp. 655-698 in A. M. C. Sengor (ed.),
Tectonic Evolution of the Tethyan Region: Kluwer
Academic Publishers, Dordrecht, The Netherlands, 698 p.
INTERNATION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Ancient collisional continental margins in the Canadian Shield:
Geophysical signatures and derived crustal transects
M.D.THOMAS
Geological Survey of Canada, 1 Observatory Crescent, Ottawa KIA OY3, Ontario, Canada
(received March 1,1989; revision accepted September 7,1989)
ABSTRACT
Geophysical studies have made a fundamental contribution to the recognition of plate tectonics as a key
mechanism in the evolution of the Canadian Shield and simultaneously have promoted the viability of
Precambrian plate tectonics. Paleomagnetic studies, though limited by a lack of resolution and often
controversial, suggest that large-scale displacements of now contiguous terranes had occurred, thereby
stimulating the search for appropriate sutures. Distinctive gravity signatures pointed to the presence of deep
crustal breaks along boundaries between structural provinces, and seismic studies indicated differences in
crustal thickness on either side. The breaks were interpreted as collisional sutures, supporting evidence for
which included juxtaposition of sedimentary facies typical of passive continental margins with mafic igneous
rocks interpreted as oceanic crust, and a marked contrast in ages and structural characteristics across the
proposed sutures. Further support was provided by geologically recent analogues such as the Peruvian Andes,
Zagros suture zone, and India/Tibet collision. Compelling evidence for plate collision comes from the Cape
Smith Belt, where ophiolitic thrust sheets complete with a diagnostic sheeted dike complex and representing
the oldest known example (1998 Ma) of oceanic crust in North America were recently identified. Geophysical
transects across several suture zones provide insights into the large-scale crustal morphology of sutured
margins. Examples include the Grenville Front, Labrador Trough, Cape Smith Belt, and Wopmay Orogen.
INTRODUCTION
Continental margins are now widely recognised
as fundamental elements within the tectonic frame-
work of the Precambrian Canadian Shield. Growing
recognition of these features paralleled gradual
acceptance, during the last two decades, of plate
tectonics as a viable mechanism for Precambrian
crustal development; this was accompanied by a
complementary demise of geosynclinal theory.
Credit for recognising continental margins in Pre-
cambrian terranes is due Dietz (1966). He made
major contributions to the emerging concept of
global plate tectonics with arguments for sea-floor
spreading and formation of geosynclines by collapse
of continental rises (Dietz, 1961,1963). Spreading
forces mantle material (= sima) to shear beneath a
continent, leading to collapse of the continental rise
and consequent formation of an orogen some 30 km
thick. Dietz (1963) considered this to be a mechan-
ism for crustal growth and used the annular geo-
logical zonation of North America to support his
theory that collapsed continental rises could explain
continental accretion and growth of continents
(Dietz, 1966). He cited the Grenville Province as an
early example of such accretion, peripherally em-
placed against an older section of the Canadian
Shield.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 5-25. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
5
6
M.D. Thomas
Dietz (1963) compared mantle shearing (his
model) with an apparently similar phenomenon
(now called subduction) occurring around the Pacific
Ocean. He suggested that the Andean margin of
South America might represent a late orogenic
phase of continental rise collapse where a eugeo-
syncline had already been compressed against the
continent. Ultimately, it was Dewey (1969) who
incorporated continental rise theory, with its
geosynclines, unequivocally into a plate tectonic
scheme using the Appalachian Caledonian Orogen
as an example.
In the early 1970's, with geosynclines only
recently proposed as elements of a plate tectonic
framework (Dewey, 1969; Dewey and Bird,1970;
Dickinson, 1971), it is not surprising that a dicho-
tomy in opinion regarding the significance of
geosynclines prevailed. The contrast is exemplified
by papers presented at two different meetings. In
Ibadan (Nigeria) in 1970, Burke and Dewey (1972)
invoked plate tectonics to explain the development
of several Precambrian terranes in Africa; contin-
ent/continent collisions ranging in age from perhaps
> 3000 Ma to 600 Ma were proposed. Contrasted
with this revolutionary thinking were traditional
viewpoints expressed at a meeting on "Basins and
Geosynclines in the Canadian Shield", where such
major Proterozoic sequences as those in the Coro-
nation and Circum-Superior geosynclines were still
being described as "geosynclines," developed within
a continental domain (Hoffman and others, 1970;
Dimroth and others, 1970, respectively). Gibb and
Walcott (1971), the first to consider plate tectonics
in the Canadian Shield, advocated the concept that
the Circum-Superior geosyncline consisted of rocks
of the plate accretion environment (e.g., ser-
pentinized peridotites, pillow basalts, cherts)
juxtaposed with rocks of the plate consumption
environment (eugeosynclinal sediments). The issue
of Precambrian plate tectonics was keenly debated
for the next decade but eventually gained wide-
spread acceptance. In this paper, I describe the
contribution of geophysics in promoting Precam-
brian plate tectonics, using examples drawn from
the Canadian Shield. Recognition of ancient con-
tinental margins is intimately linked to recognition
of ancient plate boundaries.
In the search for ancient continental margins, a
logical starting place is to examine the geosynclines
as already defined, such as on a map of the main
basins and geosynclines in the Canadian Shield
(Figure 1). Gibb and Walcott (1971) adopted this
approach in evaluating the Circum-Superior geo-
syncline as a collisional suture zone. This geosyn-
cline lies along the boundary between two major
structural provinces of the shield: the Archaean
Superior Province and the Proterozoic Churchill
Province. Juxtaposition of these two contrasting
terranes is a reminder that many continental
margins are transformed into suture zones. Thus
ancient margins may be sought along boundaries
between contrasting structural and age domains. In
some instances supracrustal elements of continental
margins, which provide most diagnostic clues of a
margin status, may be highly metamorphosed or
largely eroded. In such cases identification of the
former existence of a margin relies heavily on
geophysics.
GEOPHYSICAL SIGNATURES OF CONTINENTAL
MARGINS
Paleomagnetic Signature
In 1968, Hess ventured that the Grenville Pro-
vince was a product of continent-continent collision
(in Vine and Hess, 1970). He suggested that
paleomagnetic tests could be used to determine
whether or not terranes to either side of a colli-
sional belt such as the Grenville Province had
moved significant distances before collision. Such
tests were considered difficult but not impossible,
considering both refinement in radiometric dating
and recognition of rock types having extraordinarily
stable remanent magnetization. A paleomagnetic
study of the Grenville Province by Irving and others
(1972) indicated that the province was located thou-
sands of kilometres from the rest of the Canadian
Shield before colliding with it about 1200-1100 Ma.
Using more data, Irving and others (1974) defined a
broad zone along the northern margin of the Gren-
ville Province within which a suture was expected to
lie; the collision date was revised to "" 1000 Ma.
These studies were based on a comparison of
Grenville poles with Keweenawan poles, the ages of
magnetization being interpreted to be approxi-
mately contemporaneous. Subsequent radiometric
studies showed that relevant Grenville poles are
younger than Keweenawan poles, causing some
researchers to favour a one-plate rather than a two-
plate model. Poles used in the two-plate model
advanced by Irving and others (1972,1974) were
dated initially (K-Ar) as 1130-1120 Ma, roughly the
Ancient collisional continental margins in the Canadian Shield 7
Os
0: J""'"
E5 l''"''
_ F_ FAULT
km 500
I ! I I
Figure 1. Map of Canadian Shield showing distribution of principal belts of supracrustal and granitic rocks within and
bordering Proterozoic structural provinces (index of province names in upper right inset). Legend: 1, Archaean cratons;
2, Archaean supracrustal rocks; 3, Aphebian supracrustal rocks (contiguous dot pattern == assumed former distribution of
belt); 4, Aphebian granitic rocks; 5, Helikian supracrustal rocks; B, belt; F.B., foldbelt; Gp, group; lines 1 to 10, gravity profiles
and models, and seismic section (line 2), in Figures 7 to 11 and 13; lines labelled P, gravity profiles and models in Figure 12;
profiles and models for lines 5 and 6 also appear in Figure 12.
same as Keweenawan rocks. However, Ar dating of
Grenville rocks by Berger and others (1979) indi-
cated that they were considerably younger, ",,950
Ma. This was interpreted as evidence for a one-plate
model. In spite of new evidence for the Grenville
Province being an integral part of the Canadian
Shield 1130-950 Ma ago, the work of Irving and
others stimulated a search for other signs of colli-
sion along the Grenville Front leading to recogni-
tion of a characteristic gravity signature (discussed
in more detail in the next section) as a collision
phenomenon.
Burke and others (1976) also used paleo-
magnetic evidence to support collision between
other structural provinces in the Canadian Shield.
They plotted poles for what they considered, on the
basis of independant evidence, to be suture-bounded
blocks. They showed that the Slave and Superior
Provinces probably had a common apparent polar
wander path (APW) after 1600 Ma, but different
paths prior to that date back to "" 2700 Ma (Figure
2a). An implication is that they became united, by
plate collision, in a proto continent that included the
intervening Churchill Province. This result con-
8
M. D. Thomas
a
180
0
b
, ........
i 650\
....... "',.. \LC
...... 1 \.
KEWEENAWAN
2147
>2555
Figure 2. a) Apparent polar wander paths for Superior, Slave and Wyoming provinces suggesting that they had been
amalgamated into a single continent at '" 1650 Ma (after Burke and others, 1976). b) Apparent polar wander paths for Slave
and Superior provinces indicating separate but convergent paths prior to "'2100 Ma (after Irving and others, 1984). These
authors contend that these two provinces had attained their present relative positions by '" 1750 Ma.
tradicted earlier opmlOn (McGlynn and others,
1975) which advocated that the Slave and Superior
Provinces shared a common APW 2200-1800 Ma
ago. Although in that same year McGlynn and
Irving (1975) did suggest that the Superior and
Slave Provinces recorded separate yet close APW
paths during this time interval, and noted that the
similarity was consistent with opening and closing
of an ocean. Irving and others (1984) used addi-
tional data in proposing separate tracks for the
Slave and Superior Provinces prior to "",2100 Ma
(Figure 2b). Amalgamation into a single continent
by 1750 Ma is also suggested within the interval
2100-1750 Ma.
Not all paleomagnetic studies dealt with
sutures at boundaries of structural provinces.
Cavanaugh and Seyfert (1977) and Seyfert and
Cavanaugh (1978) presented arguments for collision
within the western Churchill Province, indicating
that APW paths for the Slave and Superior were
independant before"'" 1750 Ma. They believed that
poles determined from the northwestern Churchill
Province fell on the Slave APW path, whereas those
from the southeastern part lay on a common path
with poles from the Superior, Wyoming, and Nain
Provinces. They defined a narrow 1750 Ma collision
zone extending northeastward across the central
part of the western Churchill Province (Figure 3).
It follows closely, in part, a line of suturing ori-
ginally proposed on the basis of prominent gravity
anomalies (Walcott and Boyd, 1971), and also the
Snowbird tectonic zone, a possible collisional suture
separating the recently defined Rae and Hearne
Provinces (Hoffman, 1988).
This brief examination of paleomagnetic stu-
dies demonstrates the role they played in encour-
aging workers to think of the Canadian Shield in
terms of a mosaic of amalgamated terranes brought
together by plate tectonics. Initial conclusions, not
always correct, invited comment which contributed
to the plate tectonic debate. Not every paleo-
magnetist has been convinced that the plate option
is necessarily the right one, some maintaining a
belief in the concept of a fixed supercontinent
throughout the Proterozoic (Piper, 1982).
Gravity Signature
Tanner (1969) drew attention to a characteristic
gravity signature associated with parts of the
Circum-Superior geosyncline and with the Gren-
ville Front. It takes the form of a paired posi-
tive/negative gravity anomaly with the following
Ancient collisional continental margins in the Canadian Shield 9
Figure 3. Structural provinces in Canadian Shield with Churchill Province subdivided into Rae, Hearne, and Trans-Hudson
Orogen following Hoffman (1988); SS, Snowbird Suture; a, b, and c are positions of sutures proposed by Walcott and Boyd
(1971), Gibb and Halliday (1974), and Cavanaugh and Seyfert (1977), respectively. Ages of last major orogeny affecting
various provinces are identified by pattern, see legend.
characteristics: the gravity field decreases from a
background level over the older of two adjacent
structural provinces towards the boundary zone
where the negative of the pair is located, it then
increases sharply within the border zone of the
younger province and peaks to form the comple-
mentary positive (Figure 4). Modelling of these
signatures indicated that significant changes in
density of the crust occurred at the boundaries
(Tanner, 1969), suggesting that surface differences
in radiometric age and structure were paralleled at
depth by changes in composition. Paleomagnetic
identification of a broad suture zone along the
northwestern margin of the Grenville Province
(Irving and others, 1974) allowed the gravity
anomaly to be evaluated one step further, in terms
of a geodynamic process. Thomas and Tanner (1975)
proposed that the steep gradient between negative
and positive components marked the trace of a
cryptic collisional suture, formed"" 1000 Ma (Irving
and others, 1974), at the time of the Grenvillian
Orogeny. Berger and others (1979) challenged this
hypothesis on the basis of new radiometric evidence
and instead favoured a one-plate model in which the
Grenville Province was linked with Interior Laur-
entia at the time of the proposed 1000 Ma collisional
event. Subsequently, Thomas (1985) argued that
the gravity signature is probably related to a colli-
sion taking place much earlier (1750-1450 Ma). A
major deformational, metamorphic/plutonic event,
1680-1640 Ma (Nunn and others, 1988), produced
the Labradorian Orogen along the northwestern
margin of the Grenville Province northeast of the
Labrador Trough (Wardle and others, 1986). Pos-
sibly this orogen is a product of the collision event
implied by the gravity signature.
Initial recognition of paired gravity anomalies
as signatures of collision was fortuitous, in light of
the reappraisal of Grenville poles in terms of a one-
plate model (Berger and others, 1979). But know-
ledge subsequently accumulated on Precambrian
structural boundaries validated the collision sig-
nificance, even though originally identified on the
basis of spurious evidence. The distinctive gravity
10 M.D. Thomas
signature is present along a number of structural
boundaries in the Canadian Shield. In addition to
the Grenville Front example, paired signatures are
found: (1) along the Labrador Trough, Cape Smith
Belt, and eastern Hudson Bay segments of the
Circum-Superior geosyncline; and (2) along the
Thelon Front which marks the boundary between
the Slave and Churchill Provinces. The various
gravity and tectonophysical models derived for
these boundaries are summarised by Gibb and
others (1983). The paired gravity anomaly sig-
nature is found worldwide in Precambrian terranes
and also is associated with Phanerozoic collisional
orogens, a good example of the latter is that along
the Appalachian Orogen (Figure 4) (Thomas, 1983).
Polarity of the anomaly pair with respect to
relative ages of sutured terranes is consistent. The
negative component always straddles the terrane
120
mGal
o
... .............
.... .
.......
.... .... .
boundary or lies largely within the older terrane,
whereas the positive invariably falls within the
younger terrane (Figure 4). A number of large-
scale geological relationships maintain a parallel,
uniform polarity: (1) miogeosynclinal deposits pass
into eugeosynclinal counterparts in passing from
the older terrane towards the younger; (2) meta-
morphic grade increases in the same direction and
very often culminates in granulite-facies meta-
morphism in the younger terrane; (3) structural
vergence is towards the older terrane. A simple type
model derived from the signature comprises two
crustal blocks of contrasting density and thickness
in juxtaposition along a boundary interpreted as a
suture. The model is constrained by requiring the
two blocks to be in relative isostatic equilibrium
according to Airy-Woollard isostasy (Wilcox, 1976),
which permits variable crustal density as well as
...... STRUCTURAL VERGENCE
MIOGEOSYNCLINE ...... EUGEOSYNCLINE
GREENSCHI ST GRANULITE
Figure 4. Four gravity profiles across boundaries of various Precambrian structural provinces or terranes and one profile
across Appalachian Orogen. Profile 1 crosses eastern margin of West African Craton (== 1700 Ma) into Pan African Orogen
( == 600 Ma). Profile 2 is mean curve for 5 profiles crossing boundaries in Canadian Shield that juxtapose variously == 2560 Ma
and == 1800 Ma, 2560 Ma and == 1000 Ma, and == 1800 Ma and == 1000 Ma terranes (after Gibb and Thomas, 1976). Profile 3 is
mean curve for 9 profiles crossing Appalachian Orogen (Thomas, 1983) and may reflect ==450 Ma collision. Profile 4 crosses
from Archaean Karnataka Craton into Proterozoic Eastern Ghats terrane (== 1600-1300 Ma) of Indian Shield. Profile 5, in
Australia, crosses from> 2300 Ma Yilgarn Block into the Fraser Orogenic Domain affected by a series of Proterozoic orogenic
events commencing with high-grade metamorphism 1950-1800 Ma (Plumb, 1979). A type model based on mean Canadian
Shield profile is also illustrated. A supracrustal foldbelt is schematically incorporated and major geological features of suture
zones are indicated.
Ancient collisional continental margins in the Canadian Shield 11
variable thickness; in the model the younger block
is denser and thicker.
Development of such a model may be reconciled
with continental collision. Consumption of oceanic
crust beneath the overriding continent (the active
margin) may be followed by limited subduction of
continental crust of the underriding continent as
convergent forces continue to operate. Thickened
crust in the area of overlap becomes relatively more
buoyant, leading ultimately to enhanced isostatic
recovery, which in turn may carry denser lower
crust of the overriding plate closer to the surface
where it gives rise to the positive gravity anomaly.
The complementary low, associated mainly with the
passive margin, probably reflects a residual crustal
root, initially depressed by the overriding active
margin, which now is isostatically preserved by
overthrusted denser material, e.g., oceanic crust of
the Cape Smith Belt.
Realisation that the paired gravity anomaly
signature apparently provided a new tool, a kind of
remote sensor, for identifying ancient suture zones
(Thomas and Tanner, 1975) led to studies of paired
gravity anomalies at other structural province
boundaries in the Canadian Shield. Kearey (1976)
interpreted paired anomalies along the Labrador
Trough in terms of continent/continent collision,
thereby strengthening the earlier arguments of
Gibb and Walcott (1971) and demonstrating the
merit of plate tectonic models. This opinion
contrasted with that of Dimroth (1972) who had
rejected application of plate mechanisms to this
structure.
Thomas and Gibb (1977) carried the argument
of paired gravity signatures as indicators of collision
to another part of the Circum-Superior foldbelt, the
Cape Smith foldbelt. Baer (1977) refuted this
proposal, claiming that available data contributed
little to the more fundamental question of Pre-
cambrian plate-tectonics. A later evaluation of the
Cape Smith foldbelt by Hynes and Francis (1982)
raised doubt that the foldbelt marked a major suture
zone.
While apparently subscribing to Precambrian
plate-tectonics, these authors felt that the foldbelt
more likely represented a continental rift zone, pos-
sibly in which true oceanic crust was generated, in
contrast to Thomas and Gibb's (1977) contention
that it represented obducted oceanic crust. The
identification of an ophiolite sequence by St-Onge
and others (1988), dated at 1998 Ma (Parrish, 1989),
coupled with interpretation of the foldbelt as a
klippe (Hoffman, 1985) provides strong support for
the obduction model. Mukhopadhyay and Gibb
(1981) had previously extended the interpretation of
paired gravity anomalies to the section of the
Circum-Superior foldbelt in eastern Hudson Bay.
Perhaps the best example of the utility of paired
gravity anomalies in identifying a Precambrian
suture zone involves the Thelon Front separating
the Archaean Slave Province from the Churchill
Province. This boundary resembles the Grenville
Front in that extensive supracrustal deposits, which
may provide clues regarding possible continental
margin development, are largely absent. Those that
are present, in the Kilohigok Basin and Athapuscow
Aulacogen, were at one time not closely linked with
the Thelon Front. Gibb and Thomas (1977) inter-
preted this boundary as a collisional suture based on
the paired anomaly signature, suggesting that
collision took place 1900-1700 Ma on the evidence of
available K-Ar ages in the western Churchill Pro-
vince. Hoffman (1988), using U-Pb zircon dating,
suggested collision at = 1960 Ma. The collision
model was supported by an indentation collision
analogue drawn between the Slave/Churchill
example and that ofIndialTibet (Gibb, 1978).
Since papers proposing collision were published
in the late 1970's, various geological evidence
consistent with the hypothesis has been obtained.
The Kilohigok Basin is identified as a foreland basin
containing a sequence of shallow shelf sediments
(passive margin) which are overlain successively by
deep-water flysch and shallow marine and fluvial
molasse representing a foredeep environment
(Grotzinger and Gall, 1986); the sequence thickens
eastward towards the Thelon Front. A flexural
bulge (Gordon Bay Arch) parallel to the front is
recognized in the sedimentary record. The arch and
the basin are ascribed to crustal loading produced by
convergent tectonics along the Thelon Front.
Loading probably took the form of stacked thrusts;
recognition of a thin-skinned, thrust and fold belt
characterized by NW-verging folds and a basal
decollement in the eastern part of the Kilohigok
Basin (Tirrul, 1985) is consistent with this inter-
pretation. Hoffman (1987), following Gibb (1978),
has set Slave/Churchill collision in the context of an
India/Tibet analogue and explains several large-
scale features in the region, such as the McDonald
fault, the Thelon magmatic arc along the margin of
the Churchill Province, the Athapuscow Aulaco-
gen, and the Queen Maud uplift east of the Thelon
magmatic arc.
12 M. D. Thomas
Seismic Signature (Imaging of Crustal
Thickness)
The picture of relative crustal densities and
thicknesses across structural boundaries in the
Canadian Shield established from gravity modelling
is substantiated by seismic studies. Refraction
results near the Grenville Front indicate that the
Grenville Province is thicker (39 km) than the
Superior Province (35 km) (Berry and Fuchs, 1973).
Seismic wave velocities in the upper Grenville crust
are = 0.15 kmls faster than in the Superior Province
suggesting that crust in the Grenville Province is
denser. Two seismic profiles crossing the Superior/
Churchill boundary at the Nelson Front also yield a
picture of thicker, denser crust within the younger
(Churchill) province: along one, thicknesses in-
crease from 30-35 km (Superior Province) to 40-50
km (Churchill Province) (Mereu, 1968); along the
other, they increase from 30 to 34 km across the
front (Hall and Hajnal, 1973). Hall (1977) proposed
that the gravity signature along the latter profile
was akin to the typical paired gravity-anomaly sig-
nature proposed by Gibb and Thomas (1976). Green
and others (1980) investigated the Superior/Chur-
chill boundary where it lay hidden by Phanerozoic
deposits southwest of the Nelson Front and deter-
mined that here, too, the Churchill crust was
thicker (46 km) than that of the Superior Province
(41 km); a denser layer at the base of the Churchill
Province was indicated by higher seismic velocities.
Deep seismic sounding in southern India (Kaila
and others, 1979) indicated that crust forming the
Proterozoic Eastern Ghats is thicker (45 km) than
that of the adjacent Archaean Karnataka craton
(35-40 km), and is characterized by higher upper
crustal seismic velocities. The type paired gravity
anomaly (Figure 4) is identified with this example
(Subrahmanyam,1978). Australian seismic experi-
ments indicate that the margin of the Archaean
Pilbara craton is thinner (30-33 km) than the adja-
cent margin ofthe Proterozoic Capricorn Orogen (40
km); however, the Archaean Yilgarn craton which
also abuts the Capricorn Orogen is thicker (50-55
km) (Drummond, 1979, 1981; Drummond and
Shelley,1981). Here then is an example of the older
craton being thicker. Apparently, the Yilgarn cra-
ton is also thicker (34 km) than the bordering
Proterozoic Fraser Range (32 km) to the east
(Mathur, 1974), although Mathur (1976) cautions
that the seismic data are "inadequate and incon-
clusive in this area." Paired gravity anomalies are
associated with both of these Australian examples;
the Yilgarn/Fraser example is shown in Figure 4.
Although a pattern of thicker and denser crust
in younger orogens where paired gravity anomalies
are present appears well established, the Yilgarn
example dictates that universal application of the
model must be tempered with caution. Similar
reservations are outlined by Johnson and others
(1984) in their study of gravity anomalies along the
southern margin of the Archaean Wyoming craton
in the southwestern U.S.A. There, gravity values
are lower over the thicker juxtaposed Proterozoic
terrane (48-54 km; seismically determined) to the
south than over the Wyoming Province (37-41 km).
Johnson and others (1984) suggested that the
difference in gravity signature may result from two
differing kinds of collision; signatures such as that
along the margin of the Wyoming craton may
signify continent/island-arc collision, whereas those
described by Gibb and Thomas (1976) may reflect
continent/continent collision.
Magnetic Signature
Magnetic signatures at structural boundaries
in the Canadian Shield have received comparatively
little attention, although Hoffman (1987,1988)
noted that such signatures exist and used them to
establish tectonic models involving ancient contin-
ental margins. He observed a correlation between
belts of relatively positive magnetic anomaly and
calcalkaline magmatic arcs comprising dominantly
magnetite-series plutons. Two belts are located
along margins of younger orogens adjacent to the
Slave Province (Figure 5a). To the east is the
Thelon magmatic arc which probably is related to
eastward subduction prior to Slave/Churchill
collision. While to the west the 1875-1840 Ma Great
Bear magmatic arc in the Bear Province may have
evolved above another east-dipping subduction
system that plunged beneath the western (active)
margin of the Slave Province (Hildebrand and
others, 1987).
Ancient passive margins often preserve supra-
crustal components that provide valuable insights
into structure and sedimentary environments and
related aspects of margin development and sub-
sequent orogenesis. Active margins apparently are
often preserved as high-grade metamorphic ter-
ranes where supracrustal components are more
difficult to identify and decipher. The positive mag-
Ancient collisional continental margins in the Canadian Shield 13
netic signature associated with such margins
affords a means of identifying them and, through
modelling, obtaining an idea of their deep structure.
Hoffman (1987, 1988) used magnetic signatures to
propose the existence of the Proterozoic Fort Simp-
son magmatic arc beneath Phanerozoic sedimentary
cover west and southwest of the Slave and Bear
Provinces. Parallelism of this arc with an apparent
Figure 5. a) Grey-tone aeromagnetic map of Slave Province and adjacent regions of Bear and Churchill Provinces illustrating
paired magnetic signatures at boundaries. Areas of negative anomaly (lows) and positive anomaly (highs) indicated by Land
H, respectively. CG, Coronation Gulf; GBL, Great Bear Lake. b) Grey-tone aeromagnetic map of region spanning the
Grenville Front southwest of the Labrador Trough; Land H as for (a).
14
M.D. Thomas
positive/negative paired gravity anomaly (the nega-
tive component is only locally well developed) to the
east, prompted Hoffman to suggest that the arc
developed above a west-dipping subduction zone.
This example demonstrates particularly well the
utility of geophysics in deciphering ancient ter-
ranes. Ross and others (1988) used aeromagnetic
anomalies to trace magmatic arcs in their tectonic
study of Precambrian basement beneath the In-
terior Plains of Alberta.
The belt of positive aeromagnetic anomalies
associated with the Thelon magmatic arc is flanked
to the west by a belt of negative anomalies that
straddles the Thelon Front. Unlike the positive belt
which can be explained in terms of a single geo-
logical feature (magmatic arc), the negative belt
coincides with a region of variable geology.
Archaean granitoid and greenstone rocks occupy the
eastern margin of the Slave Province and foliated
granitoid rocks comprise the adjacent margin of the
Churchill Province (Henderson and others, 1987).
In explaining this anomaly it should be noted that a
single geological body produces a magnetic anomaly
having both a negative and a positive component-
a consequence of the dipolar nature of magne-
tization. Thus, in the case of the Thelon Front
where the negative magnetic anomaly cannot be
readily correlated with a specific geological unit, it
may be interpreted as complementary to the
positive anomaly over the Thelon magmatic arc and
hence may also be attributed to the arc. A negative
magnetic anomaly, forming a pair with the Great
Bear Batholith magnetic high, is located also along
the western margin of the Slave Province. It falls
mainly over the Bear Province where it correlates
essentially with Proterozoic metasediments and
meta volcanics formed in an initial rift/continental
slope environment. This may also be a negative
component of a dipolar signal, in this case generated
by the Great Bear magmatic arc.
A paired magnetic anomaly signature is ob-
served also along part of the Grenville Front
(Figure 5b). The polarity follows the sense of the
paired gravity anomaly along the front and indeed
the respective negatives are more or less coincident.
The positive component is located within the
interior of the Grenville Province, :s 100 km from
the Grenville Front. It is much less pronounced
than highs bordering the Slave Province. The asso-
ciated terrane is formed of high-grade metamorphic
gneisses; as yet, however, there is no evidence of it
representing a magmatic arc. The low, a more dis-
tinct feature, corresponds with the Grenville Front
Tectonic Zone (GFTZ) where metamorphic assem-
blages indicate some of the highest pressures
recorded in the Grenville Province and suggest that
former deep levels were structurally elevated by
NW-overthrusting (Wynne-Edwards, 1972). Al-
though within the Grenville Province, the GFTZ
contains a significant proportion of Archaean rocks
that doubtlessly represent an extension of Superior-
type crust into the younger province. The GFTZ is
thus characterized by uplifted Archaean rocks
which correlate with the negative magnetic ano-
maly belt. This situation apparently differs from
the one along the eastern margin of the Slave
Province where Archaean rocks correlating with the
negative anomaly belt are continuous with those
located well within the Slave Province (Henderson
and others, 1987), suggesting that uplift has not
been extensive.
The positive component of paired magnetic
anomalies associated with collision zones can often
be explained in terms of a magmatic arc developed
above a subduction zone. The origin of the negative
component is less clear, the few described examples
seemingly correlating with different geology in
each case. If the source is not lithological, an
alternative explanation is that the low is a dipolar
effect linked to a single structure producing a com-
plementary high. Regional magnetic signatures
clearly deserve more rigorous examination and
analysis.
Electrical Conductivity Signature
Electrical conductivity or electromagnetic sur-
veys have not been carried out systematically across
sutured margins in the Canadian Shield, so that
knowledge concerning the electromagnetic signa-
ture of such sutures is essentially lacking. The
North American Central Plains conductivity ano-
maly (NACPCA) (Alabi and others, 1975) (Figure
6), however, is one conductivity anomaly that fol-
lows closely a gravity-defined suture (Thomas and
others, 1987) between the Wyoming Province and
the Trans-Hudson Orogen. This geographical asso-
ciation supports the suggestion of Camfield and
Gough (1977) that the NACPCA is associated with a
Proterozoic plate boundary. The NACPCA is, how-
ever, generally poorly defined by widely spaced
magnetometer stations. Detailed magneto-telluric
surveys (Jones and Savage, 1986) indicate that it
Ancient collisional continental margins in the Canadian Shield 15
may be located up to "" 150 km east of the boundary
defined by gravity and thus its significance as a
suture-related phenomenon is questionable. Fur-
thermore, it may not be as continuous as tradition-
ally portrayed (Figure 6). Detailed studies may
reveal it to be a series of anomalies related to
discontinuous, en echelon conductors not individual-
ly resolved by coarse magnetometer spacing
(Thomas and others, 1987). The exact form and
significance of this intriguing geophysical anomaly
remain to be evaluated.
GEOPHYSICAL TRANSECTS ACROSS
COLLISIONAL SUTURED MARGINS
Following is a presentation of several transects
across suture zones in the Canadian Shield (Fig-
ures 7-13) selected to display similarities in some
cases and differences in others. Descriptions accom-
panying the figures are brief, details being available
in the quoted references. Many of the deep crustal
Figure 6. Relationship of the North
American Central Plains conductivity ano-
maly (NACPCA) to the western boundary of
the Trans-Hudson Orogen as defined by pat-
terns of gravity trends (Thomas and others,
1987), shown as fine lines: J, position of
NACPCA determined by Jones and Savage
(1986); B, Black Hills; MGH, mid-continent
gravity high; WCB, Wyoming-Churchill
boundary.
sections across suture zones have been derived by
modelling gravity anomalies using constraints
provided by seismic sounding where available (Gibb
and others, 1983). In most cases, the models consist
simply of two juxtaposed crustal blocks separated by
a boundary interpreted as a suture. Models inter-
preted for various parts of the Circum-Superior
suture include additional bodies representing supra-
crustal deposits. The lack of detail in the models is
related to two main factors: (1) many models were
derived when existing geological mapping was of a
reconnaissance nature; and (2) available seismic
data were obtained by the refraction method which
provided only a gross picture of crustal structure.
Many of the regions investigated now have been
geologically mapped in some detail and subjected to
various specialized studies. For example, structural
studies near the Grenville Front have led to recog-
nition of kinematic indicators (Davidson, 1986), and
sedimentological studies in the Kilohigok Basin
have revealed crustal flexuring attributed to con-
vergent tectonics along the Thelon Front (Grot-
16
M. D. Thomas
a
b
C
-100
GRENVILLE Profile 1
GF
I
COlculoted -of
o 40 so 120 160 200 240 280 320
o Km J. j J I I I , l J .. .I I I. 1 L.a ..t. 1 W_J----'-'-........ ...J....1.,.-
GRE VILLE
SUPERIOR
20
006
40

_Oq7---
0 100
0
"
SUPERIOR
20
284
40
324
ZAGROS
o SW
NE
.......
d mGoj -- - '--- ' _ COll ected " ...
BouQuer _._:'::::--: ..
-200 80uguer - - --- - .-.
IMBRICATED 6
o Km 100 TRENCH ZONES ZAGROS THRUST

EURASIAN
e Km
50
Figure 7. a) Profile 1: observed and calculated gravity
profiles across Grenville Front (GF) for model in (b) (see
Figure 1 for location; this applies to all subsequent profiles
shown in Figures 8-13). b) Gravity model: lateral density
contrasts are in units of g/cm
3
; horizontal bars mark
position of Moho determined by seismic refraction (Berry
and Fuchs, 1973); solid bars, true hoizontal position;
hollow bar, position extrapolated from just outside limit of
model. c) Model of (b) above with addition (schematic) of
sedimentary basin north of Grenville Front (foreland zone)
and deep crustal faults in the Grenville Front Tectonic
Zone (GFTZ), based on analogy with Zagros collisional
suture zone (e). Mean crustal and mantle densities are in
units of g/cm
3
. d) Observed Bouguer gravity profile across
Zagros suture zone. Combined negative gravity effect of
Persian Gulf and Central Iranian Plateau is also shown, as
is a Bouguer profile corrected for this effect. Note the
similarity of the latter with (a). e) Gravity model across
Zagros suture zone (after Bird, 1978). Mean crustal and
mantle densities are in units of g/cm
3
.
zinger and Gall, 1986). Another factor bearing on
improved understanding of suture zones is applica-
tion of high-resolution reflection seismic profiling.
Such an experiment conducted in Lake Huron
across an extension of the Grenville Front produced
spectacular results (Green and others, 1988,1989),
with many reflectors at various crustal levels being
imaged. With these diverse advantages available, it
should be possible to upgrade many previous inter-
pretations.
Grenville Front
In evaluating the tectonic significance of a
gravity model across the Grenville Front, Thomas
(1985) drew an analogy with the Zagros collision
zone and thus incorporated steep thrusts sche-
matically in the Grenville Front Tectonic Zone
(GFTZ) (Figure 7). This picture is substantiated by
reflection seismic profiling carried out in Lake
Huron (Green and others, 1988,1989). The crustal
image so obtained shows steep east-dipping re-
flectors in a zone lying on strike with the GFTZ
(Figure 8). This zone extends deep into the crust to
a depth of "'" 25 km. The most westerly reflector
coincides with a mylonite zone at the Grenville
Front. The reflectors probably represent highly
strained contacts between gneissic and migmatitic
rocks. The seismic results indicate that the lower
crust north of the Grenville Front contains many
shallow-dipping reflectors. This lower crust is
viewed as stretched crust of the Sl1perior passive
margin, upon which Huronian rocks were deposited
in a continental shelf and rise environment. The
thickness of the lower crust is similar to crustal
thicknesses observed at modern continental mar-
gins. Strong reflectors at the top of this lower crust
are equated with a decollement produced by
Penokean (1890-1820 Ma) overthrusting of the
Manitoulin Terrane from the south.
Labrador Trough
Gibb and Walcott (1971) proposed that iden-
tification of rocks of the Circum-Superior geo-
syncline as a suture "relies on the juxtaposition of
serpentinites and pillow basalts of the plate accre-
tion environment and eugeosynclinal sediments of
the plate consumption environment." They did not
specifically describe the miogeosynclinal/eugeosyn-
Ancient collisional continental margins in the Canadian Shield 17
w
MA I TOUllN
TEAR' E
GF
I GRENViLLE FAO T I
I TECTON' C ZONE I
E
BR' TT TERRA E
280
1
0 2400 : 2000 '600
____ ______ ____ -L __ ____ ____ ______ ____ -L ____ -. O
3600 3200 1200 400 CRP' S 800
E '_
2 C : ' : :_ "
4 -1;1 ' "-.. , ;.;.:s '''-:::-.-.. ' <. G
:-. .r-
; ' - " '-> ' : A, i( (,:l", - '- ,. -k'
G
8
'0
' M '2
'2 M
,4
M
14
M
"
'6 ?

MAN ITOULIN
E
-"
LINE J (F -K Migra ted)
BRIT T
DOMAIN

'-
20
'- --- __ -,.- .-M- 40
?---
? / '
' '-.. - ?./
o 10 20 30km
I I I 1
Figure 8, Profile 2: line drawing derived from seismic reflection profile crossing extension of Grenville Front in Lake Huron
and interpreted crustal section (after Green and others, 1989), Vertical scale in upper section is in units of two-way travel
time (seconds), CRP, common reflection point; M, Moho; other letters in upper section are keyed to text of Green and others
(1989),
clinal couple (Dimroth and others, 1970) in terms of
Dietz' (1963) continental rise model. Kearey's
(1976) appraisal of the Labrador Trough as a colli-
sional suture, likewise, did not recognise its deve-
lopment on a passive continental margin; instead, it
was portrayed as developing on the leading edge of
an active plate margin (Churchill Province) above
an east-dipping subduction zone, Ultimately, fol-
lowing consumption of all oceanic crust, the Labra-
dor Trough became juxtaposed with the Superior
Province as the latter collided with the Churchill
Province.
Recent geological studies indicate that the
Circum-Superior geosyncline fits into a continental
rise scenario (Dietz, 1963). For example, St-Onge
and others (1988) described the Cape Smith foldbelt
in terms of the evolution of an epicontinental rift
that eventually developed oceanic crust; continental
rift, transitional crust and ophiolitic suites are
documented, Thus, the miogeosyncline of Dimroth
and others (1970) represents a continental platform
environment, and the eugeosyncline, which is dom-
inated by mafic volcanic rocks and sills, is charac-
teristic of an oceanic environment.
Dimroth (1970) presented a schematic compo-
site deep structural section across the Labrador
Trough based on sections from various localities. In
this, the eastern eugeosynclinal component attains
a maximum preserved thickness of "" 22 km, of
which mafic igneous rocks constitute "" 18 km. A
gravity model by Kearey (1977) suggests that this is
a considerable overestimate - the maximum thick-
ness of igneous rocks more likely being "" 9 km
(Figure 9), Such geophysical information may be
critical to understanding the internal structural
assemblage of the igneous package. Thomas and
Kearey (1980) derived a structural transect across
the ancient active continental margin bordering the
Labrador Trough (Churchill Province) using gravity
modelling and an analogy with the Peruvian Andes
18
M. D. Thomas
mGoI Prohl. 3
100 0 30
, !
t.ftGI LITE
o
sw
50 ,----,
I ,
I "
" J((AR'EY" ....
I "
I ,
, , \
',-/ \ ".,.,
LABRADOR TROUG H ,_I
Figure 9. Structural section across Labrador Trough
(after Dimroth, 1970) and its computed gravity anomaly
(after Kearey, 1977). Kearey's model of the geometry of
the mass of mafic igneous rocks interpreted from an
observed gravity profile (Profile 3) is shown for com-
parison. Density contrasts for arenites and mafic igneous
rocks relative to the background density of Superior and
Churchill gneisses are given in units of g/cm
3
. Argillitic
rocks are similar in density to the gneisses thus precluding
modelling of these rocks.
sw
a Prohle 4
(Figure 10). Andean evolution has been strongly
influenced by block-faulting that may be related to
intense coupling between the Nazca and South
American plates. The faulting probably controlled
basin development and batholith emplacement. The
Peru/Chile trench is equated with the eastern
(eugeosynclinaD margin of the trough, the Andean
Coastal Batholith has an equivalent in a calc-
alkaline granite belt now referred to as the De Pas
Batholith (Hoffman, 1988), the East Peruvian
Trough has an analogue in the Dorset (sedimentary)
foldbelt, and east-verging compressional tectonics
along the ChurchilllNain boundary are compared
with compressive tectonics in the Sub-Andean
ranges that are thought to mark the distal limit of
influence of the shallow-subducting Nazca plate.
This analogue provides a broad framework for un-
derstanding the structure and tectonics of ancient
active continental margins.
NE

- 20 "BO.Rs" Of. s
CompreUIYI' T OU ti PAL,:'[O- ORA"' ! E

1tC;IOl'hC;S
b
Yo-EST TROUGH

COAST.l
8AlHO I H
Figure 10. a) Gravity profile (Profile 4) and derived crustal model across eastern Churchill Province (after Thomas and Kea-
rey, 1980) . Legend: 1, metasediments; 2, mafic igneous rocks; 3, granite; 4, mixed gneisses and migmatites; 5, granulite.
Lateral density contrasts (in units of g/cm3) are relative to mean density of the Superior Province above dashed line and to
mean density of mantle below the line. b) Schematic crustal section across the central Andes'" 100 Ma (after Myers, 1975).
Legend: 1, basic intrusions; 2, the rising Coastal Batholith.
Ancient collisional continental margins in the Canadian Shield 19
Cape Smith Foldbelt
Thomas and Gibb (1977) proposed that the Cape
Smith foldbelt marked a collisional suture (Figure
11). This hypothesis is strongly supported by recent
geological studies, particularly the discovery of an
ophiolite complex (St-Onge and others, 1988). In
their gravity model, Thomas and Gibb placed the
Superior/Churchill suture at the northern margin of
the foldbelt along the contact between mafic igneous
fill of the belt and Churchill gneisses to the north;
mafic rocks extend some distance northward be-
neath the gneisses. This characteristic of mafic
igneous rocks of the Circum-Superior geosyncline
extending below adjacent gneisses in the direction of
the Churchill Province is also observed in the
Labrador Trough (Kearey, 1976) and the eastern arc
of Hudson Bay (Figure 12) (Mukhopadhyay and
Gibb, 1981). Hoffman (1985) proposed a slightly
different model for the Cape Smith foldbelt, viewing
it as a klippe having a basal decollement rooted in a
suture 30-90 km north of the belt.
The distribution of rocks of the Circum-
Superior geosyncline in eastern Hudson Bay is in
marked contrast to those in other parts of the
geosyncline, where they are confined to a relatively
narrow belt; in the Labrador Trough and in the
Cape Smith foldbelt they attain a maximum width
of ""'100 and ""'80 km, respectively. Maximum
thicknesses of the fill are similar at "'" 12 km
(Mukhopadhyay and Gibb, 1981). Chandler and
Schwarz (1980) suggested that the enhanced
preservation, :5300 km wide, in the eastern arc of
Hudson Bay may be related to a re-entrant in a
continental margin conforming to two successful
arms of a triple junction. The failed arm is now
represented by the Richmond Gulf graben that
strikes into the Superior Province at a high angle to
the eastern arc of Hudson Bay. Chandler and
Schwarz proposed that convergent tectonics thick-
ened the miogeoclinal fill by compressing the
foldbelt into the re-entrant and forming an east-
facing, arcuate salient. Alternatively, Beals (1968)
speculated that the eastern arc of Hudson Bay
might reflect the signature of a meteoritic impact.
The East Arm of Great Slave Lake, which contains
an Early Proterozoic sequence (1930-1850 Ma;
Hoffman, 1987), also has a remarkably arcuate
northern margin. The positioning of large sedimen-
tary basins with arcuate boundaries at the margins
of Archaean cratons raises the question of meteori-
M I Prohl.6
:00
80
60
40
20
o
OA A
GAP ___
............................... ..
SE
-I
Km NW FOLD8ELT SE
o i , > . . ' !
20 :- 0 OJ : -006 :'009
: i --'--: 0 3 SUPERIOR
: M- A- N-T-lE:---
2
Figure 11. Crustal models interpreted along profiles
crossing Cape Smith foldbelt (Profile 6) and its hidden
extension in Hudson Bay (Profile 5) (after Thomas and
Gibb, 1977). Lateral density contrasts relative to mean
density of Superior Province (above dashed line) and to
mean mantle density (below) in g/cm
3
.
tic impact as a contributing mechanism in the
breakup of proto continents.
Wop may Orogen
The W opmay Orogen of the Bear Province may
have evolved on an Andean- or Cordilleran-type
margin during the interval 2100-1800 Ma in
response to east-directed subduction (Hoffman,
1980). From west to east, the orogen comprises the
following:
The Hottah Terrane is made up of metamor-
phosed sedimentary and volcanic assemblages cut
by 1914-1902 Ma calcalkaline plutons, interpreted
as a magmatic arc originally located on the western
edge of the Slave Province (Hildebrand and others,
1987).
20 M. D. Thomas
The calcalkaline Great Bear Magmatic Zone is
a largely unmetamorphosed volcanic-plutonic as-
semblage intruded by 1875-1843 Ma plutons; the
Hottah Terrane may form much of the basement
beneath the Great Bear Zone.
The Calderian wedge is a west-facing contin-
ental margin prism consisting of bimodal submarine
volcanics overlain by siliciclastic and carbonate
rocks that apparently developed within an exten-
sional interarc basin formed east of the Hottah arc
following cessation of arc magmatism (Hildebrand
and others, 1987). The sedimentary sequence
drapes over the western margin of the Slave Pro-
vince. The basin was deformed by thin-skinned
eastward thrusting onto the western margin of the
Slave craton, and it was simultaneously intruded by
peraluminous to metaluminous plutons (1896-1878
Ma) of the Hepburn intrusive suite to form the
Calderian accretionary wedge.
Figure 12. Fence diagram of crustal
sections interpreted from gravity
profiles across buried section of
Circum-Superior foldbelt in eastern
Hudson Bay (after Mukhopadhyay
and Gibb, 1981). Profiles 1 and 2
correspond closely to Profiles 5 and 6
of Figure 11; profiles 3 to 5 cor-
respond to lines labelled P in Figure
1. Geology legend: 1, paragneiss;
2, granodioritic gneiss; 3, granulite
(all within Churchill Province north
of foldbelt); 4, amphibolite and horn-
blende gneiss; 5, diorite and grano-
diorite; 6, basalt, andesite and vol-
canic breccia; 7, sedimentary rocks
(all within foldbelt); Pz, Paleozoic.
Profiles legend: 1, Churchill crust;
2, Superior crust; 3, volcanic and
sedimentary rocks of foldbelt;
4, granite.
The Wopmay Orogen may differ, therefore, in
its mode of origin from that of the Circum-Superior
foldbelt. In the latter, a collisional, sometimes cryp-
tic, suture can be identified on the basis of a gravity
signature; this suture sometimes coincides with the
oceanward limit of mafic igneous rocks of oceanic
affinity. Cryptic collisional sutures have also been
proposed for the Grenville and Thelon fronts. A
collisional suture is not identified in the Wopmay
Orogen. This genetic difference seems to be paral-
leled by a difference in gravity signature - the
typical paired positive/negative anomaly of other
structural boundary zones being absent. A strong
positive anomaly does lie along the western margin
of the Slave Province in the expected position of the
positive of paired anomalies, i.e., within the younger
province; it correlates mainly with the Hepburn
metamorphic/plutonic belt (Figure 13). The com-
plementary negative is not as readily identified,
km 100
1 I
Ancient collisional continental margins in the Canadian Shield 21
mG.,
. " WEST
20
.m
SITI'tOK
GRAHnE
EAST
Profile 10
'.
WOOTZEl.
FAULT
I
Figure 13. Crustal model interpreted along Profile 10 crossing Wop may Orogen. Density contrasts in units of g/cm3, relative
to a mean density of 2.64 g/cm
3
estimated for granitoid basement of Slave Province. Dashed line labelled "geological estimate"
provided by the late Rein Tirrul.
however, although areas of negative anomaly do
exist. The interpretation of the positive anomaly in
the case of W opmay Orogen is radically different.
Rather than being associated with crust of a former
active margin, the gravity high in the Wopmay
Orogen is ascribed to a westward-thickening meta-
sedimentary wedge of the Slave passive margin
(Thomas, 1984); initial rift metasediments and
metavolcanics form the western part of the wedge.
These supracrustal rocks are more dense than base-
ment rocks of the Slave Province. Gravity model-
ling indicates that the wedge attains thicknesses of
= 7.5 km and = 11 km, respectively under and west
of the Hepburn Belt (Figure 13). Kearey (1977) did
not model argillaceous and calcareous sections of
the metasedimentary sequence in the Labrador
Trough because their mean densities were very
similar to that of basement gneisses. The Great
Bear Batholith, although less dense than the Slave
basement, is associated with a gravity field that is
significantly more positive (= 16 mgal). This
suggests that basement to the batholith, the Hottah
Terrane (Hildebrand and others, 1987), is more
dense than the Slave basement.
CONCLUSIONS
Geophysics has made a major contribution to
the realisation that plate tectonics was an instru-
mental factor in development of the Canadian
Shield. The acceptance of Precambrian plate-
tectonics was accompanied by recognition of various
tectonic elements within the shield that together
comprise the plate framework; among these are
continental margins. Apart from its role in recog-
nising these margins, geophysics has provided
insights into the deep structural aspects, which in
turn have ramifications for tectonic and geodynamic
processes. As more detailed geological maps become
available and relatively new geophysical applica-
tions, such as high-resolution seismic reflection
profiling, are more widely utilised, an increasingly
finer picture of crustal structure should emerge.
Thus, new frameworks and constraints will be made
available for re-evaluating gravity and aeromag-
netic data and an understanding of continental
margins should be further enhanced. A new phase
of geophysical investigations in the Canadian
Shield along these lines is essentially already
underway in the form of the LITHOPROBE initia-
tive (Lithoprobe Steering Committee, 1986).
ACKNOWLEDGMENTS
I thank my colleague R. A. Gibb and referees H.
Helmstaedt and E. S. Robinson for various helpful
comments which contributed to an overall im-
provement of this paper. Draftsmen L. Campbell
and A. Rafeek and photographer R. Delaunais are
thanked for their combined efforts in producing the
figures.
22
M. D. Thomas
REFERENCES
ALABI, A. 0., P. A. CAMFIELD, and D. I. GOUGH, 1975,
The North American Central Plains conductivity anomaly:
Geophysical Journal of the Royal Astronomical Society, v.
43, p. 815-833.
BAER, A. J., 1977, Comment on "Gravity anomalies and
deep structure of the Cape Smith foldbelt, northern
Ungava, Quebec": Geology, v. 5, p. 651.
BEALS, C. S., 1968, Theories of the origin of Hudson Bay,
Part 1: On the possibility of a catastrophic origin for the
great arc of eastern Hudson Bay; pp. 985-999 in C. S. Beals
(ed.), Science, History and Hudson Bay, Volume 2:
Department of Energy, Mines and Resources, Ottawa,
Ontario, Canada, 1057 p.
BERGER, G. W., D. YORK, and D. J. DUNLOP, 1979,
Calibration of Grenvillian palaeopoles by 40 Ar!39 Ar dating:
Nature, v. 277, p. 46-48.
BERRY, M. J., and K. FUCHS, 1973, Crustal structure of
the Superior and Grenville Provinces of the northeastern
Canadian Shield: Bulletin of the Seismological Society of
America, v. 63, p. 1393-1432.
BmD, P., 1978, Finite element modeling of lithosphere
deformation: The Zagros collision orogeny: Tectono-
physics, v. 50, p. 307-336.
BURKE, K. C., and J. F. DEWEY, 1972, Orogeny in Africa;
pp. 583-608 in T. F. J. Dessauvagie and A. J. Whiteman
(eds.), African Geology, [badan 1970: Ibadan University
Press, Ibadan, Nigeria, 668 p.
__ , J. F. DEWEY, and W. S. F. KIDD, 1976, Precambrian
palaeomagnetic results compatible with contemporary
operation of the Wilson cycle: Tectonophysics, v. 33, p. 287-
299.
CAMFIELD, P. A., and D. I. GOUGH, 1977, A possible
Proterozoic plate boundary in North America: Canadian
Journal of Earth Sciences, v. 14, p. 1229-1238.
CAVANAUGH, M. D., and C. K. SEYFERT, 1977, Apparent
polar wander paths and the joining of the Superior and
Slave Provinces during Early Proterozoic time: Geology, v.
5, p. 207-211.
CHANDLER, F. W., and E. J. SCHWARZ, 1980, Tectonics of
the Richmond Gulf area, Northern Quebec - A hypo-
thesis; pp. 59-68 in Current Research, Part C: Geo-
logical Survey of Canada, Paper 80-1 C, 248 p.
DAVIDSON, A., 1986, New interpretations in the south-
western Grenville Province; pp. 61-74 in J. M. Moore, A.
Davidson, and A. J. Baer (eds.), The Grenville Province:
Geological Association of Canada, Special Paper 31, 358 p.
DEWEY, J. F., 1969, Evolution of the Appalachian! Cale-
donian orogen: Nature, v. 222, p. 124-129.
and J. M. BIRD, 1970, Plate tectonics and geo-
synclines: Tectonophysics, v. 10, p. 625-638.
DICKINSON, W. R, 1971, Plate tectonic models of geo-
synclines: Earth and Planetary Science Letters, v. 10, p.
165-174.
DIETZ, R S., 1961, Continent and ocean basin evolution by
spreading of the sea floor: Nature, v. 190, p. 854-857.
__ ,1963, Collapsing continental rises: An actualistic
concept of geosynclines and mountain building: Journal of
Geology, v. 71, p. 314-333.
__ , 1966, Passive continents, spreading sea floors, and
collapsing continental rises: American Journal of Science,
v. 264, p. 177-193.
DIMROTH, E., 1970, Evolution of the Labrador Geosyn-
cline: Geological Society of America Bulletin, v. 81, p.
2717-2741.
__ , 1972, The Labrador geosyncline revisited: Ameri-
can Journal of Science, v. 272, p. 487-506.
__ , W. R A. BARAGAR, R BERGERON, and G. D.
JACKSON, 1970, The filling of the Circum-Ungava geo-
syncline; pp. 45-142 in A. J. Baer, (ed.), Basins and Geo-
syncUnes of the Canadian Shield: Geological Suruey of
Canada, Paper 70-40, 265 p.
DRUMMOND, B. J., 1979, A crustal profile across the
Archaean Pilbara and northern Yilgarn cratons, north-
west Australia: BMR Journal of Australian Geology and
Geophysics, v. 4, p. 171-180.
__ ,1981, Crustal structure of the Precambrian terrains
of northwest Australia from seismic refraction data: BMR
Journal of Australian Geology and Geophysics, v. 6, p. 123-
135.
__ , and H. M. SHELLEY, 1981, Isostasy and structure of
the lower crust and upper mantle in the Precambrian
terrains of northwest Australia, from regional gravity
studies: BMR Journal of Australian Geology and Geo-
physics, v. 6, p.137-143.
GIBB, R. A., 1978, Slave-Churchill collision tectonics:
Nature, v. 271, p. 50-52.
__ , and HALLIDAY, D. W., 1974, Gravity Measure-
ments in Southern District of Keewatin and South-
eastern District of Mackenzie, N. W. T.: Earth Physics
Branch, Ottawa, Grauity Map Series Nos. 124-131,36 p.
__ , and M. D. THOMAS, 1976, Gravity signature offossil
plate boundaries in the Canadian Shield: Nature, v. 262, p.
199-200.
__ , 1977, The Thelon Front: A cryptic suture in the
Canadian Shield?: Tectonophysics, v. 38, p. 211-222.
Ancient collisional continental margins in the Canadian Shield 23
__ , and R. I. WALCOTT, 1971, A Precambrian suture in
the Canadian Shield: Earth and Planetary Science Letters,
v. 10, p. 417-422.
__ , M. D. THOMAS, P. L. LAPOINTE, and M.
MUKHOPADHYAY, 1983, Geophysics of proposed Protero-
zoic sutures in Canada: Precambrian Research, v. 19, p.
349-384.
GREEN, A. G., O. G. STEPHENSON, G. D. MANN, E. R.
KANASEWICH, G. L. CUMMING, Z. HAJNAL, J. A. MAIR,
and G. F. WEST, 1980, Cooperative seismic surveys across
the Superior-Churchill boundary zone in southern Canada:
Canadian Journal of Earth Sciences, v. 17, p. 617-632.
__ , B. MILKEREIT, A. DAVIDSON, C. SPENCER, D.R.
HUTCHINSON, W. F. CANNON, M. W. LEE, W. F. AGENA,
J.C. BEHRENDT, and W. J. HINZE, 1988, Crustal structure
of the Grenville front and adjacent terranes: Geology, v.
16, p.788-792.
__ , W. F. CANNON, B. MILKEREIT, D. R. HUTCHINSON,
A. DAVIDSON, J. C. BEHRENDT, C. SPENCER, M. W. LEE,
P. MOREL-A-L'HurSSIER, and W. F. AGENA, 1989 , A
"GLIMPCE" of the deep crust beneath the Great Lakes; pp.
65-80 in R. F. Mereu, S. Mueller, and D. M. Fountain (eds.),
Properties and Processes of Earth's Lower Crust: Am-
erican Geophysical Union/International Union of Geology
and Geophysics Volume 6; Geophysical Monograph 51, 338
p.
GROTZINGER, J. P., and Q. GALL, 1986, Preliminary
investigations of Early Proterozoic Western River and
Burnside River Formations: Evidence for foredeep origin
of Kilohigok Basin, District of Mackenzie; pp. 95-106 in
Current Research, Part A: Geological Survey of Canada,
Paper 86-1A, 826 p.
HALL, D. H., 1977, Overview of geophysics, Churchill
Province; pp. 6-23 in Annual Report - 1977, Research:
Centre for Precambrian Studies, University of Manitoba,
Winnepeg, Manitoba, Canada, 138 p.
__ , and Z. HAJNAL, 1973, Deep seismic crustal studies
in Manitoba: Bulletin of the Seismological Society of
America, v. 63, p. 885-910.
HENDERSON, J. B., P. H. McGRATH, D. T. JAMES, and R. 1.
MacFIE, 1987, An integrated geological, gravity and
magnetic study of the Artillery Lake area and the Thelon
Tectonic Zone, District of Mackenzie; pp. 803-814 in
Current Research, Part A: Geological Survey of Canada,
Paper 87-1A, 946p.
HILDEBRAND, R. S., P. F. HOFFMAN, and S. A. BOWRING,
1987, Tectonomagmatic evolution ofthe 1.9-Ga Great Bear
magmatic zone, Wopmay orogen, northwestern Canada:
Journal of Volcanology and Geothermal Research, v. 32, p.
99-118.
HOFFMAN, P. F., 1980, Wopmay Orogen: A Wilson cycle of
Early Proterozoic age in the northwest of the Canadian
Shield; pp. 523-549 in D. W. Strangway (ed.), The Con-
tinental Crust and its Mineral Deposits: Geological
Association of Canada, Special Paper 20, 804 p.
__ , 1985, Is the Cape Smith Belt (northern Quebec) a
klippe?: Canadian Journal of Earth Sciences, v. 22, p.
1361-1369.
__ , 1987, Continental transform tectonics: Great Slave
Lake shear zone (ca. 1.9 Ga), northwest Canada: Geology,
v. 15, p. 785-788.
__ ,1988, United plates of America, the birth of a craton:
Early Proterozoic assembly and growth of Laurentia:
Annual Review of Earth and Planetary Sciences, v. 16, p.
543-603.
__ , J. A. FRASER, and J. C. McGLYNN, 1970, The
Coronation geosyncline of Aphebian age, District of Mac-
kenzie; pp. 201-212 in A. J. Baer (ed.), Basins and Geo-
synclines of the Canadian Shield: Geological Survey of
Canada, Paper 70-40, 265 p.
HYNES, A., and D. M. FRANCIS, 1982, A transect of the
Early Proterozoic Cape Smith foldbelt, New Quebec:
Tectonophysics, v. 88, p. 23-59.
IRVING, E., A. DAVIDSON, and J. C. McGLYNN, 1984,
Paleomagnetism of gabbros of the Early Proterozoic Blach-
ford Lake intrusive suite and the Easter Island dyke, Great
Slave Lake, NWT: Possible evidence for the earliest con-
tinental drift: Geophysical Surveys, v. 7, p. 1-25.
__ , R. F. EMSLIE, and H. UENO, 1974, Upper Pro-
terozoic paleomagnetic poles from Laurentia and the
history of the Grenville structural province: Journal of
Geophysical Research, v. 79, p. 5491-5502.
__ , J. K. PARK, and J. L. ROY, 1972, Paleomagnetism
and the origin of the Grenville Front: Nature, v. 236, p.
344-346.
JOHNSON, R. A., K. E. KARLSTROM, S. B. SMITHSON, and
R. S. HOUSTON, 1984, Gravity profiles across the Chey-
enne Belt, a Precambrian crustal suture in southeastern
Wyoming: Journal of Geodynamics, v. 1, p. 445-472.
JONES, A. G., and P. J. SAVAGE, 1986, North American
Central Plains conductivity anomaly goes east: Geophysi-
cal Research Letters, v. 13, p. 685-688.
KAlLA, K. L., K. ROY CHOWDHURY, P. R. REDDY, V. G.
KRISHNA, H. NARAIN, S. I. SUBBOTIN, V. B. SOLLOGUB,
A. V. CHEKUNOV, G. E. KHARETCHKO, M. A. LAZAR-
ENKO, and T. V. ILCHENKO, 1979, Crustal structure along
Kavali-Udipi profile in the Indian Peninsular Shield from
deep seismic sounding: Journal of the Geological Society of
India, v. 20, p. 307-333.
24 M.D. Thomas
KEAREY, P., 1976, A regional structural model of the
Labrador Trough, northern Quebec, from gravity studies,
and its relevance to continent collision in the Precambrian:
Earth and Planetary Science Letters, v. 28, p. 371-378.
__ , 1977, A gravity survey of the central Labrador
Trough, northern Quebec: Canadian Journal of Earth
Sciences, v. 14, p. 45-55.
LITHOPROBE STEERING COMMITTEE, 1986, LITHO-
PROBE, the Third Dimension, a Coordinated National
Geoscience Project: Phase II Proposal. Canadian Geo-
science Council, Edmonton, Alberta, Canada, 140 p.
MATHUR, S. P., 1974, Crustal structure in southwestern
Australia from seismic and gravity data: Tectonophysics,
v. 24, p. 151-182.
__ , 1976, Relation of Bouguer anomalies to crustal
structure in southwestern and central Australia: BMR
Journal of Australian Geology and Geophysics, v. 1, p. 277-
286.
McGLYNN, J. C., and E. IRVING, 1975, Paleomagnetism of
early Aphebian diabase dykes from the Slave Structural
Province, Canada: Tectonophysics, v. 26, p. 23-38.
__ , E. IRVING, K. BELL, and G. PULLAIAH, 1975,
Palaeomagnetic poles and a Proterozoic supercontinent:
Nature, v. 255, p. 318-319.
MEREU, R F., 1968, Curvature effects on Hudson Bay and
Project Early Rise seismic refraction results; pp. 315-325 in
P. J. Hood (ed.), Earth Science Symposium on Hudson
Bay: Geological Survey of Canada, Paper 68-53, 386 p.
MUKHOPADHYAY, M., and R A. GIBB, 1981, Gravity
anomalies and deep structure of eastern Hudson Bay:
Tectonophysics, v. 72, p. 43-60.
MYERS, J. S., 1975, Vertical crustal movements of the
Andes in Peru: Nature, v. 254, p. 672-674.
NUNN, G. A. G., C. F. GOWER, and A. T. THOMAS, 1988, A
synthesis of the ca. 1680-1640 Ma Labradorian Orogeny
[abstract): Geological Association of Canada, Minera-
logical Association of Canada, and Canadian Society of
Petroleum Geologists Joint Annual Meeting, Program with
Abstracts, v. 13, p. A91.
PARRISH, R. R, 1989, U-Pb geochronology of the Cape
Smith Belt and Sugluk block, northern Quebec: Geoscience
Canada, v. 16, p. 126-130.
PIPER, J. D. A., 1982, The Precambrian palaeomagnetic
record: The case for the Proterozoic Supercontinent: Earth
and Planetary Science Letters, v. 59, p. 61-89.
PLUMB, K. A., 1979, The tectonic evolution of Australia:
Earth-Science Reviews, v. 14, p. 205-249.
ROSS, G. M., M. E. VILLENEUVE, R R PARRISH, and S. A.
BOWRING, 1988, Tectonic assembly of the Canadian Shield
in the Alberta subsurface: Integrated potential field map-
ping and UIPb geochronology [abstract): Geological Society
of America, Abstracts with Programs, v. 20, no. 7, p. A326.
SEYFERT, C. K., and M. D. CAVANAUGH, 1978, Reply to
comment on "Apparent polar wander paths and the joining
of the Superior and Slave provinces during early Protero-
zoic time": Geology, v. 6, p. 133-135.
ST-ONGE, M. R, S. B. LUCAS, D. J. SCOTT, N. J. BEGIN, H.
HELMSTAED, and D. M. CARMICHAEL, 1988, Thin-
skinned imbrication and subsequent thick-skinned folding
of rift-fill, transitional-crust, and ophiolite suites in the 1.9
Ga Cape Smith Belt, northern Quebec; pp. 1-18 in Current
Research, Part C: Geological Survey of Canada, Paper
88-IC, 372 p.
SUBRAHMANYAM, C., 1978, On the relation of gravity
anomalies to geotectonics of the Precambrian terrains of
the South Indian Shield: Journal of the Geological Society
of India, v. 19, p. 251-263.
TANNER, J. G., 1969, A Geophysical Interpretation of
Structural Boundaries in the Eastern Canadian
Shield: PhD dissertation, University of Durham, Dur-
ham, England, UK, 194 p.
THOMAS, M. D., 1983, Tectonic significance of paired
gravity anomalies in the southern and central Appala-
chians; pp. 113-124 in R. D. Hatcher, Jr., H. Williams, and
I. Zietz (eds.), The Tectonics and Geophysics of Moun-
tain Chains: Geological Society of America, Memoir 158,
223p.
__ , 1984, Structural significance of some gravity ano-
malies in Wopmay Orogen [abstract): Geological Associa-
tion of Canada, Mineralogical Association of Canada, Joint
Annual Meeting, Program with Abstracts, v. 9, p. 111.
__ , 1985, Gravity studies of the Grenville province:
Significance for Precambrian plate collision and the origin
of anorthosite; pp. 109-123 in W. J. Hinze (ed.), The Utility
of Regional Gravity and Magnetic Anomaly Maps:
Society of Exploration Geophysicists, Tulsa, Oklahoma,
454p.
__ . , and R. A. GIBB, 1977, Gravity anomalies and deep
structure of the Cape Smith foldbelt, northern Ungava,
Quebec: Geology, v. 5, p. 169-172.
__ , and P. KEAREY, 1980, Gravity anomalies, block-
faulting and Andean-type tectonism in the Eastern Chur-
chill Province: Nature, v. 283, p. 61-63.
__ , and J. G. TANNER, 1975, Cryptic suture in the
eastern Grenville Province: Nature, v. 256, p. 392-394.
Ancient collisional continental margins in the Canadian Shield 25
__ , V. L. SHARPTON, and R. A. F. GRIEVE, 1987,
Gravity patterns and Precambrian structure in the North
American Central Plains: Geology, v. 15, p. 489-492.
TIRRUL, R., 1985, Nappes in the Kilohigok Basin and their
relation to the Thelon Tectonic Zone, District of Mac-
kenzie; p. 407-420 in Current Research, Part A: Geo-
logical Survey of Canada, Paper 85-1A, 802 p.
VINE, F. J., and H. H. HESS, 1970, Seafloor spreading; pp.
587-622 in A. E. Maxwell (ed.), The Sea, Volume 4, Part
II: John Wiley and Sons, New York, New York, 664 p.
WALCOTT, R. 1., andJ. B. BOYD, 1971, The Gravity Field
of northern Alberta, and Part of Northwest Terri-
tories and Saskatchewan: Earth Physics Branch, Otta-
wa, Gravity Map Series 103-111, 13 p.
WARDLE, R. J., T. RIVERS, C. F. GOWER, G. A. G. NUNN,
and A. THOMAS, 1986, The northeastern Grenville Pro-
vince: New insights; pp. 13-29 in J. M. Moore, A. David-
son, and A. J. Baer (eds.J, The Grenville Province: Geo-
logical Association of Canada, Special Paper 31, 358 p.
WILCOX, L. E., 1976, Airy-Woollard isostasy; pp. 53-57 in
G. H. Sutton, M. H. Manghnani, and R. Moberly (eds.l,
The Geophysics of the Pacifze Ocean Basin and Its
Margin: American Geophysical Union, Geophysical Mono-
graph 19, 480 p.
WYNNE-EDWARDS, H. R., 1972, The Grenville Province;
pp. 263-334 in R. A. Price and R. J. W. Douglas (eds.),
Variations in Tectonic Styles in Canada: Geological
Association of Canada, Special Paper 11, 688 p.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Tectonothermics of modern and ancient continental margins
MALCOLM J. DRURY
Geological Survey of Canada, Energy, Mines and Resources Canada, 601 Booth Street, Ottawa, Ontario, KIA OE8, Canada
(received November 28,1988; revision accepted July 5,1989)
ABSTRACT
Characteristic patterns of heat flow are observed across present-day convergent and constructional
margins. At active oceanic subduction margins heat flow varies systematically from normal oceanic values to
low values, over the cold subducting slab, to high values over the zone of volcanism associated with slab
melting, to normal continental values. A similar pattern is observed over active continent-continent collision
zones. In both, hydrothermal circulation cells associated with the high heat-flow zone lead to ore deposition.
In old continent-continent collision terrains the heat-flow pattern is broadly similar to that over active
collision zones, although for quite different reasons. Low heat flow occurs over volcanic belts associated with
subduction, because of their low radiogenic heat production. Higher heat flow occurs over plutons injected
during subduction, because they have higher heat generation than normal crust. A zone of low heat flow may
occur over middle-lower crustal terrains, now exposed by deep erosion, because of their low heat generation.
The variation of heat flow can be used to indicate the direction of the ancient subduction.
At active spreading ridges heat flow decreases systematically away from the axis of magma injection. The
heat flow pattern across old continental rift systems is quite different, with bands of higher-than-usual heat
flow delineating the edges of the rift, the result of thermal refraction at the boundary between low thermal
conductivity rift-fill material and higher conductivity felsic crust. If the rift contains substantial volumes of
low heat generation volcanic rocks, heat flow over the rift may be lower than that outside.
Examples are taken from the Canadian Shield. Heat flow and heat-generation patterns across the
boundary between the western Superior Province and Churchill Province imply subduction of Churchill crust
under Superior, contrary to some previous models. Heat flow patterns across the Mid-Continent Rift in
eastern Lake Superior contradict recent interpretations of seismic reflection data and suggest it is an
asymmetrical rift containing a relatively small volume of volcanic rocks.
INTRODUCTION
Heat is the fundamental driving force of most
dynamic systems in the earth's crust and mantle,
from large-scale processes such as plate tectonics to
regional-scale phenomena such as hydrothermal
circulation in permeable crust. Its study is therefore
of paramount importance in the earth sciences.
Heat-flow studies are most often used to examine
the present internal state of the earth; for example,
the variation of heat flow across active spreading
and subduction zones is a key indicator of the nature
and spatial extent of those processes. The profound
importance of fluids in crustal evolution (Fyfe, 1986)
has considerable bearing on heat-flow studies, as
fluid flow and heat flow are usually linked.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 27-36. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
27
28
M. J. Drury
Heat flow (Q) is detennined from the product of
the temperature gradient and thennal conductivity.
In addition, it is desirable to measure heat genera-
tion (A) from the natural decay of radioactive
isotopes of U, Th, and, usually of lesser importance,
K in the rock. For several different terrains, heat
flow plotted as a function of heat generation shows
an apparent linear relationship expressed as:
Q = Qo + AoD
[1]
in which Qo is the "reduced heat flow" (Roy and
others, 1968) and is interpreted to be the component
of heat flow from the mantle and lower crust; D,
which has the dimension of depth, is a scale depth
for the vertical distribution of the heat-producing
isotopes. The significance of the relationship has
been discussed by many authors (Drury, 1989a, and
references therein). It implies partitioning of the
crust with respect to heat generation, with the lower
crust having unifonnly low heat generation and the
upper crust having higher and more variable heat
generation.
This paper is concerned principally with ap-
plication of heat-flow studies to characterisation of
ancient continental margins associated with past
cycles of ocean basin opening and closing, encom-
passing continental collision and continental rift-
ing. Geothennics is an important component in the
identification and delineation of such features.
HEAT-FLOW PATTERNS ACROSS
CONTINENTAL MARGINS
Present-Day Convergent Margins
Characteristic patterns of heat-flow variations
are observed across present-day convergent contin-
ental margins. Figure IA shows schematically the
heat-flow pattern across a zone of subduction of an
oceanic slab beneath continental crust. The slab is
initially cooler than the sub-continental mantle into
which it descends. Hence, immediately landward of
the trench, heat flow is low. Thennal conduction,
friction arising from the slipping of the slab, and
some adiabatic heating combine to raise the tem-
perature of the slab. As the slab sinks deeper into
the mantle, it begins to melt and very high heat flow
is measured above the melting zone. Both active
volcanism and pluton emplacement occur.
An excellent series of heat flow measure-ments
across a convergent ocean-continent plate boundary
is that across the subducting Juan de Fuca plate
landward to the Canadian Cordillera (Hyndman,
1976; Lewis and Jessop, 1981; Davis and Lewis,
1984; Lewis and others, 1985, 1988) and Cascade
range of the northwestern United States (Blackwell
and others, 1982). In the Canadian profile, heat
flow decreases from = 50 m W 1m
2
on the continental
shelf, to 25 mW/m
2
over the cool subducting slab, to
>80 mW/m
2
over and seaward (by =20 km) of the
Garibaldi Volcanic Belt. Lewis and others (1988)
postulated that dehydration of subducting oceanic
crust at =450C absorbs heat and produces water
that flows updip above the subducting crust, thereby
redistributing the heat seaward.
If the subducting plate carries a block of con-
tinental crust, there is eventually a continental
collision. The characteristic thennal profile across
an active continent-continent collisional margin is
shown in Figure IB and is based on modelling
discussed by Bird and others (1975). It has a fonn
similar to that of the ocean-continent collisional
margin. The cool subducted crust produces a zone of
low heat flow down-dip of the hinge zone, but fric-
tional heating of the upper surface of the slab leads
to a high heat-flow zone further down-dip.
Thennal effects of fluids in metamorphic pro-
cesses associated with continent continent collisions
are of great importance (Fyfe, 1986; Peacock, 1987).
Release of fluids during orogenic activity in thrust
zones has been discussed by, for example, Fyfe
(1986). He estimated the amount of water released
by prograde metamorphism from a crustal thrusting
event on the scale of the Himalayan tectonic zone.
Approximately 1.6 X 10
19
kg of water is liberated
over an area of =4.5X 10
6
km
2
Hoisch (1987) pre-
sented evidence that heat was transported by fluids
during Late Cretaceous regional metamorphism in
southeastern California (USA). He has suggested
that peak temperature increases, produced solely by
rising fluids, were = 300
0
K at mid-crustal depths.
That implies a potential peak increase of surface
heat flow by a factor of 2 or so. The time required for
an equilibrium thennal gradient to be established
from a perturbation at a depth of 15-20 km is = 7
million years, so although the short-tenn thennal
perturbation may be high, it will no longer be
detectable for Precambrian or Paleozoic events.
However, it results in hydrothennal deposition of
ores, metamorphism, and resetting of isotopic and
Tectonothermics of modern and ancient continental margins 29
fission-track dates - each of which may be used to
reconstruct tectonic history.
Ancient ColliswnalMargins
Figure Ie shows schematically the theoretical
heat-flow pattern across an ancient continent-
continent collisional margin that has been
considerably eroded. The middle and lower parts of
the formeriy overriding crust have been exposed at
surface, and remnants of the oceanic slab exist in
volcano-sedimentary terrains trending parallel to
the strike of the subduction hinge zone. The situa-
tion is appropriate for Precambrian terrains; a
possible example from the Canadian Shield is dis-
cussed below. The heat flow is generally quite low,
because ofthe age of the terrain, but does show some
variability. For example, heat flow is lower in
volcano-sedimentary belts because volcanic rocks
are generally of significantly lower heat production
than felsic rocks (Jessop and Lewis, 1978). Simi-
larly, heat flow is reduced over the exposed lower
crust since it also has a low heat production. Heat
generation in lower crustal rocks of granulite grade
is typically 0.1-0.6 pW/m
3
, in contrast to upper
crustal, felsic rocks in which heat generation is
more variable and much higher, > Ip W/m
3
(Drury,
1989).
Note that Figure lC also indicates a peak in
heat flow, which is associated with the remains of
radiogenic plutons intruded during early stages of
subduction. High heat production plutons are not
unusual in Precambrian terrains (Jessop and Lewis,
1978; Drury and Lewis, 1983; Drury, 1990). Be-
cause they must must originally have produced
much more heat than now, they may have served to
focus hydrothermal circulation and, therefore, ore
deposition. In normal plutons, hydrothermal cir-
culation is driven by heat associated with the initial
intrusion, by heat from any subsequent tectono-
thermal events, and perhaps from topographic
relief. That heat may be responsible for driving
hydrothermal circulation in large volumes of the
crust (lateral distances of '" 30 km, depth of 5-7 km,
as deduced from studies of the Idaho batholith by
Criss and Taylor, 1983). In radiogenic plutons, the
heat produced by radioactive decay can prolong
hydrothermal circulation long after cooling has
ended (Fehn and others, 1978; Brown and others,
1980).
a
(AI
a
(el
a
(CI
120

30
eo
60
30
f-- ',"1
,,- CONTINENTAL
r---l CRUST
L--.l LOWER
&j SEOIMENTS
---------
---------
VOLCANO-
SEDIME TARY ROCK
G
D PLUTONS
a HEAT FLOW (mW/m21 , FLUIO FLOW
Figure 1. Variation of heat flow across convergent
terrains. A) Oceanic crust being subducted beneath con-
tinent; B) continent-continent collision, with one crustal
block underriding another, resulting in uplift; C) deeply
eroded remnant of former continent-continent collision
zone, with lower crust exposed at surface in formerly
uplifted terrain. Ranges of heat flow values shown are
approximate.
Comparison of Figures lA and IB shows that
heat-flow patterns across present-day subduction
and continental collision zones are broadly similar.
The principal characteristic is a paired heat-flow
anomaly, with the direction of subduction being
indicated by the low-high heat-flow pattern. This
pattern is similar to that of paired gravity ano-
malies at collisional sutures, which may also be
used to indicate former subduction directions
(Thomas, 1983).
30 M.J. Drury
In deeply eroded Precambrian terrains the
heat-flow pattern may be different, especially if
formerly mid-lower crust is exposed, when there
would be a second low heat flow anomaly corres-
ponding to (the low-heat generating lower crust).
However, the former subduction direction may be
inferred if remnants of radiogenic plutons produce a
definable high heat-flow anomaly. The high heat
flow would be down-dip (i.e., dip of the subducted
slab) of the accreted volcanic terrane (Figure IC),
and up-dip of the exposed lower crust.
Present-Day Constructive Margins
High heat flow is associated with basalt injec-
tion at spreading ridges; it decreases away from the
axis of rifting. Rifts that have progressed to the
proto-ocean stage are expected to be characterised
by heat flow patterns similar to those across oceanic
spreading ridges. For the Red Sea, heat flow is high,
though variable, over the central trough, and de-
creases with distance from the spreading centre
(Evans and Tammemagi, 1974), as shown sche-
matically in Figure 2a. A review of heat-flow
studies in continental rifts was presented by
Morgan (1983). In active rifts that have not yet
progressed to the spreading stage, heat flow is found
to be high and variable, with large convective com-
ponents. The East African Rift valley, for example,
is characterised by a major geothermal energy
resource exemplified by large numbers of hot
springs (Crane and O'Connell, 1983).
k: ':'.: co TINE TAL CRUST
_ VOLCANICS
C:=J SEDIMENTS
lo)oeli rift
Figure 2. Variation of heat flow across Ca) an active rift
zone and Cb) an inactive rift.
Igneous activity related to rifting can also
produce characteristic patterns of metallogenesis.
For example, the Bergslagen ore belt in central
Sweden, which comprises a complex of Middle
Proterozoic volcano-sedimentary and plutonic units
formed in a rift environment, contains ore deposits
related to sub-seafloor hydrothermal convection
systems driven by heat associated with the rifting
(Oen, 1987).
Ancient Constructive Margins
FIgure 2b shows schematically the expected
variation of heat flow across an ancient rift in which
thermotectonic activity has long since ceased, so
that no thermal anomaly arising from the process of
rifting remains. The heat-flow pattern is charac-
terised by peaks associated with the edges of the rift.
The peaks arise because of the contrast in thermal
conductivity between low conductivity volcanic and
sedimentary rift-fill material and higher conducti-
vity crystalline crust (Saltus and Lachenbruch,
1987). The thermal refraction effect may be quite
pronounced - 50% or more - depending on the con-
ductivity contrast. A possible example is seen in the
Liaohe Rift Basin in China. The basin consists of
two parallel rifts underlain by Archaean metamor-
phic rocks and separated by a horst of the same
crustal material. Rift fill materials are clastic
sediments and volcanic rocks. Heat flow at the
outer edges of the double graben system, and along
the central horst, is 50% higher than in the cen-
tral parts of the grabens (Wang and Wang, 1988).
Heat flow in non-active rifts may be lower than
in the bounding felsic crust because of low heat
production of volcanic material. The quantitative
difference would depend largely on the amount of
volcanic material: if sediments derived from crystal-
line crust were a major component of the rift fill the
difference may be small, as the sediments would
have heat generation similar to the felsic crust.
EXAMPLES FROM PRECAMBRIAN TERRANES
OF NORTH AMERICA
Tectonothermal studies of ancient continental
margins must consider both current heat-flow pat-
terns and those deduced from other indicators of
former thermal activity, such as metamorphism and
the products of hydrothermal circulation. In this
Tectonothermics of modern and ancient continental margins 31
section, some examples of different kinds of Pre-
cambrian continental margins are discussed in
terms of their present heat-flow patterns. The
examples are from Canadian Shield, an assemblage
of Archaean and Proterozoic terrains in which 54
borehole heat-flow measurements have been
reported, and summarised by Drury (1991).
A Precambrian Collisional Margin: The
Superior-Churchill Boundary
Heat-flow measurements in the western part of
the Churchill Province of the Canadian Shield were
reported by Drury (1985). Measurements from the
adjoining Superior Province were presented by
several authors, principally Jessop and Lewis
(1978), and summarised by Drury (1991). Figure 3
shows the heat-flow pattern across parts of the
western Superior and Churchill Provinces. A single
borehole value (51 mW/m
2
) measured entirely from
the Phanerozoic sedimentary cover has been
reported by Jessop and Vigrass (1989); the value of
38 mW/m
2
(Figure 3) was derived from a measure-
ment in basement underlying the sediments.
The Archaean Superior Province is bounded on
the west by a wide Proterozoic mobile belt (PMB,
Figure 3), a terrain that was considerably affected
by the Hudsonian Orogeny (= 1.9-1.7 Ga). Imme-
diately west of the Superior Province is the
Thompson Belt, in which a single heat-flow mea-
surement (59 mW/m
2
) has been made. The Thomp-
son Belt consists of Aphebian metasedimentary and
metavolcanic supracrustals and ultramafic rocks
tightly infolded in Archean basement. Two major
volcanic belts are prominent features: the Flin
Flon-Snow Lake belt, in which three heat-flow
measurements, ranging from 38 to 46 mW/m2, have
been reported; and the La Ronge-Lynn Lake belt,
with two heat-flow measurements (23 and 41
mW/m2). Immediately east of the Thompson Belt,
within the Superior Province, is an area of inferred
exposed lower crust, the Pikwitonei granulite
terrain, which has undergone =20 km of uplift
(Hubretgse, 1980). No heat-flow measurements
have been made in this terrain, but heat-generation
measurements on surface samples were made by
Fountain and others (1987). Heat generation in the
high-grade terrain was found to be low, in the range
0.14-0.3711W/m3. Fountain and others proposed a
lithological model for the Superior Province crust
and calculated for it the Q-A relationship (Equation
60
CffJRCHILL
j:!rO! fI'Ol QoC
.. tI.'la
. 51
Figure 3. Detail of western Superior Province and Chur-
chill Province. WB, Wathaman batholith, PMB, Protero-
zoic Mobile Belt. Borehole heat flow sites shown by solid
dots. Features of PMB are identifiable by heat flow values:
Thompson Belt, 59 mW/m2; Flin Flon-Snow Lake belt, 38
to 46 mW/m2; La Ronge-Lynn Lake belt, 23 mW/m
2
.
1); predicted surface heat flow in the Pikwitonei
terrain would be = 30 m W 1m
2
, based on this model.
Insufficient heat-flow data exist for the area of
Figure 3 to warrant construction of a profile. How-
ever, line AB, which crosses all terranes of the area,
indicates the variation of heat flow. In the west,
over the Archean part of the Churchill Province, the
single value of 54 mW/m
2
appears to be typical of
that terrane when compared with measurements
from the Phanerozoic cover (Majorowicz and others,
1986). Substantially lower heat flow is measured in
the volcanic belts of the PMB; a high value (59
mW/m
2
) is found in the Thompson Belt, and low
heat flow (30 mW/m2) is inferred for the Pikwitonei
terrain. The eastern end of the profile is in typical
Superior Province terrane in which heat flow is
inferred to be =45 mW/m
2
(Drury, 1991). The
variation of heat flow along line AB is qualitatively
very similar to that shown for the hypothetical
model of Figure lC (left to right). The Thompson
Belt heat-flow site was in a gneissic body with mean
heat generation of 1.0411W/m
3
-not high enough to
explain the high heat flow, whose measurement was
32
M. J. Drury
Figure 4. Two stages in the possible development of the
Proterozoic Mobile Belt (PMB) and edge of the Superior
craton. Intial rifting at ""2.4 to 2.3 Ga was followed by
ocean closure and continent-continent collision, with the
Churchill craton underriding the Superior. Fluids carried
downward by subduction led to redistribution of heat-
producing isotopes throughout the crust of the PMB.
very reliable (Drury, 1985). Some mechanism pro-
ducing a high heat flow is sought; it may be a sub-
surface radiogenic pluton, although there is no
independent evidence for this. There are occur-
rences of unusually high heat flow, attributed at
least in part to the presence of buried radiogenic
plutons, along the western margin of the Superior
craton (Drury, 1991).
Several authors have attempted to reconstruct
the tectonic history of the southwestern Churchill
Province. All proposed reconstructions involve
components of present-day plate tectonics. Lewry
(1981) considered the volcanic Flin Flon-Snow Lake
and La Ronge-Lynn Lake terranes to be the result of
subduction processes at "" 1.9 Ga years ago, and the
Thompson Belt to have been a transform fault.
Green and others (1985) presented a detailed hypo-
thetical model for the tectonic history of the region
2.4-1.6 Ga. Their model involved initial rifting, sea-
floor spreading, and ocean basin closure, with sub-
duction in a westerly direction.
Drury (1985) found that the Q-A relationship
(Equation 1) for the PMB is different from that of
the Superior Province. Whereas the parameter D is
not significantly different between the two, the
reduced heat flow in the PMB is sUbstantially
higher, 374 mW/m2, than in the Superior Pro-
vince, 271 mW/m
2
(Jessop and Lewis, 1978).
Drury (1985) suggested that the deep crust of the
PMB contains a higher concentration of heat-
producing isotopes of U, Th, and K than the Super-
ior Province crust. He proposed a scheme for the
tectonic development of the region during in the
Archaean that could have led to enrichment, in
terms of radioisotopes, of the lower crust of the PMB
(summarised in Figure 4). The major elements are:
(a) rupturing of craton, at ""2.4-2.3 Ga, to form an
ocean basin into which was deposited high-heat-
production sediments from the surface of the craton;
(b) cessation of rifting, closure of the ocean basin
and subduction of the oceanic crust, including the
sediments (Karig and Kay, 1981); (c) widespread
redistribution of radiogenic isotopes from sediments
by fluid circulation above the subducting slab (see
also Figure lA).
A further stage is now proposed: (d) continent-
continent collision, with uplift of the Superior
craton, implying the direction of subduction to be
Churchill under Superior crust. The justification for
assuming subduction in that direction is the simi-
larity of both geological structure and pattern of
heat flow between the PMB and exposed lower crust
of the Pikwitonei terrain and the hypothetical
crustal profile (Figure lC).
A model very similar to that described by Drury
(1985) was proposed by Paktunc and Baer (1986) on
the basis of quite different evidence. They examined
the geothermometry and geobarometry across the
Pikwitonei terrain and derived a model in which
Archaean crust of the Superior craton was thrust
over the Churchill craton in mid-Proterozoic time.
Subsequently, isostatic uplift of the thickened crust
exposed the Pikwitonei terrain. Furthermore,
Paktunc and Baer also suggested that water needed
for hydrous retrogression of the granulites may
have come from underthrusted sediments of the
Churchill crust. As noted above, Drury (1985)
suggested deep circulation of fluids to redistribute
radioisotopes from sediments of the underthrusting
Churchill crust.
This agreement between tectonic models
derived from significantly different information
suggests that the models have merit. Green and
others (1985) had difficulty in accounting for the
Pikwitonei terrain in their model, which involved
subduction in the opposite direction to that des-
cribed here. They invoked "an unusually thin upper
crust near to the eventual continental margin," but
without independent justification. If subduction
Tectonothermics of modern and ancient continental margins 33
were as described here and by Paktunc and Baer
(1986), no such invocation is necessary.
A Precambrian Rifted Margin: The Mid-
Continent Rift
The Mid-Continent Rift is a major Middle Pro-
terozoic feature of North America, extending"'" 2000
km from Kansas, curving through Lake Superior
and into central Michigan. In and around Lake
Superior, it is associated with thick sequences of
Keweenawan basaltic lavas and volcanogenic sedi-
mentary rocks. Seismic reflection profiles in the
lake were reported by Behrendt and others (1988).
Magnetic and gravity data for Lake Superior were
summarised by Hinze and others (1982). The
picture that emerges of the Lake Superior part of
the Mid-Continent Rift is one of thickened crust, of
which the upper part contains substantial thick-
nesses (> 10 km) of volcanic and sedimentary
sequences. The potential field signature for the
southwestern part of the Mid-Continent Rift is
different from that in Lake Superior. In the south-
west, the gravity profiles across the strike of the
Mid-Continent Rift are characterised by strong
central positive anomalies; the anomalies are more
complex in central Lake Superior, and in the east-
ern part of the lake they are significantly subdued
(Hinze and others, 1982). Magnetic anomalies also
exhibit variations, with the correlation between
magnetic and gravity anomalies being stronger in
the southwestern Mid-Continent Rift than in
eastern Lake Superior.
Figure 5. Variation of heat
flow in and around the Mid-
Continent Rift in Lake Super-
ior.
92'
.46
90'
A large number of lake bottom measurements
of heat flow have been made in Lake Superior by the
oceanic probe technique (Steinhart and others,
1968), with stations being reasonably well spaced
across the lake. Figure 5 shows contours of heat
flow in the lake, based on the contouring shown by
Steinhart and others (1968). Figure 6 shows gra-
vity, magnetic, and heat-flow profiles across the
eastern part of Lake Superior (profile A-B in Figure
5). A seismic reflection survey along the profile was
reported by Behrendt and others (1988), whose
interpretation of the seismic data is shown at the
bottom of Figure 6. They concluded that consider-
able thicknesses - <32 km - of volcanic and sedi-
mentary rocks occur in a symmetrical rift (their
Figure 2), with the upper 12-14 km consisting of
predominantly clastic sedimentary rocks, and the
lower section consisting principally of Keweenawan
volcanic rocks, with minor intercalated sediments.
They further suggested that the crust is uniformly
"'" 50 km thick beneath the lake, although their
positioning of the Moho based on the reflection data
seems to be highly speculative.
The thermal data do not support the conclusions
of Behrendt and others (1988). The heat-flow low at
the northern end of the profile is consistent with
thermal refraction at a dipping interface with a
lower conductivity - in this case, volcanic rocks -
to the south. No corresponding belt of low heat flow
marks the southern boundary of the graben in the
location postulated by Behrendt and others (1988).
Further, the uniform heat flow across the profile
(beyond the low at the northern end) is inconsistent
with the presence, across part of the profile, of a
.42
86'
HEAT FLOW
._____50- mW/m2
Borehole
(corrected)

84'
34 M.J. Drury
A
o ....
aOUilar .... ,
;25 Al"lomoly .......
(mGoIJ
--
-'0
-70.L...-----------------'

400
o
o 20
...............
km
60,-----------------,
H.-OI Flow
...................... ,
40 "'. " ......
....
20.L...-----------------'

"
.......... - - .............. .
. . . . . . . - ............... .
.................. . .. 50
-- ?
TwoO - Woy ApPfOJ:lmClle
Time h) OfP1h t km)
______________________
D
CL.ASTIC
SEDIMENTS
D
..
0,
RIFT
\lOL.CANfCS
AR'CHAEAN
BASEMENT
Figure 6. Variation of potential field signatures and heat
flow across transect AB of Figure 5. The lowermost part of
figure shows an interpretation (Behrendt and others, 1988)
of seismic reflection data from the profile AB.
thick pile of low-heatgeneration volcanic rocks.
Although the increased thickness of the crust may
compensate to some extent for the low-heat-
generation material suggested for the centre of the
profile, the heat flow beyond the profile should
increase if crustal thickness remains constant (as
implied by Behrendt and others). This is not ob-
served. The heat-flow data suggest an asymmetri-
cal rift system in eastern Lake Superior, perhaps
containing thick sequences of volcanic materials
with a major proportion of intercalated non-volcanic
sediments.
CONCLUSIONS
Attempts to unravel the tectonic development
of old terranes depend on information from diverse
sources: geological field mapping, geophysical sur-
veys, analysis ofP-T-t curves, and many other tech-
niques in the earth sciences. Unfortunately, it is
rare that tectonothermics is incorporated into
multidisciplinary studies. The importance of heat
in active tectonic regimes is obvious. Moreover, it
has been shown that observations of present-day
patterns of heat flow can be used to develop models
for ancient tectonic developments. Further, the
effects of former hydrothermal activity in ancient
terranes can be recognised indirectly -- as, for ex-
ample, in the Proterozoic Mobile Belt of the
Churchill Province, or directly -- as with ore depo-
sits. In the examples of heat-flow patterns across
ancient terrains presented here, the inferences to be
drawn solely from the tectonothermal studies are
different from inferences drawn by other authors
from other kinds of information. This should not be
taken as meaning that the tectonothermal models
are correct and others wrong: the models of Figure
1 are hypothetical and not open to independent
verification. It should simply be taken as an
indication of the importance of thermal studies as
part of a true multidisciplinary attack on geological
and geophysical problems.
REFERENCES
BEHRENDT, J. C., A. G. GREEN, W. F. CANNON, D. R.
HUTCHINSON, M. W. LEE, B. MILKEREIT, W. F. AGENA,
and C. SPENCER, 1988, Crustal structural of the Mid-
Continent Rift System - Results from GLIMPCE deep
seismic reflection profiles: Geology, v. 16, p. 81-85.
BIRD, P., M. N. TOKS6z, and N. H. SLEEP, 1975, Thermal
and mechanical models of continent-continent convergence
zones: Journal of Geophysical Research, v. 80, p. 4405-
4416.
BLACKWELL, D. D., R. G. BOWEN, D. A. HULL, J. RICCIO,
and J. L. STEELE, 1982, Heat flow, arc volcanism, and
subduction in northern Oregon: Journal of Geophysical
Research, v. 87, p. 8735-8754.
BROWN, G. C., J. CASSIDY, E. R. OXBURGH, J. PLANT, P.
A. SABINE, and J. V. WATSON, 1980, Basement heat flow
and metalliferous mineralisation in England and Wales:
Nature, v. 288, p. 657-659.
CRANE, K., and S. O'CONNELL, 1983, The distribution and
implications of heat flow from the Gregory Rift in Kenya:
Tectonophysics, v. 94, p. 253-275.
Tectonothermics of modern and ancient continental margins 35
CRISS, R. E., and H. P. TAYLOR, 1983, An
18
0/
16
0 and D/H
study of Tertiary hydrothermal systems in the southern
half of the Idaho batholith: Geological Society of America
Bulletin, v. 94, p. 640-663.
DAVIS, E. E., and T. J. LEWIS, 1984, Heat flow in a back-
arc environment: Intermontane and Omineca Crystalline
Belts, southern Canadian Cordillera: Canadian Journal of
Earth Sciences, v. 21, p. 715-726.
DRURY, M. J., 1985, Heat flow and heat generation in the
Churchill Province of the Canadian Shield and their
palaeotectonic significance: Tectonophysics, v. 115, p. 25-
44.
__ ,1988, Tectonothermics of the North American Great
Plains basement: Tectonophysics, v. 148, p. 299-307.
__ , 1989, The heat flow-heat generation relationship:
Implications for the nature of continental crust: Tectono-
physics, v. 164, p. 93-106.
__ , 1991, Heat flow in the Canadian Shield and its rela-
tion to other geophysical parameters; pp. 317-337 in V.
Cermak and L. Rybach (eds.), Terrestrial Heat Flow and
the Structure of the Lithosphere: Springer-Verlag, Ber-
lin, Germany.
__ , and T. J. LEWIS, 1983, Water movement within Lac
du Bonnet batholith as revealed by detailed thermal stu-
dies of three closely spaced boreholes: Tectonophysics, v.
95, p. 337-35l.
EVANS, T. R, and H. Y. TAMMEMAGI, 1974, Heat flow and
heat production in northeast Africa: Earth and Planetary
Science Letters, v. 23, p. 349-356.
FEHN, D., L. M. CATHLES, and H. D. HOLLAND, 1978, Hy-
drothermal convection and uranium deposits in abnormal-
ly radioactive plutons: Economic Geology, v. 73, p. 1556-
1566.
FOUNTAIN, D. M., M. H. SALISBURY, and K. P. FURLONG,
1987, Heat production and thermal conductivity of rocks
from the Pikwitonei-Sachigo continental cross section,
central Manitoba: Implications for the thermal structure
of Archean crust: Canadian Journal of Earth Sciences, v.
24, p. 1583-1594.
FUMERTON, S. L., M. R. STAUFFER, and J. F. LEWRY,
1984, The Wathaman batholith: Largest known Precam-
brian pluton: Canadian Journal of Earth Sciences, v. 214,
p.1082-1097.
FYFE, W. S., 1986, Fluids in deep continental crust; pp. 33-
39 in M. Barazangi and L. Brown (eds.), Reflection Seis-
mology: The Continental Crust: American Geophysical
Union, Geodynamics Series 14, 339 p.
GREEN, A. G., Z. HAJNAL, and W. WEBER, 1985, An
evolutionary model of the western Churchill Province and
western margin of the Superior Province in Canada and
the north-central United States: Tectonophysics, v. 116, p.
281-322.
HINZE, W. J., R J. WOLD, and N. W. O'HARA, 1982,
Gravity and magnetic anomaly studies of Lake Superior;
pp. 203-221 in R J. Wold and W. J. Hinze (eds.), Geology
and Tectonics of the Lake Superior Basin: Geological
Society of America, Memoir 156, 280 p.
HOISCH, T. D., 1987, Heat transport by fluids during Late
Cretaceous regional metamorphism in the Big Maria
Mountains, southeastern California: Geological Society of
America Bulletin, v. 98, p. 549-553.
HUBREGTSE, J. J. M. W., 1980, The Archean Pikwitonei
Granulite Domain and Its Position at the Margin of the
Northwestern Superior Province (Central Manitoba):
Manitoba Geological Survey, Paper GP80-3, 16 p.
HYNDMAN, RD., 1976, Heat flow measurements in the
inlets of southwest British Columbia: Journal of Geophysi-
cal Research, v. 81, p. 337-349.
JESSOP, A. M., and T. J. LEWIS, 1978, Heat flow and heat
generation in the Superior province of the Canadian
Shield: Tectonophysics, v. 50, p. 55-77.
__ , and L. W. VIGRASS, 1989, Geothermal measure-
ments in a deep well at Regina, Saskatchewan: Journal of
Volcanology and Geothermal Research, v. 37, p. 151-166.
KARIG, D. E., and R W. KAY, 1981, Fate of sediments on
the descending plate at convergent margins: Philosophical
Transactions of the Royal Society of London, v. A301, p.
233-25l.
LEWIS, T. J., and A. M. JESSOP, 1981, Heat flow in the
Garibaldi Volcanic Belt, a possible Canadian geothermal
energy resource: Canadian Journal of Earth Sciences, v.
18, p. 366-375.
__ , A. M. JESSOP, and A. S. JUDGE, 1985, Heat flux
measurements in southwestern British Columbia: The
thermal consequences of plate tectonics: Canadian Jour-
nal of Earth Sciences, v. 22, p. 1262-1273.
__ , W. H. BENTKOWSKI, E. E. DAVIS, RD. HYNDMAN,
J. G. SOUTHER, andJ.A. WRIGHT, 1988, Subduction of the
Juan de Fuca Plate - Thermal consequences: Journal of
Geophysical Research, v. 93, p. 15207-15225.
LEWRY, J. F., 1981, Lower Proterozoic arc-microcontinent
collisional tectonics in the western Churchill Province:
Nature, v. 294, p. 69-72.
36 M.J.Drury
MAJOROWICZ, J. A., F. W. JONES, and A. M. JESSOP,
1986, Geothermics of the Williston Basin in Canada in
relation to hydrodynamics and hydrocarbon resources:
Geophysics, v. 51, p. 767-779.
MORGAN, P., 1983, Constraints on rift thermal processes
from heat flow and uplift: Tectonophysics, v. 94, p. 277-
298.
OEN, 1. S., 1987, Rift-related igneous activity and metallo-
genesis in SW Bergslagen, Sweden: Precambrian Re-
search, v. 35, p. 367-382.
PAKTUNC, A. D., and A. J. BAER, 1986, Geothermo-
barometry of the northwestern margin of the Superior
Province - Implications for its tectonic development:
Journal of Geology, v. 94, p. 281-394.
PEACOCK, S. M., 1987, Thermal effects of metamorphic
fluids in subduction zones: Geology, v. 15, p. 1057-1060.
ROY, RF., D. D. BLACKWELL, and F. BIRCH, 1968, Heat
generation of plutonic rocks and heat flow provinces:
Earth and Planetary Science Letters, v. 5, p. 1-12.
SALTUS, R W., and A. H. LACHENBRUCH, 1987, Two-
Dimensional Finite Element Models of the Variation
of Heat Flow with Depth Caused by Refraction at a
Low Conductivity Graben: U.S. Geological Survey, Open
File Report 87-618, 11 p.
STEINHART, J. S., S. R HART, and T. J. SMITH, 1968, Heat
flow: Carnegie Institute Year Book, p. 360-367.
THOMAS, M. D., 1983, Tectonic significance of paired
gravity anomalies in the southern and central Appala-
chians; pp. 113-124 in R D. Hatcher, H. Williams, and I.
Zietz (eds.), Contributions to the Tectonics and Geo-
physics of Mountain Chains: Geological Society of Am-
erica, Memoir 158, 223 p.
WANG, JI-YANG, and JI-AN WANG, 1988, Thermal struc-
ture of the crust and upper mantle of the Liaohe Rift basin,
north China: Tectonophysics, v.145, p.293-304.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Geochronological studies of fault-related rocks
PAULD. FULLAGAR
Department of Geology, University of North Carolina, Chapel Hill, North Carolina 27599-3315, USA
(received October 27,1988; revision accepted May 11, 1989)
ABSTRACT
Review of geochronological studies of mylonites and cataclasites indicates that results have been
equivocal. For example, mylonites from the Brevard fault zone in the southern Appalachians have been
sampled at different locations and dates have been obtained using several techniques: 460 Ma, U-Pb zircon;
384,381,348,273 Ma, Rb-Sr whole-rock isochron. This fault zone has a complex history and records several
episodes of ductile and brittle deformation and metamorphism. Investigators concluded that several of these
dates are times of specific events (e.g., mylonitization). It also was suggested that some Rb-Sr dates may be
spurious, perhaps due to the mixing of two components; this could result in mixing lines rather than isochrons.
Interpretation of data from Brevard zone mylonite samples is discussed here, along with examples from other
shear zones.
Rb-Sr whole-rock systems may be reset (i.e., undergo isotopic re-equilibration or homogenization) during
mylonitization, especially if the process is accompanied by fluid transport. The scale of re-equilibration can
vary: several hundred meters or even considerably more, but often several meters or less. Sampling for age
determinations should be restricted to those rocks which are most likely to have undergone isotopic re-
equilibration. Petrographic and microprobe analyses may indicate whether or not equilibration is likely to
have occurred, as well as the scale ofresetting. Survival ofporphyroclasts from the original rock suggests lack
of equilibration.
U-Pb systems of zircon generally seem to be more resistant to resetting than Rb-Sr whole-rock systems.
In cases where mylonitization did reset the zircon age, or new zircon formed, analyses of single crystals would
yield the time of this event. K-Ar and Rb-Sr analyses ofpseudotachylites also may provide meaningful ages.
INTRODUCTION
Fault zones commonly are a major aspect of
orogenic belts. Determining the time offormation of
mylonites and other fault-related rocks would add
useful time constraints to our information about the
evolution of orogenic belts. This paper reviews past
efforts to date rocks from fault zones, presents new
Rb-Sr data, and suggests ways to enhance the
likelihood of obtaining useful dates for these rocks.
The terminology used here for fault-related rocks is
that suggested by Wise and others (1984).
Most attempts to date the time of formation of
pre-Pleistocene fault zone rocks have used multiple-
sample Rb-Sr whole-rock techniques. For these
attempts to be successful, all of the fault-related
rocks sampled must have attained the same Sr iso-
topic composition during faulting. This isotopic re-
equilibration would reset the age of the rocks at the
time of faulting. Several investigators have tried to
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 37-50. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
37
38
P. D. Fullagar
date fault-zone rocks by V-Pb analyses of zircon. If
the zircon formed in response to the faulting event,
or if existing zircon lost all radiogenic Pb during
formation of the fault-related rocks, the V-Pb age of
the zircon would be set, or reset, to zero. If different
fractions of pre-existing zircon lost different
amounts of radiogenic Pb in response to the faulting
episode, evaluation of the V-Pb data using an
episodic Pb-Ioss model should yield the age of this
event. Rb-Sr and K-Ar studies of other mineral
phases (e.g., micas) can yield useful age information,
but these usually provide only a minimum age for
the faulting episode because deep-seated, fault-
produced rocks are at elevated temperatures; min-
erals in such rocks will not accumulate daughter
products until the rocks have cooled due to uplift
and erosion. The extremely fine grain size of many
mylonitic rocks, often considerably less than 50
microns, may make separation of mineral phases
difficult.
ISOTOPIC DATING OF FAULT ZONE ROCKS
Rb-Sr Dates of Mylonitic Rocks
Early Efforts to Date Fault Zones. Dietrich
and others (1969) investigated the effects offaulting
on isotopic systems by obtaining Rb-Sr (and K-Ar)
dates on Precambrian gneisses from the Blue Ridge
province of Virginia and on Cambrian Rome
Formation (shaly siltstone) from the adjacent Valley
and Ridge Province. Samples were obtained from
both within and outside fault zones. In retrospect,
too few samples were analyzed from the Grayson
Gneiss outside the fault zone to satisfactorily
establish the isotopic age of these rocks; subsequent
Rb-Sr analyses by Fullagar and Odom (1973) in-
dicated that this gneiss has an age of "" ll50 Ma.
[Note: All ages and dates reported in this paper
have been calculated or recalculated using the now
conventional decay constants (Steiger and Jager,
1977).] The mylonitic fault zone samples of the
Precambrian gneisses yielded approximately mid-
Paleozoic Rb-Sr and K-Ar model dates. [A model
date means that values were assumed for the initial
87Sr and 4Ar contents of the samples.] Dates for the
Rome Formation samples were somewhat incon-
sistent, although the fault-zone samples again were
younger than the unsheared rocks; model dates for
these sheared samples also were mid-Paleozoic.
Because analyses of the mylonitic rocks gave geo-
logically reasonable dates, Dietrich and others
(1969) concluded that the isotopic ratios may be re-
equilibrated in response to fault zone processes so
that the time of the faulting might be determined by
analyzing appropriate samples.
Rb-Sr systematics of the Roffna gneiss in
Switzerland were studied by Hanson and others
(1969). Although zircon 207Pb_
206
Pb dates indicated
a significantly older time of crystallization, most of
the Rb-Sr whole-rock data suggested a period of
recrystallization and resetting of the Rb-Sr data at
"" ll3 Ma. Hanson and others suggested that the
Roffna gneiss underwent cataclasis and recrystal-
lization of the mylonitic matrix at this time. They
also suggested that the variable degree of re-
crystallization of the Roffna gneiss was controlled
by the amounts of water available in different parts
of the system; where there was adequate water, the
water combined with MgO to help form schistose
phengitic gneiss at "" ll3 Ma.
Mylonitic rocks from the boundaries of a Pre-
cambrian greenstone belt in southeastern Manitoba
have a Rb-Sr whole-rock isochron date of 2295 100
Ma, according to Turek and Peterman (1971). They
interpreted the date as a minimum age for the
development of the mylonite zones, as subsequent
fault movement might reset the Rb-Sr isotopic clock.
They noted that the relatively large uncertainty for
their mylonite date could reflect differences in the
isotopic compositions of the parent rocks. Given
that the samples apparently were collected over a
distance of "" 100 km, and that field evidence sug-
gests the mylonites were derived from sediments,
granite, and quartz diorite, it certainly would be
possible for the initial isotopic compositions of the
samples to differ.
Rb-Sr isotopic systematics of a Precambrian
mylonite-bearing shear zone in the Front Range of
Colorado were investigated by Abbott (1972), who
analyzed suites of mylonitic samples collected on
three different scales: 10 cm-wide layer, 3 m-wide
outcrop, and over a distance of about 12 km. The
first group of samples probably was derived from
biotite schist, the others from biotite schist, peg-
matite, quartz monzonite, and perhaps amphibolite.
Most samples have had their Rb-Sr dates lowered
relative to "" 1700-1800 and 1400 Ma dates for rocks
considered to be unsheared equivalents of the mylo-
nitic rocks. Five of six samples from the 10 cm-wide
mylonite layer plot on a 1000 Ma isochron; the
anomalous sample, with a model age of "" 1700 Ma,
was obtained immediately adjacent to relatively
Geochronological studies of fault-related rocks 39
un sheared rocks; Abbott suggested that the isotopic
homogenization was effective only over short
distances. The four samples from the 3 m-wide out-
crop define a 1200 Ma isochron. The samples col-
lected over distances of several kilometers give
inconsistent results, but values plot on or very near
a 1200 Ma isochron. Abbott (1972) concluded that
nearly complete isotopic homogenization occurred
over distances of at least several meters, indicating
that Rb-Sr whole-rock studies may be used to obtain
the age of shear zones. Abbott noted that additional
studies were needed to determine the conditions
under which Rb and Sr are mobilized in fault zones
and the extent to which transport occurs.
Southern Appalachians of the USA. The
Brevard fault zone in the southern Appalachians
(Figure 1) is a linear belt of cataclastic and mylo-
nitic rocks that likely records Taconic and/or Aca-
dian ductile metamorphic events, as well as several
Alleghanian ductile and brittle events (Hatcher and
Odom, 1980; Hatcher, 1981). The Brevard fault
zone rocks are well-exposed near Rosman, North
Carolina (Horton and Butler, 1986), and many
samples for geochronological studies have come
from this location.
In the Rosman area, the Henderson Gneiss is
the dominant lithologic unit in the Inner Piedmont.
Figure 1. Generalized map of the
geological belts of the southern Ap-
palachians: GA, Georgia; SC, South
Carolina; NC, North Carolina; TN,
Tennessee; VA, Virginia; KY, Ken-
tucky; WVA, West Virginia; CSB,
Carolina slate belt; RB, Raleigh belt;
GMW, Grandfather Mountain win-
dow; KMB, Kings Mountain belt;
BFZ, Brevard fault zone; LFF, Lin-
ville Falls fault; SRF, Stony Ridge
fault. Sample locations (open cir-
cles): R, Rosman, NC; OF, Old Fort,
NC; L, Linville Falls fault; S, Stony TN
Ridge fault. --'-------:,..,..--77
This augen orthogneiss has a Rb-Sr whole-rock
isochron date of 5109 Ma (data from Odom and
Fullagar, 1973; Bond, 1974); the analytical un-
certainties listed here and elsewhere in this paper
represent one standard deviation based on use of
York's (1969) regression calculation. The Hender-
son Gneiss becomes progressively mylonitized
toward the Brevard fault zone, where good ex-
posures of mylonites and ultramylonites derived
from the Henderson Gneiss are present. Six
samples of ultramylonite collected over a distance of
50 m from the exposures near Rosman provided a
Rb-Sr whole-rock isochron with a slope correspond-
ing to a date of 348 4 Ma and an initial 87Sr/86Sr of
0.7186 3 (Odom and Fullagar, 1973).
A goodness-of-fit test for isochron data points
was suggested by Brooks and others (1972). This
index parameter, mean square of weighted deviates
(MSWD), is obtained from the York (1969) regres-
sion normally used to obtain slopes (and thus ages)
of isochrons. If the observed scatter of data points is
solely due to analytical uncertainties, the MSWD
value should be s; 1.0. If> 1.0, a significant amount
of the scatter is due to geological factors. If the
calculated MSWD is significantly >2.5, the line is
best termed an errorchron, and the date obtained
from its slope may have no geologic significance.
The MSWD value for the Brevard fault zone
KILOMETERS
40 P. D. Fullagar
mylonites analyzed by Odom and Fullagar (1973) is
1.6. The 348 Ma date was interpreted as a time of
recrystallization and tectonic activity along the
Brevard fault zone. This interpretation did not
preclude earlier or later times of recrystallization
along this fault zone.
As part of a study to characterize the changes in
the Henderson Gneiss as it progresses from augen
gneiss to ultramylonite, three additional samples of
mylonite from the general Rosman area were
analyzed for Rb and Sr isotopic compositions,
together with three mylonite samples collected 34
km northeast of Rosman (Bond and Fullagar, 1973;
Bond, 1974). Combining data for these samples
yielded a date of 381 11 Ma (MSWD = 1.5). How-
ever, some of these samples contain a significant
volume of megacrysts or protolith that did not
recrystallize. For this reason, and because the
samples are from two widely separated locations
and thus could have had different initial isotopic
compositions, it seems best to place little if any
reliance on the 381 Ma date.
From the above-mentioned outcrop of Brevard
fault zone mylonite near Rosman, North Carolina,
Sinha and others (1986) obtained a block of rock
= 25 cm on each side. This block was cut into ten
2.5-cm-thick layers, parallel to tectonic layering,
and each layer was cut into three equal portions.
The 30 samples obtained varied from protomylonite
to ultramylonite. Detailed mineralogical and
chemical studies indicated that the ultramylonite
zones formed in response to focused fluid flow, which
resulted in retrogressive mineral assemblages and
bulk chemical changes. Sinha and others (1988)
have obtained Rb and Sr isotopic data for the 30
samples they obtained from the Rosman exposure of
the Brevard zone. Six of these samples gave highly
anomalous values; the remaining 24 samples
yielded a Rb-Sr date of = 425 Ma, but the MSWD
value apparently is = 17. [This MSWD value, along
with those that follow, was calculated from the Rb-
Sr data given by Sinha and others, assuming
analytical errors of 1 % and 0.025% for 87Rb/86Sr and
87Sr/86Sr, respectively.] The high MSWD value re-
flects the considerable scatter shown by the Rb-Sr
data on an isochron plot, which probably is due to
the different degrees of mylonitic recrystallization
of the 24 samples. The 425 Ma date for all 24
samples is considered to have little if any signi-
ficance.
The seven completely retrogressed samples
(ultramylonites) analyzed by Sinha and others
(1988) gave a Rb-Sr isochron date of 273 Ma with an
MSWD of 3.3. These data points are sufficiently
colinear that the date could represent a time of
isotopic homogenization of the ultramylonites, as
was concluded by Sinha and others. [The MSWD
value could be < 3.3 if actual analytical errors are
larger than those assumed above.]
The remaining 17 samples analyzed by Sinha
and others yielded a Rb-Sr date of = 380 Ma, with
an MSWD value of 3.7. They attributed this date,
plus the 348 Ma date obtained by Odom and
Fullagar (1973), to the formation of mixing lines
between the Henderson Gneiss protolith (510 Ma)
and new minerals that formed during the 273 Ma
recrystallization. This mixing line interpretation
seems plausible for the samples with = 380 Ma
dates that were analyzed by Sinha and others (1988)
and Bond (1974). Most of these samples contain a
significant percentage of unrecrystallized or partly
recrystallized protolith (megacrysts of K-feldspar-
plagioclase-quartz), and the scatter of data points
only approximates an isochron. However, the very
good fit of data points for the 348 Ma isochron places
constraints on mixing models. For example, Figure
2 shows the consequence of mixing a homogeneous
original rock and a new component. Different
amounts of mixing can lead to a linear array of
points and an apparent age older or younger than
the age of the original rock, depending on the
specific composition of the new component. As the
Henderson Gneiss protolith and the mylonites have
quite variable Rb/Sr ratios, it seems unlikely that
the simple type of mixing illustrated by Figure 2
would occur. A more realistic mixing model is
shown in Figure 3, where both the original rock and
the new component vary in composition. As
suggested in Figure 3, mixing might result in an
array of data points that suggests an isochron, but
the data points would scatter unless mixing was by
some specific and different percent for each sample.
This type of mixing seems unlikely, and thus the
348 Ma isochron should be considered to have
possible geological significance.
Another well-exposed portion of the Brevard
faul t zone occurs = 80 km northeast of Rosman, near
Old Fort, North Carolina (Figure 1). Field relation-
ships, petrography, and Rb-Sr systematics have
been studied for the rocks from this location (McGill
and others, 1980). Rb-Sr data for analyzed samples
are given in Table 1. Regression calculations
(York, 1969) for all data listed in this table use
analytical uncertainties of 0.5% and 0.025% for
Geochronological studies offault-related rocks 41
en
<0
0.74 r------------------,
MIXING OF SINGLE COMPONENT
WITH HOMOGENEOUS ROCK SAMPLES
0.73
0.72
NEW
COMPONENT
en
.....
eo
0.71
'\ ORIGINAL
ROCK
MIXING LINE
O. 70 _ __'_----'
o 2 3 4 5
87Rbl86Sr
Figure 2. Simple mixing line between rock and new
component. A, B, and C represent intermediate compo-
sitions; the actual fraction of original rock and new
component depends on their Rb and Sr concentrations. If
the y-intercept for the mixing line is higher than the initial
87Sr/86Sr for the original rock, the date obtained will be
younger than the actual age of the rock. If the intercept is
lower, the date will be too old.
en
0.78,-----------------,
0.76
MIXING OF MULTIPLE COMPONENTS
WITH SAMPLES FROM ONE UNIT
MIXING RESULTS
IN DATA POINTS
WITHIN POLYGON
0.74
en
r-
eo
0.72
0.70 '----'----'-----'----'-----'----'
o
87Rbl86Sr
Figure 3. Mixing of samples that define an isochron with
components of variable composition. Depending on mixing
proportions, the data points could plot anywhere within
the indicated polygon. (The dashed lines connect the end-
members: the original rock and the mixing component.)
Although fortuitous, the data points could result in a
mixing line (not shown) which would have a steeper slope
than the original isochron and thus would yield a date
older than that of the original rock.
Table 1. Rb-Sr data.
Rb Sr
87Sr 87Rb
-- --
Sample (ppm) (ppm) 86Sr 86Sr
Old
Fort
1472 62.3 252.4 0.71383 0.715
1473 45.8 329.8 0.71066 0.402
1474 55.9 336.5 0.71038 0.481
1475 47.7 349.1 0.70942 0.396
1476 45.3 397.8 0.70869 0.33
1477 48.3 342.1 0.71069 0.409
1466 87.7 410.6 0.70911 0.618
1467 98.8 361.8 0.71012 0.791
1468 96.6 360.2 0.71006 0.777
1469 120.3 247.3 0.71146 1.409
1470 168.9 211.1 0.71694 2.318
1471 168.1 144.5 0.72319 3.371
1796 129.6 299.7 0.71136 1.252
1797 206.4 86.3 0.74207 6.95
1798 184.9 131.9 0.72618 4.064
1799 160.7 202.7 0.71638 2.298
Secret
Peak
141-17 347.4 89.6 0.74321 11.26
141-18 258.5 93.9 0.74201 7.991
141-19 286.2 84.9 0.74243 9.792
141-20 339.1 80.5 0.74392 12.23
141-21 322.7 83.8 0.74385 11.19
175-8-1 323.9 79.1 0.74278 11.89
175-8-2 320.4 75.0 0.74314 12.41
5174-2 260.0 99.9 0.74244 7.56
5174-3 240.4 113.0 0.74186 6.176
174-4A 238.8 99.7 0.74146 6.954
175-57N 256.6 150.1 0.73905 4.963
175-57F 208.4 125.5 0.72824 4.814
175-57G 198.9 124.9 0.72810 4.615
175-57H 194.5 119.2 0.72924 4.733
175-57J 180.0 101.5 0.73202 5.142
175-57L 213.9 92.8 0.73849 6.691
87Rb/86Sr and 37Sr/86Sr, respectively. Samples 1472-
1477 represent gneisses considered to be older
(Grenville-age) basement; these gneisses show some
recrystallization in the form of a fine matrix. These
samples do not define a Rb-Sr isochron but cluster
about a 940 Ma errorchron (Figure 4). The lack of
an isochron is attributed to the partial and variable
recrystallization of the samples in response to move-
ment along the Brevard fault zone. Other samples
from the same location apparently are derived from
the same basement gneiss and exhibit differing
degrees of mylonitization. Three mylonitic gneiss
samples (1466-1468) exhibit a slightly greater de-
gree of recrystallization than do the samples of older
basement mentioned above. Rb-Sr data for these
42
P. D. Fullagar
0.730
BREVARD FAULT ZONE
OLD FORT, NC
0.725
0
0.720
GNEISS
U,
940MA

'"
<X>
0.715

r--
<X>
0.710
IR = 0.70305 34
0.705

T=3804MA
IR = 0.70430 19
0.700
0 2
87Rb/86Sr
Figure 4. Rb-Sr isotopic data from the Brevard fault zone,
Old Fort, NC. The open and filled circles represent two
suites of orthomylonites; one sample (1797) of the latter
g,r,0up is not represented because of the scale used:
7Rb/86Sr= 6.95; 87Sr/86Sr= 0.74207; T=time or date; IR=
initial 87 Sr/
86
Sr.
protomylonite samples plot (Figure 4) to the right of
the errorchron for the older basement gneiss
samples and suggest that their date has been reset
or at least partially reset to =400 Ma. Seven
samples exhibiting a still greater degree of mylo-
nitization, orthomylonites, yield a date of 3843
Ma, with initial 87Sd6Sr (IR) of 0.7041515 and
MSWD of 4.4. This errorchron is not shown on Fig-
ure 4, although six of the seven data points are
shown (one data point plots off the figure). These
orthomylonites are from two locations (that for
samples 1469-1471, and for samples 1796-1799)
separated by = 40 m. Dates and initial 87Sr/86Sr
ratios for these two sample groups are shown on
Figure 4. The difference in dates, 420 Ma versus
380 Ma, possibly could be real. The MSWD values
are 0.04 and 3.3, respectively. However, the small
number of samples (3 and 4) for each subgroup of
orthomylonites means that confidence in the dates
cannot be very great.
The Old Fort mylonite data present another
problem: the apparent initial ratios of 0.703-0.704
are quite low. Data available (Fullagar and Odom,
1973) indicate that Grenville-age basement gneisses
in the Blue Ridge often have initial ratios in this
range. However, =400 Ma, these rocks generally
would have had 87Sr/86Sr ratios significantly greater
0.730
MIXING WITH SINGLE COMPONENT:
SIMILAR PERCENTAGES D
0720
C
U,
'"
ORIGINAL
<X>


B
r--
<X>
............ :::::t{X
0.710
A
' MIXING LINE
0.700
0 2
87Rb/86Sr
Figure 5. Mixing of samples that define an isochron with
a single component. Similar percentages of mixing for
each sample have produced a line or array with a slope
similar to that of the isochron. The mixing line provides
approximately the correct age but the initial 87Sr/86Sr is
anomalously low.
than 0.710. In order for the mylonites to have
formed from typical basement gneiss at = 400 Ma,
and have initial ratios of 0.703-0.704, essentially all
radiogenic 87Sr would have had to be removed from
the recrystallizing rocks, and they must have
reacted only with fluids of low Sr content or with low
87Sr/86Sr ratios (probably no greater than 0.704).
Removal of all radiogenic 87Sr seems unlikely, and it
is especially doubtful that fluids derived from a
terrane characterized by rocks with ages ;::: 1000 Ma
would have such low 87Sr/86Sr ratios. Accordingly, it
is suggested that the apparent isochrons for the
mylonites might be mixing lines. Two types of
mixing lines are shown in Figures 2 and 3; an
additional way to obtain a mixing line is shown in
Figure 5; numerous others could be presented as
well. If similar percentages of mixing occur between
each sample (A-D) and the new component (X), the
isochron points would shift so as to cluster about a
line parallel to the original isochron (i.e., same date)
but with an anomalously low initial 87Sr/86Sr ratio.
Although there is no evidence that this has hap-
pened to the Old Fort orthomylonites, it is a possible
but not unique explanation for the low initial ratios.
Bryant's (1966) study ofphyllonite formation in
the Grandfather Mountain window area suggests
another type of mixing that could explain the data
Geochronological studies of fault-related rocks 43
for the Old Fort mylonitic rocks. The phyllonites
gained K and lost Ca during mylonitization; in
addition, presumably Rb was added and Sr
subtracted. If, prior to mylonitization, a portion of
basement gneiss had a uniform and very low Rb/Sr
ratio, addition of variable but significant amounts of
Rb would have the effect of resetting the Rb-Sr
system. Each sample would plot on or close to a
horizontal line (T = 0), regardless of whether Sr was
removed from the samples. Chemical data (see
Table 1) suggest that the Old Fort mylonitic rocks
both gained Rb and and lost Sr.
The Linville Falls fault is a major fault in the
Blue Ridge Province. It bounds the Grandfather
Mountain window (see Figure 1) and separates
medium-grade metamorphic rocks of the Blue Ridge
thrust sheet(s) from the lower grade rocks of the
window. The geology of Linville Falls has been
described by Hatcher and Butler (1986). Below the
Linville Falls fault, Lower Cambrian Chilhowee
Group sandstones (Erwin Quartzite) are exposed.
The fault zone is up to 45 cm thick, consisting of
finely laminated and crenelated ultramylonite. The
ultramylonite appears to have been derived from
the overlying "Cranberry Gneiss" (Elk Park
Plutonic Group). In the vicinity of the fault, the
basement gneiss is mylonitic (protomylonite) and,
like the fault zone ultramylonite, exhibits green-
schist-facies mineral assemblages.
Rb-Sr isotopic systematics for the mylonitic
rocks at Linville Falls were studied by Van Camp
(1982) and reported by Van Camp and Fullagar
(1982). Mylonitic gneiss was collected at about 10
cm intervals out to 150 cm from the ultramylonite,
and one additional sample was collected 20 m from
the gneiss-ultramylonite contact. Six samples of
ultramylonite were collected at approximately
regular intervals over a distance of 23 cm. Mylo-
nitic "Cranberry Gneiss" samples were analyzed for
Rb and Sr isotopic compositions. The data scatter
significantly, as shown on Figure 6 and by the
MSWD parameter of 165; thus, these data have
little or no age significance. Although unmylo-
nitized "Cranberry Gneiss" samples from near
Linville Falls have not been analyzed to establish
their Rb-Sr date, over 30 samples have been
analyzed from locations 10-to-40 km distant. These
samples yield a date of 1018 19 Ma (Full agar and
Bartholomew, 1983), and this is considered a
reasonable estimate for the age of the basement
gneiss samples at Linville Falls - before they were
affected by mylonitization. A 1018 Ma reference
1.00
0.90
0.80
LINVILLE FALLS
AND
CRANBERRY GNEISS
MYLONITE
302 MA
MYLONITIC GNEISS
10 20 30 40
87Rb/86Sr
50
Figure 6. Rb-Sr isotopic data from the Linville Falls fault
and related rocks. The Cranberry Gneiss isochron is based
on data not shown (Fullagar and Bartholomew, 1983). The
solid symbols represent ultramylonites from Linville Falls,
and they define the 302 Ma isochron; data for mylonitic
gneiss adjacent to the fault scatter about the isochron.
Data from Van Camp (1982).
isochron (no data points) is shown on Figure 6. The
six samples of ultramylonite are colinear and this
line also is shown on Figure 6. This isochron cor-
responds to an age of 3022 Ma, with an initial
87Sr/86Sr ratio of 0.7472 11 (MSWD = 1.4). The
linear distribution of data points does not require
that this 302 Ma date have geologic significance.
However, the colinear data points, the high degree
of recrystallization of this very fine-grained rock,
the near absence of megacrysts, and the limited area
that was sampled all indicate that the 302 Ma date
is the time of formation or recrystallization of the
ultramylonites as suggested by Van Camp and
Fullagar (1982). If our interpretations are correct,
progressive mylonitization resulted in an increase
in Rb/Sr ratio as shown in Figure 6; this increase in
Rb/Sr is the consequence of both an increase in Rb
and a decrease in Sr with increasing mylonitization.
The same pattern is shown by the Old Fort mylo-
nitic rocks discussed above (see Figure 5, Table 1).
Details of sampling and geochemical studies at
Linville Falls are given by Van Camp (1982).
Several additional Rb-Sr studies of mylonitic
rocks from the southern Appalachians also suggest
resetting of the Rb-Sr system in response to ductile
deformation along fault zones. Russell (1978) and
44 P. D. Fullagar
Hatcher and Odom (1980) reported a Rb-Sr whole-
rock isochron date of 368 9 Ma for mylonites
derived from the Precambrian Fort Mountain gneiss
in the Blue Ridge province of Georgia. These mylo-
nites are exposed in faults above the Great Smoky
fault at the base of the Blue Ridge thrust sheet . .In
the southernmost exposed Appalachians of Alabama
and Georgia, mylonites (phyllonites) from the Bart-
letts Ferry fault zone at the southern boundary of
the Pine Mountain belt have a Rb-Sr date of 375
11 Ma, whereas Rb-Sr data for inter layered mylo-
nitic gneiss suggest a date of "" 600 Ma (Russell,
1985). The mylonitic gneiss did not have its age
reset, and perhaps the reason for this is indicated by
a lack of hydrous phases.
Other Rb-Sr Studies. Recent studies of suites
of mylonitic rocks from other parts of the world have
yielded Rb-Sr dates generally considered to be the
time of mylonitization. Baer (1980) reported that
mylonites from a shear zone in the Timber Lake
pluton of Ontario yield a Rb-Sr isochron of "" 1000
Ma, younger than the possibly 1300 Ma emplace-
ment age for the pluton. In Labrador, the Island
Harbour granites are cut by the Kanairiktok shear
zone; Rb-Sr analyses of these granite samples plus
mylonites from the shear zone resulted in isochron
dates of "" 1800 Ma for both suites of samples (Grant
and others, 1983). They interpreted the results as
indicating that the shearing and granite emplace-
ment were closely related in time. It also could be
possible that shearing in this case had little or no
effect on the Rb-Sr systems.
Basement gneisses in the Tiinniis Augen Gneiss
Nappe of the southern Swedish Caledonides ori-
ginally were granodiorites emplaced at "" 1685 Ma
(Claesson, 1980). Near thrust contacts, the gneisses
have been transformed to protomylonite, and at
contacts the rock is ultramylonite. Samples ofmylo-
nite from one 10-m-wide outcrop were analyzed for
Rb-Sr isotopic composition. The data points scatter
but suggest a Caledonian date of "" 485 Ma (Claes-
son, 1980). Claesson also analyzed a single block of
mylonite (probably ultramylonite, based on infor-
mation given), which he first cut into 1-cm-thick
slices. Samples prepared from each of the 11 slices
were analyzed for Rb-Sr isotopic compositions. Most
of these samples defined an isochron of 485 50 Ma,
which was interpreted as probably being the time of
development of the mylonite. Claesson noted that
Sr isotopic equilibration was not complete, even
over a distance of only 10 cm. He also suggested
that Rb-Sr systems that are almost but not com-
pletely reset will yield scattered data which do not
define a good isochron but which suggest a date that
is somewhat greater than the time of the incomplete
isotopic homogenization. Claesson reviewed the
relative immobility of Sr through solids and sug-
gested that a fluid phase along grain boundaries
may be critical for Sr isotopic equilibration to occur,
even on the limited scale observed in his study.
A preliminary Rb-Sr study by Leake and others
(1983) on mylonitized felsic volcanic rocks from near
the Mannin thrust in Connemara, Ireland, resulted
in a 460 Ma date. It was noted that this could be the
time of original igneous crystallization or that of
mylonitization. The Rb-Sr data were re-evaluated
by Kennan and Murphy (1987; 1988); they sugges-
ted that the data defined two parallel isochrons and
that mylonitization actually occurred at 426 Ma.
This interpretation was questioned by Tanner and
Dempster (1988), who said that the preliminary
data were best interpreted as yielding a date of
""460Ma.
These different views seem to illustrate the
difficulty of interpreting data from mylonite zones,
especially those that have undergone a complex
history. In another example of results that have
been difficult to decipher, LaTour and Fullagar
(1986) obtained Rb-Sr data on samples from several
ultramylonite zones and sheared granites from the
Grenville front near Coniston, Ontario. Most of the
data points plot on or close to one of several 1600 Ma
errorchrons. These data were interpreted as indi-
cating that mylonitization of these rocks occurred at
"" 1600 Ma. However, the scatter of data points
makes this interpretation uncertain.
Additional studies on Rb-Sr isotopic systema-
tics of deformed rocks are pertinent, even though
the affected rocks are not described as mylonitic.
These studies by Hickman and Glassley (1984),
Page and Bell (1986), and Etheridge and Cooper
(1981) are discussed below. Each of these studies
emphasized the importance of fluids in affecting
isotopic re-equilibration or near re-equilibration.
Hickman and Glassley (1984) analyzed 130
samples of gneisses and granites (but apparently not
mylonites) from the Nordre Stromfjord shear zone in
West Greenland and found that for samples most
affected by fluid-rock interaction and recrystalliza-
tion, the Rb-Sr system was reset to approximately
the time of shear zone formation ( "" 1850 Ma). This
resetting was on the scale of tens of meters.
Samples least affected by fluid-rock interaction in
Geochronological studies offault-related rocks 45
the Proterozoic shear zone yielded Rb-Sr data which
still indicate an Archean age. Hickman and Glass-
ley pointed out that the scale of sampling can affect
the dates obtained. Thin slices of layered gneiss
generally exhibited Proterozoic dates, but they
noted some scatter of data, possibly due to lack of
complete isotopic equilibration or due to subsequent
diffusive exchange during deformation or cooling.
Poor sampling strategy could result in mixing lines
or arrays that suggest a date either older or younger
than the time of formation of the shear zone (Hick-
man and Glassley, 1984).
Investigations of the Sybella Granite of north-
ern Australia by Page and Bell (1986) resulted in
conclusions similar to those reached by Hickman
and Glassley (1984). Deformation of the Sybella
Granite locally produced well foliated and strongly
lineated gneiss; precise and consistent Rb-Sr whole-
rock dates were obtained for samples with fabrics
characterizing each separate deformational event
(Page and Bell, 1986). They suggested that isotopic
re-equilibration occurred over distances of hundreds
and perhaps thousands of meters, and that this
homogenization was associated with significant
migration of fluids through zones of rock associated
with a particular deformation.
A different view is offered by Etheridge and
Cooper (1981). Although their "retrograde schist"
from a shear zone at Broken Hill, Australia, may
not be mylonite, their discussion of mixing processes
in shear zones is very pertinent. They obtained Rb-
Sr isotopic data plus major and trace element chemi-
cal data for samples within and adjacent to a 10-m-
wide shear zone in the Potosi Gneiss. They con-
cluded that large-scale circulation of fluids intro-
duced radiogenic 87Sr to the shear zone without sig-
nificantly altering Rb/Sr ratios, and that mixing
between fluid and rock samples produced a band of
data points with a negative slope. They attributed
the mixing array to incomplete isotopic exchange
between the fluid and rock samples. Unless, by
chance, mixing results in a horizontal array of data
points, the date obtained would be spurious.
Although not stated by Etheridge and Cooper,
depending on the Rb/Sr and 87Sr/86Sr values of fluid
and rock, mixing lines with positive slopes also
could be obtained. A mixing line with negative
slope would result in a date younger than the actual
time of mixing; a positive slope would give a date
that is too old. Etheridge and Cooper (1981)
concluded that Rb/Sr dates on "this type of mylonite
zone are likely to be meaningless."
Rb-Sr Dates ofCataclastic Rocks
Cataclasis, if carried to an extreme, could homo-
genize a rock with respect to both 87Rb/86Sr and
87Sr/86Sr. Subsequent analyses of samples of such a
rock would not yield an isochron as each data point
would be isotopically identical. If a fluid containing
Rb and Sr were added to this rock during or
immediately after cataclasis, a Rb-Sr mixing line
would result provided that samples incorporated
different proportions of Rb and Sr (i.e., developed
different 87Rb/86Sr). This line would result in a
spurious date unless the 87Sr/86Sr ratios for the fluid
and rock system were identical.
If cataclasis occurred it would seem more likely
that the rock would retain variations in 87Rb/86Sr.
As a result of extensive microfracturing of mineral
grains due to brecciation, and saturation of these
grains with fluid, radiogenic 87Sr might be mobile
enough for Sr isotopic homogenization to be achie-
ved, at least on a limited scale. In this case, homo-
genization would result in a horizontal distribution
of Rb-Sr data points, and subsequent analyses
should yield the time of homogenization. If only
partial isotopic equilibration occurred, data points
would scatter and any date probably would be
meaningless.
Few Rb-Sr isotopic studies have been done on
rocks clearly identified as cataclasites (see Wise and
others, 1984, for definition), undoubtedly because of
the likelihood of obtaining spurious results.
Fullagar and Butler (1980) analyzed silicified
microbreccia samples from the Stony Ridge fault
zone in North Carolina (see Figure 1). This fault
zone consists of three en echelon segments, with the
easternmost coinciding with the Dan River fault on
the edge of the Dan River Triassic basin. Samples
appear to have originated as feldspathic rocks that
have undergone multiple episodes of microbreccia-
tion and silicification; chemical analyses of three
samples indicate that they contain 90-95% Si0
2
.
The samples exhibit brittle deformation, although
some grains are sutured, indicating limited re-
crystallization. The Rb-Sr data are plotted in
Figure 7. The data scatter; however, the samples
can be divided into two groups. First, those data
points shown by solid symbols represent samples
that do not show evidence of weathering; the
samples contain fresh pyrite and show little or no
oxidation of iron-bearing minerals. The other
samples (open symbols) contain oxidized iron-
bearing minerals and often are stained by iron
46 P. D. Fullagar
0.76
0.74
0.72
STONY RIDGE FAULT
MICROBRECCIA
o
o
o
o
o
o
T=1803MA
IR = 0.7187 4
0.70
87Rb/86Sr
Figure 7. Rb-Sr isotopic data from the Stony Ridge fault,
NC (Fullagar and Butler, 1980). Open symbols represent
weathered samples; solid symbols are for fresh samples.
The isochron is based on 6 of the 7 unweathered samples.
T = time or date; IR = initial 87Sr/86Sr.
oxides. The 180 3 Ma date obtained for six of the
seven fresh samples (MSWD = 2.3) is what one
might expect if the age of these rocks was reset by
processes associated with Late Triassic-Middle
Jurassic faulting. In another study, a Rb-Sr iso-
chron was obtained for cataclasites derived from
rhyolites that are part of the 260 Ma(?) Klondike
Schist in the Yukon Territory (Metcalfe and Clark,
1983). They interpreted the 202 Ma date as the time
of initial cataclasis.
As part of a geochronological study of the Ruby
Mountains, Nevada, Rb-Sr isotopic analyses were
made on the Secret Peak gneiss plus cataclasites
derived from the gneiss. Preliminary Rb-Sr dates
for rocks from this region were reported by Snoke
and others (1984). Rb-Sr data for the Secret Peak
gneiss and associated cataclasites are listed in
Table 1 and plotted on Figure 8. The gneiss
(samples 141-17-21, 175-8-1 and -2, 5174-2 and -3)
has a Rb-Sr date of194 Ma. Two samples (174-4A,
175-57N) exhibit a fabric transitional between the
gneiss and the cataclasites (175-57F, -G, -R, -J, and-
L). These transitional samples contain highly frac-
tured feldspar grains and highly strained quartz
grains, plus some fine-grained matrix resulting
from microbrecciation. These two samples plot close
0.78
RUBY MOUNTAINS
0.76
SECRET PEAK GNEISS
19 MA
in

<D
0.74 0
in
.....
eo
0.72
'" CATACLASITES
350 MA
0.70
0 10 12
87Rb/86Sr
Figure 8. Rb-Sr isotopic data for the Secret Peak gneiss;
open circles are for samples transitional between gneiss
and cataclasite (x). The 19 Ma isochron is based on the
Secret Peak gneiss. Mixing of gneiss and a component that
would plot on the lower portion of the 350 Ma reference
line could result in an array of points such as that for the
cataclasites.
to, but slightly below, the 19 Ma isochron. The five
cataclasites show significant grain-size reduction
with essentially no large fragments remaining.
These samples plot close to a 350 Ma reference
isochron in Figure 8. Given that the data scatter
and that a 350 Ma age is not reasonable for these
rocks, the distribution of these data points is be-
lieved due to mixing of Secret Peak gneiss with a
component lower in Rb, higher in Sr, and with a
lower 87Sr/86Sr ratio than the gneiss. [Such a com-
ponent would plot on or near the lower part of the
350 Ma reference line in Figure 8.] This component
could have been introduced by fluids moving
through the brecciated rocks.
U-Pb Zircon Dates of Mylonitic Rocks
There have not been many attempts to obtain
V-Pb dates for zircon from fault zones. Most
investigators who have done studies of this type, or
on zircon from deformed rocks not within fault
zones, have concluded that the V-Pb system in zir-
con is more resistant to resetting than is the Rb-Sr
system for whole-rock samples; the V-Pb system
often shows little or no indication of having been re-
Geochronological studies of fault-related rocks 47
set by the deformational processes (e.g., Chase and
others, 1983; Page and Bell, 1986). For example,
zircon from highly sheared plutons in a mylonite
zone on the margin of the Bitterroot dome, Idaho,
yielded U-Pb dates that are interpreted as reflecting
the time of crystallization of the plutons rather than
shearing (Chase and others, 1983).
Sinha and Glover (1978) separated and ana-
lyzed zircon from mylonitized Henderson Gneiss in
the Brevard fault zone near Rosman, North Caro-
lina (see Figure 1), to determine U-Pb dates. The
U-Pb isotopic systematics of zircon from this
location have been studied in additional detail by
Wayne and Sinha (1988). Zircon from the ultra-
mylonite apparently underwent total loss of radio-
genic Pb at =460 Ma. Microfracturing ofthe zircon
allowed fluids to enter and leach radiogenic Pb from
the radiation-damaged portions of the crystals. This
date is considered to be the time of both Taconic
prograde metamorphism and development of a my-
lonitic fabric (Sinha and Glover, 1978; Wayne and
Sinha, 1988).
Other Dating Methods
K-Ar (and 4Ar/39Ar) dates also may provide
information about the timing of fault movement.
For example, Adams (1981) reported K-Ar dates for
mylonites and pseudotachylite from the Alpine fault
zone of South Island, New Zealand. These dates
indicated that the K-Ar systems were affected by
significant fault zone movement at about 5-10 Ma.
Adams pointed out that mylonitization may not
expel all 4Ar from the original rocks. If this is the
case, the date obtained would be older than the time
of mylonitization. Also, he noted that rapid forma-
tion of pseudotachylite could result in trapping of
40 Ar, again leading to a date that is too old.
Fission-track dating procedures also may pro-
vide useful information about fault-zone histories.
Seward and Sibson (1985) and White and Green
(1986) obtained fission-track dates for pseudotachy-
lite and zircon, respectively, from the Alpine fault
zone. Only one of the five investigated pseudo-
tachylites contained sufficient undevitrified glass
for fission-track dating. A 0.43 Ma date was ob-
tained, but it is considered a minimum due to
annealing of spontaneous fission tracks (Sibson and
Seward (1985). The zircon fission-track dates are
= 80 Ma, 9 Ma, and < 5 Ma (White and Green,
1986). They interpreted these results as indicating
at least three periods of movement along this fault
zone, each of which resulted in uplift, erosion, and
cooling of the affected rocks.
CONCLUSIONS
Fault zones yield dates that are difficult to in-
terpret. This is not surprising considering the com-
plexity of processes which operate in shear zones,
including the possibility that events subsequent to
mylonitization might alter the isotopic systems
without producing easily recognized changes in the
fabric or mineral assemblage of the rock. More
dating has probably been done on rocks from the
Brevard fault zone near Rosman, North Carolina,
than on any other comparable volume of mylonitic
rocks. The U-Pb zircon and Rb-Sr whole-rock dates
are =460 Ma, 380 Ma, 350 Ma, and 270 Ma. In spite
of numerous studies, there is disagreement as to
what some of the numbers mean, although at least
several of the dates are believed to reflect times of
recrystallization in response to shearing.
Considering all dates obtained on many fault
zones, it is obvious that dating of fault movement
has met with only limited success. However, it
seems reasonable that success might be enhanced by
better sampling strategies, analysis of more
samples, and use of additional dating techniques.
The most consistent Rb-Sr whole-rock dates
have been obtained where samples had a common
proto lith and were collected over distances of less
than a few meters; this increases the likelihood that
samples achieved a common 87Sr/86Sr ratio during
mylonitization. It is important to try to ascertain
whether the isotopic system has been reset.
Interaction of large volumes of fluid with the shear
zone seems crucial if resetting is to occur (see Hick-
man and Glassley, 1984, for discussion). Petro-
graphy, trace and major element chemistry, and
mineral chemistry may provide evidence for exten-
sive fluid-rock interaction. For Rb-Sr dating, it is
desirable for recrystallization to have produced a
fine-grained ultramylonite free of megacrysts of the
protolith. At the least, megacrysts should be rare.
U-Pb dates of mylonitization could be obtained from
zircon samples that became differentially discordant
during the mylonitization event, providing all the
zircon crystals had a common initial age. Detailed
studies of zircon, such as the one by Wayne and
Sinha (1988), will help in recognizing which zircons
are likely to have lost all radiogenic Pb during mylo-
48
P. D. Fullagar
nitization and thus have had their isotopic systems
reset. These crystals, as well as zircon crystals that
formed during mylonitization, should yield the least
ambiguous U-Pb ages.
The complexity of most fault zones suggests
that interpretation of isotopic dates should not be
based on a small number of samples. Mixing lines
and meaningless dates can be obtained. Analyzing
a greater number of samples than might be studied
from a granitic pluton, for example, should increase
the likelihood of recognizing spurious results (e.g.,
scatter of data may become more obvious). Also,
standard statistical tests (Brooks and others, 1972)
should be used to carefully assess the fit of data
points to a Rb-Sr isochron or U-Pb chord. If data
scatter more than can be attributed to analytical
uncertainty, then the date may well be spurious. Of
course, it could be spurious with no scatter of the
data. If possible, dates should be verified by using
other dating techniques. At the very least, one can
consider whether the date is reasonable based on
geological constraints. Relatively new techniques
may prove very useful for materials from fault
zones. Single zircon crystal U-Pb dating techniques,
including ion microprobe mass spectrometers, will
make it easier to decipher multiple events that
affected a fault zone. Use of a laser microprobe with
4oAr/39Ar dating methods will make it possible to
analyze specific small volumes (e.g., pseudotachylite
grains). With small volumes, it is easier to obtain
material with an isotopic system that was com-
pletely reset during mylonitization.
ACKNOWLEDGMENTS
I am very grateful to numerous colleagues and
former students who have worked with me on por-
tions of these investigations, accompanied me in the
field, and provided me with much useful advice. I
especially thank P. A. Bond, J. R. Butler, T. E.
LaTour, J. W. Mies, A. L. Odom, A. W. Snoke, and S.
G. Van Camp. Helpful reviews of earlier versions of
this manuscript were provided by S. A. Goldberg, S.
A. Kish, A. L. Odom, and A. K. Sinha.
REFERENCES
ABBOTT, J. T., 1972, Rb-Sr study of isotopic redistribution
in a Precambrian mylonite-bearing shear zone, Northern
Front Range, Colorado: Geological Society of America
Bulletin, v. 83, p. 487-494.
ADAMS, C. J., 1981, Uplift rates and thermal structure, in
the Alpine fault zone and Alpine schists, Southern Alps,
New Zealand; pp. 211-222 in K. R. McClay and N. J. Price
(eds.J, Thrust and Nappe Tectonics: Geological Society
of London, Special Publication 9, 539 p.
BAER, A. J., 1980, Foliated and recrystallized granites
from the Timber Lake pluton, Ontario: Current Re-
search, Part C: Rb-Sr and U-Pb Isotopic Age Studies:
Geological Survey of Canada, Paper 80-1 C, p. 201-205.
BOND, P. A., 1974, A Sequence of Development for the
Henderson Augen Gneiss and Its Adjacent Cata-
clastic Rocks: MS thesis, University of North Carolina,
Chapel Hill, North Carolina, USA, 53 p.
__ , and P. D. FULLAGAR, 1973, Origin and age of the
Henderson Augen Gneiss and associated cataclastic rocks
in southwestern North Carolina [abstract]: Geological
Society of America, Abstracts with Programs, v. 6, p. 336.
BROOKS, C., S. R. HART, and T. L. WEND, 1972, Realistic
use of two-error regression treatments as applied to rubi-
dium-strontium data: Reviews in Geophysics and Space
Physics, v. 10, p. 551-577.
BRYANT, B., 1966, Formation ofphyllonites in the Grand-
father Mountain area, northwestern North Carolina: U.S.
Geological Survey, Professional Paper 550-D, p. D144-
D150.
CHASE, R. B., M. E. BICKFORD, and A. E. C. ARRUD, 1983,
Tectonic implications of Tertiary intrusion and shearing
within the Bitterroot dome, northeastern Idaho batholith:
Journal of Geology, v. 91, p. 462-470.
CLAESSON, S., 1980, A Rb-Sr isotopic study of granitoids
and related mylonites in the Tiinniis augen gneiss nappe,
southern Swedish Caledonides: Geologiska Foreningens i
Stockholm Forhandlinger, v. 102, p. 403-420.
DIETRICH, R. V., P. D. FULLAGAR, and M. L. BOTTINO,
1969, K-Ar and Rb-Sr dating of tectonic events in the Ap-
palachians of southwestern Virginia: Geological Society of
America Bulletin, v. 80, p. 307-314.
ETHERIDGE, M. A., and J. A. COOPER, 1981, Rb/Sr iso-
topic and geochemical evolution of a recrystallized shear
(mylonite) zone at Broken Hill: Contributions to Min-
eralogy and Petrology, v. 78, p. 74-84.
Geochronological studies of fault-related rocks 49
FULLAGAR, P. D., and A. L. ODOM, 1973, Geochronology of
Precambrian gneisses in the Blue Ridge province of north-
western North Carolina and adjacent parts of Virginia and
Tennessee: Geological Society of America Bulletin, v. 84, p.
3065-3080.
__ , and J. R. BUTLER, 1980, Radiometric dating in the
Sauratown Mountains area, North Carolina; Article II, pp.
B1-BlO in V. Price, Jr., P. A. Thayer, and W. A. Ranson
(eds.), Carolina Geological Society Guidebook: Savan-
nah River Laboratory, E. I. duPont de Nemours and Co.,
Aiken, South Carolina, USA, 198 p.
___ and M. J. BARTHOLOMEW, 1983, Rubidium-
strontium ages of the Watauga River, Cranberry, and
Crossing Knob Gneisses, northwestern North Carolina;
Article 2, pp. 1-29 in S. E. Lewis (ed.), Geologic In-
vestigations in the Blue Ridge of Northwestern North
Carolina: Carolina Geological Society Field Trip
Guidebook, 1983: North Carolina Geological Survey
Section, Raleigh, NC, USA, 236 p.
GRANT, N. K., F. R. VONER, M. S. MARZANO, M. H.
HICKMAN, and I. F. ERMANOVICS, 1983, A summary of
Rb-Sr isotope studies in the Archean Hopedale block and
the adjacent Proterozoic Makkovik Subprovince, Labrador;
pp. 127-134 in Current Research, Part B: Geological
Survey of Canada, Paper 83-1B, 456 p.
HANSON, G. N., M. GRUNENFELDER, and G. SOPTRAYA-
NOVA, 1969, The geochronology of a recrystallized tec-
tonite in Switzerland - the Roffna gneiss: Earth and
Planetary Science Letters, v. 5, p. 413-422.
HATCHER, R. D., Jr., 1981, Thrusts and nappes in the
North American Appalachian Orogen; pp. 491-499 in K. R.
McClay and N. J. Price (eds.), Thrust and Nappe Tec-
tonics: Geological Society of London, Special Publication
9,539p.
__ , and A. L. ODOM, 1980, Timing of thrusting in the
southern Appalachians, USA: Model for orogeny?: Jour-
nal of the Geological Society of London, v. 137, p. 321-327.
__ , and J. R. BUTLER, 1986, Linville Falls fault at
Linville Falls, North Carolina; pp. 229-230 in T. L. Neath-
ery (ed.), Centennial Field Guide, GSA Southeastern
Section, Volume 6: Geological Society of America, Boul-
der, Colorado, USA, 457 p.
HICKMAN, M. H., and GLASSLEY, W. E., 1984, The role of
metamorphic fluid transport in the Rb-Sr isotopic resetting
of shear zones: Evidence from Nordre Stromfjord, West
Greenland: Contributions to Mineralogy and Petrology, v.
87, p. 265-281.
HORTON, J. W., Jr., and J. R. BUTLER, 1986, The Brevard
fault zone at Rosman, Transylvania County, North Caro-
lina; pp. 251-256 in T. L. Neathery (ed.), Centennial Field
Guide, GSA Southeastern Section, Volume 6: Geologi-
cal Society of America, Boulder, Colorado, USA, 457 p.
KENNAN, P. S., and F. C. MURPHY, 1987, Tectonically
reset Rb-Sr system during Late Ordovician terrane assem-
bly in Iapetus, western Ireland: Geology, v. 15, p. 1155-
1158.
__ , 1988, Reply to comment on "Tectonically reset Rb-Sr
system during Late Ordovician terrane assembly in Iape-
tus, western Ireland": Geology, v.16, p. 763-764.
LATOUR, T. E., and P. D. FULLAGAR, 1986, Rb-Sr study of
mylonitic rocks at the Grenville front near Coniston,
Ontario: Some preliminary results; pp. 223-233 in J. M.
Moore, A. Davidson, and A. J. Baer (eds.), The Grenville
Province: Geological Association of Canada, Special
Paper 31, 358 p.
LEAKE, B. E., P. W. G. TANNER, D. SINGH, and A. N.
HALLIDAY, 1983, Major southward thrusting of the Dal-
radian rocks of Connemara, western Ireland: Nature, v.
305, p. 210-213.
McGILL, K. A., J. R. BUTLER, and P. D. FULLAGAR, 1980,
Tectonic history of the Brevard Zone and Blue Ridge east of
Asheville, North Carolina [abstract]: Geological Society of
America Abstracts with Programs, v.12, p. 480.
METCALFE, P., and G. S. CLARK, 1983, Rb-Sr whole-rock
age of the Klondike Schist, Yukon Territory: Canadian
Journal of Earth Sciences, v. 20, p. 886-891.
ODOM, A. L., and P. D. FULLAGAR, 1973, Geochronologic
and tectonic relationships between the Inner Piedmont,
Brevard Zone, and Blue Ridge belts, North Carolina:
American Journal of Science, v. 273-A, p. 133-149.
PAGE, R. W., and T. H. BELL, 1986, Isotopic and structural
responses of granite to successive deformation and meta-
morphism: Journal of Geology, v. 94, p. 365-379.
RUSSELL, G. S., 1978, U-Pb, Rb-Sr, and K-Ar Isotope
Studies Bearing on the Tectonic Development of the
Southernmost Appalachian Orogen, Alabama: PhD
dissertation, Florida State University, Tallahassee, Flori-
da, USA, 197 p.
__ , 1985, Reconnaissance geochronological investiga-
tions in the Phenix City Gneiss and Bartletts Ferry
mylonite zone; pp. 9-11 in S. A. Kish, T. B. Hanley, and S.
Schamel (eds.), Geology of the Southwestern Piedmont
of Georgia: Geological Society of America Guidebook,
98th Annual Meeting: Florida State University, Talla-
hassee, Florida, USA, 47 p.
50 P. D. Fullagar
SEWARD, D. and R. H. SIBSON, 1985, Fission-track age for
a pseudotachylite from the Alpine fault zone, New Zea-
land: New Zealand Journal of Geology and Geophysics, v.
28, p. 553-557.
SINHA, A. K., and L. GLOVER, 1978, U-Pb systematics of
zircons during dynamic metamorphism: Contributions to
Mineralogy and Petrology, v. 66, p. 305-310.
__ , D. A. HEWITI, and J. D. RIMSTIDT, 1986, Fluid
interaction and element mobility in the development of
ultramylonites: Geology, v. 14, p. 883-886.
__ , D. A. HEWITI, and J.D. RIMSTIDT, 1988, Meta-
morphic petrology and strontium isotope geochemistry
associated with the development of mylonites: An example
from the Brevard fault zone, North Carolina: American
Journal of Science, v. 288A, p.115-147.
SNOKE, A. W., R. D. DALLMEYER, and P. D. FULLAGAR,
1984, Superimposed Tertiary mylonitization of a Mesozoic
metamorphic terrane, Ruby Mountains-East Humboldt
Range, Nevada [abstractl: Geological Society of America
Abstracts with Programs, v. 16, p. 662.
STEIGER, R. H., and E. JAGER, 1977, Subcommission on
Geochronology: Convention on the use of decay constants
in geo- and cosmochronology: Earth and Planetary Science
Letters, v. 36, p. 359-362.
TANNER, G., and T. DEMPSTER, 1988, Comment on "Tec-
tonically reset Rb-Sr system during Late Ordovician ter-
rane assembly in Iapetus, western Ireland": Geology, v. 16,
p.762-763.
TUREK, A., and Z. E. PETERMAN, 1971, Advances in the
geochronology of the Rice Lake-Beresford Lake area,
southeastern Manitoba: Canadian Journal of Earth Sci-
ences,v. 8,p. 572-579.
VAN CAMP, S. G., 1982, Geochronology and Geochemis-
try of Cataclastic Rocks from the Linville Falls Fault,
North Carolina: MS thesis, University of North Carolina,
Chapel Hill, North Carolina, USA, 40 p.
__ , and P. D. FULLAGAR, 1982, Rb-Sr whole-rock ages
of mylonites from the Blue Ridge and Brevard Zone of
North Carolina [abstactl: Geological Society of America
Abstracts with Programs, v. 14, p. 92.
WAYNE, D. M., and A. K. SINHA, 1988, Physical and
chemical response of zircons to deformation: Contributions
to Mineralogy and Petrology: v. 98, p. 109-121.
WHITE, S. H., and P. F. GREEN, 1986, Tectonic develop-
ment of the Alpine fault zone, New Zealand: Geology, v.
14, p. 124-127.
WISE, D. U., D. E. DUNN, J. T. ENGELDER, P. A. GEISER,
R. D. HATCHER, S. A. KISH, A. L. ODOM and S. SCHAMEL,
1984, Fault-related rocks: Suggestions for terminology:
Geology, v. 12, p. 391-394.
YORK, D., 1969, Least squares fitting of a straight line
with correlated errors: Earth and Planetary Science
Letters, v. 5, p. 320-324.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
The importance of subduction-related margin-parallel shear
zones along transpressional convergent plate margins
VICKI L. HANSEN
Department of Geological Sciences, Southern Methodist University, Dallas, Texas 75275, USA
(received October 17, 1988; revision accepted April 28, 1989)
ABSTRACT
Structural and tectonic models of ancient convergent margins commonly assume simple orthogonal
convergence and collision; hence, they are easily represented in cross-sectional models. These simplified
models, useful for predicting first-order relationships between geologic elements, often do not consider a
margin-parallel component of relative motion. A growing body of evidence suggests that during oblique
convergence, subduction-related margin-parallel strike-slip shear zones (MPSZs) play an important role in
the evolution of many convergent margins.
MPSZs are probably most important along margins of oblique convergence characterized by shallow
subduction, an overriding plate comprised of continental crust, and an absence of salients that may inhibit
margin-parallel motion. Recognition of evidence for oblique convergence and MPSZs, as preserved in the rock
record, is the next step in refining convergent plate margin tectonic models. This is particularly important
with respect to our understanding of continental margin kinematics in the past, for which there exist few data
to constrain plate motions. The formation and geometry of elements within MPSZs depends on structural
level and bcation (i.e., distance from the trench) within the overriding plate. Detailed structural and
kinematic analysis across presumed ancient margins may reveal subduction-related MPSZs. Knowledge of
MPSZs active along the plate decollement as well as within the overriding plate may allow limits to be placed
on the direction, time, and magnitude of the translation of upper-plate slivers. Documentation of MPSZs and
incorporation of relevant motion and age data into multi-dimensional models, those which consider more than
a classical cross-orogen section, may greatly advance models of Mesozoic (and older) convergent plate-margin
evolution.
INTRODUCTION
A growing body of empirical evidence (Fitch,
1972; Saleeby, 1977; Walcott, 1978; Beck, 1986;
Karig and others, 1986; Sarewitz and Karig, 1986;
Hansen, 1987a,b, 1988), statistical study (Jarrard,
1986a,b), and theoretical models (Beck, 1986),
indicates that margin-parallel translation of upper-
plate tectonic slivers is an important process in
many modern and ancient convergent plate
margins. Rocks overlying subduction zone decolle-
ments are affected by strike-slip displacements in
addition to dip-slip shortening in cases of oblique
plate convergence. The possibility that translation
of upper-plate tectonic fragments may occur in a
plane of motion normal to the plane of an orogen-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 51-65. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
51
52 V. L. Hansen
orthogonal cross-section introduces an important
degree of freedom not allowed by classical cross-
sections. Hence, orogen-orthogonal cross-sections
are often oversimplified and may lead to incorrect
correlation of unrelated tectonic elements, a lack of
correlation of related tectonic elements, or
correlation of tectonic elements within a tectonic
framework which does not account for margin-
parallel displacement.
Zones that accommodate margin-parallel trans-
lation of upper-plate tectonic slivers are referred to
by a variety of terms, including "arc-parallel strike-
slip faults" (Fitch, 1972), "subduction-related strike-
slip faults" (Beck, 1983, 1986), or "forearc strike-slip
faults" (Jarrard, 1986a,b). Throughout this paper
these subduction-related margin-parallel strike-slip
shear zones will be referred to as margin-parallel
shear zones (MPSZs) . Although these strike-slip
faults probably have steep orientations at shallow
crustal levels, they probably shallow with depth and
may merge into the plate decollement (Hansen,
1987a,b, 1988). In addition, their attitude may
change with time and tectonic setting (e.g., Karig
and others, 1986).
The description of these structural features as
shear zones simply reflects their nature as boun-
daries of displacive strain; they may accommodate
movement principally by ductile flow, brittle fail-
ure, or a combination. Although many modern ex-
amples are commonly exposed as strike-slip faults,
these structures evolve at depth as ductile shear
zones. The modifier "margin-parallel" refers to the
direction of slip along these shear zones, parallel to
the plate margin rather than orthogonal or oblique
BACKARC
BASIN
ARC
to it. MPSZs were described earlier as a type of
transform fault (Dewey, 1980). However, as I hope
to illustrate, these shear zones should not be
considered transform faults because they probably
do not transect the lithosphere, they do not connect
zones of crustal extension or shortening, and they do
not originate as plate boundaries. Offsets along
transforms may be small, but displacements are
generally large, probably indeed larger than along
MPSZs. However, displacements along MPSZs may
be very large (see Gabrielse, 1985; Umhoefer, 1987).
Although MPSZs may form transient plate boun-
daries, they initiate and evolve primarily in the
overriding plate parallel to, but distinct from, the
local plate margin.
MARGIN-PARALLEL SHEAR ZONES
Margin-parallel shear zones (MPSZs) can be
thought of as strike-slip faults formed synchron-
ously with subduction within the upper plate of a
convergent margin system. These shear zones
develop inward from, and parallel to, the trench in
the forearc, arc, or backarc regions, and dissect the
overriding plate into narrow tectonic slivers which
can move somewhat independently from either
major plate (Figure 1) (Fitch, 1972; Dickinson,
1972; Karig and Mammericks, 1972). The tectonic
slivers bounded by MPSZs within the upper plate
undergo shortening normal to the plate boundary.
MPSZs resolve the overall oblique shear between
converging plates into two distinct components of
shear - one component at a high angle to the plate
FOREARC ACCRETIONARY
BASIN COMPLEX
Fi gure 1. Simplified three-dimensional block dia-
gram of a convergent plate margin illustrating the
form ofMPSZs along strike and with depth.
Margin-parallel shear zones at convergent plate margins 53
margin which acts along the decollement between
the converging plate and possibly along other
decollement-parallel thrust faults, and the second
component of shear parallel to the plate margin
acting along shear zones within the overriding plate
(Fitch,1972).
The geometry and structures of MPSZs are
dependent on the crustal level at which they formed.
MPSZs are simply zones of crustal shear, and hence
exhibit the mechanical behavior of all faults
(Sibson, 1977, and this volume). They are charac-
terized by zones of brittle shear deformation at
shallow crustal levels and by zones of ductile shear
deformation at depth. In addition, as with other
fault zones, the width of MPSZs, which form in part
by ductile shear, increases with increasing depth
and with an increase in factors that enhance ductile
deformation (see Sibson, 1977, and this volume).
MPSZs are generally steep at upper-crustal levels
and may shallow with depth and merge with the
subduction zone or plate decollement (Figure 1)
(Hansen, 1987a,b, 1988). Mechanical changes in
shear zones with increasing depth, and correspond-
ing changes in MPSZ orientation from steep to
shallow, should result in broadly spaced, yet rela-
tively narrow zones of brittle deformation at shallow
crustal levels, and more closely spaced, yet wider
zones of more penetrative ductile deformation at
deeper crustal levels.
Beck (1986) proposed that MPSZs form within
the weakest portion of the overriding plate, forming
first within the structurally disturbed and "soft-
ened" leading edge of the upper-plate wedge and
stepping successively closer to the arc with time
(Figure 1). Several anastomosing MPSZs may in-
teract to form a system of dominantly strike-slip
shear zones within the upper plate of a convergent
margin system. Such a suite of transcurrent shear
zones can dissect the leading edge of the overriding
plate into discrete, upper-plate tectonic slivers.
Once disconnected these slivers translate or rotate
along the length of the margin. Although each
tectonic sliver can move somewhat independently
relative to neighboring slivers and the overriding
plate from which it was derived, the motion of a
particular tectonic sliver is not random or chaotic.
The direction and amount of translation, or rotation,
of these detached pieces depend on the direction and
rate of convergence between the two major plates.
Movement along MPSZs can cause truncation, off-
set, or doubling of tectonic elements within a con-
vergent system; hence, knowledge of the direction
and amount of transport along MPSZs in ancient
rocks can lead to important constraints on ancient
plate interactions and relative movements.
PARAMETERS CONTROLLING MPSZ
FORMATION
Based on a statistical study of 39 modern
subduction systems. Jarrard (1986a) documented
that MPSZs are more common than: (1) strike-slip
faults between major plates; (2) transform faults
perpendicular to a trench; and (3) strike-slip faults
that form tear faults within fold and thrust belts.
Jarrard found that 50% of active convergent
margins show evidence of MPSZs, and that 67% of
such margins in which the upper plate is comprised
of continental crust show evidence of MPSZs.
As stated earlier, MPSZs form as a result of
resolution of oblique plate convergence into two
distinct components of shear - one component of
shear at a high angle to the plate margin and a
second parallel to the plate margin. Three main
parameters determine whether or not decomposition
of the convergence vector from uniform shear is
energetically feasible within a convergent plate-
margin system, and they are the major factors that
control formation of MPSZs: (1) angle of obliquity
between converging plates; (2) strength, and hence
composition, of the overriding plate; and (3) coup-
ling between subducting plates, dependent on the
angle of the subducting slab (Beck, 1986; Jarrard,
1986a,b).
Angle of Convergence Obliquity
The causal relationship between oblique sub-
duction and margin-parallel translation of upper-
plate tectonic slivers was first recognized by Fitch
(1972) in the western Sunda region of the Indian
Ocean. He proposed that margin-parallel dissection
and dextral translation of the Sumatra arc along the
Semangko fault resulted directly from the oblique
angle of convergence between oceanic lithosphere of
the Indian plate and the overriding Asian plate.
Resolution of the oblique slip between the Asian and
Indian plates resulted in initiation of slip parallel
(along the Semangko fault) and perpendicular to the
plate margin (Figure 2).
Fitch (1972) assumed equal friction per unit
area for the thrust boundary between plates and for
54 V. L. Hansen
- ---1,.--------11--
6
t l
( b)
A 10 6
XSECTION VIEW "'-J C
!
r
tan ( Y/ 2) > - ' tan 8
(a) rs d
Figure 2. Simplified map of the Sumatra region showing
the Indian plate subducting beneath the Asian plate (a),
schematic diagrams illustrating angle of oblique conver-
gence (b), and dip of the subduction slab (e). Equation (d)
from Beck (1986) relating subduction variables; rf = resis-
tance to slip along MPSZs; rs = resistance to slip along the
subduction zone. Modified from Beck (1986).
15
10
em/ yr
5
(a)
15
25
10
em/ yr
5
(b)
OCEANIC
Figure 3. Convergence rate and azimuth with respect to
the trench (barbed line) for modern subduction zones in
which continental crust comprises the upper plate (a) and
the oceanic crust comprises the upper plate (b); angles of 0
and 25 obliquity are shown for reference. Arrows indicate
the presence and direction (relative to the trench) of
MPSZs; dots indicate that no evidence for MPSZ formation
is documented. Diagram modified from Jarrard (1986a).
the strike-slip fault within the overriding plate. His
calculations showed that shear vector decomposi-
tion, dependent on the angle of subduction obliquity
and the dip of the subducting slab (Figure 2b,c), is
likely to occur when the angle of oblique relative
convergence exceeds 45 given a 30 dip of the sub-
ducting slab. The angle of obliquity is measured
from a vector normal to the plate margin, where
normal plate convergence equals an angle of zero
obliquity (Figure 2b). Beck (1983, 1986) revised
Fitch's geometric analysis and allowed the resist-
ance-to-slip on the margin parallel shear zone to dif-
fer from that on the subduction shear zone (Figure
2d). Beck's calculation predicted shear vector de-
composition with slightly lesser angles of conver-
gence (e.g., if slab dip =20 and obliquity =40,
degeneration will result and MPSZs will form).
However, the minimum angle of oblique con-
vergence necessary for formation of MPSZs may not
be as large as predicted by these theoretical treat-
ments. Statistical analysis of active convergent
margins (Jarrard, 1986a) revealed that a significant
number of margins that show evidence of subduc-
tion-related margin-parallel translation have con-
vergence angles much less than 40. Furthermore,
shallower angles of convergence result in margin-
parallel translation when the overriding plate is
composed of continental rather than oceanic crust
(Figure 3).
Strength of the Overriding Plate
The strength of the overriding plate directly
affects the resistance-to-slip along MPSZs. Beck
(1986) proposed that MPSZs form when the resis-
tance-to-slip along high-angle structural discon-
tinuities is less than the resistance-to-slip along the
subduction-zone plate decollement itself; he ex-
pressed a ratio relating slip resistance along these
two surfaces based on a geometric analysis of sev-
eral modern convergent-margin regimes (Figure
2d). However, Beck did not explicitly consider the
effect of crustal composition on the resistance-to-slip
along MPSZs. Jarrard (1986a,b) documented that
these shear zones form more commonly when the
overriding plate is composed of continental crust.
The ability to form MPSZs can thus be directly
related to the composition of the crust they dissect,
and hence to the strength of the overriding plate
(Jarrard, 1986a).
Margin-parallel shear zones at convergent plate margins 55
On average, continental crust is significantly
weaker than oceanic crust (Vink and others, 1984),
and therefore it is much more likely to host MPSZs
when it comprises the overriding plate. Strength of
a plate is dependent on four major variables:
lithosphere composition, geothermal gradient, pore-
fluid pressure, and crustal thickness (Brace and
Kohlstedt, 1980; Vink and others, 1984). Vink and
others, (1984) presented simplified models of oceanic
and continental lithosphere in order to illustrate the
interaction of these four major variables and their
effect on crustal strength. They modeled oceanic
lithosphere as pure olivine crust above olivine
mantle, and continental lithosphere as 30 km of
pure quartz crust above an olivine mantle. The
strengths of these two model lithospheres are
estimated by using appropriate experimental data
on crystalline creep of olivine and quartz, under con-
ditions of variable geothermal gradients and pore-
fluid pressure (Figure 4). The two-layered con-
tinental lithosphere results in an intra-lithosphere
weakness zone arising from ductile flow in the basal
quartz layer (Figure 4b), whereas the single-layer
oceanic lithosphere does not have a strength mini-
mum (Figure 4c). Dunbar and Sawyer (1988) pre-
dicted a similar zone of continental lithosphere
weakness based on a more realistic three-layered
model. Although the depth to the zone of weakness
within continental lithosphere is dependent on
geothermal gradient, pore-fluid pressure, and litho-
sphere composition, there is abundant geologic
evidence to support the presence of such a zone
within actual continental lithosphere (see Turcotte,
1987, and references cited therein).
Figure 4. Differences between maximum
horizontal stress (compressive) and vertical
stress as a function of depth, modified from
Vink and others (1984). Values for differ-
ent pore pressures shown in parentheses,
and values for the slopes 15, 20, and 25
geotherms are shown for quartz and oli-
vine. Strain rate is constant at 10 -IS/sec.
Crustal strength profile from (a) using
hydrostatic pore pressure and 20 geotherm
for model continental lithosphere (b) and
model oceanic lithosphere (c).
20
E 40
:!!.
J:
...
[t 60
o
80
The strength of continental lithosphere is also
dependent on thickness of the continental crust; the
thicker the continental crust, the weaker its litho-
sphere (Vink and others, 1984). Model continental
crust shows an order of magnitude decrease in
strength with a doubling of its thickness from 30 km
to 60 km (Vink and others, 1984).
The regional geothermal gradient has an im-
portant effect on the strength of the lithosphere.
Geotherms with low slope (high dT/dz) will induce a
shallower and broader zone of lithospheric weak-
ness. A geothermal gradient that varies in mag-
nitude across a convergent margin will strongly
affect crustal strength as a function of distance from
the trench. The geothermal gradient is generally
highest within proximity of the magmatic arc and
causes a decrease in the strength of the crust near
the arc. Because MPSZs form most easily within
portions of the overriding plate weakened by heat-
ing or structural disruption, they first form within
the leading edge of the upper plate and step suc-
cessively closer to the arc over time due to mech-
anical softening related to progressive heating of
the arc with age (Figure 1) (Beck, 1986).
Coupling Between Plates and/or Angle of the
Subducting Slab
Jarrard (1986a,b) introduced the importance of
"coupling between plates" with respect to the for-
mation of MPSZs. This concept refers to the strain
regime present within the overriding plate. Is the
strain regime compressional or extensional - that
() H - () v lMPa)
400
(b)
constant f = 10 15 Isec
(a) (c)
56
V. L. Hansen
is, is the crust undergoing shortening or extension?
Uyeda and Kanamori (1979) showed that a cont-
inuum from highly extensional (active backarc
spreading) to highly contractional (folding and
thrusting) strain regimes can exist within the
overriding plate, and they presented a three-fold
classification scheme to outline these differences.
Jarrard (1986a) presented a seven-fold classification
scheme along this continuum for modem subduction
zones. The actual value of plate coupling assigned
to a particular margin is somewhat arbitrary and
does not seem important beyond discriminating
between end-member cases. Jarrard (1986b) pre-
sented a workable three-fold strain regime classi-
fication: compressional, neutral, and extensional
subduction zones, in which compressional margins
typically show formation of MPSZs whereas
extensional margins generally do not form MPSZs.
A point to be noted is that the state of stress,
compression, or tension, with respect to convergent
margins, has become a subject of recent debate (see
Platt, 1986). Extensional and contractional tec-
tonics in accretionary wedges (Platt, 1986) are
active simultaneously: extensional at shallow
depths and contractional at greater depths.
The strain regime in the overriding plate, or the
coupling between plates as defined by Jarrard
(1986a,b, his Table 1), can be directly related to the
angle of the down-going slab (see Jarrard, 1986 a,b).
Modem compressional subduction zones (Jarrard
1986b, his Table 1) have 10-26 dips (generally <20
and averaging 18), whereas subduction zones
classified as extensional have 22-44dips (generally
> 25 and averaging 29). Therefore, it may be best
to refer discussion of MPSZ formation to the quanti-
fiable parameter of subduction angle as suggested
by Beck (1986), rather than to the more arbitrary
classification based on qualities of extension or shor-
tening. Thus, shallow angles of subduction 20)
enhance MPSZ formation whereas steep angles of
subduction (> 25) do not.
Although several factors influence the dip angle
of the down-going slab (see Cross and Pilger, 1982;
Jarrard, 1986a), three major factors promote shal-
low angles of subduction: (1) high rates of conver-
gence; (2) rapid absolute motion of the overriding
plate toward the trench; and (3) subduction of young
lithosphere.
In summary, one should expect the formation of
MPSZs if the following criteria for a particular
convergent margin, modem or ancient, are met:
(1) the continental crust comprises the overriding,
or obducted, plate; (2) the down-going slab is sub-
ducting, or was subducted, at an angle < 20; and
(3) the convergence direction is, or was, oblique to
the trench. Even a low angle of convergence obli-
quity (-10) can result in MPSZ formation if the
other criteria are met. In addition, the margin of
the overriding plate must be free of salients that
could inhibit translation of upper-plate tectonic
slivers along MPSZs (Beck, 1986).
RECOGNITION OF MPSZs IN THE ANCIENT
ROCK RECORD
Although there is much evidence to indicate
that MPSZs form within the overriding plate in
modem convergent plate systems, and that trans-
lation of tectonic slivers along these zones may be an
important factor in continental margin evolution,
few studies exist which describe MPSZs in the
ancient rock record. This dearth of documentation
is related, in part, to the fact that tectonic models of
convergent plate systems have only recently evolved
to the point where we are beginning to look for
second-order structures, such as MPSZs, related to
subduction. In additon, evidence of MPSZs can
easily be destroyed by subsequent dip-slip dis-
placements.
Detailed geologic study within a well-defined
regional tectonic framework is required to document
the presence of MPSZs and to understand their
evolution. Despite problems that may hinder the
recognition of MPSZs, recently developed tech-
niques of structural and kinemantic analysis,
geothermobarometric study, and geochronology
provide useful tools to constrain the structural and
kinematic, metamorphic (P-T), and temporal evo-
lution ofMPSZs.
Identification ofMPSZs
In looking for MPSZs within the ancient rock
record one must consider spatial location (i.e.,
distance from the arc or trench), structural level,
temporal evolution, and changes of each of these
factors with time. Although the level of structural
exposure is certainly dependent on erosional level,
in general upper-crustal levels may be more
common in Cenozoic-age margins, whereas lower-
crustal levels may be more often exposed in older
margins. The crustal level of exposure of any
Margin-parallel shear zones at convergent plate margins 57
particular margin will determine in part whether
the major structures are brittle or ductile in char-
acter.
Investigators of lithotectonic belts that may
represent ancient convergent plate margins should
look for evidence of margin-parallel discontinuities
between petrologic, metamorphic, or structural
facies. Discontinuities may be extremely difficult to
recognize, however, because tectonic and lithologic
facies often parallel the margin; therefore offset
marker units are uncommon. In these cases the
shear zones themselves must be located through
careful study of structural changes normal to the
regional trend of the belts. Major margin-parallel
structures may juxtapose out-of-sequences facies or
offset related petrofacies. These structures should
be closely examined using structural and kinematic
techniques to determine the structural evolution
and movement history of the zone(s). If possible,
determine the style of shear deformation, type of
displacement (dip-slip versus strike-slip), direc-
tion(s) of shear, relative timing of different dis-
placement or strain histories within a particular
shear zone, and the amount of offset(s) or strain.
It is not known if below a certain crustal level
the two perpendicular decomposed vector compon-
ents are expressed as a single vector oblique to the
plate margin and parallel to the plate-convergence
vector. If such a level exists, it is probable that the
location of this transition, from two perpendicular
vectors to a single oblique vector, like the "ductile-
brittle transition" is dependent on parameters that
affect the mechanical properties of rocks and is not
confined to a specific depth in the crust. These in-
clude thermal profile, fluid content and composition,
lithology, and strain rate. These parameters almost
certainly change throughout the lifetime of a par-
ticular subduction zone, and thus they may influ-
ence changes in the crustal level at which shear
vector decomposition takes place. As a result,
structural level may vary during tectonic evolution
of a particular convergent margin and hence com-
plicate the structural and kinematic record.
MPSZs are probably most easily recognized at
shallow crustal levels within relatively young mar-
gins. Structural discontinuities marked by margin-
parallel zones of brittle shear deformation may be
relatively easy to recognize even if offset marker
units are lacking. At shallow crustal levels MPSZs
are probably relatively steeply oriented zones of
brittle deformation recording strike-slip displace-
ment. MPSZs may form a single discrete shear zone
or a suite of anastomosing shear zones. The brittle
deformation history recorded in shallow MPSZs is
almost certainly complex. Local structures and
geometries depend on the host rock rheology and
pre-existing structures.
MPSZs formed at deeper crustal levels and
marked by zones of ductile deformation are probably
more difficult to recognize. With increasing depth
the spacing between distinct MPSZs narrows as
they shallow and merge into the plate decollement.
As with any zone of crustal shear (Sibson, 1977, and
this volume), the width of an individual MPSZ
marked by ductile deformation will broaden with
depth. In addition, discrete higher crustal MPSZ
splays may merge into common zones of shearing at
depth. The tectonic slivers bounded by such zones
also narrow at depth and probably also deform
internally by ductile deformation processes. Ductile
deformation in the tectonic slivers could be expected
to be dip-slip rather than strike-slip deformation
marking MPSZs. The direction of tectonic transport
preserved by elongation lineations may be one of the
few ways to distinguish dip-slip zones from strike-
slip zones that formed at similar crustal levels. As
MPSZs join with the plate decollement - the boun-
dary between the hanging-wall block and sub-
ducting footwall block, their mutual structural and
kinematic evolution, and the structural and kine-
matic evolution of the bounded tectonic slivers, may
become quite complex.
In the tectonically dynamic environment of a
plate decollement within an evolving convergent
margin system, only the youngest structures may be
preserved. A record of older deformation is pre-
served only if younger, more brittle shear zones
form discrete zones around blocks or domains con-
taining structures related to more penetrative duc-
tile shear. As more examples of MPSZs from vari-
ous crustal levels are documented in the ancient
rock record, multi-dimensional models of their
formation will emerge.
During oblique subduction, early formed strike-
slip shear zones may be progressively overprinted
by zones of dip-slip motion. If resolution of oblique
shear takes place at greater depth, MPSZs might
dissect zones that record initial penetrative dip-slip
movement. In this case, one would expect the pre-
served portions of the ductilely deforming subduc-
tion complex to record dip-slip motion and to show
evidence of overprinting by younger shear zones
related to lateral translation. The latest preserved
lateral shear zones will have generally formed at
58
V. 1. Hansen
higher crustal levels (resulting in deformation along
discrete zones) than the dip-slip shear zones they
dissect. As a result, one would predict that these
MPSZs should be thinner and less diffuse than
deeply formed, more penetratively developed dip-
slip shear zones that they deform. In essence,
ductile MPSZs that are preserved in the rock record
must have formed later, and under lower ductility
conditions, than penetrative dip-slip fabrics. Para-
meters affecting decomposition of the convergence
vector not only vary with depth but vary spatially
with proximity to the arc. This spatial variance
increases the potential complexity of structures
which may form within a convergent margin system
in which MPSZs play an active role.
Analysis ofMPSZs
In order to interpret the kinematic evolution of
MPSZs we must first identify MPSZs and study
these zones of deformation in detail. As with study
of any shear zone, one must study the boundaries
between the shear zone and the undeformed host
rock, as well as the shear zones themselves. Struc-
tural and kinematic analysis of MPSZs provide the
initial tectonic framework and relative temporal
constraints with which to examine their evolution.
Tools of modern structural analysis (Ramsey and
Huber, 1983, 1987) provide means to evaluate strain
histories. In addition, numerous newly developed
techniques are available for determining directions
of tectonic transport in brittle (Chester and Logan,
1987; Gammond, 1987; Petit, 1987) and ductile
shear zones (Berthe and others, 1979a,b; Bouchez
and others, 1983; Simpson and Schmid, 1984; Lister
and Snoke, 1984; Simpson, 1986; Passchier and
Simpson, 1986; Cobbold and Gapais, 1987, and in-
cluded papers; Mawer, this volume).
In undertaking a study of a potentially complex
shear zone such as an MPSZ, one should pay atten-
tion to the following relationships.
Determine the character of the structural fab-
ric. What are the elements that define the fabric
foliation and lineation? Are lineations slickenside,
elongation, mineral alignment, intersection linea-
tion, or all of the above? Determine the relative
brittle-ductile character of each fabric element. At
what scale(s) are the various fabric elements pene-
trative? Define the spatial and temporal interaction
of each fabric element.
Define zones of simple shear-dominated defor-
mation within the MPSZ. Determine the type of
displacement within these zones, dip-slip versus
strike-slip, and the relative timing between differ-
ent displacement histories. Determine relative
directions of displacement or shear - normal or
reverse, dextral or sinistral. If possible, estimate
the amount of strain or displacement associated
with anyone shear zone (Ramsey and Huber, 1983).
Interpret possible changes in the displacement
history through time as a result of cross-cutting
relationships. Faults, vein-filled fractures, and
strain shadows may preserve a record of the chang-
ing strain regime that affected a particular package
of rock through time (Choukroune, 1971; Wickham,
1973; Durney and Ramsey, 1973; Ramsey and
Huber, 1983; Etchecopar and Malavielle, 1987).
Consider structural level of deformation, and
the deformation history. Constrain P-T conditions
of shear zone formation and changes in P-T con-
ditions with structural evolution. Recall the model
of a shear zone with respect to crustal level (Sibson,
1977, and this volume) in order to interpret relative
crustal level of cross-cutting shear structures. In
addition, directly constrain P-T conditions using
fluid inclusions (Roedder, 1984) and, when possible,
geothermobarometry (Bostick, 1974; Laird and
Albee, 1981a,b; Ferry, 1982; Spear and others, 1984;
Hodges and Crowley, 1985; Bohlen and Liotta, 1986;
and many others). Complementary petrologic study
of vein composition and volumes may also constrain
changes in P-T and fluid composition through time.
Constrain relative and absolute timing rela-
tionships between structural fabric elements, ther-
mal evolution, and uplift through determination of
cooling ages for minerals associated with specific
aspects of the structural or thermal evolution of the
MPSZ. Different minerals close to different isotopic
systems at a wide range of temperatures (Parrish
and Roddick, 1985). Cooling ages for a range of
temperatures derived from a variety of minerals or
isotopic systems combined with P-T constraints
allow construction of P-T time paths and calculation
of cooling and uplift rates.
If MPSZs formed within a terrane that has
subsequently been accreted to a larger contintal
mass, it is imperative to try to constrain the timing
of MPSZ formation relative to accretion. This
timing may have important effects on the tectonic or
accretionary history of the terrane. In addition, as
discussed below, the timing of MPSZ formation
Margin-parallel shear zones at convergent plate margins 59
relative to terrane accretion is important with
respect to the final distribution of upper-plate
tectonic slivers. Karig and others (1986) have
shown through detailed study of MPSZs within the
Philippines archipelago that MPSZs form within a
terrane prior to its accretion. Similar relations are
shown by Hansen (1987a,b,c, 1989). In these cases,
MPSZs within the overriding plate are reactiviated
as zones which accommodate at least a portion of
subsequent terrane accretion. Hansen (1988) out-
lines a model by which Cordilleran terrane accre-
tion along MPSZs may take place. Margin-parallel
deformation may also post-date terrane emplace-
ment, as in the case of Jurassic (?) and Cretaceous to
Tertiary age MPSZs within the western North
American Cordillera (Ave Lallemant and Oldow,
1988).
Study of upper-crustal MPSZs requires careful
detailed study with special attention to structural
evolution and sedimentation. Recent advances in
the study of soft-sediment deformation enable
workers to distinguish between autokinematic and
allokinematic (tectonic) deformation (Jones and
Preston, 1987). By employing methods of careful
field ovservation and laboratory analysis of oriented
samples one can differentiate between pre- and post-
lithification faulting (Laville and Petit, 1984; La-
baume, 1987; Owen, 1987, Petit and Laville, 1987).
Interpretation of the results of detailed studies
within a regional structural framework developed
from field mapping and structural analysis should
enable workers to develop a coherent model for the
upper-crustal evolution of MPSZs and their role in
convergent margin processes.
TECTONIC IMPLICATION OF MPSZs
Evidence that MPSZs played an important role
in the tectonic evolution of the ancient margin of
western North America stems from geological data,
which indicate both doubling (Supppe, 1970; Page,
1982) and truncation of Mesozoic terranes (Cowan,
1982; Johnson, 1984; Pavlis, 1987), and paleomag-
netic data, which show that western North America
is composed of a series oflatitudinally allochthonous
crustal blocks (Irving, 1979; Beck, 1980, 1986; Um-
hoefer, 1987, and references cited therein). MPSZs
may have also played an important role in the
tectonic evolution of Proterozoic and Archean
convergent margins (e.g., Bidwell and Bauer, 1987;
Huddleston and others, 1988).
Karig and others (1986) described how oblique
convergence in the Philippines results in margin-
parallel translation of tectonic slivers within the
hanging wall and eventual emplacement of hang-
ing-wall rocks onto the footwall block along these
same shear zones. Hansen (1988) outlined a model
in which MPSZs, formed within upper-plate arc-
complex rocks of an allochthonous terrane, may be
responsible for pre-accretion, margin-parallel, dis-
tribution of petrotectonic facies within the
overriding plate, followed by final emplacement of
the distended terrane onto continental crust of the
down-going slab.
Dewey (1980) outlined a variety of geometric
and kinematic configurations for zones of margin-
parallel translation and their related tectonic
slivers within the overriding plate. Although the
tectonic complexities that can result from these
configurations are infinite, a few general conse-
quences can be expected from translation of upper-
plate tectonic slivers along MPSZs, including:
(1) lateral distribution of genetically related con-
vergent-margin petrofacies along intracontinental
MPSZs (Beck, 1986); (2) preservation or enhance-
ment of such laterally distributed assemblages after
terrane-continent collision and resultant terrane
accretion (Karig and others, 1986; Sarewitz and
Karig, 1986; Hansen, 1987a,b, 1988); and (3) doub-
ling of tectonic elements to form enlarged terranes,
or truncation and juxtaposition of apparently un-
related terranes, without evidence of a large-scale
suture zone.
Orogen-parallel distribution of convergent-
margin facies along MPSZs can result directly from
translation of tectonic slivers within the upper plate
of a subduction system (Figure 5) (Beck, 1986) or
from combined margin-parallel translation followed
by later collision and emplacement of a tectonically
dissected terrane (Figure 6) (Hansen, 1988). The
simplest case of lateral margin-parallel distribution
occurs along an obliquely convergent margin in
which the upper plate is variably transected by one
or more MPSZs along which tectonic slivers are
progressively displaced in a direction sympathetic
with oblique convergence (Figure 5). The upper
plate in such a case may contain continental crust
and any of the following arc-related features: back-
arc basin, magmatic arc, forearc, and accretionary
prism. Trenchward tectonic slivers are offset pro-
gressively further than continentward slivers, re-
gardless of the timing of movement along individual
MPSZs (Figure 5c). Translation of tectonic slivers
60
V. 1. Hansen
along MPSZs may result in thin linear belts of
lithofacies distibuted over distances much greater
than their original length. Although MPSZs are
shown as straight shear zones for simplicity in
Figure 5, they can anastomose and merge with one
another along strike (see Figure 1). Offset along
anastomosing MPSZs could result in doubling,
truncation, or entire removal of a lithotectonic
facies belt along a particular portion ofthe margin.
The tectonic distribution of convergent-margin
facies becomes more complex if the upper plate
consists of an arc terrane (or fragment of buoyant
crust) which collides with continental crust attached
to the down-going slab (Figure 6). As continental
crust attached to the down-going slab encounters
the leading edge of the accretionary prism of the
overriding plate, trenchward MPSZs may become
successively active as thrust faults as individual
tectonic slivers are obducted onto continental crust
of the subducting slab (Figure 6d). Hence, these
slivers become accreted terranes. The resultant
geometry of the accreted terrane (or multiple
terranes as they may later be interpreted from the
rock record) is a sub-linear distribution of arc
lithofacies parallel to the ancient margin. Therefore
the size of the arc terrane as estimated from the
distribution of anyone of the lithofacies could
greatly overestimate the size of the arc prior to its
emplacement if the motions along MPSZs were not
considered. In addition, each tectonic sliver may
later be interpreted (incorrectly) in the rock record
as an individual tectonic terrane.
Continued convergence across a collisional belt
similar to that illustrated in Figure 6d may lead to
initiation of a new subduction zone beneath the
accreted arc terrane. If the relative convergence
direction remains the same with only a reversal in
the vergence of subduction (Figure 6), margin-
parallel translation of upper-plate slivers will
continue in the same direction as translation prior
to terrane accretion. Complications may arise, how-
ever because the newly accreted terrane and under-
lying continental crust may now become involved in
MPSZ deformation and translation. If, however,
convergence obliquity changes along with the ver-
gence of plate subduction, the direction of margin-
parallel upper-plate translation may also change
accordingly. If great enough so as to cause obliquity
reversal, such changes may result in back transla-
tion of upper-plate tectonic slivers. Back trans-
lation of sufficient magnitude may return the upper-
plate slices to a position very near where they ori-
gin ally formed, as has been recently proposed for a
portion of the western North American Cordillera
(Ave Lallemant and Oldow, 1988).
The orientation of the continental margin (cra-
tonal edge) on the down-going slab with respect to
the convergent margin (trench) will also affect the
syn- to post-accretion distribution of arc lithofacies
within the overriding plate, as well as the mode of
terrane accretion (Hansen, 1988). Figure 6 illus-
trates the simplest situation in which continental
crust is oriented parallel to the convergent margin;
however, if the trend of the cratonal edge is not
parallel to the convergent margin, additional com-
plications may result.
The concept of suspect terranes (Coney and
others, 1980) has led many North American Cor-
dilleran geologists to emphasize differences in litho-
logy, metamorphism, styles of deformation, and
timing in order to delineate distinct tectonic ter-
ranes (Jones and others, 1983; Coney and Jones,
1985). Although this approach has proven useful in
identifying some of the complexities of Cordilleran
tectonic evolution, the similarities between "dis-
tinct" lithotectonic terranes, and possible correla-
tions of these terranes in space, time, or tectonic
environment has not always been given sufficient
consideration. The models presented here for
margin-parallel translation of upper-plate terranes
may be applicable to a number of outstanding pro-
blems in Cordilleran tectonics. Such models account
for translation of "distinct terranes" along a conver-
gent margin without requiring each terrane to form
thousands of kilometers outboard of its present
location and in a manner "distinct and separate"
from neighboring terranes. In addition, such models
present a mechanism by which a terrane can be
accreted or emplaced onto pre-existing continental
crust in a long linear belt without evidence of a
major suture, and they suggest a plate geometry by
which the polarity of subduction beneath an arc may
reverse itself during continued convergence.
Kinematic models based on regional geologic
data of the Tintina-Rocky Mountain Trench-Fraser-
Straight Creek fault system in southen British
Columbia (Gabrielse, 1985; Price and Carmicheal,
1986), and the Tintina-Denali fault system in the
Yukon and Alaska (Meisling and others, 1987;
Panuska, 1987) indicate significantly less post-mid-
Cretaceous displacement of terranes than is inter-
preted from paleomagnetic data (Irving and others,
1985; Umhoefer, 1987). It is possible that the solu-
tion to this problem lies in understanding the dyna-
Margin-parallel shear zones at convergent plate margins 61
cc SA ARC FA AP OC
(a)
,.;,
.. '

(b)
-
I I I
'l'
I I
I I
I I
:

-
I I- II
I
:-11
I
I
(c)
--
I I I
Figure 5. Simplified plan-view models of a convergent-
margin system showing relative offset of upper plate litho-
tectonic facies. a) Strip map of starting geology; lithotec-
tonic facies: CC, continental crust, BA, backarc, ARC,
magmatic arc, FA, forearc, AP, accretionary prism and
OC, oceanic crust - any of which may be absent within a
given margin; solid barbed line represents the trench,
barbs are on the upper plate. b) Initiation of oblique sub-
duction (heavy arrow) of oceanic crust beneath the con-
tinent and associated arc facies; solid lines mark sites for
MPSZs and form boundaries of the individual tectonic
slivers. c) Margin-parallel offset of upper-plate tectonic
slivers with continued oblique convergence; displacement
is represented by arrow above each sliver; dashed lines
mark extension of each sliver along strike. Along strike
anastomosoing of MPSZ is possible (not shown). Amount of
the offset along individual MPSZs may vary; however,
trenchward tectonic slivers are progressively further offset
relative to continentward slivers regardless of the amount
or timing of movement along individual MPSZs. The
result of oblique subduction will be to distend upper-plate
convergent-margin facies in a margin-parallel fashion.
Right-lateral translation should continue unless conver-
gence angle changes. Orthogonal convergence will result
in no further margin-parallel translation, whereas NW-
(top-left) directed subduction will result in left-lateral
translation of upper-plate tectonic slivers, possibly return-
ing them to their original location offormation.
(a)
(b)
(c)
(d)
OC ARC F A A
I
I
, ,
Figure 6. Simplified plan-view models of a convergent-
margin system in which the upper plate is composed of an
arc terrane and oceanic crust and the down-going plate is
composed of subducting oceanic crust attached to contin-
ental crust (patterns and symbols as in Figure 5) . a) Strip
map of starting geology. b) Initiation of oblique subduc-
tion of oceanic crust beneath associated arc facies.
c) Margin-parallel offset of upper-plate tectonic slivers
with continued oblique convergence. Displacement is re-
presented by arrow above each sliver; dashed lines mark
extension of each sliver along strike. Anastomosing
MPSZs are not considered. d) Map view of collision and
accretion of distended arc terrane with and onto con-
tinental crust attached to down-going slab. MPSZs pro-
gressively become thrust faults (open barbs with barbs on
upper plate) beginning with most trenchward MPSZs.
Heavy dashed line indicates location ofleading edge of the
craton beneath the arc terrane. Continued convergence
across the illustrated margin with oceanic crust behind the
arc may lead to initiation of a new subduction zone, and a
change in subduction polarity, beneath the accreted arc
terrane and continental crust, shown by dark arrow and
filled barbs. Convergence that simply changes subduction
polarity without reversing the angle of convergence will
continue to distend upper plate tectonic slivers in a similar
fashion, dextral in the case shown. However, as discussed
for Figure 5, a reverse of convergence obliquity will result
in a sympathetic change in the translation direction of
upper-plate tectonic slivers.
62 V. L. Hansen
mic evolution of these faults as potential MPSZs
related to Mesozoic and Cenozoic subduction of the
Farallon and Kula plates beneath North America.
SUMMARY
In conclusion, a growing body of evidence
indicates tht margin-parallel translation of upper-
plate tectonic slivers along MPSZs is an important
tectonic process which contributes to the dissection,
displacement, and accretion of tectonic slivers with-
in convergent plate-margin systems - processes
which themselves greatly affect the architecture of
continental crust. MPSZs are most important along
margins of oblique convergence which are charac-
terized by (1) shallow subduction, (2) an overriding
plate comprised of continental crust, and (3) boun-
daries that are free of salients which may inhibit
margin-parallel motion. Recognizing evidence for
oblique convergence and MPSZs as preserved in the
rock record is the next step in refining convergent
plate-margin tectonic models and in expanding our
understanding of continental-margin kinematics in
the past for which there exist few data to constrain
plate motions. The formation and geometry of
elements within MPSZs depend on structural level
and location (i.e., distance from the trench) within
the overriding plate. Documentation of MPSZs in
the rock record requires detailed structural and
kinematic analysis within a regional tectonic frame-
work, and may be integrated with complementary
geothermobarometric study and geochronologic
analysis in order to place limits on the direction,
timing, and magnitude of the translation of upper-
plate slivers. Further study of MPSZs recognized
within the rock record, particularly those docu-
mented as having formed at various crustal plate
margins.
ACKNOWLEDGMENTS
This work was supported by U.S. National
Science Foundation grants (EAR85-07953 and EAR-
8715911). I thank J. W. Goodge and H. G. Ave
Lallemant for reviewing and vastly improving ear-
lier versions of the manuscript.
REFERENCES
AVE LALLEMANT, H. G., and J. S. OLDOW, 1988, Early
Mesozoic southward migration of Cordilleran transpres-
sional terranes: Tectonics, v. 7, p.l057-1075.
BRACE, W. F., and D. L. KOHLSTEDT, 1980, Limits on
lithospheric stress imposed by laboratory experiments:
Journal of Geophysical Research, v. 85, p. 6248-6252.
BECK, M. E., Jr., 1980, Paleomagnetic record of plate
margin tectonic processes along the western edge of North
America: Journal of Geophysical Research, v. 84, p. 7115-
7131.
__ ,1983 On mechanism of tectonic transport in zones of
oblique subduction: Tectonophysics, v. 93. p. 1-11.
__ , 1986, Model for late Mesozoic-early Tertiary tec-
tonics of coastal California and western Mexico, and specu-
lation on the origin of the San Andreas fault: Tectonics, v.
5, p. 49-64.
BERTHE, D., P. CHOUKROUNE, and D. GAPAIS, 1979a,
Orientations preferentielles du quartz et orthogneissi-
fication progressive en regime cisaillant: I'example du
cisaillement sud-Amoricain: Bulletin de Mineralogie, v.
102, p. 265-272.
__ , P. CHOUKROUNE, and P. JEGOUZO, 1979b, Ortho-
gneiss, mylonite and noncoaxial deformation of granite:
The example of the South American shear zone: Jounal of
Structural Geology, v. 1, p. 31-42.
BIDWELL, M. E., and R R BAUER, 1987, Ductile
shearingand coeval folding during dextral transpression,
central Vermillion district, NE Minesota [abstract]:
Geological Society of America Abstracts with Programs v.
19, p. 588.
BOHLEN, S. R, and J. J. LIOTTA, 1986, A barometer for
garnet amphibolites and garnet granulites: Journal of
Petrology, v. 27, p. 1025-1034.
BOSTICK, N. H., 1974, Phytoclasts as indicators of thermal
metamorphism, Franciscan assemblage and Great Valley
sequence (upper Mesozoic), California; pp. 1-17 in R R
Dutcher, P. A. Hacquebard, J. M. Schops, and J. A. Simon
(eds.), Carbonaceous Materials as Indicators of Meta-
morphism: Geological Society of America, Special Paper
153,108p.
BOUCHEZ, J. 1., G. NANTES, S. LISTER, and A. NICOLAS,
1983, Fabric asymmetry and shear sense in movement
zones: Geologische Rundschau, v. 72, p. 401-419.
CHESTER, F. M., and J. M. LOGAN, 1987, Composite
planar fabric gouge from the Punchbowl Fault California:
Journal of Structural Geology, v. 9, p. 621-634.
Margin-parallel shear zones at convergent plate margins 63
CHOUKROUNE, P., 1971, Contribution a l'etude desmech-
anisms de la deformation avec schistositie grace aux
cristallisation syncinematiques dans les "zones abritees"
("pressure shadows"): Bulletin de la Societe Geologique de
France, v. 7, p. 257-271.
COBBOLD, P. R., and D. GAPAIS., 1987, Shear criteria
inrocks: an introductory review: Journal of Structural
Geology, v. 9, p. 521-524.
CONEY, P.J., D.L. JONES, and J. W. H. MONGER, 1980,
Cordilleran suspect terranes: Nature, v. 188, p. 329-333.
__ , and D. L. JONES, 1985, Accretion tectonicsand
crustal structure in Alaska, Tectonophysics, v. 119, p. 265-
283.
COWAN, D. S., 1982, Geological evidence for post-40 m.y.
B.P. large-scale northwestward displacement of part of
southeastern Alaska: Geology, v. 10, p. 309-313.
CROSS, T. A., and R. H. PILGER, Jr., 1982, Controls on
subduction geometry, locations of magmatic arc and back-
arc regions: Geological Society of America Bulletin, v. 93,
p.545-562.
DEWEY, J. F., 1980, Episodicity, sequency and style at
convergent plate boundaries; pp. 553-573 in D. W. Strange-
way (ed.), The Continental Crust and Its Mineral De-
posits: Geological Association of Canada, Special Paper
20,804p.
DICKINSON, W. R., 1972, Evidence for plate-tectonic
regimes in the rock record: American Journal of Science, v.
272, p. 551-576.
DUNBAR, J. A., and D. S. SAWYER, 1988, Continental rift-
ing at pre-existing lithosphere weaknesses: Nature, v. 333
p.450-452.
DURNEY, D. W., and J. G. RAMSEY, 1973, Incremental
strains measured by syntectonic crystal growth; pp. 67-96
in K. A. DeJong and R. Scholten (eds.), Gravity and Tec-
tonics: Wiley, New York, New York, USA, 502 p.
ETCHECOPAR, A., and J. MALAVEILLE, 1987, Computer
models of pressure shadows: A method for strain measure-
ment and shear sense determination: Journal of Struc-
tural Geology, v. 9, p. 667-678.
FERRY, J. M. (ed.), 1982, Characterization of Metamor-
phism through Mineral Equilibria: Mineralogical So-
ciety of America, Reviews in Mineralogy 10, 397 p.
FITCH, T. J., 1972, Plate convergence, transcurrent faults
and internal deformation adjacent to Southeast Asia and
the western Pacific: Journal of Geophysical Research, v.
77, p. 4432-4461.
GABRIELSE, H., 1985, Major dextral transcurrent dis-
placements along the northern Rocky Mountain Trench
and related lineaments in north-central British Columbia:
Geological Society of America Bulletin, v. 96, p. 1-14.
GAMMOND, J. F., 1987, Bridge structures as sense of
movement on fault surfaces in brittle rocks: Journal of
Structural Geology, v. 9, p. 609-620.
HANSEN, V .. L, 1987a, Structural, Metamorphic, and
Geochronologic Evolution of the Teslin Suture Zone,
Yukon: Evidence for Mesozoic Oblique Convergence
Outboard of the Northern Canadian Cordillera: PhD
dissertation, University of California, Los Angeles, Cali-
fornia, USA, 221 p.
__ , 1987b, Tectonic correlation of the Yukon-Tanana
(YT) and Slide Mountain (SM) terranes [abstract]: Geo-
logical Society of America Abstracts with Programs, v. 19,
p.386.
__ ,1988, A model for terrane accretion: Yukon-Tanana
and Slide Mountain terranes, northwestrn North America:
Tectonics, v. 7, p. 1167-1177.
__ , 1989, Structural and kinematic evolution of the
Teslin suture zone, Yukon: Record of an ancient transpres-
sional margin: Jounal of Structural Geology, v. 11, p. 717-
733.
HODGES, K. V., and P. D. CROWLEY, 1985, Error
estimation and empirical geothermobarometry for pelitic
systems: American Mineralogist, v. 70, p. 702-709.
HUDDLESTON, P. J., D. SCHVLZELA, and D. L. SOUTH-
WICK, 1988, Transpression in an Archean greenstone belt,
northern Minnesota: Canadian Journal of Earth Sciences,
v. 7, p. 1060-1068.
IRVING, E., 1979, Paleopoles and paleolatitudes of North
America and speculation about displaced terranes: Cana-
dian Journal of Earth Sciences, v. 16, p. 669-694.
__ , P. J. WOODSWORTH, P. J. WYNNE, and A. MOR-
RISON, 1985, Paleomagnetic evidence for displacement
from the south of the Coast Plutonic Complex, British
Columbia: Canadian Journal of Earth Science, v. 22, p.
584-598.
JARRARD, R. D., 1986a, Relations among subduction para-
maters: Reviews of Geophysics, v. 24, p. 217-284.
___ , 1986b, Terrane motion by strike-slip faulting of
forearc slivers: Geology, v.14, p. 780-783.
JOHNSON, S. Y., 1984, Evidence for a margin-truncating
transcurrent fault (pre-late Eocene) in western Washing-
ton: Geology, v.11, p. 538-541.
64 v. 1. Hansen
JONES, D. L., D. G. HOWELL, P. J. CONEY, and J. W. H.
MONGER, 1983, Recognition, character and analysis oftec-
to no stratigraphic terranes in western North America; pp.
21-35 in M. Hashimto and S. Uyeda (eds.), Accretion
Tectonics in the Circum-Pacific Regions: Terra Scien-
tific, Tokyo, Japan, 358 p.
JONES, M. E., and R. M. F. PRESTON (eds.), 1987, Defor-
mation of Sediments and Sedimentary Rocks: Geologi-
cal Society of London, Special Publication 29,350 p.
KARIG, D. E., D. R. SAREWITZ, and G. D. HAECK, 1986,
Role of strike-slip faulting in the evolution of alloch-
thonous terranes in the Philippines: Geology, v. 14, p. 852-
855.
__ , and J. MAMMERICKX, 1972, Tectonic frame-work of
the New Hebrides island arc: Marine Geology, v. 12, p. 187-
205.
LABAUME, P., 1987, Syn-diagenetic deformation of a
turbidite succession related to submarine gravity nappe
emplacement, Autapie Nappe, French Alps; pp. 147-163 in
M. E. Jones and R. M. F. Preston (eds.), Deformation of
Sediments and Sedimentary Rocks: Geological Society
of London, Special Publication 29,350 p.
LAIRD, J., and A. L. ALBEE, 1981a, High-pressure meta-
morphism in mafic schist from northen Vermont: Ameri-
can Journal of Science, v. 281, p. 97-126.
__ , and A. L. ALBEE, 1981b, Pressure, temperature and
time indicators in mafic schist: Their application to re-
constructing the polymetamorphic history of Vermont:
American Journal of Science, v. 281, p. 127-175.
LAVILLE, E., and J. P. PETIT, 1984, Role of synsedi-
mentary strike-slip faults in the formation of Moroccan
Triassic basins: Geology, v. 12, p. 424-427.
LISTER, G. S., and A. W. SNOKE, 1984, S-C mylonites:
Journal of Structural Geology, v. 6, p. 617-638.
MEISLING, K. E., M. C. GARDENER, G. W. CUSHING, and
S. C. BERGMAN, 1987, A reconstruction of southern Alaska
at 75 Ma [abstract]: Geological Society of America Ab-
stracts with Programs, v. 19, p. 769.
OWEN, G., 1987, Deformation processes in unconsolidated
sands; pp. 11-24 in M. E. Jones and R. M. F. Preston (eds.),
Deformation of Sediments and Sedimentary Rocks
Geological Society of London, Special Publication 29, 350 p.
PAGE, B. M., 1982, Migration of Salin ian composite block,
California, and disapppearence of fragments: American
Journal of Science, v. 282, p. 1694-1734.
PANUSK, B. C., 1987, The paleomagnetism-geology con-
flict in southern Alaska tectonics: A review of constraints
and conjectures [abstract): Geological Society of America
Abstracts with Programs, v. 19, p. 799.
PARRISH, R., and J. C. RODDICK, 1985, Geochronology
and Isotope Geology for the Geologist and Explora-
tionist: Geological Association of Canada, Cordilleran
Section, Short Course No.4, 77 p.
PASSCHIER, C. W., and C. SIMPSON, 1986, Porphyroclast
systems as kinematic indicators: Journal of Structural
Geology, v. 8, p. 831-844.
PAVLIS, T. L., 1987, Significance of the Border Ranges
fault system in the Mesozoic tectonics of the northern
Cordillera [abstract): Geological Society of America Ab-
stracts with Programs, v. 19, p. 80l.
PETIT, J. P., 1987, Criteria for the sense of movement on
fault surfaces in brittle rocks: Journal of Structural
Geology, v. 9, p. 597-608.
__ , and E. LAVILLE, 1987, Morophology and micro-
structures of hydroplastic slickensides in sandstone; pp.
107-121 In M. E. Jones and R. M. F. Preston (eds.), De-
formation of Sediments and Sedimentary Rocks: Geo-
logical Society of London, Special Publication 29, 350 p.
PLATT, J. P., 1986, Dynamics of orogenic wedges and the
uplift of high-pressure metamorphic rocks: Geological
Society of America Bulletin, v. 97, p. 1037-1053.
PRICE, R. A., and D. M. CARMICHAEL, 1986, Geometric
test for Late Cretaceous-Paleogene intracontinental trans-
form faulting in the Canadian Cordillera: Geology, v. 14, p.
468-471.
RAMSEY, J. G. and M. I. HUBER, 1983, The Techniques
of Modern Structural Analysis, Volume 1: Strain Ana-
lysis: Academic Press, New Y ork, New York, USA, 307 p.
ROEDDER, E., 1984, Fluid Inclusions: Mineralogical
Society of America, Reviews in Mineralogy 12, 644 p.
SALEEBY, J., 1977, Fracture zone tectonics, continental
margin fragmentation, and emplacement of the Kings-
Kaweah ophiolite belt, southwest Sierra Neveda, Cali-
fornia; pp. 141-160 in R. Coleman and W. P. Irwin (eds.),
North American Ophiolites: Oregon Department of Geo-
logy and Mineral Industries, Bulletin 95, 183 p.
SAREWITZ D. R., and D. E. KARIG, 1986, Processes of
allochthonous terrane evolution, Mindoro Island, Philip-
pines: Tectonics, v. 5, p. 525-552.
SmSON, R. H., 1977, Fault rocks and fault mechanisms:
Journal of the Geological Society of London, v. 133, p. 191-
213.
Margin-parallel shear zones at convergent plate margins 65
SIMPSON C., 1986, Determination of movement sense in
my lonities: Journal of Geological Education, v. 34, p. 246-
261.
__ , and S. M. SCHMID, 1984, An evaluation of the sense
of movement in sheared rocks: Geological Society of Am-
erica Bulletin, v. 94, p. 1281-1288.
SPEAR, F., S. J. SELVERSTONE, D. HICKMOTT, P.
CROWLEY, and K. V. HODGES, 1984, P-T paths from gar-
net zoning: A new technique for deciphering tectonic pro-
cesses in crystalline terranes: Geology, v. 12, p. 87-90.
SUPPE, J., 1970, Offset of late Mesozoic basement terranes
by the San Andreas fault system: Geological Society of
America Bulletin, v. 81, p. 3253-3258.
TURCOTTE, D. L., 1987, Rheology of the oceanic and
continental lithosphere; pp. 61-67 in K. Fuchs and C.
Froidevaux (eds.), Composition, Structure and Dyna-
mics of the Lithosphere-Asthenosphere System: Am-
erican Geophysical Union, Geodynamics Series 16, 327 p.
UMHOEFER, P. J., 1987, Northward translation of "Baja
British Columbia" along the Late Cretaceous to Paleocene
margin of western North America: Tectonics, v. 6, p. 377-
394.
UYEDA, S., and H. KANAMORI, 1979, Back-arc opening
and the mode of subduction: Journal of Geophysical Re-
search, v. 84, p. 1049-1061.
VINK, G. E., W. J. MORGAN, and W. L. ZHAO, 1984,
Preferential rifting of continents: A source of displaced
terranes: Journal of Geophysical Research, v. 89, p.
10,072-10,076.
WALCOTT, R. 1., 1978, Geodetic strains and large
earthquakes in the axial tectonic belt of North Island, New
Zealand: Journal of Geophysical Research, v. 83, p. 4419-
4429.
WICKHAM, J. S., 1973, An estimate of strain increments in
a naturally deformed carbonate rock: American Journal
Science, v. 273, p. 23-47.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Kinematic indicators in shear zones
CHRISTOPHER K. MA WER
Department of Geology, University of New Mexico, Albuquerque, New Mexico 87131, USA
(Current address: M.I.M. Exploration Pty. Ltd., GPO Box 1042, Brisbane, QLD 4001, Australia)
(received August 17, 1988; revision accepted November 17, 1989)
ABSTRACT
Shear zones are fundamental components of all continental margin orogenic belts, irrespective of style or
age. They can occur at any depth in the lithosphere and may be cataclastic, transitional, or mylonitic. They
dominate much large- and small-scale lithospheric deformation and control the formation and siting of many
economic deposits. Kinematic indicators are asymmetric structures developed within the zones during
shearing. They can indicate: (1) presence of a shear zone; (2) lithospheric depth of shearing; (3) direction of
shear; (4) sense of shear; and possibly (5) absolute amount of shear. Typical kinematic indicators are Riedel
fracture assemblage (cataclastic), deformed sigmoidal and en echelon vein arrays (transitional), and SoC
textures and shear-band foliation (mylonitic). Kinematic indicators are developed in sheared rocks of any
original composition, texture and protolith; they are not confined to sheared granitoids as commonly thought.
SHEAR ZONES AND KINEMATIC INDICATORS
Shear zones are major architectural elements in
deformed lithosphere (e.g., Ramsay 1980; Kirby
1985), especially within continental margin oro-
genic belts. Major normal, reverse, and strike-slip
shear zones are observed or interpreted in the upper
crust, the lower crust, and the upper mantle (e.g.,
Tchalenko, 1970; Ramsay, 1980; Percival and Card,
1983; Wernicke, 1985; Brown, 1986; Mawer, 1987a;
Brodie and Rutter, 1987). They can date from the
Archaean (e.g., Friend and others, 1987) to the
Recent (e.g., Karig and others, 1986) and can have a
protracted and complex history (e.g., Watterson,
1975; White and others, 1986; Mawer and White,
1987). Many major shear zones are thought to be
developed along reactiviated ancient basement-
shear-zones, which act as stress guides (e.g., White
and others, 1986). The shear-zones geometry, with
all attendant features, can occur at any scale (e.g.,
Tchalenko, 1970; Mawer and White, 1987) and,
indeed, shear zones can be described in terms of
fractal geometrical concepts (due to self-similarity;
Mandelbrot, 1977; Allison, 1982; Mawer and White,
1987).
Shear zones act as migration pathways and
structural traps for ore-forming fluids (e.g., Cox and
others, 1986; Mawer, 1987b), such as in meso-
thermal gold-quartz deposits (e.g., Red Lake,
Timmins, Hemlo, the Golden Mile, and the Mother
Lode; Roberts, 1987; Sibson and others, 1988; and
many others). Major ore deposits have been dis-
sected and dismenbered by shear zones, and know-
ledge of where the rest of the deposit has gone may
be desirable.
Without a well-understood, obvious stratigra-
phy in an area, even the recognition of shear zones
can be non-trivial. It can be difficult or impossible
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 67-81. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
67
68
C. K. Mawer
to tell the direction and sense of shearing across a
shear zone if (as is commonly the case) piercing
points are not found. These problems can be largely
resolved by recognition and correct interpretation of
kinematic indicators.
This contribution concentrates on common and
useful kinematic indicators developed at outcrop to
thin-section scale, from cataclastic, transitional,
and mylonitic shear zones, which can be developed
in any protolith. Emphasis is on recognition and
correct shear-sense determination, and possible
pitfalls in interpretation. All of the indicators
discussed are found at shear-zone margins or within
the zones.
SOME DEFINITIONS
The definitions given below are offered to com-
plement other published definitions, in order to
make discussion of kinematic indicators more pre-
cise.
Shear zone - a long, narrow, broadly tabular
zone of concentrated inhomogeneous deformation
across which one block of rock is displaced with
respect to a second (see also Ramsay and Graham,
1970; Figure 1). The deformation may be true
simple shearing, but it more commonly consists of
simple shearing plus a component of either shor-
tening or extension normal to the shear zone
boundaries, with or without a rotation about an axis
normal to the boundaries. This definition is scale-
independent; shear zones can be developed at any
scale from the single grain to that of major litho-
spheric plates. Shear zones may be broadly clas-
sified into cataclastic, transitional, or mylonitic
types (Figure 1), depending on dominant syn-
shearing deformation micromechanisms (e.g., Sib-
son, 1977). Cataclastic shear zones are developed by
microfracturing and fragmentation, perhaps invol-
ving stress-corrision cracking, and may show
solution-transfer features. Transitional shear zones
show cyclic overprinting of brittle, solution-transfer
and crystal-plastic deformation mechanisms, as
shown by the development of arrays of quartz veins
(fracturing and fracture healing), subsequent solu-
tion-transfer and crystal-plastic deformation (by
quartz dissolution and reprecipitation, twinning,
kinking, dislocation glide and climb), and then
overprinting by younger quartz veins (fracturing).
Commonly, some mineral types will show crystal-
plastic features (e.g. , quartz with subgrains), some
A
Figure 1. Schematic displacement - parallel cross-
sections of shear zones: A) catac1astic (deformation
dominated by fracturing and comminution), BI and
B2) transitional (deformation showns cyclic overprinting
of brittle, solution-transfer, and crystal-plastic processes),
C) mylonitic (deformation dominated by crystal-plastic
processes).
will be fractured (e.g., feldspars), while others will
develop solution-transfer phenomena (e.g., mica
beards on pyrite grains). Not surprisingly, these
transitional shear zones commonly give the impres-
sion of great structural complexity. Mylonitic shear
zones develop by crystal-plastic phenomena (various
dislocation and diffusional deformation mechan-
isms, associated dynamic recovery, dynamic recrys-
tallization and neocrystallization) and contain
mylonites, which are generally strongly foliated and
lineated, finer-grained than their host rocks, and
contain distinctive kinematic indicators (e.g., Tullis
and others, 1982; Mawer, 1986). Shear zones of
whatever sort can be developed in rocks of any
original composition and fabric.
Kinematic indicators - structures developed at
shear zone boundaries and within shear zones that
are asymmetric with respect to both the boundaries
and any foliation developed within the zones (e.g. ,
Berthe and others, 1979; Lister and Snoke, 1984;
Passchier and Simpson, 1986; White and others,
Kinematic indicators in shear zones 69
1986; Cobbold and others, 1987; Mawer, 1987a;
Mawer and White, 1987). They can be developed at
any scale from a single grain to (at least) tens of
kilometers in diameter (e.g., Gallagher, 1981;
Sorensen, 1983). There is a diverse assemblage of
kinematic indicators, and their asymmetry is direct-
ly related to the direction and sense of displacement
across the shear zone.
Principal plane of observation - kinematic
indicators must be studied and interpreted in the
plane perpendicular to shear zone boundaries or
foliation and parallel to displacement direction or
extension lineation (Figure 1).
The extension lineation is a critical symmetry
indicator in shear zones. It can be unambigously
recognized in outcrop, hand specimen, or thin
section. Extension lineations are generally defined
by elongated mineral grains (especially where those
minerals are usually equant, such as quartz), bou-
dinaged mineral grains (such as fractured and
pulled-apart sillimanite, tourmaline, or biotite,
which may have quartz or calcite pressure shadows
formed in the boudin necks), stretched crystal
aggregates (such as recrystallized trails of feldspar,
garnet, or amphibole), and stretched pebbles or
fragments (in deformed conglomerates, breccias, ag-
glomerates, and other clastic rocks).
EXAMPLES OF KINEMATIC INDICATORS
Many different asymmetric structures have
been proposed and used as kinematic indicators
(e.g., White and others, 1986; Cobbold and others,
1987). All kinematic indicators present in a given
sample must be studied and interpreted before a
sense of shear is attached to that particular piece of
rock.
Cataclastic Shear Zones
Cataclastic shear zones (Figure 1, zone A) show
ubiquitous fracturing, comminution, and formation
of porosity and permability. Because of the greatly
increased surface area available for chemical altera-
tion, these zones weather readily and commonly do
not crop out, except where silicified by precipitation
of quartz from solution.
Cataclastic shear zones generally form at high
crustal levels [surface down to levels where the
ambient temperature is 200C or so for continental
crust (Sibson, 1986); for a geothermal gradient of
25CIkm, this corresponds to "'" 8 km], though they
can form at any depth if fluid pressures are suffi-
ciently high or strain rates rapid. Such zones are
characterized by fractured rock, which may be
either incohesive (gouge, breccia) or cohesive (gen-
eric cataclasites; Sibson, 1977). Rock fragments are
generally angular and may either be chaotic or
define a foliation (e.g., Chester and others, 1985).
The protolith is usually hydrated, with formation of
clays, silification, or other chemical alteration
effects observed. Both the pronounced fracturing
and hydration are abruptly bounded by discrete
fracture surfaces at the shear zone boundaries,
although commonly a less intense alteration zone is
developed outside the shear zone.
The fractured rock within the shear zone
commonly hosts a distinctive asymmetric family of
small shear fractures, collectively called the Riedel
fracture assemblage (Riedel, 1929). This assemblage
of fractures and associated foliation is a common,
interpretable, and useful kinematic indicator in
cataclastic shear zones (e.g., Riedel, 1929; Tcha-
lenko, 1970; Logan and others, 1979; Chester and
others, 1985; Mawer and White, 1987). The assem-
blage is shown completely developed in Figure 2A.
The presence of this family of fractures is indicative
of a cataclastic shear zone and shows the sense and
direction of displacement across the zone. The pat-
tern offractures shown in Figure 2A is seen when
the plane of observation is approximately perpen-
dicular to the displacement direction. The fracture
types can be distinguished by noting their angular
relationships, the sense of displacement across
individual fractures, and their relationship to the P
foliation if developed - this holds even for wide
cataclastic shear zones (e.g., Mawer and White,
1987). The displacement direction can be found
fairly exactly by measuring the orientation of
several fractures from each set and determining
their intersection direction using the stereo graphic
projection (Mawer and White, 1987). The displace-
ment direction is perpendicular to this intersection
and parallel to the shear zone boundaries.
The entire fracture assemblage may not be
developed. Commonly, only the Y and Rl fractures
are well developed, with or without the P foliation
(e.g., Figures 2B and 3; see Mawer and White,
1987). In any case, the resulting asymmetric array
offractures is a reliable shear-sense indicator.
As deformation proceeds in cataclastic shear
zones, most significant displacement becomes loca-
70
C. K. Mawer
Figure 2. The Riedel fracture assemblage in catac1astic
shear zones. A) Diagram showing the assemblage in its
fullest development (after Logan and others, 1979). B) Dia-
gram showing a more typical development in natural shear
zones (after Mawer and White, 1987). Note that in natural
examples commonly only Y and Rl fractures develop. The
X fracture is sometimes developed, and the P orientation, if
expressed, is usually a foliation defined by elongated
fragments of fractured rock. The Riedel assemblage is a
reliable kinematic indicator.
lized on one or few discrete faults, generally in the Y
orientation, from which it can be difficult to derive
the sense of shear. However, the Riedel fracture
assemblage is commonly preserved within less-
N
!
deformed lozenges in the shear zone, or towards the
shear zone margins, so that shear sense can still be
determined (Mawer and White, 1987). Even though
fractures will rotate somewhat during shearing
(e.g., Riedel, 1929; Tchalenko, 1970), angular rela-
tionships are not completely obliterated before the
through-going Y shears are developed (Tchalenko,
1970; Logan and others, 1979; Mawer and White,
1987).
Lineations developed in the fractured rock, such
as grooves and striations on slickensides and fibrous
crystal growths along fractures (e.g., Ramsay, 1967;
Hancock and Barka, 1987; Petit, 1987; Figure 4A),
have been used to determine displacement direction.
These generally indicate last-stage movement direc-
tions or local displacements, which may not be re-
presentative of the movement history of long-lived
zones. For zones with a simple history they seem to
be quite reliable, however (e.g., Hancock and Barka,
1987; Petit, 1987). It is commonly observed that
sliding jumps from discrete fracture to fracture in
these zones, and several different (probably only
local and minor) displacement directions are recor-
ded by striations or fibrous crystals in a small thick-
ness of shear zone (Figure 4B).
Mylonitic Shear Zones
Mylonitic shear zones (Figure 1, zone C) gen-
erally form at moderate to deep crustal levels and
below, at depths in the continental crust where the
ambient temperature is above 350C (Sibson, 1986;
for a geothermal gradient of 25Clkm this corres-
ponds to "'" 14 km), though they can form at very
shallow depths if the temperature is high enough
(Kirby and others, 1989). These shear zones can
200m
Q.:;:::::
Figure 3. Map of natural catac1astic shear zone (the Cobequid fault zone of Nova Scotia, Canada), showing development of a
large-scale Riedel fracture assemblage, which indicates dextral displacement (after Mawer and White, 1987). a, dextral,
b, sinistral.
Kinematic indicators in shear zones 71
develop in rocks of any bulk composition and
original texture; they are characterized by plastic
deformation at all scales down to that of single
mineral grains. This does not preclude grain-scale
fracturing of certain minerals such as feldspar or
amphibole, or fracturing developed as a result of a
major influx of fluids.
The mylonitic foliation is commonly fine,
planar, and laterally continuous, and usually forms
an anastomosing mesh around ellipsoidal lozenges
of less-deformed rock (at all scales, and in all rock
types). Sometimes the transition from host-rock
through progressively better foliated and finer
grained mylonite can be traced. Commonly, pre-
existing markers, (e.g., dykes, xenoliths) will be
progressively bent towards the shear zone boun-
daries (note that this is not a good kinematic
indicator, as it depends on the initial orientation of
the marker with respect to the shear zone; Hill,
1959; Ramsay, 1980; and below).
One of the most important structures associated
with mylonitic shear zones is the extension lineation.
This develops progressively towards the shear zone
and is a critical geometrical feature, as it parallels
the displacement direction across the zone. The
extension lineation gives the trend but not the sense
of displacement.
Sense of displacement in mylonitic shear zones
is given by a variety of asymmetric structures.
Perhaps the most reliable kinematic indicators are
those formed by foliation pairs within the
mylonites. There are two main types of these, S-C
texture and shear-band foliation (Figure 5), and it is
useful to distinguish the two as they represent
different stages in a progressive deformation,
different states of the deforming material, and have
different mechanical significance (cf Lister and
Snoke, 1984; and see Mawer and White, 1987).
Berthe and others (1979) studied the pro-
gressive deformation of a series of granites, con-
centrating on the development and evolution of
foliations within shear zones. As the shear zone
boundaries are reached and crossed, two foliations
are observed. These are interpreted as having
developed more-or-Iess simultaneously. One of
these lies at a relatively high angle (ca. 45) to the
shear zone boundary, the second essentially paral-
lels it. In sections parallel to the extension
lineation, the high-angle foliation (S) is tilted back
against the shear direction and the parallel foliation
(C) has the appearance of very small shear zones
with the same sense of displacement as the major
A.
B.
s-- ----R
7
Figure 4. A) Diagram showing the "rough-smooth" rule
for shear fractures with fibrous crystal growth. The
surface feels smooth if a hand is moved left on it, indicating
that the removed block moved left with respect to the lower
block. B) Diagram showing trends of fibrous crystals on
three shear fractures separated by only millimetres.
Trends are different, and no one trend may be the same as
the overall displacement trend of the shear zone, because of
inhomogeneous deformation. Such fibrous growths indi-
cate only the last, minor, and generally local components of
overall deformation. Scale bar = 2 cm.
zone (Figure 5). These two foliations form the S-C
texture. Note that the asymmetric S-C texture
reliably gives the sense of shearing, as it is not only
a kinematic indicator but is in fact a mechanism of
shearing that accommodates the bulk deformation
(Cobbold and Gapais, 1987; for examples see Lister
and Snoke 1984, Weijermars and Rondeel, 1984,
Mawer and White, 1987).
With increased shear strain accumulation, the
grains that define the S foliation are progressively
deformed; the S foliation rotates towards the C
foliation and loses its integrity (Berthe and others,
1978). The product at this stage is a mylonite with
just one well-developed foliation, which lies at a
very low angle to the shear zone boundary.
72
C. K.Mawer
SHEAR BANDS

shear
band
Figure 5. S-C texture (upper diagram) and shear-band foliation (lower diagram) developed in mylonitic shear zones. S-C
texture is developed in initial stages of deformation of an homogeneous protolith, whereas shear-band foliation develops in the
most deformed and foliated mylonites (or from an originally highly foliated protolith). Right-hand side of both diagrams shows
sketches of common aspects of both composite foliations. Note that both are somewhat irregular in spacing, orientation, and
continuity. Both of these composite foliation geometries are excellent kinematic indicators.
As shearing intensity further increases, yet
another foliation tends to develop in the shear zone.
This has the appearance of a set of small synthetic
shears with approximately the same orientation as
the Rl shears in the cataclastic case. This is shear
band foliation (sometimes known as C' foliation). In
sections parallel to the extension lineation it tilts at
a shallow angle (ca. 30 or less) in the direction of
shearing, thus giving the shear sense (Figure 5).
Shear band foliation is also a reliable shear sense
indicator.
Several criteria exist to distinguish either the
S-C texture or the C-shear band foliation pair from
two overprinting foliations, perhaps associated with
folding, developed at different times.
Both are confined to relatively long, narrow
zones which, as shown by their association with
other asymmetric structures and by the presence of
mylonites, are shear zones.
No folds are mapped or observed with their
axial surfaces parallel to C or the shear band folia-
tion (the presence of sheath folds may complicate
this).
The S-C texture is confined to the least foliated
and deformed parts of the shear zone, whereas shear
band foliation is confined to the most deformed and
foliated parts.
Both C in the S-C texture and shear band
foliation are penetrative throughout the parts of the
shear zones where they develop, but they are
discontinuous. Further, continued development ofS
can warp C in the S-C texture, and C and shear band
foliation can overprint each other locally with no
consistent sense of overprinting. This indicates that
the foliation pairs evolve synchronously during
shearing.
The acute angles between Sand C, and between
C and shear band foliation, always point in the same
sense - i.e., there is the same angular relationship
between the two foliations in each pair, even though
the absolute trends of the pairs are slightly differ-
ent.
The acute angle between Sand C is "" 45 or
less, and that between C and shear band foliation
"" 30 or less. Two foliations intersecting at angles
Kinematic indicators in shear zones 73
A. o

8.
.... _'"
c.
Figure 6. Asymmetric tailed porphyroclasts in mylonitic shear zones, showing both elongated (A, upper porphyroclast, Band
C, small porphyroclastsl and equant (A, lower porphyroclast, C, large porphyroclastl varieties. Stippling in all cases indicates
dynamically recrystallized porphyroclast tails. Note that elongated porphyroclasts have this recrystallized material dragged
off upper leading and lower trailing tips, whereas equant porphyroclasts have tails wrapped around them due to rotation of the
core. Also note in B and around large porphyroclast in C the small asymmetric folds, which independently confirm the
sinistral sense of shear in all cases. D illustrates a possible history of development of equant from elongated porphyroclasts by
continued deformation and dynamic recrystallization. A, B, and C traced from polished slabs of mylonites from the Grenville
Province, Ontario, Canada, cut parallel to principal plane of observation. Scale bars in A and B = 1 cm; scale bar in C = 2 cm.
Asymmetric porphyroclasts are reliable kinematic indicators if distant from any major textural inhomogeneity in the
mylonite.
>45
0
are almost certainly of different ages and are
superimposed.
The minerals defining Sand C, or C and shear
band foliation, indicate the same metamorphic
grade with no pronounced retrogression associated
with one foliation. This does not say they have to be
composed of the same mineral species, however.
In mylonites large, less-deformed augen or
single crystals are commonly set in the fine-grained
mylonitic groundmass, generally with tails of finer-
grained material. These porphyroclasts (Figure 6)
can be very useful shear-sense indicators. Local
flow perturbations can give ambiguous or reversed
shear senses in individual porphyroclasts, however,
and shear sense must be extracted from a statistical
sample of porphyroclasts. Porphyroclasts, embed-
ded in the mylonite matrix but distant from any
other porphyroclast, or dyke or obvious area of tex-
tural difference, should be analysed to minimize
interference effects.
Asymmetric tailed porphyroclasts fall into two
shape groups (Passchier and Simpson, 1986; Mawer
1987a): elongate crystals and equant crystals.
Elongate crystals (sigma-type of Passchier and
Simpson, 1986) - these are elongated parallel or at
a slight angle to the mylonitic foliation and gen-
erally show fine-grained tails attached to their lead-
ing and trailing edges, indicating the shear sense
(Figure 6A, top crystal; 6B,C, small crystals).
These porphyroclasts are commonly developed from
single crystals of plagioclase, mica, amphibole, and
hematite. Spectacular examples of this type of
asymmetric porphyroclast are the mica "fish,"
apparently ubiquitous in mylonites of the Basin and
Range Province of the southwestern USA (Lister
and Snoke, 1984). The fine-grained porphyroclast
74
C. K. Mawer
tails should be the same material as its body, or
perhaps its retrograde products. Ifnot, the apparent
tails may in fact be pressure shadows, and an
incorrect sense of shear interpreted (e.g., Simpson
and Schmid, 1983).
2) Equant crystals (delta-type of Passchier and
Simpson, 1986) - these are rotated, as shown by
attached fine-grained tails (of the same material)
that wrap around the body of the porphyroclast to a
greater or lesser extent, indicating the shear sense
(Figures 6A, lower crystal; 6C, large crystal).
These types of porphyroclasts commonly develop
from K-feldspar, garnet, and pyrite, and from de-
graded remnants of the elongated porphyroclasts
(Figure 6D). Spectacular examples of K-feldspar
"snails" are seen in the high-grade mylonites of the
Grenville Province of Canada (Figure 6; Mawer,
1987a).
Tailed porphyroclasts, observed in isolation to
reduce interference effects, generally prove to be
reliable shear-sense indicators.
Asymmetric extensional structures, such as bou-
dins and pinch-and-swell structures (Figure 7), are
common in mylonitic shear zones and form kine-
matic indicators that are consistent with other
types. They show distinct shapes, depending on the
relative strengths of pulled-apart layer and matrix
during the shearing, and whether the layer is dis-
membered or is still whole. Boudins show a rotation
of their tips and a dragging-out of leading and
trailing tips synthetic with the shear sense (Figure
7, upper layer), similar in this respect to elongated
porphyroclasts. Pinch-and-swell structures, formed
from layers only a little more competent than the
matrix, rotate antithetically (back-rotate) with
respect to the shear sense (Figure 7, lower layer).
They are therefore geometrically and mechanically
similar to shear-band foliation, mentioned above. A
model for these rotations has been proposed by
Hanmer (1986).
Asymmetric folds include both folds within the
shear zone and "drag folds" at the zone boundaties
caused by the bending of pre-existing markers
towards the boundaries.
Intrafolial folds are enclosed by the mylonitic
foliations, are generally asymmetric, and can be
used as shear-sense indicators. They seem to form
by buckling of foliation or compositional layering or
by inhomogeneous shear-strain rates across the
shear zone (Figures 8-10). If the amount of shear-
ing is great, these folds can be over-sheared and
show an apparent reverse sense of shear (Figure
lOA; Sanderson, 1979; Ramsay and others, 1983).
Note that these intrafolial folds commonly do not
have straight hinges; they may be sheath folds
elongated parallel to the extension lineation with
highly curved hinges (Quinquis and others, 1978;
Figure 11). Abundant sheath folds are a good in-
dication of the presence of a mylonitic shear zone
(rarely, they may form elsewhere, for example in
intense constrictional straining environments;
Skjernaa, 1985), though they can be difficult to in-
terpret in terms of shear sense.
A second problem with intrafolial folds relates
to scale of observation (Figure lOB). The small-
scale folds seen in outcrop may be parasitic to larger
scale folds which are not directly observed. It is thus
posible to interpret an incorrect shear sense as cri-
tical information (the asymmetry, and sometimes
even existense, of the larger folds) is not known.
It is also relatively common to find folds in
mylonitic shear zones with hinges that are close to
parallel to the extension lineation, but that appear
as simple cylindroidal asymmetric folds - i.e., not
sheath folds. These apparently indicate shearing
perpendicular to the extension lineation, if inter-
preted as drag folds. It is not yet known how such
folds develop in mylonitic shear zones, and they
should not be used to interpret shear sense.
Folds at shear zone margins are commonly in-
terpreted as due to pre-existing markers such as
either dykes or foliations associated with shear zone
formation being dragged along the shear zone boun-
dary; they are very commonly used to indicate shear
sense. They can be extremely misleading for at
least two reasons: (1) the folds may be related to
much earlier folds that have had one limb sheared
out; and (2) if they are associated with shear zone
formation, their present shape and asymmetry is
due to the original orientation of the now-folded
layer with respect to the shear zone - and does not
simply relate to the displacement vector (Figure 12;
Ramsay, 1980; Hill, 1959, for the brittle case).
Thus many of the kinematic indicators used in
mylonitic shear zones can be ambiguous or difficult
to interpret correctly. There may be reasons, such
as local flow perturbations around less-deformed
bodies or unrecognized folds, why some indicators
are ambiguous or indicate the opposite sense of
shear. Individual kinematic indicators must be con-
sidered in terms of their reliability and compared to
other types of indicators, before making a final
decision on shear sense.
Kinematic indicators in shear zones
Figure 7. Experimental deformation of pre-formed
boudins and pinch-and-swell structures in a simulated
mylonitic shear zone (after Hanmer, 1986). Note the
asymmetry of the structures after shearing, both indica-
ting sinistral shear zense. Scale bar = 2 cm. These
asymmetric extensional structures are reliable kine-
matic indicators in mylonitic shear zones (e.g., Mawer,
1987).
Figure 8. Asymmetric intrafolial folds at a boudin tip in
a mylonitic shear zone (field sketch from a shear zone in
the Grenville Province, Ontario, Canada). Boudin =20
cm thick at thickest part. Asymmetric intrafolial folds
can be used as kinematic indicators (as here), but care
should be exercised - see Figure 10 and text.
profile open
hnge line _IICCI
prdlle IIghI
hnge lJOe plungong
75
prof lie ,sochnal
hlOge line , horlzonlal
Figure 9. Sketch of progressive fold tightening and hinge rotation in dextrally sheared rocks near the Chedabucto fault zone
of Nova Scotia, Canada. Note that in this case the intrafolial fold asymmetry correctly indicates dextral shearing, confirmed
by SoC texture and shear band foliation in adjacent mylonites. These folds have an amplitude of about 2-7 cm.
76 C. K. Mawer
A.
--


,--S_


10
----
B.
..... ..
..........
.' I

.J .
..
... -
",..
Figure 10. A) Diagram illustrating how intrafolial folds
can be oversheared and apparently indicate an incorrect
sense of shear (after Ramsay and others, 1983). This is
fairly common in mylonitic shear zones, and shows the
danger of interpreting such structures as "drag" folds.
B) Diagram showing that asymmetric folds may in fact be
parasitic on a larger hidden fold and may be incorrectly
interpreted to indicate, in this case, a dextral shear sense.
Asymmetric intrafolial folds should not be interpreted as
kinematic indicators in the absense of independent cri-
teria.
Transitional Shear Zones
Moderate continental crustal levels ("" 8-14 km)
with ambient temperatures "" 200-350C for a
geothermal gradient of 25Clkm (Sibson, 1986; note
provisions in previous two sections) show very com-
plex deformational behavior. Transitional shear
zones (Figure 1, zones B1 and B2) form with brittle,
solution-transfer, and crystal-plastic deformation
processes operating synchronously within the whole
shear zone and cyclically in any particular volume
of rock in the zone. Different mineral species may
show very different deformational behavior, from
crystal-plastic (quartz) to brittle (feldspars). Many
of the features and kinematic indicators described
above for catacIastic and mylonitic shear zones are
seen (e.g., fracture sets, mylonites with asymmetric
structures) and they can be interpreted the same
way in terms of shear sense. The passage of large
amounts of fluids at some stage(s) in the history of
the shear zone generally alters and retrogresses the
Figure 11. Sketch showing development of sheath folds by
inhomogeneous deformation. Sheath folds are good indica-
tors of the presence of a mylonitic shear zone but generally
are not good kinematic indicators.
TRUE SLIP IS'
I-LODR
2 - Ls.
3 - LoDN
4 - Nds
5 - RoDN
Figure 12. Diagram showing the impossibility of
constraining displacement direction and sense across a
shear zone from simple consideration of the "drag" folds at
shear-zone (SZ) boundaries; see text for further explana-
tion: L, left, R, right, ob, oblique, ss, strike slip, ds, dip slip,
R, reverse sense, N, normal sense. See also Hill (1959).
Kinematic indicators in shear zones 77
shear-zone rocks; ubiquitous veins and mineral-
filled fractures are developed and in turn deformed.
The shear-zone rocks were thus both highly porous
and permeable at least intermittently during their
active history.
Transitional shear zones form and accommo-
date most deformation in a temperature range
where many mineral solubilities (including quartz
and gold) show abrupt decreases. For example, if
gold-saturated fluids percolate up through the
porous and permeable transitional shear zone and
cool as they ascend, then at some level in the shear
zone the gold should precipitate (solubility data
indicate this should occur over a very narrow tern
perature interval - i.e .. , restricted depth range).
The porous, permeable transitional shear zones can
also act as conduits for mineralized fluids, effi-
ciently delivering them into structural traps at
higher levels. This seems true for Archaean lode-
gold deposits (e.g., Roberts, 1987; Sibson and others,
1988) and slate-belt gold deposits (e.g., Mawer,
1987b).
Transitional shear zones are loci for large
earthquakes (M
L
> 5.5); they straddle the level of
peak crustal shear resistance and bracket the seis-
mogenic regime in continental crust (Sibson, 1977,
1986). Study of these shear zones has relevance to
earthquake analysis and to understanding crustal
rheology and tectonic stress levels.
It can be difficult to determine whether a given
shear zone is truly of the transitional type - that is,
one in which brittle, solution-transfer and crystal-
plastic processes all occurred at the same time.
Shear zones are places where deformation tends to
be concentrated for long periods of time (e.g.,
Watterson, 1975), and one may therefore have early
plastic structures overprinted by brittle structures
at a much later time, or vice versa.
Shear zones loosely termed "transitional" or
"brittle-ductile" may develop in several ways.
(1) Brittle, solution-transfer, and crystal-plastic
processes may occur sychronously and cyclically -
either a plastically deforming rock becomes tran-
siently hardened (by various strain-hardening
processes) and/or brittle (by locally increasing strain
rate, or increasing applied stress), or effective nor-
mal stresses are momentarily lowered by increasing
fluid pressures. These are the only true transitional
shear zones. (2) Mylonitic shear zones may be up-
lifted and cooled - deformation continues and
brittle structures overprint plastic structures.
(3) An old inactive zone of plastic deformation may
be reactivated at a much later time in an
unconnected deformational episode (with or without
some fluid influx).
For truly transitional shear zones, diagnostic
structures occur which show that brittle, solution-
transfer, and crystal-plastic processes were syn-
chronous and cyclical. One example is that cata-
clastic rocks may be plastically deformed and then
overprinted by more fracturing. Another criterion,
and one that forms a series of good kinematic
indicators, is deformed asymmetric veins (Figures
13-17). These are ubiquitous in transitional shear
zones, commonly are composed of either quartz or
calcite, and have become the focus for much study in
recent years (e.g., Cox and others, 1986; Mawer,
1987b; Sibson and others, 1988).
The shapes and asymmetry of such deformed
quartz veins are commonly used to determine shear
sense (e.g., Beach, 1975; Knipe and White, 1979;
Ramsay, 1980). The veins typically form as infilled
extension fractures oblique to the shear-zone boun-
daries, and with continued deformation they become
folded and rotated to sigmoidal shapes (Figure 13).
Several sets of veins commonly overprint one
another in these shear zones, to form an asymmetric
vein array, which is also a good kinematic indicator
(Figures 13 and 14). These types of vein arrays
represent the essence of transitional shear zones.
New veins and vein segments develop oblique to the
shear-zone boundaries by hydraulic fracturing and
precipitation of vein material from solution. For
simple shearing, the obliquity is 45. Shearing
continues, veins and segments are deformed and
rotated by both dissolution-precipitation and crys-
tal-plastic processes; as shearing proceeds, further
new hydraulic fractures form and more veins and
segments develop (Figure 14). The synchronous
and cyclical operation of brittle/solution-trans-
fer/crystal-plastic deformational mechanisms is
particularly clear in these cases.
Other vein-associated kinematic indicators may
be developed in these situations. These depend on
the length to width ratio of the veins, and several
are associated with the vein walls. Some of these
effects are subtle but can generally be observed in
the field with a hand lens.
Long, thin veins be.come folded by buckling as
they rotate during shearing (Figure 15) and under-
go a shear of the reverse sense to the overall zone,
resulting in minor folds of the vein with apparently
the wrong asymmetry for the major zone. These
sorts of structures are relatively common in
78
C. K.Mawer
Figure 13. Sketches of deformed sigmoidal quartz veins in
transitional shear zones at Cornwall, England, both
indicating dextral shear sense (traced after photographs of
Ramsay and Huber, 1987). Note that in A the shear zone
has increased volume considerably (by introduction of the
quartz veins), whereas in B the shear zone has either
maintained constant volume or has lost volume (shown by
the presence of sigmoidal dissolution seams trending lower
left to upper right). Scale bars both = 2 cm. These sorts of
asymmetric deformed veins and vein arrays are reliable
kinemetic indicators.
greenschist-facies shear zones. Short, thick veins
tend to rotate as relatively rigid bodies, a process
that can result in en echelon arrays which ap-
parently indicate a sense of shear opposite to the
correct one. However, vein walls studied in detail
show that the vein includes phyllosilicate grains
which originally grew parallel to the wall-rock
foliation (Figure 16; Durney and Ramsay, 1973;
Mawer, 1987b). As the vein rotates, this foliation is
folded at the vein wall, and asymmetry of these
folds, the included phyllosilicates, and the wall-rock
foliation form a useful kinematic indicator. Several
other small-scale kinematic indicators also form as
a result of this vein rotation during shearing
(Figure 17), including shear bands in the veins,
shear-band foliation at the vein walls, and offset of
original marker layers across the veins (Mawer and
Williams, 1986).
A.
....
s. ,

c ,
.-----_ _ -----, it
...
Figure 14. Diagram illustrating developmental sequence
for asymmetric deformed veins and vein arrays (after
Durney and Ramsay, 1973). Newest veins and vein seg-
ments at any stage are black, older are stippled. Note
overprinting and asymmetry of individual veins, and
overall asymmetry of vein array.
....
J
.......
Figure 15. Formation of asymmetric folds by buckling and
rotation of long, thin vein in dextral shear (after a field
example, Nova Scotia, Canada). See text for discussion.
Kinematic indicators in shear zones 79
A
B.
--
.!l . .

--
10m
Figure 16. Formation of asymmetric structures (folds in
overgrown phyllosilicates and vein-margin foliation) at
margins of short, thick veins due to their rotation in
dextral shear. A) Diagram traced from polished slab and
thin section (inset), both cut parallel to principal plane of
observation, transitional shear zone in Nova Scotia,
Canada. B) Diagram showing sequence of development of
this kinematic indicator. See text for discussion.
CONCLUSIONS
Shear zones are important architectural ele-
ments in the Earth's lithosphere (e.g., Ramsay,
1980; Kirby, 1985). Their shear sense can be un-
ambiguously interpreted by the judicious use of
kinematic indicators. There are many suitable
types of asymmetric structures; doubtless more will
be recognized in future years.
This contribution has concentrated on the most
common reliable meso- and micro scale kinematic
indicators. The principles used in this paper can be
directly applied to macroscopic features and to other
types of kinematic indicator not described above.
ACKNOWLEDGMENTS
I thank both G. Mitra and an anonymous
reviewer for their painstaking reviews, even though
they disagree with some of my statements. Re-
search was partly supported by Sandia National
Figure 17. Asymmetric structures associated with
rotation of short, thick quartz vein in dextral shear, a
summary of features developed in rocks adjacent to the
Chedabucto Fault Zone of Nova Scotia, Canada. 1, Asym-
metric arrangement of mica grains in matrix; 2, rotated
syntectonic porphyroblasts; 3, offset of marker bed across
rotated vein; 4, folds at vein margin (see Figure 16);
5, shear band transecting vein; 6, shear-band foliation
developed at vein margin during its clockwise rotation.
Sequence of vein rotation as Figure 16. See text for dis-
cussion.
Laboratories (via SURP 06-5629, Task 7) and the
U.S. National Science Foundation (grants EAR-
8721264 and EAR-8816402).
REFERENCES
ALLISON, 1., 1982, On the geometric similarity and scale of
shear zones: Mitteilungen aus dem Geologischem Institut
der ETH Zurich, Neue Folge 239 a, p. 12-14.
BEACH, A., 1975, The geometry of en echelon vein arrays:
Tectonophysics, v. 28, p. 245-263.
BERTHE, D., P. CHOUKROUNE, and P. JEGOUZO, 1979,
Orthogneiss, mylonite and non-coaxial deformation of
granites: The example of the South Armorican shear zone:
Journal of Structural Geology, v.l, p. 31-42.
BRODIE, K. R., and E. R. RUITER, 1987, Deep crustal
extensional faulting in the Ivrea zone of northern Italy:
Tectonophysics, v. 140, p. 193-212.
80
C. K. Mawer
BROWN, L. D., 1986, Aspects of COCORP deep seismic
profiling; pp. 209-222 in M. Barazangi and L. Brown (eds.),
Reflection Seismology: A Global Perspective: Ameri-
can Geophysical Union, Geodynamics Series 13, 311 p.
CHESTER, F. M., J. FRIEDMAN, and J. M. LOGAN, 1985,
Foliated cataclasites: Tectonophysics, v. 111, p. 139-146.
COBBOLD, P. R, D. GAPAIS, W. D. MEANS and S. H.
TREAGUS (eds.), 1987, Shear criteria in rocks: Journal of
Structural Geology, v. 9, p. 521-778.
COX, S. F., M. A. ETHERIDGE, and V. J. WALL, 1986, The
role of fluids in syntectonic mass transport, and the
localization of metamorphic vein-type ore deposits: Ore
Geology Reviews, v. 2, p. 65-86.
DURNEY, D. W., and J. G. RAMSAY, 1973 Incremental
strains measured by syntectonic crystal growths; pp. 67-96
in K. A. DeJong and R Scholten, (eds.), Gravity and
Tectonics: John Wiley and Sons, New York, New York,
USA,502p.
FRIEND, C. R L., A. P. NUTMAN, and V. R McGRE-
GOR,1987, Late-Archean tectonics in the Faeringhavn-Tre
Brodre area, south of Bukesfjorden, southern West Green-
land: Journal of the Geological Society of London, v. 144, p.
369-376.
GALLAGHER, .J. J., Jr., 1981, Tectonics of China: Con-
tinental scale cataclastic flow; pp. 259-273 in N. L. Carter,
M. Friedman, J. M. Logan, and D. W. Stearns (eds.),
Mechanical Behavior of Crustal Rocks: American Geo-
physical Union, Geophysical Monograph 24,326 p.
HANCOCK, P. L., and A. A. BARKA, 1987, Kinematic
indicators on active normal faults in western Turkey:
Journal of Structural Geology, v. 9, p. 573-584.
HANMER, S. K., 1986, Asymmetrical pull-aparts and
foliation fish as kinematic indicators: Journal of Struc-
tural Geology, v. 8, p. 111-122.
HILL, M. L., 1959, Dual classification of faults: American
Association of Petroleum Geologists Bulletin, v. 43, p. 217-
237.
KARIG, D. E., D. R SAREWITZ, and G. D. HAECK, 1986,
Role of strike-slip faulting in the evolution of alloch-
thonous terranes in the Philippines: Geology, v. 14, p. 852-
855.
KillBY, S. H., 1985, Rock mechanics observations pertinent
to the rheology of the continental lithosphere and the
localization of strain along shear zones: Tectonophysics, v.
119, p. 1-27.
__ , C. MAWER, G. ITURRINO, and N. CHRISTEN-
SEN,1989, Ductile shear zones in Layer 3 gabbroic rocks:
Their deformation structures, Vp anisotropy and candi-
dacy for the curvilinear reflection structures of the lower
crust: EOS, Transactions of the American Geophysical
Union,v.70,p.460.
KNIPE, R J., and S. H. WHITE, 1979, Deformation in low-
grade shear zones in the Old Red Sandstone, S.W. Wales:
Journal of Structural Geology, v. 1, p. 53-66.
LISTER, G. S., and A. W. SNOKE, 1984, S-C mylonites:
Journal of Structural Geology, v. 6, p. 617-638.
LOGAN, J. M., M. FRIEDMAN, N. HIGGS, C. DENGO, and T.
SHIMAMOTO, 1979, Experimental studies of simulated
gouge and their application to studies of natural fault
zones; pp. 305-343 in Analysis of Actual Fault Zones in
Bedrock: U.S. Geological Survey, Open File Report 79-
1239.
MANDELBROT, B. B., 1983, The Fractal Geometry of
Nature: W. H. Freeman and Company, New York, New
York, USA, 468 p.
MANDL, G., 1988, Mechanics of Tectonic Faulting:
Elsevier, Amsterdam, The Netherlands, 407 p.
MAWER, C. K., 1986, What is a mylonite?: Geoscience
Canada, v. 13, p. 33-34.
__ , 1987a, Shear criteria in the Grenville Province,
Ontario, Canada: Journal of Structural Geology, v. 9, p.
531-539.
__ , 1987b, Mechanics of formation of gold-bearing
quartz veins, Nova Scotia, Canada: Tectonophysics, v. 135,
p.99-119.
__ , and J. C. WHITE, 1987, Sense of displacement on the
Cobequid-Chedabucto fault system, Nova Scotia, Canada:
Canadian Journal of Earth Sciences, v. 24, p. 217-223.
__ , and P. F. WILLIAMS, 1986, Structural study of
highly deformed Megma phyllite and granite, vicinity of
White Head Village, S.E. Nova Scotia: Maritime Sedi-
ments and Atlantic Geology, v. 22, p. 51-64.
PASSCHIER, C., and C. SIMPSON, 1986, Porophyroclast
systems as kinematic indicators: Journal of Structural
Geology, v. 8, p. 831-843.
PERCIVAL J. A., and K. D. CARD, 1983, Archean crust is
revealed in the Kapuskasing uplift, Superior Province,
Canada: Geology, v. 11, p. 323-326.
PETIT, J. P., 1987, Criteria for the sense of movement on
fault surfaces in brittle rocks: Journal of Structural
Geology, v. 9, p. 597-608.
QUINQUIS, H., C. AUDREN, J. P. BRUN, and P. R.
COBBOLD, 1978, Intense progressive shear in the Ile de
Groix blueschists and compatibility with subduction or
obduction: Nature, v. 273, p. 43-45.
RAMSAY, J. G., 1967, Folding and Fracturing of Rocks:
McGraw-Hill, New York, New York, USA, 568pp.
Kinematic indicators in shear zones 81
__ , 1980, Shear zone geometry: A review: Journal of
Structural Geology, v. 2, p. 83-99.
__ , M. CASEY, and R. KLIGFIELD, 1983, Role of shear in
development of the Helvitic fold-thrust belt of Switzerland:
Geology, v. 11, p. 439-442.
__ , and R H. GRAHAM, 1970, Strain variation in shear
belts: Canadian Journal of Earth Sciences, v. 7, p. 786-813.
__ , and M. 1. HUBER, 1987, Techniques of Modern
Structural Geology, Volume 2: Folds and Fractures:
Academic Press, London, England, UK, 700 p.
RIEDEL, W., 1929, Zur Mechanik geologischer Brucher-
scheinungen: Zentralblatt fur Mineralogie, Geologie and
Palaontologie, v. 1929B, p. 354-368.
ROBERTS, R G., 1987, Ore deposit models #11. Archean
lode gold deposits: Geoscience Canada, v. 14, p. 37-52.
SANDER, B., 1948, Einfuhrung In Die Gefugekunde
DerGeologischen Korper, Erster Teil: Springer-Verlag,
Vienna and Innsbruck, Austria, 215 p.
SANDERSON, D. J., 1979, The transition from upright to
recumbent folding in the Variscan fold belt of southwest
England: A model based on the kinematics of simple
shear: Journal of Structural Geology, v. 1, p. 171-180.
SrnSON, R H., 1977, Fault rocks and fault mechanisms:
Journal of the Geological Society of London, v. 133, p. 190-
213.
__ , F. ROBERT, and K. H. POULSEN, 1988, High-angle
reverse faults, fluid-pressure cycling, and mesothermal
gold-quartz deposits: Geology, v. 16, p. 551-555.
SIMPSON, C., and S. M. SCHMID, 1983, An evaluation of
criteria to deduce the sense of movement in sheared rocks:
Geological Society of America Bulletin, v. 94, p. 1281-1288.
SKJERNAA, L., 1985, Some possible ways to develop sheath
folds [abstract]: Abstracts of the International Conference
on Tectonic and Structural Processes (Utrecht, The
Netherlands), p. 123.
SORENSEN, K., 1983, Growth and dynamics of the Nordre
Strornfjord shear zone: Journal of Geophysical Research, v.
88, p. 3419-3487.
TCHALENKO, J. S., 1970, Similarities between shear zones
of different magnitudes: Geological Society of America
Bulletin, v. 81, p. 1625-1640.
TULLIS, J., A. W. SNOKE, and V. R TODD, 1982, Signi-
ficance and petrogenesis of mylonitic rocks (Penrose
Conference Report): Geology, v. 10, p. 227-230.
WATTERSON, J., 1975, Mechanism for the persistence of
tectonic lineaments: Nature, v. 253, p. 520-522.
WEIJERMARS, R, and H. E. RONDEEL, 1984, Shear band
foliation as an indicator of sense of shear: Field observa-
tion in central Spain: Geology, v. 12, p. 603-606.
WERNICKE, P., 1985, Uniform-sense normal simple shear
of the continental lithosphere: Canadian Journal of Earth
Sciences, v. 22, p. 108-125.
WHITE, S. R., P. G. BRETAN, and E. R. RUTTER, 1986,
Fault-zone reactivation: Kinematics and mechanisms:
Philosophical Transactions of the Royal Society of London,
v. A317, p. 81-97.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Zone of weakness concept: A review and evaluation
JOHN JAMES PRUCHA
Department of Geology, Syracuse University, Syracuse, New York 13244-1070, USA
(received August 5,1988; revision accepted January 19, 1989)
ABSTRACT
The zone of weakness concept implies that there exist within the earth's crust relatively weak zones
which when differentially stressed will fail more readily than will the surrounding rock. Such zones are often
inferred to determine the extent, orientation and style of younger deformation. Although the concept has been
around a long time and is intuitively attractive, our ability to predict and explain younger deformation
patterns based upon knowledge of older fabrics is poor. The concept is not rooted in adequate knowledge of
specific strain mechanisms.
Basement fabric elements which have been inferred to produce weak zones include faults, joints, foliation,
cleavages, folds, shear zones, plate sutures, lithologic discontinuities, and basement terrane boundaries.
Within this array of fabric elements one must consider a multiplicity of potential operating mechanisms:
shear and extension fractures, crystal gliding, cataclastic flow, crystal dislocations, recrystallization, pressure
solution. Each is conditioned by one or more environmental factors: temperature, confining pressure, stress
fields, pore pressures, strain rates. Experimental deformation of basement rocks reveals complex
relationships and mechanisms. Mechanisms indigenous to the ductile realm do not operate so well, if at all, in
the brittle realm. Thus inferences of syntectonic basement mechanisms being reactiviated to propagate
structures at higher levels may be unwarranted.
Reactiviation of some structures is well-established by field studies, but control of younger structures by
older fabric anisotropies is not. The complexity of tectonic histories, and of the dynamic systems which
determine them, limits our ability to predict younger deformation patterns based upon knowledge of older
fabrics.
Plate tectonics offers a conceptual framework within which confirmed dynamic principles provide a basis
for inferring younger strain patterns and mechanisms within the upper crust. In spite of many complicating
factors, the ability to predict strain patterns and structural assemblages of a given generation, and to
extrapolate accurately from limited data, is more readily achievable within the context of plate tectonics
modeling than it is by calling upon the tenuous zone of weakness concept.
INTRODUCTION AND EARLY CONCEPTS
This essay reviews some early antecedent ideas
of the zone of weakness concept and examines
various aspects of the concept itself. The essay
concludes that, at present, the concept is not useful
as a working tool because it is not rooted in
adequate knowledge of specific strain mechanisms.
Our understanding of the mechanisms and controls
of reactivation of older structures is too limited to
warrant uncritical application of the concept in
detailed interpretation of regional strain history.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 83-92. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
83
84 J. J. Prucha
Systematic through-going fractures in the
earth's crust were noted by many early workers
(Hobbs, 1911, p. 163-165; Sonder, 1938, 1956; Umb-
grove, 1947, p. 294-320; Vening Meinesz, 1947;
Cloos, 1948). Such features were generally con-
sidered to be of primordial origin and were believed
to determine the loci of periodic rejuvenation and
displacement during the secular history of the
crusted earth.
Hans Cloos (1948) believed that early in the
earth's history the crust was divided by "geo-
fractures," or "geosutures," into polygonal blocks.
The shape of the blocks and renewed movement on
the ancient deep-seated breaks were considered to
have influenced subsequent episodes of deformation.
Sonder (1956) discussed global strike-slip dis-
placements, the production of which he called
"regmagenesis" (Rheomagenese). Regmatic shear
zones were believed to form a world-wide network
(rhegmatische Klu{tnetz) in response to simple glo-
bal stress patterns.
DeSitter (1964, p. 161-165) recognized the oc-
currence worldwide of what he called "great funda-
mental faults." These he considered to be great
cracks in the earth's crust along which repeated
displacements occurred in such a way as to obscure
the original dynamic character of the breaks. He
suggested that perhaps they never did fall into one
of the distinct categories - normal, thrust, and
wrench. Examples which he gave and elucidated
include the North Pyrenean fault zones of south-
eastern France and the Insubric-Tonale-Pusteria
line of the Alps. He considered such features to be
" .. .long fundamental cracks in the earth's crust
penetrating the whole upper crust and characterized
by frequent movements. They occasionally con-
stitute a way of access to the surface for basic or
granite magma. Their particular tectonic function
at a particular moment in geological history
depends upon the stress field prevailing at that
moment. They may facilitate radial displacements,
or they may act as wrench faults or as great zones of
compression, as the stress field dictates."
Consonant with these ideas is a somewhat later
contribution by Wise (1969). Working with 250
plastic relief maps which he illuminated by low-
angle incident light, he depicted topographic linea-
ments belonging to one or another of six to eight sets
which persisted over distances of up to 200 miles
each. He concluded that complex fracture networks
occur with constant orientation over vast areas of
Europe, Iceland, and North America. Some ele-
ments of the networks correlate with known frac-
ture or fault systems; others do not. The linears
extend without change to the edges of continents;
they occur in the youthful crust of Iceland; they are
independent of local curvatures and local geome-
tries of the mountain ranges in which they occur. In
North America they continue from the ancient
Appalachians to the youthful Coastal Plain.
Wise concluded that these relationships suggest
either that near-modern stress trajectories in the
crust occur with constant orientation over very
large areas or that a "tectonic heredity" of older
strain patterns was propagated upward through
younger sedimentary cover. Earlier (1958, p. 24),
Wise applied the term "tectonic heredity" to the
concept of a " ... controlling relation between ancient
basement anistropism and Laramide structural
trends ... " as it pertained to the foreland structures of
the Middle Rocky Mountains of Wyoming and
Montana.
According to Wise (1969, p. 176), "tectonic here-
dity implies that a pattern of numerous potential
fracture directions can be set into basement soon
after crustal solidification." He suggested that suc-
cessive events of regional crustal extension would be
particularly conducive to reactivation of earlier
fracture trends.
In both of his papers cited above, Wise empha-
sized that the concept of tectonic heredity has very
limited applicability in elucidating younger strain
patterns in sedimentary cover rocks lying above
deformed basement. More recently (D. U. Wise,
1988, written communication, 1988) he reaffirmed
his view that the regional fracture patterns of the
Mediterranean region " ... had to be reflections of
near-modern stress trajectories rather than indica-
tors of tectonic heredity from basement."
The notion of tectonic heredity is consistent
with the clearly stated views of Hodgson (1965), who
concluded that structures in the Paleozoic and
Mesozoic sedimentary rocks of selected areas in the
Colorado Plateau and the Wyoming foreland follow
elements of an ancient fracture pattern that was
established in Precambrian time within crystalline
rocks of the basement.
Hoppin and Palmquist (1965, p. 1000), in cont-
rast, found in the Bighorn Mountains of Wyoming,
that " ... there is a difference between fracture pat-
terns of the basement and those of the adjacent
cover rocks, that this difference is highly signifi-
cant, and that it has an important bearing on the
determinations of the origins and ages of the frac-
Zones of weakness concept: A review and evaluation 85
tures." They conclude that basement anisotropy in
some cases may be responsible only for minor fea-
tures which modify the principal structures that
were determined by stress-field orientations and the
three-dimensional geometry of the younger cover
rocks.
These are but a few early examples of inter-
pretations of the basic theme. Today's geologic
literature is replete with countless repetitions of the
same idea, though few offerings attempt any critical
analysis or present substantial numbers of data in
support of the zone of weakness concept.
NATURE AND ROLE OF THE BASEMENT
The zone of weakness concept is one of several
dominant themes that run through the subdisci-
pline of basement tectonics. The habitat of such
zones is usually inferred to be the basement, which
is older and typically subjacent to younger rocks
whose own deformation occurred following estab-
lishment of the underlying structural fabric. On the
one hand, the fact that so much of the basement is
covered by younger rocks makes elucidation of
cause-and-effect relationships difficult; on the other
hand, it makes it easier and more convenient to
make the intuitively appealing inference that the
control of younger deformation was determined by
pre-existing basement structures.
Definition
The term basement has been defined variously,
but I still prefer the definition which describes it as
"".those igneous and metamorphic rocks which
underlie unconformably the unmetamorphosed,
dominantly sedimentary rocks of a region. The un-
metamorphosed rocks overlying the basement do
not everywhere consist exclusively of sedimentary
rocks, but they may in some places include volcanic
flows, pyroclastic beds, and concordant and dis-
cordant hypabyssal rocks" (Prucha and others, 1965,
p 967). The term basement thus lacks an age
connotation (though in North America it is mostly
Precambrian), but it does indirectly carry a mech-
anical connotation. At least above any brittle/duc-
tile transition which may exist at some level within
the crust, the basement generally behaves as a
strong, brittle material (Borg and Handin, 1966)
relative to a somewhat more ductile sequence of
overlying layered rocks. The basement/sedimentary
rock interface is thus a mechanical discontinuity of
considerable importance.
In the Rocky Mountains foreland of south-
western Montana, Schmidt and Garihan (1983, p.
290-291) described well-foliated Archean metasedi-
mentary sequences of gneisses and amphibolites
which have been folded concordantly with Paleozoic
sedimentary rocks. They properly raise the ques-
tion of whether these metamorphic rocks meet the
mechanical connotation proposed here for the base-
ment in contrast with overlying unmetamorphosed
sedimentary rocks. Their clear discussion of the
problem underscores the difficulty of applying pre-
cise limiting definitions to parts of the continuum
which is nature.
Mechanical Behavior
The mechanical discontinuity of the basement/
sedimentary rock interface is a matter of some
significance in any attempt to infer reactivation of
strain mechanisms that are genetically related to
basement structural fabrics. Every strain mech-
anism in deforming rocks is related both to the
inherent structure of the constituent minerals and
to the ambient conditions, or environment, at the
time of operation. The most important environmen-
tal parameters are temperature and confining
pressure, whose effects may be modified by strain
rates and by abnormal pressures of pore fluids.
Reactivation of ancient basement zones of yielding
or rupture typically is inferred for later times in
which the environment is characterized by signi-
ficantly lower energy levels. Mechanisms that are
indigenous to the ductile realm of regional meta-
morphism in which the original basement rocks and
their fabrics were produced may not operate so well,
if at all, in the brittle realm.
Experimental work by Borg and Handin (1966)
is pertinent. Triaxial compression tests on 18 typi-
cal basement rocks were run at conditions simu-
lating depth of burial of the basement in the Rocky
Mountains foreland during the Laramide orogeny
(100 MPa confining pressure, 150C). The rocks
generally behaved similarly and as strong and
brittle materials. Maximum stress differences sus-
tained prior to the faulting were in the range of 500
to 800 MPa, and strain did not exceed 4% before
failure. The similarity of mechanical response and
failure characteristics was established in spite of the
86 J.J.Prucha
fact that the rocks tested ranged widely in mineral
composition and fabric. Conspicuous fabric aniso-
tropies and lithologic inhomogeneities established
under the deep-seated conditions that originally
produced the basement rocks did not show signi-
f1cant differences in mechanical response to differ-
ential stress at the simulated depths of shallower
burial. Exceptions were the results of tests run on
highly micaceous rocks, in which lower shear
strength and higher ductility are probably related to
the low critical resolved shear stress (ca. 10 MPa)
needed to produce intragranular slip in biotite.
Fabric Elements
Basement fabric elements which at one time or
another have been inferred to produce zones of
weakness include faults, joints (both mega- and
micro-), foliation, cleavages, folds, shear zones, plate
sutures and transform faults, lithologic discon-
tinuities, terrane boundaries. Within this array of
fabric elements one must consider a multiplicity of
potential operating mechanisms: shear and exten-
sion fracturing, crystal gliding, cataclastic flow,
crystal dislocations, recrystallization, pressure solu-
tion. Each is conditioned by one or more environ-
mental factors. Thus, in specific cases characterized
by large differences between original environment
and environment of reactivation, it may be un-
warranted to infer that pre-existing syntectonic
basement mechanisms propagated structures to
higher crustal levels.
PRE-EXISTING ANISOTROPIES
Analysis of Fault Reactivation
On of the most common examples used in sup-
port of the zone of weakness concept is reactivation
of a pre-existing fault within the brittle realm. For
such examples it is tacitly assumed that the pre-
existing fault (typically within the basement)
consists of a simple surface across which all cohesion
has been lost. Thus the condition for renewed slip
on the surface would be
L = all [1]
in which L is the tangential component of stress on
the surface, a is the normal component of stress
across the surface, and 11 is the coefficient of sliding
friction for rock on rock. This condition for renewed
faulting by frictional sliding contrasts with the
condition for initial fracturing, which is
L = LO + a tan <1> [2]
where tan <1> is th so-called internal friction, which is
characteristic of the rock for the given environmen-
tal conditions of termperature and confining pres-
sure, and LO is the cohesive shear strength. Older
faults which have been cemented by secondary
mineralization, or which have been frictionally
locked by comminuted rock ranging from gouge to
fault breccias, may be in an analogous condition
best expressed by
L = LO* + a tan <1>* [3]
where LO* and tan <1>* are the new cohesive shear
strength and the new coefficient of internal friction,
respectively.
Stearns and others (1981, p. 216-220) empha-
sized the importance of these relationships to the
question of fault reactivation and went on to cite
earlier experimental work (Handin and Stearns,
1964; Handin, 1969) which showed that whether
reactivation of an older fault or the creation of a new
fault occurs in a given instance depends upon the
inherent properties of the rocks involved and upon
the orientation of the principal stresses relative to
pre-existing S surfaces. This latter point is in accord
with deSitter's inferences about the dynamic rela-
tionships of reactivated "great fundamental faults,"
cited above.
Experimental Studies
Even the simplest kind of stress-strain tests in
experimental deformation reveal complex relation-
ships and mechanisms. To be understood they must
be considered in the context of varying strain rates,
temperatures, confining pressures, and angular
relationships of stress fields. Futhermore, the
mechanisms operating in response to the defined
environmental parameters change sequentially dur-
ing the total strain event.
Early experimental work by F. A. Donath has
special relevance for the zone of weakness concept.
In an experimental study of the development of kink
bands in brittle anisotropic rock (Donath, 1969, p.
491), he summarized results which showed that for
the Martinsburg slate" ... modes of deformation ... are
clearly related to the inclination of anisotropy to the
direction of maximum compression (al), as well as to
the environmental conditions at the time of de-
formation. At confining pressures below about 1000
Zones of weakness concept: A review and evaluation 87
bars for slate, faulting across the anisotropy occurs
in specimens inclined more than 60
0
to 01. For
orientations between a few degrees to about 60
0
to
01, simple gliding on the anisotropy occurs."
Donath (1969, p. 488) pointed out that the for-
mation of kink bands in experiments in which the
foliation of the Martinsburg slate was oriented 15
0
from the axis of maximum compression was not just
the result of simple gliding on the cleavage:
"Rather, the development of kink bands in brittle
anisotropic rocks appears to reflect the operation of
several mechanisms - namely, gliding on cleavage
accompanied by cataclasis, definition of kink planes
along planes of high shear stress, and rotation of the
foliation segments between the kink planes, with
slip, cataclasis, and dilatation all occurring within
the kink band."
In experimental studies of faulting in Martins-
burg slate, Donath (1961) showed that at low
confining pressures (3-50 MPa) the slaty cleavage
has a significant effect upon the angle of shear
fracture. The angle of shear fracture was dependent
upon the angle of inclination of the foliation with
respect to the orientation of the principal stresses.
The shear fractures developed parallel to the slaty
cleavage when the greatest principal pressures was
oriented in the 15
0
, 30
0
, and some 45
0
positions
relative to the foliation. Some effect of the cleavage
was present in the 60
0
and 70
0
orientations but not
in the 90
0
position. Except for the 90
0
orientations,
the strike of the shear fracture was parallel to that
of the cleavage. The effect of the anisotropy on the
angle of the shear fractures was reduced by in-
creased confining pressure.
In other experiments with Martinsburg slate,
Donath (1964, p. 284-288) determined that in failure
by shear fracture at low confining pressures (3.5,
10.5,35 MPa) the foliation had a marked effect upon
the breaking strength. The effect of the planar an-
isotropy was also pronounced at 50, 100, and 200
MPa confining pressures, although in some cases
deformation was by faulting without loss of cohe-
sion. He clearly showed that the breaking strength
of the material depends upon the orientation of the
anisotropy relative to the axis of greatest com-
pression.
Borg and Handin (1966, p. 249) showed that
under conditions of deep burial (500 MPa c.p. and
500C) the Fordham Gneiss (Precambrian) and the
Mettawee Slate (Cambrian) " ... are appreciably
weaker when the load is applied at 45
0
to the planar
anisotropy defined by the statistical parallelism of
the micas, and the faults lie close to the foliation and
the 45
0
position of maximum shear stress." Bend
gliding on the basal pinacoid of biotite within the
foliation surfaces is the specific deformation mech-
anism. In noting that the strengths of the gneiss
and slate when loaded obliquely to the planar an-
isotropies are lower than for loads normal and
parallel to the foliation, Borg and Handin (1966, p.
317) attributed the difference to the surfaces of low
internal friction which are provided by the statis-
tically parallel alignment of the platy mica grains.
Field Studies
Field studies in areas of basement outcrops with
partial cover of sedimentary rock provide oppor-
tunities to test the possible validity oflong-standing
ideas about cause-and-effect relationships between
older basement fabrics and younger deformation.
Granted, this approach has many built-in limita-
tions. These include: (1) limited access to the third
dimension; (2) poor understanding of relative times
of formation of different structural elements;
(3) lack of knowledge of stress-field orientations at
various levels and times within the earth's crust;
and (4) the inability, generally, to assess the degree
of mechanial anisotropy provided by a given fabric
element developed in a former environment. (Geo-
physical approaches, such as gravity and magnetic
surveys, have their own built-in limitations.)
Hoppin and Palmquist (1965) provided a thoughtful
assessment of the difficulties inherent in resolving
the problem by detailed field studies.
My colleagues and I reported long ago (Prucha
and others, 1965, p. 689-690) our inability to estab-
lish on the basis of detailed field studies in the
Rocky Mountains foreland a consistent relationship
between Laramide structures and basement struc-
tures of Precambrian age. Subsequent work has
reinforced that view. In spite of widely different
basement lithologies and fabrics in such widely
separated areas as the Front Range of Colorado, the
Sangre de Cristo Range of northeastern New Mexi-
co, the Shirley Mountains of Wyoming, and the
Uncompahgre Uplift of Colorado and Utah, the
basic style of Laramide deformation is the same
throughout.
Our conclusions about control of regional
structural style by older basement fabrics are
consistent with those presented by Rodgers (1987, p.
667-668) in a recent discussion of the Rocky Moun-
88 J.J.Prucha
tains foreland: "The relation of the Laramide struc-
tures to preexisting structures in the underlying
Precambrian basement is controversial. The curi-
ously irregular plan of many of the ranges and such
phenomena as the right-angle bends in the mono-
clines suggest control from beneath, for nothing in
the sedimentary cover could provide the necessary
anisotropies or inhomogeneities. On the other hand,
taken as a whole the eastern Rockies form a broad
arc from the salients in front of the Western Rockies
in the northern United States toward the salient in
what may be the same front at the north end of the
Sierra Madre Oriental of Mexico, and the arc cuts
with superb indifference across a whole series of
well-established Precambrian age provinces, whose
boundaries trend northeast or east-northeast and
can in some cases be followed under the platform
cover to the east until they crop out again in the
Canadian shield." He concludes "".that many de-
tails of the pattern of the ranges are controlled by
older structures, but that the overall pattern is not."
In spite of ambiguities which often stem from
field work, some detailed field studies yield con-
vincing evidence for reactivation along older struc-
tures. A good example is Ratcliffe's (1971) study of
the Ramapo fault system in New York State and
adjacent northern New Jersey.
The Ramapo fault forms the northwestern
boundary of the Newark basin, and for much of its
length it separates the Late Triassic rocks from the
Precambrian complex of the Hudson Highlands.
Ratcliffe concluded, principally on the bases of field
evidence, that the fault is part of a fundamental
crustal fracture system that was intermittently
active during as much as 700 million years of Earth
history. He proposed for the Ramapo fault the
following tectonic history (1971, p. 125 and 139):
1. Late Precambrian faulting with concomitant
intrusion of a quartz-monzonite and granodio-
rite pluton followed by post-tectonic intrusion of
part of a dioritic pluton.
2. Early Ordovician block faulting (SE-side-down)
resulting in an unconformity between Precam-
brian gneisses and a Middle Ordovician gneiss-
boulder conglomerate.
3. Renewed normal (?) faulting and intrusion of
diorites in Middle to Late Ordovician.
4. Right-lateral wrench faulting in post-Middle
Ordovician to pre-Middle Devonian rocks.
5. Repeated normal movements during the Late
Triassic to produce coarse clastic fanglomerates
of the Newark Series. This episode may have
extended into the Jurassic.
6. Syndepositional faulting followed by later
downdropping along the Ramapo fault south of
the Hudson River.
7. Recent seismic activity, which suggests that the
fault is still active.
Ratcliffe pointed out (1970, p. 139) that, with
the possible exception of the late Precambrian right-
lateral displacement, the history of movement on
the Ramapo fault is generally in accord with the
plate-tectonics model presented by Bird and Dewey
(1970) to explain the origin of the northern Appala-
chians.
Another good example offield-documented fault
reactivation is summarized by Schmidt and Garihan
(1983, p. 273-276) for the Rocky Mountains foreland
in southwestern Montana. There, a Proterozoic
antecedence is established for a system of NW-
trending, steeply dipping Laramide faults that were
variously active from Late Cretaceous to late Paleo-
cene. Evidence for Precambrian displacement in-
cludes (a) separations of Archean metamorphic
units which are significantly different from separa-
tions of Paleozoic sedimentary rocks, (b) distinctly
different metamorphic assemblages on opposite
sides of a fault (Carmichael fault) immediately
below the unconformable Flathead Sandstone, and
(c) intrusion along the faults by diabase dikes which
have been radiometrically dated as Middle Protero-
zoic (Wooden and others, 1978). Evidence for com-
parative slip on the reactivated faults is not as
compelling, but it seems clear that in this example
too the displacement dynamics changed over time.
Based upon examination of the recent geologic
literature, Swanson (1986) concluded that the Meso-
zoic basins of the Appalachians are almost uni-
versally controlled by late Paleozoic and older faults
that were reactivated during Late Triassic-Early
Jurassic rifting of the continental margin. He em-
phasized that discrete crustal discontinuities, rather
than simply the general structural grain of the
orogen, controlled rift faulting. He attributed the
control to pre-existing faults of high-angle reverse
and dextral strike-slip nature. He proposed, how-
ever, that in some cases "basin formation would
proceed through the direct reactivation of listric
splay thrusts or in the development of new listric-
normal fault surfaces that would merge with the
sole thrust at depth" (p. 421).
Aggarwal and Sykes (1978) found that contem-
porary earthquake shocks in southern New York
Zones of weakness concept: A review and evaluation 89
State and northern New Jersey are produced by
reverse faulting. Hence, the present sense of move-
ment in at least part of the Newark basin is opposite
to that which occurred during Mesozoic rifting.
Faulting Rejuvenation and Inversion
The confounding complexity of tectonic his-
tories, and of the dynamic systems which determine
them, is seen in the fact that on old faults that have
been rejuvenated the sense of slip commonly has
changed from one deformation event to the next.
This is so commonplace as to be more nearly the rule
rather than the exception, as was recently empha-
sized by Muehlberger (1986).
He brought this circumstance into sharp focus
when he cited an extensive list of major fault zones
whose sense of slip demonstrably changed from
event to event over the course of respective struc-
tural histories. The Ramapo fault, which was
studied by Ratcliffe (1971) and referred to above,
reinforces Muehlberger's own documentation of this
understanding. Muehlberger (1986, p. 79) conclu-
ded that "major faults do not die; they will move
again in whatever direction is necessary to accom-
modate the new stresses that are being imposed."
This conclusion is consonant with de Sitter's (1964,
p. 161-165) understanding of the behavior of "great
fundamental faults" (mentioned earlier).
Fault inversion, or reactivation with an oppo-
site or significantly changed sense of displacement,
is well documented for many areas. It is perhaps the
best confirmation that some deep-seated faults
determine loci of subsequent deformation. Excel-
lent examples of fault inversion are documented
with seismic data in A. W. Bally's atlas (1983). Spe-
cial attention is merited for individual contributions
by Plawman (1983), Ziegler (1983), Harding (1983),
and Davis (1983). Entire sedimentary basins may
be inverted and brought to view by reactivation of
bounding faults with sense-of-slip reversal.
Geometry-Based Inferences
In some cases antecedent fault patterns have
been inferred from purely geometric relationships.
Friedman (1964, p. 487-488) cited umpublished
work by F. B. Conger to provide an example of this
approach. Conger, using U.S. Geological Survey
1:250,000 topographic maps, compiled the topo-
graphic trends of 410 range-front segments within a
50,000 square mile area of the Basin and Range
province in Nevada. His basic assumption was that
the range fronts, typically straight for great dis-
tances, were fault-controlled. The statistically
most-prominent trends were north-south, N200E,
and N35W. The geometry of this array, Friedman
pointed out, suggests a Coulomb fracture assem-
blage consisting of a N-S extension fracture approxi-
mately bisecting the dihedral angle between con-
jugate shear fractures striking N200E, and N35"W,
respectively. Whether this is a valid example of
"tectonic heredity" in which Mesozoic faults were
somehow controlled by an earlier regional wrench-
fault system is highly speculative, but there exist
many known examples of assumed mimetic fault-
ing.
Donath (1961) presented in detail a good ex-
ample of inferred mimetic faulting from the Basin
and Range province of south-central Oregon. Here,
late Tertiary basalt flows are cut by numerous
nearly vertical dip-slip faults occurring in two
principal sets, N35W and N20E. The two sets were
interpreted as being essentially contemporaneous in
origin, based upon mutual offsetting relationships.
The latest movement on the faults was dip-slip.
Donath (1961, p. 1) summarized his interpretation
this way: "The contemporaneous origin of the two
fault sets, the intersection angle of approximately
50, and the nearly vertical dips indicate that the
faults originally developed as conjugate strike-slip
shears in a stress system characterized by a north-
south maximum principal stress and an east-west
minimum principal stress." He pointed out (p. 14)
that the latest faults with dip-slip displacement
" ... were not caused by and therefore do not reflect
the boundary forces responsible for the initial
development of the fractures. Dip-slip movements
occurred on previously formed strike-slip shears
through a redistribution of surface forces acting on
individual fault blocks."
CONCLUSIONS
Limited Predictability
There is no doubt that certain zones of major
crustal displacement have long histories of renewed
and intermittent movement. The reality of such
features, however, is far too complicated to consider
the zone of weakness concept to be an everyday tool
90 J. J. Prucha
for structural or tectonic analysis. The test of the
concept as a working tool is whether derivative
structures and their cognate rock fonnations (e.g.,
syntectonic sedimentary rocks and/or igneous intru-
sives) can with a fair measure of confidence and
reliability be predicted from known antecedent
structures.
In practice it usually works the other way
around. Antecedent structures are inferred from
and sometimes defined by careful and systematic
mapping of the derivative structures. Effect-and-
cause inferences are far less useful as analytical
tools than are established cause-and-effect relation-
ships. Our inability to detennine cause and effect in
a given instance does not necessarily invalidate the
concept, but it does limit its practical value. Its
predictive reliability is minimal. It is important
always to make a distinction between the possibility
that a cause-and-effect relationship exists and the
validity and pertinence of the data which are mar-
shalled in support of such an interpretation.
A principal objective of research in basement
tectonics, and indeed in all other subdisciplines of
structural geology, is to establish basic principles of
dynamics and mechanics that will make it possible
to take the limited data which are available for the
resolution of a particular problem and extrapolate
them to establish a sound and verifiable understan-
ding of a strain pattern and strain history. Intuitive
interpretations of complex structural patterns are
not good enough to substitute for more definitive
explication, although they may serve admirably to
develop multiple working hypotheses. It does not
advance understanding of basement tectonics to
assert that "it is clear (or obvious) that this younger
faulting was controlled by an older fabric in the
basement" when in fact it is neither clear nor
obvious. In this matter, it is not satisfactory to
dispense with a complex problem by employing a
simplistic explanation.
Summary of Present Knowledge
What do we really know about so-called zones of
weakness? Most definitively we know that in many
places of the world there exist surfaces or zones of
crustal displacement which were reactivated one or
more times during the geologic history of the
encompassing area. We know that for many faults
the sense of slip was inverted one or more times
during reactivation. We know that broad regional
patterns of younger defonnation do not generally
match regional patterns of older structures, al-
though some younger structures may match specific
older structures. We know that over wide areas the
style of younger defonnation does not correlate well
with rock type or structural grain of the basement.
We know from experimental studies that for some
strongly foliated rocks preferred orientation of
constituent micas serves to localize faulting within
a limited angular range of the S surfaces relative to
the greatest principal pressure. We know that for
most crustal defonnation in Phanerozoic time the
basement behaved in a brittle way and that many of
the deep-seated mechanisms that yielded basement
fabrics in the first instance were inoperative during
younger, high-level defonnation.
In summary, we seem to know quite a lot in a
general way about factors pertinent to an under-
standing of the zone of weakness concept, but we do
not know enough to make the concept a real work-
ing tool in most instances. Our ability to predict
younger defonnation patterns based upon know-
ledge of older fabrics is very limited. What can we
do about it?
Conceptual Framework of Plate Tectonics
Plate tectonics theory is well enough estab-
lished to provide a conceptual framework within
which confinned dynamic principles provide a bet-
ter basis for inferring younger strain patterns and
mechanisms within the upper crust. Detailed
studies accumulating from many parts of the world
are beginning to provide an understanding of
structural assemblages characteristic of the several
kinds of plate boundaries - compressional, exten-
sional, and strike-slip, or trench, ridge, and trans-
fonn, respectively. Each kind of plate boundary is
characterized by respective regional stress patterns.
These yield dynamically compatible structural as-
semblages with orientations that are consistent
with the stress trajectories which existed at the time
of deformation. As we refine our knowledge of past
and present plate movements, we will improve our
ability to predict structural assemblages that have
resulted from plate interaction. Our knowledge of
such movements is based upon such diverse methods
and criteria as paleomagnetic reconstructions,
dating with increased precision discrete episodes of
deformation within broad orogenic histories, recog-
nition of suture zones, distribution of clastic wedges
Zones of weakness concept: A review and evaluation 91
of post-convergence derivation, and direct measure-
ment of present-day plate movement directions and
rates using Very Long Baseline Interferometry
(VLBI).
This approach might well be fairly simple and
reliable if plate geometries were simple and regular
and if crustal materials which the plates comprise
were physically homogeneous. Such, of course, is
not the case. Plate thicknesses range substantially,
and the constituent rocks typically are mechanically
heterogeneous. Geometrically irregular plate boun-
daries in compression result in progressive or se-
quential contacts with correlative progression in the
orientation of stress trajectories. Tapponier and
others (1982) have suggested that more-rigid plates
may act as indentors which during collision with a
more-yielding plate may cause lateral extrusion of
crustal slices ("squished like a watermelon seed,"
someone has suggested!). Even relatively simple
wrench faulting in transform zones may result in
complicated assemblages which form transpres-
sional and transtensional zones, respectively
(Christie-Blick and others, 1985).
In spite of these and similar complicating fac-
tors, I believe that the ability to predict strain pat-
terns and structural assemblages of a given genera-
tion and to extrapolate accurately from limited data
is, in the foreseeable future, more readily achievable
within the context of plate-tectonics modeling than
it is by calling upon the tenuous zone of weakness
concept. Continuation of research on both ap-
proaches is warranted and important. Diligence
and perspicacity may one day provide a workable
model that integrates the two ideas.
ACKNOWLEDGMENTS
The manuscript was reviewed by M. J. Aldrich
and D. U. Wise. Their comments and suggestions
were very helpful and are gratefully acknowledged.
REFERENCES
AGGARWAL, Y. P., and L. R. SYKES, 1978, Earthquakes,
faults, and nuclear power plants in southern New Y ork-
northern New Jersey: Science, v. 200, p. 425-429.
BALLY, A. W. (ed.), 1983, Seismic Expression of
Structural Styles (Volume 3): American Association of
Petroleum Geoligists, Studies in Geology Series 15, 406 p.
BIRD, J. M., and J. F. DEWEY, 1970, Lithosphere plate-
continental margin tectonics and the evolution of the Ap-
palachian orogen: Geological Society of America Bullletin,
v. 81, p. 1031-1060.
BORG, 1., and J. HANDIN, 1966, Experimental deformation
of crystalline rocks: Tectonophysics, v. 3, p. 249-368.
CHRISTIE-BLICK, N., and K. T. BIDDLE, 1985, Deforma-
tion and basin formation along strike-slip faults; pp. 1-34
in K. T. Biddle and N. Christie-Blick (eds.), Strike-Slip
Deformation, Basin Formation and Sedimentation:
Society of Economic Paleontologists and Mineralogists,
Special Publication 37, 386 p.
CLOOS, H., 1948, The ancient European basement blocks
- Preliminary note: American Geophysical Union Trans-
actions, v. 29, p. 99-103.
DAVIS, P. N., 1983, Gippaland basin, southeastern Aus-
tralia; pp. 3.3-19-3.3-24 in A. W. Bally (ed.), Seismic
Expression of Structural Styles (Volume 3): American
Association of Petroleum Geologists, Studies in Geology
Series 15, 406p.
DE SITIER, L. U., 1964, Structural Geology, 2d Edition:
McGraw-Hill, New York, New York, USA, 551 p.
DONATH, F. A., 1961, Experimental study of shear failure
in anisotropic rocks: Geological Society of America Bulle-
tin, v. 72, p. 985-990.
__ ,1962, Analysis of Basin and Range structure, south-
central Oregon: Geological Society of America Bulletin, v.
73, p 1-16.
__ , 1964, Strength variation and deformational be-
havior in anisotropic rock; pp. 281-297 in W. R. Judd (ed.),
State of Stress in the Earth's Crust: American Elsevier,
NewYork,NewYork, USA, 732p.
__ , 1969, The development of kink bands in brittle
anisotropic rock; pp. 453-493 in L. H. Larsen, M. Prinz, and
V. Manson (eds.), Igneous and Metamorphic Geology
(Poldervaart Volume): Geological Society of America.
Memoir 115, 561 p.
FRIEDMAN, M., 1964, Petrofabric techniques for the
determination of principal stress directions in rocks; pp.
451-550 in W. R. Judd, (ed.), State of Stress in the
Earth's Crust: American Elsevier, New York, New York,
USA,732p.
HANDIN, J., 1969, On the Coulomb-Mohr failure criterion:
Journal of Geophysical Research, v. 74, p. 5343-5348.
__ , and D. W. STEARNS, 1964, Sliding friction of rocks
[abstract): EOS, Transactions of the American Geophysical
Union, v. 45, p. 103.
HARDING, T. P., 1983, Structural inversion at Rambutan
oil field, South Sumatra basin; pp. 3.3-13 - 3.3-18 in A. W.
Bally (ed.), Seismic Expression of Structural Styles
92 J.J.Prucha
(Volume 3): American Association of Petroleum Geolo-
gists, Studies in Geology Series 15, 406 p.
HOBBS, W. H., 1911, Repeating patterns in the relief and
structure of the land: Geological Society of America
Bulletin, v. 22, p. 123-176.
HODGSON, R A., 1965, Genetic and geometric relations
between structures in basement and overlying sedimen-
tary rocks, with examples from Colorado Plateau and
Wyoming: American Association of Petroleum Geologists
Bulletin, v. 49, p. 935-949.
HOPPIN R A., and J. C. PALMQUIST, 1965, Basement
influence on later deformation: The problem, techniques of
investigation, and examples from Bighorn Mountains,
Wyoming: American Association of Petroleum Geologists
Bulletin, v. 49, p. 993-1003.
MUEHLBERGER, W. R., 1986, Different slip senses of major
faults during different orogenies: The rule?; pp. 76-81 in
M. J. Aldrich and A. W. Lauglin (eds.), Proceedings of the
Sixth International Conference on Basement Tect-
onics (Santa Fe, New Mexico, 1985): International
Basement Tectonics Association, Salt Lake City, Utah,
USA,208p.
PLAWMAN, T. L., 1983, Fault with reversal of displace-
ment, central Montana; pp. 3.3-1-3.3-2 in A. W. Bally (ed.),
Seismic Expression of Structural Styles (Volume 3):
American Association of Petroleum Geologists, Studies in
Geology Series 15, 406p.
PRUCHA, J. J., J. A. GRAHAM, and R P. NICKELSEN,
1965, Basement-controlled deformation in Wyoming pro-
vince of Rocky Mountains foreland: American Association
of Petroleum Geologists Bulletin, v. 49, p. 956-992.
RATCLIFFE, N. M., 1971, The Ramapo fault system in New
York and adjacent northern New Jersey: A case study of
tectonic heredity: Geological Society of America Bulletin,
v. 82, p. 125-142.
RODGERS, R, 1987, Chains of basement uplifts within
cratons marginal to orogenic belts: American Journal of
Science, v. 287, p. 661-692.
SCHMIDT, C. J., and J. M. GARIHAN, 1983, Laramide
tectonic development of the Rocky Mountain foreland of
southwestern Montana; pp. 271-294 in J. D. Lowell (ed.),
Rocky Mountain Foreland Basins and Uplifts: Rocky
Mountain Association of Geologists, Denver, Colorado,
USA,392p.
SONDER, R A., 1938, Die Lineamenttektonik und ihre
Probleme: Eclogae Geologicae Helvetiae, v. 31, p. 199-238.
__ ,1956, Die Rhegmagenese; pp. 110-131 in Mechanik
der Erde: E. Schweizerbart'sche Verlagsbuchhandlung,
Stuttgart, Germany, 291 p.
STEARNS, D.W., G. D. COUPLES, W. R JAMISON, and J. D.
MORSE, 1981, Understanding faulting in the shallow
crust: Contributions of selected experimental and theo-
retical studies; pp. 215-229 in N. L. Carter, M. Friedman,
J. M. Logan, and D. W. Stearns (eds.), Mechanical Be-
havior of Crustal Rocks (Handin Volume): American
Geophysical Union, Geophysical Monograph 24, 326 p.
SWANSON, M. T., 1986, Preexisting fault control for Meso-
zoic basin formation in eastern North America: Geology, v.
14, p. 419-422.
SYKES, L. R, 1978, Intraplate seismicity, reactivation of
preexisting zones of weakness, alkaline magmatism and
other tectonism post-dating continental fragmentation:
Reviews of Geophysics and Space Physics, v. 16, p. 621-688.
TAPPONIER, P., G. PELTZER, A. Y. LEDAIN, and R
ARMIJO, 1982, Propagating extrusion tectonics in Asia:
New insights from simple experiments with plasticine:
Geology, v. 10, p. 611-616.
UMBGROVE, J. H. F., 1947, The Pulse of the Earth: M.
Nijhoff, The Hague, The Netherlands, 358 p.
VENING MEINESZ, F. A., 1947, Shear patterns of the
Earth's crust: American Geophysical Union Transactions,
p. 1-61.
WISE, D. U., 1958, The relationship of Precambrian and
Laramide structures in the southern Beartooth Mountains,
Wyoming; pp. 24-30 in D. L. Ziegler (ed.), Beartooth
Uplift and Sunlight Basin: 9th Annual Field Confer-
ence Guidebook: Billings Geological Society, Billings,
Montana, USA, 108 p.
__ , 1969, Regional and subcontinental sized fracture
systems detectable by topographic shadow techniques; pp.
175-199 in A. J. Baer and D. K. Norris (eds.), Kink Bands
and Brittle Deformation: Proceedings of the Confer-
ence on Research in Tectonics (Ottawa, 1968): Cana-
dian Department of Energy of Mines and Resources, Otta-
wa, Ontario, Canada, 373 p.
WOODEN J. L., C. J. VITALIANO, S. W. KOEHLER, and P.
C. RAGLAND, 1978, The late Precambrian mafic dikes of
the southern Tobacco Root Mountains, Montana: Geo-
chemistry, Rb-Sr geochronology and relationship to belt
tectonics: Canadian Journal of Earth Science, v. 15, p. 467-
479.
ZIEGLER, P. A., 1983, Inverted basins in the Alpine
foreland; pp. 3.3-3-3.3-12 in A. W. Bally (ed.), Seismic
Expression of Structural Styles (Volume 3): American
Association of Petroleum Geologists, Studies in Geology
Series 15, 406 p.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Petrogenetic evaluation of trace element
discrimination diagrams
FRANCIS 6. DUDAs
Geological Survey of Canada, 601 Booth Street, Ottawa, Ontario, KIA OE8, Canada
Geological Survey of Canada Contribution No. 25489
Current address: Department of Geological Sciences, Old Dominion University, Norfolk, Virginia 23529-0496, USA
(received November 4,1988; revision accepted September 29,1989)
ABSTRACT
Diagrams of trace element proportions in igneous rocks are commonly used in interpreting tectonic
environments of magma generation. The reliability of these tectonic discrimination diagrams is discussed for
basaltic and granitic rocks.
Discrimination diagrams are reliable if five criteria are satisfied: (1) if analytical uncertainty is small
compared to the size of the discriminant fields; (2) if post-crystallization element mobility is negligible; (3) if
tectonic environments represented in data used to construct discrimination diagrams are appropriate ana-
logues of environments to be identified; (4) if petrogenetic processes are primarily related to specific tectonic
environments; and (5) if magma source compositions are primarily related to specific tectonic environments.
The first two criteria are usually satisfied, but the last three may be suspect.
Trace element distributions calculated from petrogenetic models appropriate to mid-ocean ridge basalt
(MORB), volcanic arc basalts (VAB), and within-plate basalts (WPB) are plotted on discrimination diagrams
and compared with discriminant fields. Model results are sensitive to source compositions, source mineralogy,
distribution coefficients, and degree of partial melting. Natural variations in these parameters probably
extend the spread of natural compositions beyond those calculated in models and lead to more extensive
overlap. For continental WPB and granitic rocks, particularly, inheritance of apparent tectonic charac-
teristics from source compositions is likely and may lead to misclassifications. Thus geochemical data alone
are usually inadequate for reliable tectonic classifications.
INTRODUCTION
". . . fields on the discriminant diagrams strictly reflect
source regions (and melting and crystallization histories)
rather than tectonic regimes" (Pearce and others, 1984, p.
978).
Igneous rocks from different tectonic environ-
ments may have distinguishable major element,
trace element, and mineral compositions, and vari-
ous empirical methods have been used to infer tec-
tonic setting from geochemical characteristics.
Using geochemistry to infer paleotectonic settings of
rocks whose present geologic environment differs
from that in which they were formed is desirable, if
methods for tectonic discrimination are reliable.
Geochemical discrimination techniques are review-
ed here, and petrologic criteria are examined for
their reliability, focusing primarily on trace ele-
ment characteristics of basaltic rocks.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 93-127. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
93
94 F. O. Dudas
Geochemical variations in igneous rocks depend
on (1) compositions of magma sources, (2) minera-
logy of solid phase assemblages coexisting with
melts, and (3) processes which affect melts during
migration from sources to final destinations. Com-
positional variations of igneous rocks should not be
expected to be discriminants of tectonic setting
because none of these features is a unique function
of tectonic environment. The essential requirement,
then, for compositional criteria to be successful
discriminants of tectonic environment is for certain
sources or processes to be related primarily to spe-
cific settings.
The historical development of tectonic discrim-
ination diagrams echoes the maturation of igneous
petrology. In the mid-1960's, experimental petro-
logy provided the foundation for a genetic classifica-
tion of basaltic magmas (Yoder and Tilley, 1962;
Green and Ringwood, 1967), superseding the em-
pirical, though useful, classifications of the first half
of the century (Peacock, 1931; Kennedy, 1933).
Simultaneously, acceptance of the mobilist hypo-
thesis provided a conceptual framework for revising
extant petrologic province models (Mediterranean
= potassic; Atlantic = sodic, alkalic; Pacific = cal-
cic, calcalkalic; Harker, 1897; Niggli, 1923; Ritt-
mann, 1962). Though trace element analyses had
been done prior to the 1960's (e.g., Wager and
Mitchell, 1951), theoretical, process-oriented meth-
ods for interpreting data were not available, nor
were analyses sufficiently precise to yield unam-
biguous interpretations. Coincident with improve-
ments in analytical capabilities, major advances
toward theoretical treatment of trace elements came
in the late 1960's (Schilling and Winchester, 1967;
Gast, 1968; Shaw, 1970) and built on pioneering
work by Neuman and others (1954).
When Pearce and Cann (1971, 1973) published
the first tectonic discrimination diagrams, geo-
chemical understanding of trace element systema-
tics was almost exclusively empirical. Though the
first studies of trace element partitioning had
appeared several years earlier (Schnetzler and
Philpotts, 1968), critical data for treatment ofTi, Zr,
and Y variations, in particular, were not available.
In this context, empirical description of trace
element covariations, and of their possible linkage
with newly defined tectonic environments, was a
valuable contribution to igneous petrology. Twenty
years later, our appreciation of the complexity of
tectonic and petrogenetic processes motivates re-
evaluation of relationships between magma com-
position and environment of magma generation.
The most recent approach to tectonic discrimina-
tion, involving not only geochemical data, but also a
variety of other geological and geophysical obser-
vations (Pearce, 1987), reflects the need for a
broader context in tectonic interpretation of igneous
events.
TECTONIC DISCRIMINATION DIAGRAMS
Basic statistical concepts relating to construc-
tion of discrimination diagrams are discussed by
Pearce (1976, 1987). Discrimination diagrams are
constructed with data from rocks of known tectonic
environment. On these diagrams, areas in which
samples of specific environments predominate are
delineated either by a mathematical algorithm
which minimizes overlap or by visual inspection.
These fields are then used to classify rocks of un-
known tectonic setting. Such an empirical proce-
dure usually produces discriminant fields which
have "soft" boundaries - that is, different environ-
ments are rarely if ever completely separated: some
overlap is unavoidable. The magnitude of the "soft
boundary" problem is, to some extent, reflected in
changes made to various diagrams over time (e.g.,
compare Ti-Zr of Pearce and Cann, 1973, with
Pearce and others, 1981; Zr/Y -Zr of Pearce and
Norry, 1979, with Pearce, 1983; Hf-Ta-Th of Wood
and others, 1979b, with Wood, 1980; Th/Yb-Ta/Yb
of Pearce and others, 1981, with Pearce, 1983).
These changes suggest, also, that an extended trace
element database could lead to significant modi-
fications to discrimination diagrams.
Two types of misc1assification are implicit in
the diagrams. Rocks can be assigned to environ-
ments in which they do not belong and, conse-
quently, can be left out of environments in which
they do belong. The "success" rate of discrimination
diagrams is usually reported as the percentage of
correct classifications: a 93% success rate suggests
that, of 100 mid-ocean ridge basalt (MORB)
samples, 93 plot within the MORB or ocean floor
basalt (OFB) field. A 93% success rate does not
suggest that, among samples of unknown tectonic
environment, the discrimination diagram will cor-
rectly classify 93% of all samples. In cases where
two separate data sets of known tectonic setting
have been used, one to construct the diagram and
the other to test its reliability, success rates for test
data are 5-10% lower than for "construction" data.
Petrogenetic evaluation of trace element discrimination diagrams 95
Basaltic Rocks
Most discrimination diagrams for basic rocks
are based on trace element abundances, although a
few major element discriminants have also been
proposed (Table 1). The most sensitive major
element discriminant is K
2
0 (Pearce, 1976). Major
element discriminant functions (Pearce, 1976)
successfully classified about 85% of the samples in a
test group, although lower success rates were ob-
tained in trying to separate continental and ocean-
island basalts. Ti0
2
-K
2
0-P
2
0
5
proportions (Pearce
and others, 1975) discriminate between oceanic and
non-oceanic basalts with a success rate of 93% for
MORB but only 14% for the Deccan continental
tholeiite suite. Variations of Ti0
2
and P
2
0
5
have
been used to discriminate settings of ophiolitic rocks
(Hawkins, 1980). By contrast, a review of K-Ti-P
abundances in basaltic rocks suggests that no sys-
tematic discrimination of various basalt types
should be expected (Wilkinson and LeMaitre, 1987).
Success rates of 64-82% were achieved for a MgO-
FeO*-AI
2
0
3
discrimination diagram that included
four different eruption settings (Pearce and others,
1977).
Among the trace element discrimination dia-
grams, the ones most widely used are based on
relative abundances of high field strength elements
(HFSE; Ti, Zr, Hf, Nb, Ta, and Y). These elements
are generally depleted, compared to MORB, in
magmas formed in subduction zones (Perfit and
others, 1980) and therefore are expected to provide a
relatively robust discrimination of oceanic and plate
margin magmatism. Combinations of HFSE with
alkali, alkaline earth, and 3-d transition metals, or
with rare earth elements (REE) and large ion
lithophile elements (LILE) , account for numerous
discrimination diagrams listed in Table 1. Some of
these diagrams are shown in Figures 1-7. For
many of these plots, success rates near 90% have
been reported. Diagrams of La/Yb-SclNi and
La/Yb-Th have been suggested for andesitic rocks
(Bailey, 1981) and, with the Ce/Sr-Cr diagram
(Pearce, 1982), are the only ones that are not depen-
dent in some way on HFSE.
Spidergrams (Pearce, 1982, 1983) were not ini-
tially designed as discrimination diagrams. Quanti-
tative interpretations of spidergrams, however, pur-
port to apportion trace element abundances between
contributions from MORB, lithosphere, and subduc-
tion zone components (Pearce, 1983). Spidergrams
are included here because these components are
identified with tectonically defined geochemical
reservoirs and because spidergrams have been used
to infer tectonic settings for basaltic rocks (Holm,
1985).
Granitic Rocks
Few major element discriminants have been
designed for rocks of broadly granitic composition
(Table 2). A series of major element discrimination
diagrams can potentially distinguish tectonic en-
vironments of granitoid genesis (Maniar and Piccoli,
1989), but individual diagrams show significant
overlaps along discriminant boundaries.
Trace element discrimination diagrams for
granitoids are similar to those suggested for basaltic
rocks in that they rely primarily on HFSE. The Ta-
Th-Hf diagram (Wood and others, 1979b) is appli-
cable to both basic and acidic rocks. For granitoids,
the best discriminant elements include Y, Nb, Ta,
Yb, and Rb, although Si0
2
content is also useful
(Pearce and others, 1984). Diagrams of Nb-Y, Ta-
Yb, Rb-Y + Nb, and Rb-Yb + Ta discriminate four
different tectonic settings with a success rate of
about 90% (Pearce and others, 1984). A Rab-Hf-Ta
diagram (Harris and others, 1986) is useful for
identifying collision-related granitoids. Post-colli-
sional or anorogenic granitoids plot in distinct fields
on diagrams involving Ga/AI and a variety of major
and trace elements (Whalen and others, 1987;
Sylvester, 1989); composite axes involving several
trace (Zr+Nb+Ce+ Y) and major ([K20+Na20]1
CaO) elements provide good separation of anoro-
genic granites from other granitoids (Whalen and
others, 1987).
Spidergrams of granitoid rocks from distinct
tectonic environments differ somewhat in shape
(Pearce and others, 1984) but have not been widely
used in tectonic interpretations of granitoid rocks.
Mineral Composition Discriminants
Tectonic discrimination based on mineral com-
positions has been investigated for a number of
phases. Because mineral compositions are usually
dependent on bulk rock composition, there can be
considerable overlap between compositions of min-
erals from different environments. Minerals may be
most useful in characterizing magma type (e.g., al-
kalic or subalkalic, tholeiitic or calcalkalic; Morris,
96 F. O. Dudas
Table 1. Summary of discrimination diagrams for basaltic and andesitic rocks.
Major Elements Environments Reference
F1-F2-F3 WPB, SHO, OFB, CAB, LKT Pearce, 1976
Ti0
2
-K
2
O-P
2
0
5
OVC Pearce and others, 1975
Ti0
2
-P
2
0
5
lAT, OIB, ALKOIB Hawkins, 1980
MgO-FeO*-AI20 3 OIB, OFB, CONT, ORO, PL Pearce and others, 1977
FeO*/MgO-AI20 3 lAT, MORB, WlP Glassley, 1974
Trace Elements Environments Reference
Ti-Zr CAB, LKT, OFB, WPB Pearce and Cann, 1973
Pearce and others, 1981
Ti-Zr-Y CAB, LKT, OFB, WPB Pearce and Cann, 1973
ZrlY-Ti/Y PM,WPB Pearce and Gale, 1977
ZrlY-Zr lAB, MORB, WPB Pearce and Norry, 1979
Pearce, 1983
Ti/Y-NblY TH, TR, ALK, VAB, WPB Pearce, 1982
Nb-Zr-Y WBP, EMORB, NMORB, V AB Meschede, 1986
Ti-Zr-Sr OFB, LKT, CAB Pearce and Cann, 1973
Ti-Cr lAB,OFB Pearce, 1975
Pearce and Gale, 1977
Ti-V lAT, OFT, OIB Shervais, 1982
Ti/Cr-Ni lAT,OFT Beccaluva and others, 1979
Cr-Y VAB, MORB, WPB Pearce and others, 1981
Ti-Zr/P ALK,TH Floyd and Winchester, 1975
Nb/Y-Zr/P ALK,TH Floyd and Winchester, 1975
Ti-P-Mn CAB, lAT, MORB, OIT, OIA Mullen, 1983
Hf-Ta-Th lAY, NMORB, EMORB, WPB Wood and others, 1979b
Wood,1980
ThlYb-TalYb TH, CA, SHO, TR, ALK Pearce, 1982
CelYb-TalYb TH, CA, SHO, TR, ALK Pearce, 1982
KIYb-TalYb TH, CA, SHO, TR, ALK Pe3rce, 1982
Cr-Ce/Sr lAY, MORB, WPB Pearce, 1982
Ce-Sr-Sm NMORB, EMORB, BABB, lAT, OIT, OIA lkeda, 1990
Nb-Ti-Th PM,WPB Holm, 1985
Zr-V OFB,lAT Hawkins, 1980
LalYb-SclNi LKlAV, lAV, CAlAV, A Bailey, 1981
LalYb-Th LKlAV, lAV, CAlAV, A Bailey, 1981
Spidergrams DM, LM, SZ, OIT, CT, OFT, LKT, EMORB, BAT, RT Pearce,1982,1983
Holm, 1985
Key to Abbreviations
A = Andean lAB = island arc basalt OIB =ocean island basalt
ALK = alkalic lAT =island arc tholeiite ORO = orogenic
ALKOIB = alkalic OIB lAV = island arc volcanic OVC = oceanic vs continental
BABB = back-arc basin basalt IRT = initial rift tholeiite PL =plume
BAT = back-arc tholeiite LKlAV =low-KlAV PM =plate margin
CA =calcalkalic LKT =low-K tholeiite SHO = shoshonite
CAB = calcalkalic basalt LM = lithospheric mantle SZ = subduction zone
CAIAV =calcalkalic lA V MORB = mid-ocean ridge basalt TH =tholeiite
CONT = continental NMORB = normal MORB TR = transi tiona I
CT = continental tholeiite OFB = ocean floor basalt VAB =volcanic arc basalt
DM = depleted mantle OFT = ocean floor tholeiite WlP = within-plate
EMORB = enriched MORB OIA =ocean island alkalic WPB =witnin-plate basalt
Petrogenetic evaluation of trace element discrimination diagrams 97
Table 2. Summary of discrimination diagrams for felsic rocks.
Major Elements Environments Reference
Various lAG, CAG, CCG, POB, RRG, CEUG, OP Maniar and Piccoli, 1989
Trace Elements Environments Reference
Hf-Ta-Th IA V, NMORB, EMORB, WPB Wood and others, 1979b
Wood, 1980
Nb-Y WPG, ORG, VAG, SYN-COLG Pearce and others, 1984
Ta-Yb WPB,ORG, VAG,SYN-COLG Pearce and others, 1984
Rb-Y +Nb WPG, ORG, VAG, SYN-COLG Pearce and others, 1984
Rb-Yb+Ta WPB,ORG, VAG,SYN-COLG Pearce and others, 1984
Rb-Hf-Ta WP,VA,OF Harris and others, 1986
Rb-Tb-Ta SYN-COLG, POST-COLG, SYN-SUB, ANP, AP Thieblemont and Cabanis, 1990
Gal Al-others POST-COLG, A Whalen and others, 1987
Zr+Nb+Ce+Y- POST-COLG, A Whalen and others, 1987
others
Spidergrams WPG, ORG, VAG, SYN-COLG Pearce and others, 1984
Key to Abbreviations
A = anorogenic
ANP = anorogenic, non-peralkaline
AP = anorogenic, peralkaline
CAG = continental arc granite
CCG = continental collision granite
CEUG = continental epeirogenic uplift granite
EMORB = enriched MORB
lAG = island arc granite
lAV = island arc volcanic
MORB = mid-ocean ridge basalt
NMORB = normal MORB
OF = ocean floor
1988). Clinopyroxene is the most successful dis-
criminant (Nisbet and Pearce, 1977; Leterrier and
others, 1982), providing correct results for 75-80% of
the cases. In clinopyroxene, the best discriminant
elements commonly are minor elements (MnO,
Ti0
2
, Na20, and Cr203), so that uncertainties rela-
ted to bulk rock chemistry are magnified by
analytical imprecision (Leterrier and others, 1982).
Spinel structured oxides (Dietrich and others, 1983)
have also been tested as tectonic discriminants.
Lack of data on spinels, amphiboles, and biotites
from a variety of different settings makes them less
attractive as discriminant minerals; as with clino-
pyroxene, their compositions also vary with that of
the bulk rock.
OP = oceanic plagiogranite
ORG = ocean ridge granite
POG = post-orogenic granite
POST-COLG = post-collisional granite
RRG =rift-related granite
SYN-COLG =syn-collisional granite
SYN-SUB =syn-subduction granite
VA = volcanic arc
VAG = volcanic arc granite
WP = within-plate
WPB = within-plate basalt
WPG = within-plate granite
CRITERIA FOR RELIABILITY OF
DISCRIMINATION DIAGRAMS
The weaknesses of discrimination diagrams
have long been recognized (Thompson and others,
1980; Holm, 1982; Prestvik, 1982; Zeck and
Morthorst, 1982; Arculus, 1987). No attempt has
been made, however, to evaluate whether these
weaknesses are specific to individual diagrams or
whether they result from failure of fundamental
assumptions.
Reliability of discrimination diagrams might be
evaluated by three approaches. The first involves
case-by-case testing of the "reasonableness" of the
discriminant results, and this is the approach most
98 F. G. Dudas
commonly used (Holm, 1982; Prestvik, 1982; Zeck
and Morthorst, 1982; Duncan, 1987; Marsh, 1987).
In these cases, difficulties become apparent when
discrepancies in interpretations exist based on
different diagrams, when results conflict with field
observations, or when rocks from a single magmatic
event are classified into several different tectonic
environments.
A second approach is to assess whether geo-
chemical data represent a sufficient number of
independent variables to allow discrimination of a
number oftectonic environments. Because chemical
compositions of igneous rocks are a function of
multiple sources and multiple processes, and not
directly of tectonic environments, geochemical data
may not contain a sufficient number of independent
variables to facilitate petrogenetic as well as
tectonic interpretation. Note, for example, that
concentrations of individual REE cannot be treated
as independent sources of information because their
behavior in magmatic processes is strongly cor-
related. The constant sum effect makes major and
minor element data difficult to treat independently,
and even trace elements might be affected by
closure (a limited number of substitution sites or
vacancies exists in any magmatic system; J. C.
Griffiths, personal communication, 1983). Further-
more, any single diagram is subject to the geometric
limitation that a diagram of (n) dimensions can
resolve at most (n + 1) components; a field that lies,
geometrically, between two others on any diagram
can, arguably, represent a mixture of components
from adjacent fields rather than a unique tectonic
setting. It would be tectonically significant only if
the mixing process itself were unique to a specific
environment.
The third approach, the one followed here,
examines whether conceptual models of petrogene-
tic processes and sources lead to conclusions that are
consistent with those implied by discrimination
diagrams. If discrimination diagrams are generally
valid, the patterns they exhibit should be inter-
pretable in general petrologic terms. This approach
compares two distinct model systems: discrimina-
tion diagrams represent empirical models of rela-
tionships between geochemical features and tectonic
settings, whereas petrologic models are experimen-
tally constrained attempts to describe sources and
processes which produce specific geochemical
features. Inconsistencies between petrologic models
and discrimination diagrams do not necessarily
invalidate either model system. To the extent that
petrologic models are better constrained, inconsis-
tencies would indicate that empirical discrimination
diagrams lack general applicability. From a differ-
ent perspective, discrimination diagrams might
suggest where refinements of petrogenetic models
might be made.
Five conditions need to be satisfied if discrim-
ination diagrams are to be successfully applied. In
increasing order of importance, they are as follows:
1) Analytical uncertainty and bias must be signifi-
cantly smaller than discriminant fields.
2) Element abundances must remain unchanged
after crystallization of the magma.
3) Tectonic environments represented in "con-
struction" data sets must be reasonable ana-
logues of environments sought in data to be
classified.
4) Processes affecting element abundances must
be primarily associated with specific tectonic
environments.
5) Source characteristics must be primarily asso-
ciated with specific tectonic environments.
Inconsistencies between discrimination diagrams
and petrologic models are likely to occur when
either of the last three conditions is not met. If
petrogenetic processes or magma sources are not
uniquely associated with specific tectonic environ-
ments, these fundamental assumptions are suspect.
The following discussion of these conditions focuses
on basaltic rocks.
Condition 1: Analytical Uncertainty
Analytical uncertainty does not contribute sig-
nificant ambiguity to interpretations of discrimina-
tion diagrams. Most analytical procedures can yield
data that are precise to within 10% (relative). Un-
certainties exceeding 10% are likely only at low
concentrations. Some elements (Ta and Nb, espe-
cially) are difficult to analyze precisely, and uncer-
tainties up to 50% are possible (Meschede, 1986).
Calculated uncertainty envelopes for depleted
MORB compositions (Figure 1), in which uncertain-
ties might be most pronounced, show that the pre-
cision of routine analyses introduces no ambiguity
in discrimination diagrams. For analyses near dis-
criminant boundaries, uncertainty envelopes contri-
bute to "soft" boundaries caused by overlap. In such
instances, analyses of multiple samples are clearly
needed.
Petrogenetic evaluation of trace element discrimination diagrams 99
0
30
10
zrjv
lAB
,
0.3
3
10 100 1000 0.01 0.1
Zr, ppm
Ta/Vb

Ti/100
@
Ti/100

Th
Figure 1. Calculated effect of analytical uncertainty on discrimination diagrams. All uncertainty envelopes assume
independent variations in element concentrations for the NI-MORB average (Viereck and others, 1989). See Table 1 for
abbreviations. a) 10% and 20% uncertainty envelopes for Zr and Y (Pearce and Norry, 1979). b) 10% uncertainties for Ce and
Yb and 20% uncertainty for Ta (Pearce, 1982). c) 10% uncertainty for Zr and Y, with no uncertainty in Ti, produces the
smaller uncertainty envelope, whereas the larger envelope assumes 10% uncertainty for Ti and 20% for Zr and Y (Pearce and
Cann,1973). LKT, OFB, and CAB are included in the subduction-related basalt field. d) 10% uncertainty envelope for Ti, Zr,
and Sr (Pearce and Cann, 1973). e) 10% uncertainty in Hf with 20% uncertainties in Ta and Th produce the uncertainty
envelope (Wood and others, 1979b; Wood, 1980). f) The smaller uncertainty envelope assumes 10% uncertainty in Y and 20%
in Zr and Nb; the larger envelope assumes a 50% uncertainty in Nb (Meschede, 1986).
100 F. O. Dudas
Condition 2: Element Mobility
Discrimination diagrams rely predominantly
on elements which are little affected by processes
occurring after magma crystallization. Alkali and
alkaline earth metals (RB, Sr) and other LILE are
notable exceptions to this generalization, and dia-
grams involving these elements are suspect for
samples showing any evidence of alteration. HFSE
are generally unaffected by hydrothermal alteration
and metamorphism below middle amphibolite
grade, although some HFSE mobilization occurs
during carbonate-rich, greenschist-facies metamor-
phism (Murphy and Hynes, 1986). Although REE
are also considered immobile (Menzies and others,
1979), the light REE (La, Ce) can be redistributed
during hydrothermal alteration (Ludden and
Thompson,1978). The 3-d transition metals (Ni, Cr)
appear to be immobile at least to greenschist-facies
(Humphris and Thompson, 1978). Because Mn has
multiple oxidation states, it can be very mobile
during hydrothermal alteration (Dudas and others,
1983), making the Mn-Ti-P diagram (Mullen, 1983)
suspect. Analytical uncertainty calculations (Fig-
ure 1) also suggest the magnitude of effects which
could result from element mobility.
Condition 3: Adequate Analogues
Tectonic settings included in discrimination
diagrams are shown in Tables 1 and 2; Pearce
(1987) enumerates 12 potentially distinguishable
environments of basalt generation for which ana-
logues need to be identified.
Confusion of petrogenetic and tectonic nomen-
clature and lack of precise definition of tectonic
settings have contributed to ambiguities in inter-
preting discrimination diagrams. The fields on
various diagrams are not consistently labeled.
Subduction-related environments (Table 1), for
example, are represented by fields for shoshonite
(SHO), calcalkalic basalt (CAB), low-K tholeiite
(LKT), island arc tholeiite (IAT), orogenic basalt,
island arc basalt (IAB), volcanic arc basalt (VAB),
island arc volcanics (IA V), continental arc volcanics,
and "Andean-type" volcanics. Some of these cate-
gories are relatively exclusive (e.g., shoshonite),
whereas others are very general (volcanic arc
basalt, orogenic basalt). These labels, furthermore,
mix tectonic and petrologic classes: low-K tholeiite,
for example, identifies a petrologic class, not a
tectonic setting, yet is used as a discriminant field
adjacent to within-plate basalts (WPB) and MORE.
In some cases, analogues of petrogenetic classes may
not be adequate tectonic analogues; conversely,
almost any subduction-related basalt would qualify
as an adequate analogue for "orogenic basalt."
Petrologic diagrams used to characterize rock series
- AFM diagrams or Jensen (1976) cation plots -
should not be confused with tectonic discrimination
diagrams.
For environments that have been sites of ig-
neous activity throughout earth history (MORB,
V AB), modem sample suites represent appropriate
analogues. The only caveats (Condie, 1989) may be
that WPB are rare or absent in the Archean and
that, because MORB may not easily be preserved,
there is bias in the geological record. Tectonic
discrimination of Archean rocks could be less reli-
able than classifications of younger rocks, especially
if significant secular changes occurred in magma
composition or in the predominant processes of
magma generation (Condie, 1989).
Many discrimination diagrams subdivide broad
tectonic classes. MORB, for example, are divided
into enriched (E-MORB, or plume-type P-MORB)
and normal (N-MORB) categories, and sometimes a
transitional group (T-MORB) is recognized. Whe-
ther such refinements can successfully be extended
to rocks formed in unknown tectonic settings is
unclear. In some instances, recognizable tectonic
environments have no modem examples from which
discriminants might be built (e.g., fore-arc basin
spreading centers; Pearce, 1987), and no evidence
suggests that rocks from such settings would be
geochemically distinct.
Some diagrams are constructed with data from
older rocks whose tectonic settings cannot now be
verified directly (e.g., Maniar and Piccoli, 1989).
Though the resulting discriminants may be correct,
the validity of the diagrams is linked to the reli-
ability of interpretations made for the "construc-
tion" data set, and there is circularity in the tectonic
interpretations.
Condition 4: Petrogenetic Processes
Four classes of processes potentially affect dis-
tributions of elements in igneous rocks: (1) partial
melting, (2) differentiation, (3) mixing and assimi-
lation, and (4) alteration (previously discussed). Of
the first three, only some features of partial melting
Petrogenetic evaluation of trace element discrimination diagrams 101
and assimilation may be strongly linked to tectonic
environments.
Conditions under which partial melting occurs
- e.g., temperature, pressure, and bulk composition
(including volatile components) of the material
being melted - control compositions of parental
magmas. Basaltic magmas form over a relatively
limited temperature range (e.g., 1000-1300C), inde-
pendent of tectonic environment. Pressure during
melting can vary significantly, from < 10 kb for
some MORB to 30-40 kb for some alkalic rocks of
broadly basaltic composition. Because temperature
and pressure control the stabilities of minerals in
the source region, they determine both the extent of
melting and the distribution of elements between
solid residue and melt. Thus, a source of specific
bulk composition can yield different basaltic melts
depending on temperature and pressure.
Differentiation processes are not uniquely
linked with tectonic settings. For many magmatic
systems, fractional crystallization is assumed to be
the most important differentiation mechanism.
Basaltic rocks, regardless of tectonic setting, cry-
stallize a restricted mineral assemblage that can
include olivine, orthopyroxene, clinopyroxene, pla-
gioclase, and spinel-structured oxides; accessory
minerals such as zircon, apatite, and ilmenite are
rarely important in crystallization of basaltic
compositions. Even for alkalic basalts, the early
history of the magma is controlled by an olivine-
clinopyroxene assemblage, and feldspars and feld-
spathoids crystallize later than the mafic minerals.
Only for lamprophyres and lamproites might amp-
hibole and mica figure prominently in differen-
tiation within the basaltic composition range.
Mixing and assimilation are not restricted to
specific tectonic environments but are best docu-
mented in arc and intraplate rocks, where lithologic
complexity is greatest and where crustal melts may
be involved. Mixing is a process in which two pre-
existing melts combine into a single magma,
whereas assimilation is the process by which a melt,
injected into a foreign host, causes melting and
eventual consumption of the host by the magma.
Mixing processes can, but do not necessarily, in-
volve perturbations in the thermal and crystalliza-
tion history of a magma. Assimilation is always
accompanied by loss of magmatic heat to cooler wall
rocks and, consequently, by partial crystallization of
injected melt. Mixing of separate batches of MORB
may be chemically undetectable, whereas mixing of
basaltic and rhyolitic magmas may produce pat-
terns that approximate those of fractional cry-
stallization. The chemical effects of assimilation are
commonly more pronounced because assimilated
material is usually a minimum melt derived from
the host rocks.
Condition 5: Magma Sources
Source composition is probably the single most
important determinant of magma chemistry. A
broad correlation exists between compositions of
magma sources and tectonic settings, and it sug-
gests that trace element discrimination of tectonic
settings is possible. However, significant exceptions
exist to any generalizations about tectonic depen-
dence of source compositions; sources are not
uniquely related to settings.
The bulk compositions of magma sources (Fig-
ure 2 and Appendix) are less constrained than
their mineralogy and can most conveniently be
discussed by reference to an assumed, primitive,
undifferentiated mantle (or bulk silicate earth)
composition that has chondritic trace element pro-
portions but twice chondri tic trace element abun-
dances. Primitive mantle (PM) has not been recog-
nized as a magma source. Mantle that is depleted or
infertile by comparison with primitive mantle is the
source of most N-MORB; this source composition
has been estimated from analyses of xenoliths, from
inversion of N-MORB compositions, and from ana-
lyses of experimental residues. Depleted mantle
(DM) contains relatively low concentrations of
incompatible elements - i.e., elements enriched in
silicate melts relative to a coexisting solid mineral
assemblage.
Mantle enriched in incompatible elements,
relative to primitive mantle, usually differs from
primitive mantle both in absolute abundances of
trace elements and in proportions of various ele-
ments. Most enriched sources consist of garnet
lherzolite; strongly enriched sources might also
contain amphibole, phlogopite, carbonate, or apa-
tite. Enriched mantle (EM) sources contribute to
most ocean island and continental basalts and
represent, essentially, the within-plate field of
many discrimination diagrams. Enriched mantle
sources are also detected in some MORB (even N-
MORB) and volcanic arc basalts.
102
F. a.Dudas
zr/v

<---=-_J
PS
....
Zr, ppm 0.1 Ta/Vb
Ti/100
Figure 2. Compositions of major geochemical reservoirs and magma sources plotted on discrimination diagrams. Mantle
compositions (Table AI) include depleted mantle (DM), primitive mantle (PM), and enriched mantle (EM). Crustal
compositions (Table A5) are lower crust (LC), total crust (TC), and upper crust (UC) from Taylor and McLennan (1985), Pacific
sediment (PS) from Hole and others (1984), and the North American Shale Composite (NASC) from Gromet and others (1984).
PETROGENETIC MODELS AND
DISCRIMINATION DIAGRAMS
In this section, trace element distributions
calculated from simple petrogenetic models (see
Appendix for details of calculations) are plotted on
discrimination diagrams. Comparison of petro gene-
tic model results for MORB, V AB, and WPB demon-
strates that tectonic fields on discrimination dia-
grams are not uniquely related to tectonic environ-
Petrogenetic evaluation of trace element discrimination diagrams 103
ments. Model results for each environment overlap,
to some degree, with those of the others and overlap
multiple fields on discrimination diagrams. The
sensitivity of model results to variations in model
parameters (residue mineralogy, source composi-
tion, distribution coefficients, extent of fractiona-
tion) is assessed only for the MORB model but is
comparable in other calculations. A summary of
model results is presented first (Figure 3), followed
by more detailed interpretations of individual
models (Figures 4-7) and a discussion of spider-
grams.
Basaltic Rocks
Summary. The compositions generated from
MORB, V AB, and WPB models are shown in Figure
3. V AB and MORB discriminant fields overlap on
Zr/Y-Zr, CelYb-TalYb, Ti-Zr-Y, and Nb-Zr-Y (Fig-
ure 3a,b,c,f). Distinct compositions are thus not
expected for V AB and MORB model magmas in
these plots and, as expected, MORB model composi-
tions are almost completely overlapped by V AB
model compositions in ZrIY-Zr, CeIYb-Ta/Yb, and
Nb-Zr-Y. MORB model compositions are only part-
ly overlapped by VAB in Ti-Zr-Y (Figure 3c) and,
although the OFB field is also considered a V AB
composition field (Pearce and Cann, 1973), it is not
covered by VAB model results. MORB model
compositions are almost completely overlapped by
VAB results in Ti-Zr-Sr (Figure 3d), where no
overlap is predicted. MORB results are unique in
the Hf-Ta-Th diagram (Figure 3e).
V AB compositions are distinct from other model
results on Ti-Zr-Y (Figure 3c) and Hf-Ta-Th (Fig-
ure 3e). Model compositions, however, cover only
small portions of V AB fields on these diagrams. On
other plots, areas unique to V AB model results are
relatively restricted. V AB data are overlapped both
by MORB and WPB.
Some WPB model results are overlapped by
VAB and MORB on ZrIY-Zr, Ti-Zr-Y, and Nb-Zr-
Y, but regions of exclusively WPB compositions also
exist on each of these diagrams. On Ti-Zr-Sr, no
WPB field is recognized (WPB compositions should
be excluded by prior screening of data; Pearce and
Cann, 1973); complete overlap is expected and ob-
served for unscreened results. WPB and V AB over-
lap in part on Ce/Yb-Ta/Yb and Hf-Ta-Th.
Model calculations suggest that WPB magmas
are most recognizable in the discrimination dia-
grams: WPB model results have the least overlap,
and V AB and MORB model data rarely plot within
WPB discriminant fields. WPB results do not al-
ways, however, plot within the WPB discriminant
fields. WPB results cover the whole of the SHO field
and half of the CA field on CelYb-Ta-Yb (Figure
3b) and plot primarily within the VAB field on Hf-
Ta-Th (Figure 3e), suggesting that WPB magmas
might easily be misclassified as trace element-rich
VAB. Only on Hf-Ta-Th are VAB and MORB com-
positions potentially distinguishable (Figure 3e).
The model trace element distributions result
from interplay of assumed source compositions and
calculated D (distribution coefficient). Multiple
diagrams involving the same element (e.g., Zr or Y),
or involving elements of closely similar geochemical
behavior (e.g., Zr and Hf or Nb and Ta), provide
constraints on the extent to which either source
composition or D might be modified to fit expected
discriminant fields. Typically, an improvement of
fit on one diagram is accompanied by increased
overlap on others. Though the set of source com-
positions that was tested is limited, the patterns
observed in the results are general: reasonable
source compositions will produce similar results,
and overlapping fields, which produce ambiguity in
tectonic classifications, are unavoidable.
MORB. The petrogenetic model adopted for N-
MORB (Appendix) assumes a depleted mantle
source, a residue mineralogy of either plagioclase or
plagioclase-spinel lherzolite, and melting propor-
tions consistent with experimental and mass bal-
ance constraints (Viereck and others, 1989).
Trace element distributions predicted from 10-
20% partial melting of plagioclase lherzolite (model
1, Figure 4) fall within N-MORB or OFB fields on
four of six diagrams. Where model predictions fail
to match discrimination diagrams (Figure 4a,c),
the mismatch is due to excess Y: the DM source sug-
gested by Viereck and others (1989) lies outside the
OFB field, and low Dy results in partial melts hav-
ing proportionately higher Y content.
Fractional crystallization paths (Appendix) for
20% partial melts of model 1 show that differentia-
tion has a negligible effect on trace element distri-
butions. Insets (Figure 4a,b,c) show the changes
expected during crystallization of a parental MORB
magma. Only for Zr-Zr/Y (Figure 4a) might frac-
tionation significantly affect the range of model
compositions, bringing some low Zr/Y melts into the
104
F. O. Dudas

ZVV
TI/ l00
z,
Figure 3. Summary diagrams of model results show extensive overlap. MORB model results are stippled, VAB model results
are vertically lined, and WPB model results are horizontally lined. Note lack of coverage of E-MORB and WPB fields on Hf-
Ta-Th (e) and Nb-Zr-Y (f).
Figure 4. Diagrams of MORB model results, showing effects of variations in source composition, residue mineralogy, and
fractional crystallization. Modell is for plagioclase lherzolite; model 2 is for plagioclase-spinel lherzolite (see Appendix).
Tick marks are for 1, 5, 10, and 20% partial melting, and for 5, 10, 20, and 30% fractional crystallization. Partial melting
curves are for D and P calculated from low values in Table A2; fractional crystallization curves include variations of D.
a) Fractional crystallization (fc) calculations for 10% and 20% partial melts (pm; inset). b) Inset shows effect of 30%
fractional crystallization of 20% partial melt composition. c) Inset shows effect of 30% fractional crystallization (fc) of 20%
model 1 partial melt (pm). Curves labeled a utilize low D from Table A2; curves labeled b utilize high D from Table A2 and
a range of fractionating mineral assemblages. d) Modell and model 2 show distinct partial melting curves because
Petrogenetic evaluation of trace element discrimination diagrams
z,

Zt . ppm
Figure 4 (continued)
of differing proportions of plagioclase. Partial melting curves
and point labeled V are based on Viereck's and others (1989)
DM composition; LR curves and point are for LeRoex's (1987)
DM composition. Area labeled fc encloses compositions gener-
ated by :530% fractional crystallization of a 20% model 1 par-
tial melt of the Viereck source. e) 30% fractional crystalliza-
tion has effects that are within the point size of DM sources on
Hf-Ta-Th. f) As for (e), 30% fractional crystallization effects
are within the point size of DM sources on Nb-Zr-Y.
Ti
Zr
Y
Sr
Hf
Ta
Th
Ce
Yb
Nb
r o/ 100
Modell
D P
0.0970 0.2069
0.0210 0.0416
0.0161 0.0634
0.1351 1.027
0.0349 0.104
0.0654 0.1324
0.0017 0.0065
0.0075 0.0498
0.0236 0.0812
0.0434 0.0336
105
Model 2
D P
0.1265 0.1849
0.0283 0.0492
0.0237 0.0542
0.0738 0.8188
0.0512 0.0996
0.0864 0.1088
0.0016 0.0054
0.0063 0.0400
0.0306 0.0658
0.0500 0.0290
106 F. O. Dudas
MORB field when the distribution coefficient (Dy) is
varied by about a factor of 10.
Changing residue mineralogy and melting pro-
portions from plagioclase lherzolite to plagioclase-
spinel lherzolite (model 2) has little effect on cal-
culated trace element distributions. The proportion
of Sr (Figure 4d) is strongly affected by changes in
the role of plagioclase, which has high D
sr
. The
model MORB magmas remain outside the MORB
field (Figure 4a,e).
Changing DM source compositions affects the
range of calculated magma compositions but, for 10-
20% partial melting, generally yields trace element
distributions within the MORB fields of the discrim-
ination diagrams. The most depleted MORB source
(Hole and others, 1984) produces melts outside the
MORB field (Figure 4a,b,e), which is consistent
with the suggestion (Hole and others, 1984) that
this extremely depleted source composition is a
component of V AB rather than MORB. MORB
fields at Zr!Y>3 (Figure 4a), at proportions of Y
(fY) < 35 (Figure 4e), at ITh < 11 (Figure 4e), and at
fY < 53 (Figure 4f) are not accessible from model
calculations using the sources identified in the Ap-
pendix. All partial melts of the LeRoex (1987)
source fall outside the OFB field (Figure 4d).
The greatest changes in model values result
from varying D. Bulk D for models 1 and 2 (cal-
culated from data in Table A2, Appendix) are
inconsistent with the order of element compatibility
expected for garnet lherzolite sources (Pearce,
1983); they reflect the change from garnet-bearing
to spinel- and plagioclase-bearing residues, as well
as uncertainties in D. In general, model calcula-
tions for Figure 5 incorporate changes of D up to a
factor of 10 above and below values in Table A2;
this range represents an unreasonable maximum, if
only because it potentially perturbs ratios of D by
100.
Diagrams involving Y (Figure 4a,e,f) were the
only ones for which model compositions consistently
differed from those expected. Increasing Dy by a
factor of 15 (Green and others, 1989) produces model
1 distributions compatible with some discrimination
diagrams (Figure 5a,e,f) but leads to within-plate
characteristics for some source compositions on
others (Figure 5a,e). Bulk Dy might thus be
greater than indicated from Table A2 but less than
suggested by Green and others (1989). On the re-
maining diagrams, large changes of Dare insuffi-
cient to move MORB or OFB fields.
MORB models and discrimination diagrams
potentially conflict in three areas: (1) compositions
of model melts outside of the 10-20% partial melting
range are, in most cases, not MORB-like; (2) some
compositions on the diagrams are not accessible
from any realistic combination of model parameters;
Figure 5. Diagrams of MORB results including effects
of varying D in partial melting models. Tick marks
indicate 1, 5, 10, and 20% partial melting. The range for
10-20% partial melting is stippled.
a) Unfractionated model 1 partial melts cover the stip-
pled area enclosed by solid lines; 30% fractional cry-
stallization extends MORB melts to the stippled area
inside the dashed line.
D P
Curve a: Zr 0.0210 0.04164
Y 0.0161 0.0634
Curves b: Zr 0.0210 0.0416
Y 0.2404 0.3820
b) Curves extending to high CelYb are for high Dyb,
with DCe and DTa unchanged.
Curve a:
Curves b:
Ce
Yb
Ta
Ce
Yb
D P
0.0075
0.0236
0.0654
0.0075
0.2360
0.0498
0.0812
0.1324
0.0498
0.8120
Ta 0.0654 0.1324
c) The partial melt field for MORB probably extends
over a range of Dy, stretching between curves a and b.
For all curves, DTi = 0.0970, P
Ti
= 0.1265. Curve c is
unrealistic, as DTi will probably never be < DZr and Dy.
D P
Curves a: Zr 0.0210 0.0416
Y 0.0161 0.0237
Curves b: Zr 0.0210 0.0416
Y 0.2404 0.3820
Curve c: Zr 0.2089 0.2160
Y 0.2404 0.3820
d) Partial melting curves with varying D. DTi = 0.0970,
PTi = 0.1265, and DZr = 0.0210, P
Zr
= 0.0416 for all
curves; DSr (calculated from Table A2) = 0.1351, P
Sr
=
1.027 (Figure 4). There are additional tick marks at 15%
melting on curves a, c, and d. Curves a and b - Vier-
eck source, model 1; curve c - LeRoex source, model 1;
and curve d- LeRoex source, model 2.
Curve a: Sr
Curveb: Sr
Curve c: Sr
Curved: Sr
D P
0.2702 2.054
0.0676 0.5135
0.2702 2.054
0.2702 2.054
Petrogenetic evaluation of trace element discrimination diagrams 107

@
zrjv

Zt , ppm
Figure 5 (continued)
e) Partial melting curves with varying D. DTh = 0.0017,
P
Th
= 0.0065 for all curves.
D P
Curves a: Hf 0.0349 0.104
Ta 0.0065 0.013
Curves b: Hf 0.349 1.04
Ta 0.0065 0.013
Curvesc: Hf 0.349 1.04
Ta 0.065 0.1324
0.1 r.,Nb
f) Partial melting curves with varying D. Dy = 0.2404,
Py = 0.3820 for all curves.
D P
Curves a: Nb 0.0043 0.0034
Zr 0.0210 0.0416
Curves b: Nb 0.0434 0.0336
Zr 0.0210 0.0416
Curves c: Nb 0.2583 0.2235
Zr 0.0210 0.0416
Curves d: Nb 0.0043 0.0034
Zr 0.2089 0.2160
108 F. O. Dudas
and (3) N-MORB model melts are misclassified as
CAB or WPB on several diagrams. In the first case,
melts that are petrogenetically identical with
MORB in all but the extent of melting are poten-
tially misclassified as WPB (Figure 5a,c,f) or as
VAB (Figure 5b,e,f). In the second case, rocks that
do not share a petrogenetic linkage with N-MORB
are classified as N-MORB (Figure 5e). In the third
case, N-MORB models give rise to WPB or VAB
compositions (low ZrIY sources of Figure 5a; low Y
sources of Figure 5c, using Dy from Green and
others, 1989; low Sr sources of Figure 5d; and low
Ta sources of Figure 5e). Given the known range of
MORB source compositions and a realistic range of
D, the conclusion is inescapable that some N-MORB
compositions fall outside the delineated discrimina-
tion fields and will be misclassified.
V AB. Five sources potentially contribute to
magmatism near destructive plate margins: (1) de-
pleted mantle, (2) sediment, (3) oceanic crust,
(4) subcontinental mantle, and (5) continental crust
(Kay, 1980). Though relative contributions of these
sources vary in individual arcs, HFSE depletion and
LILE enrichment (Perfit and others, 1980) appear to
be universal features of arc magmas. HFSE deple-
tion is probably inherited from a DM source (Hole
and others, 1984; Arculus and Powell, 1986).
Ryerson and Watson (1987) specifically exclude
residual oxide phases as the cause of HFSE deple-
tion, whereas LILE enrichment is usually ascribed
to components derived from subducted crust.
Models adopted here for arc magmatism consider
two scenarios: (1) partial melting of sources formed
by bulk mixing of DM with sediments, and (2) par-
tial melting of sources formed from DM and melts
derived from subducted crust. Parental magma
compositions from these sources are plotted on
Figure 6.
Depleted mantle alone is not an appropriate
source composition for arc magmas. Mixed model
sources cover a significantly larger range of element
ratios and allow model melts to cover 30% of the arc-
related fields on all of the diagrams tested (Figure
6). Essentially 100% of the arc fields are covered in
ZrIY -Zr (Figure 6a) and Nb-Zr-Y (Figure 6f) dia-
grams. For 10-20% partial melting, model melts do
not adequately cover the arc fields on CeIYb-TalYb
(Figure 6b), Ti-Zr-Y (Figure 6c), Ti-Zr-Sr (Fig-
ure 6d), and Hf-Ta-Th (Figure 6e). If lower de-
grees of melting are considered, the coverage of arc
fields can be improved but more extensive overlap
with non-arc fields occurs.
Areas which are not accessible to model V AB
melts on CeIYb-TaIYb would require Dce1D
Yb
and
D
Ta
::;; D
Yb
; this implies that Dee < D
Ta
, contrary to
the expected order of compatibility (Pearce, 1983).
Note also that high D
Yb
probably requires garnet in
the source region. No plausible mantle/crust mixing
model is known that would produce sources with
CeIYb and TaIYb plotting left of the 10% partial
melting line (Figure 6b), nor is a viable model
known for selective enrichment of Ce, as the high
CeIYb fields (Figure 6b) require.
Coverage of both CAB and LKT fields on the Ti-
Zr-Y diagram (Figure 6c) seems to require con-
flicting changes in DZr and D
Ti
. Extending model
melts to high proportions of Zr (f(Zr)) in Figure 6c
would require decreasing D
Zn
increasing DTi (i.e.,
DZr< D
Ti
), or mixing with larger fractions of a
NASC-like component. Doubling DTi allows com-
plete coverage of the CAB field and is within
uncertainty of DTi; at the same time, however,
"" 50% of the melt field would then lie outside the
discriminant fields of Figure 6c. By contrast, the
LKT field (Figure 6c) would only be accessible to
model melts if DTi<Dzr""Dy. A similar difficulty
exists for LKT in Ti-Zr-Sr (Figure 6d). Rather
than indicating unusual D, the LKT fields suggest,
instead, a source composition with f(Zr) < 20% and
f(TO;::: 50%. Such a source, not represented among
those tested here, would also cover the OFB field
with V AB model melts, consistent with original
interpretations of Ti-Zr-Y data (Pearce and Cann,
1973). No plausible mixture of DM with crustal
materials can generate such a source composition,
nor is it clearly a DM source. Interpretation of LKT
compositions is thus problematic; neither realistic
variations ofD nor recognizable source compositions
are involved in modeling LKT.
Areas that are inaccessible to V AB model melts
on Hf-Ta-Th are not significant and include only
extreme compositions (f(Th) > 85%, f(Ta) < 5%, and
f(Hf) >75%). Rocks with f(Hf75% may require a
distinct source inasmuch as DHf<D
Th
is unlikely.
Overlap of model compositions into non-arc
fields is significant on all diagrams except CeIYb-
TaIYb. For diagrams involving Y, much of the
overlap is due to melt models having high Dy; as for
MORB models, a range of Dy is indicated. Any
model with Dzr<Dy will produce overlap on ZrIY-
Zr (Figure 6a), whereas complete coverage of the
CAB field on Ti-Zr-Y is impossible if DZr;::: Dy:
Petrogenetic evaluation oftrace element discrimination diagrams
0_1
T.;vb
109
Figure 6. Diagrams of V AB model results. Curves labeled a are for bulk mixing of DM with PS; curves labeled b are for bulk
mixing of DM with NASC; open circles or curves labeled c are for mixing of DM with partial melts of 0 +S (oceanic crust plus
10% PS); 20% partial melts ofO+S are shown as solid circles. Source mixing lines have tick marks for 1 and 5% mixtures, and
the range of possible source compositions is outlined. Partial melting curves have tick marks at 1, 5, 10, and 20% partial
melting; The 10-20% partial melting range, calculated only for NASC and PS mixed sources, is outlined and stippled. The
partial melting range includes models with varying D, as indicated in captions to Figures 4 and 5. Diagram (d) shows D
Ti
, P
Ti
,
DZr> and P
Zr
as indicated for Figure 5d. For high D, DSr = 0.2702, P
Sr
= 2.054; for low D, DSr = 0.0676, PSr = 0.5135. Curve p
- Viereck source + 1% PS, low D; curve q - Viereck source + 1% PS, high D; curve r - LeRoex source + 5% NASC, low D;
curve 8 - Viereck source + 5% PS, low D; curve t - LeRoex source + 5% NASC, high D; and curve u - LeRoex source + 5%
PS,highD.
110 F. o. Dudas
overlap is unavoidable in the model results.
Overlap into the MORB field is unavoidable on most
diagrams, partly because the discriminant fields
overlap (Figure 6a,b,c,f) and partly because the
VAB and N-MORB models converge as the extent of
source mixing decreases. The tholeiite field of
CeIYb-TaIYb includes both MORB and arc tholei-
ites, and no distinction of these rock types is possible
on Figure 6b. Overlap with MORB is a serious
concern on the Hf-Ta-Th diagram (Figure 6e)
because DM mixtures with PS lie primarily within
the MORB field and, at reasonable extents of
melting, magmas from these mixed sources retain
feTal above those indicated for V AB. To decrease
f(Ta) , DTa would have to increase unrealistically.
Overlap with WPB occurs in all diagrams involving
Y and is due to high Dy in model calculations.
Model results suggest that PS is not generally
useful as a mixing component, though it provides a
good fit to some REE data (Hole and others, 1984).
Mixtures with NASC, or with compositions between
NASC and PS (e.g., for Ti-Zr-Y, Ti-Zr-Sr, and Nb-
Zr-Y) , yield more appropriate sources. Mixtures
with melts of subducted crust (oceanic crust plus
10% sediment; OS on Figure 6) are also less suc-
cessful than NASC mixtures; the Zr/Y (Figure 6a)
and f(Ta)(Figure 6e) of mixed DM-melt sources is
too high and feY) is too low (Figure 6c,f), increasing
the overlap with non-arc fields. Mixed DM-melt
sources have higher TaIYb than DM-NASC mix-
tures and can produce model melts that cover the
high TaIYb portion of the calcalkalic field on CeIYb-
TaIYb (Figure 6b). In no case is mixing with a
lower crustal composition indicated.
The greatest discrepancy between model pre-
dictions and discrimination diagrams lies in the
inability of petrogenetic models to generate LKT
(Figure 6c,d) and SHO (Figure 6b) compositions.
The former almost certainly requires a significantly
different source composition, with little or no sedi-
ment admixture. The latter requires either low
degrees of melting (:5 5%) or garnet-bearing source
compositions, neither of which is generally char-
acteristic of arc environments. Extensive high-
pressure fractionation (Meen, 1987) alone is
unlikely to produce CeIYb in the SHO field, espe-
cially if Dee and DYb are almost identical as Meen
suggests.
WPB. WPB model compositions cover larger
areas on discrimination diagrams (Figure 7) than
do V AB and MORB model results, primarily be-
cause the range of partial melting for WPB is 1-20%,
compared to 10-20% for the other models. However,
the most significant geochemical distinction of the
WPB models is that garnet- and amphibole-bearing
sources are evaluated in addition to spinel
lherzolites. Garnet-bearing WPB models have high
Dy and DYb during partial melting, leading to
correspondingly low fey) and high Zr/Y and Ce/Yb
in model results. The WPB environment is recog-
nizable on several discrimination diagrams because
garnet-bearing sources have such characteristic
trace element distributions; discrimination is
possible not because of tectonic environment but be-
cause of source mineralogy. Amphibole lherzolite
sources produce melts very similar to spinel lherzo-
lites but usually with lower absolute trace element
contents.
Minimal overlap exists between WPB model
results and non-WPB fields on ZrIY-Zr (Figure 7a),
Ti-Zr-Y (Figure 7c), and Nb-Zr-Y (Figure 7f),
primarily because model compositions have low Y.
On the other three diagrams, WPB results overlap
VAB fields, covering much of the CA and SHO fields
on CeIYb-TalYb (Figure 7b) and the high f(Th)
portion of the IA V field on Hf-Ta-Th (Figure 7e).
Overlap on Ti-Zr-Sr (Figure 7d) is not considered
important inasmuch as that diagram was not de-
signed to discriminate WPB; if, however, screening
of analyses (Pearce and Cann, 1973) does not suc-
cessfully eliminate all WPB, they can be mis-
classified as V AB or OFB on Ti-Zr-Sr.
Some model WPB results plot outside of com-
positional ranges observed in real rocks, especially
on diagrams involving Y and Yb. Dy and D
Yb
in
these models (Green and others, 1989) may be too
high. As for V AB and MORB models, a range of Dy
and DYb is indicated. Models using low Dy are
needed to generate low Zr/Y WPB (Figure 7a), but
Figure 7. Diagrams ofWPB model results. Source com-
positions (EM) are shown as solid diamonds; uncertainty
in Ti for Frey's and others (1978) source is indicated with
open diamonds. Curves have tick marks at 1, 5, 10, and
20% partial melting. Melt-composition range is stippled.
Calculations used varying D, as indicated in captions to
Figures 4 and 5 and as shown in Table A2. Curve a -
Frey source, spinel lherzolite residue mineralogy, low D
from Table A2 (as for Figure 4); curve b - Frey source,
garnet lherzolite, high D values from Table A2, mostly
based on Green and others (1989); curve c - Frey
source, amphibole lherzolite, high D; curve d- Wood
Petrogenetic evaluation of trace element discrimination diagrams 111

2" ppm 1000
0-01
To/ .OO
Figure 7 (continued): and others (1979a) source, amphibole lherzolite, high D; curve e - Wood source, spinel lherzolite, high
D; curve f- Wood source, garnet lherzolite, low D; curve g- Frey source, garnet lherzolite, low D; curve h- Wood source,
amphibole lherzolite, low D; and curve i-Wood source, garnet lherzolite, high D. b) The dashed curve at high TalYb is for
DCe = 0.0131, twice D used elsewhere; DYb = 0.3112, equal to Dy and < 0.5D
Yb
used elsewhere; DTa = 0.0047, O.ID suggested
by Green and Pearson (1987). c) Uncertainty of Ti in the Frey source results in large areas for potential melt compositions.
d) Uncertainty of Ti in the Frey source results in large areas for potential melt compositions. e) Curves x and yare
unconstrained by measured D and indicate range of D required to cover WPB compositions from tested sources. For curve x-
DHf = 0.0876, DTa = 0.0047, DTh = 0.0014; curve y - DHf = 0.0438, DTa = 0.0047, DTh = 0.0141. Values for curve x are in
the right order of compatibility for a garnet lhezolite source, but DTa and DTh are both too low. f) Curves x and yare
unconstrained by measured D and indicate range of D required to cover WPB compositions from tested sources. For curve x-
DNb = 0.0047 (=D
Ta
assumed for Figure 7b,e), DZr = 0.109 (= high D suggested by Green and others, 1989), Dy = 0.3
(",DmaxfromTableA2);curvey-DNb = 0.0047,D
Zr
= 0.218,Dy = 0.218.
112
F. O. Dudas
compositions for such models on Ti-Zr-Y (Figure
7c) almost completely overlap the CAB field and
hardly reach the Zr-rich, Y -poor WPB on Nb-Zr-Y
(Figure 7f).
Models using calculated D do not cover WPB
fields on CelYb-Ta/Yb (Figure 7b), Hf-Ta-Th (Fig-
ure 7e), and Nb-Zr-Y (Figure 7f). For the source
compositions considered, unusual D values (near
0.005) are assumed for Ta and Nb if the Nb- and Ta-
rich WPB fields in these diagrams are to be covered
by model results. Alternately, source compositions
rich in Nb and Ta could be assumed. One EM source
composition used in WPB calculations is based on
data from Iceland (Wood and others, 1979a) and
should allow coverage of E-MORB discriminant
fields. E-MORB are not adequately covered by WPB
results, suggesting that a broader range of source
compositions is needed. Sources for E-MORB and
WPB on Hf-Ta-Th would have to be relatively
depleted in Th, whereas on Nb-Zr-Y they would
have to have low Zr/Y. These features are unex-
pected in enriched sources.
Spidergrams
Multi-element variation diagrams, extended
rare earth diagrams or spidergrams, have become
increasingly popular. Their major advantages are
easy presentation of a large amount of data and
rapid comparisons between elements and samples.
Their major drawback is lack of uniformity in ele-
ment order and choice of normalizing values. The
most commonly used element orders (Thompson,
1982; Pearce, 1982, 1983) are based roughly on the
expected compatibility of elements in basaltic
magma equilibranted with garnet peridotite, al-
though Pearce (1983) also considered the expected
mobility of elements during hydrothermal altera-
tion. Both compatibility and mobility are depen-
dent on field strength, the ratio of cation charge to
ionic radius. Because element mobility varies as a
function of (1) temperature, (2) rock composition,
(3) fluid composition, and (4) fluid-to-rock ratio,
element orders based on element mobility are, in
part, SUbjective.
Two different approaches have been used in
extracting tectonic information from spidergrams.
The first approach (Pearce, 1982, 1983) uses MORB-
normalized element abundances and apportions
them between DM, within-plate (WP; in many
cases, this is equivalent to EM, or to lithospheric
OM
Figure 8. MORB-normalized spidergram of a hypothetical
rock composition similar to that of some high-K volcanics
from Italy (Rogers and others, 1985). Normalizing values
and element order from Pearce (1982) . The DM field
represents the contribution of depleted mantle (MORB
source); LM represents contribution of lithospheric mantle,
or enriched subcontinental mantle; SZ, which includes
peaks for Ce and Sm, is ascribed to subduction zone
processes. Line separating SZ from LM to the left of Ta is
dashed because its location is arbitrary and materially
affects inferred contributions ofSZ and LM components.
mantle, LM), and subduction-zone (SZ) components.
The most compatible of the incompatible elements,
Y and Yb, plot to the right of Pearce spidergrams
(Figure 8) and are assumed to have abundances
controlled by the DM component. A line drawn
horizontally from and through Y and/or Yb to the
ordinate then represents the estimated contribution
of DM to all other elements. Because HFSE are
depleted in arc volcanics and LILE are enriched, a
line drawn above the DM line and approximately
through Nb, Ta, Zr, Hf, and Ti is assumed to
separate WP (LM) and SZ components (Figure 8).
Contributions from these components can thus be
quantified. Depending on the shape of individual
spidergrams, one or more components might be
missing; MORB spidergrams, for example, are
virtually horizontal on Pearce spidergrams and
require only a DM component. For some arc rocks,
HFSE abundances are approximately equal to those
ofDM, and no WP component is recognizable.
Quantitative interpretation of spidergrams in-
volves numerous assumptions that should be tested
on a case-by-case basis before Pearce's (1983)
Petrogenetic evaluation of trace element discrimination diagrams 113
methodology is used. There are four objections to
Pearce's approach: (1) in the absence of an indepen-
dent determination of the roles of DM, WP, and SZ
in particular samples or sample suites, no reason
exists to restrict the interpretation to these com-
ponents; (2) the assumption that Y and Yb content
are controlled by DM is suspect, particularly for
basaltic rocks derived from garnet-bearing sources;
(3) the distinction of SZ and WP components is
arbitrary, especially for Sr, Rb, Ba, and Th (Figure
8), and is partly a function of element order [note
that Thompson (1982) places Sr in the middle of the
spidergram]; and (4) because differentiation and
partial melting can increase anomalies associated
with HFSE (model calculations show that melts can
have anomalies 20-50% larger than their sources),
significant overestimates of SZ contributions are
likely for all but unfractionated, primitive melts
that form by high degrees of partial melting. The
precision of this approach is difficult to determine;
although Pearce reports component contributions to
the nearest percentage point (Pearce, 1983, p. 244),
analytical imprecision alone may lead to greater
uncertainties.
The second approach to tectonic interpretation
of spidergrams is qualitative. Using data for rocks
of known tectonic setting, Holm (1985) constructed
reference spidergrams for several different tectonic
environments and suggested that tectonic settings
could be identified by comparisons of spidergram
shapes. This approach is similar to that used for
most discrimination diagrams but, because it
employs a larger array of data for each sample, it
avoids the geometric limitations of bivariate or
triangular discrimination diagrams. A modification
of this approach (Myers and Breitkopf, 1989)
involves normalizing all analyses to the value for a
strongly incompatible element (e.g., Nb) in one of
the samples being tested; this effectively removes
element differences due to fractionation and partial
melting and allows comparisons of source composi-
tions. Among all the discrimination techniques,
spider grams hold the greatest promise for consis-
tency and reliability.
Granitic Rocks
Petrogenetic interpretation of discrimination
diagrams for granitoid rocks (Si0
2
2:: 65 wt%) is more
complicated than that for basaltic compositions. At
least four factors contribute to petrogenetic com-
plexity among granitoids: (1) source mineralogy
and composition is variable; (2) a large range of D
exists for most trace elements; (3) accessory min-
erals are potentially stable during partial melting;
and (4) processes associated with variable volatile
contents can have profound impact. Because several
discrimination diagrams for granitoids are based in
part on Rb, alteration and element mobility are
significant concerns (Lidiak and Ceci, 1991).
Petrogenetic interpretation of granitoid dis-
crimination diagrams in Pearce and others (1984)
assumes that the mantle is the ultimate source of
granitic magmas and that granitoids form through
extensive fractionation of basaltic melts. A variety
of experimental, isotopic, and trace element studies,
however, demonstrates that most granitoids,
excluding perhaps ocean-ridge granites (ORG), form
by partial melting of crustal sources at crustal
pressures (Hildreth, 1981; Clemens and Vielzeuf,
1987; Wall and others, 1987; Conrad and others,
1989; but compare Thompson, 1983), rather than by
extreme fractionation. Mantle-like isotopic compo-
sitions in granitic suites probably reflect partial
melting of mantle-derived basaltic protoliths, rather
than fractionation of mafic magmas (e.g., Shirey and
Hanson, 1986). Thus, petrogenetic interpretation of
granitoid discrimination diagrams must consider
melting of crustal materials.
Results for melting models based on an as-
sumed lower crustal composition, on NASC, and on
OS are shown in Figures 9, 10, and 11 (see Ap-
pendix for details of calculations). These models
are not assumed to represent specific tectonic en-
vironments, unlike the models calculated for basal-
tic compositions. Model melts with amphibolitic,
granulitic, and eclogitic residues have distinctive
compositional trends on discrimination diagrams
but, for reasonable degrees of partial melting (2::
10%), rarely extend into a discriminant field differ-
ent than that of the source composition. The limited
range of partial melt compositions is due to the
relatively high D that characterizes low tempera-
ture, silica-rich, and potentially hydrous magmas
and to the relatively large fractions of partial melt-
ing, which force melt compositions to converge near
the source. The inferred tectonic environment of
model granitoids is thus very sensitive to source
composition.
Average crustal sources are not consistently
classified on the discriminant diagrams. For ex-
ample, lower crust (LC) plots in the volcanic arc
granite (VAG) or syn-collisional granite (COLG)
114 F. O. Dudas
1000
COlG WPO
/w

100

Rb VAG
5;
,
I- Te
>0/
"
"

10
,oJ
ORO
.
lC
P5
os
1
1 10 100 1000
Y+ Nb
Figure 9. Rb-Y + Nb discrimination diagram for granitic
rocks (Pearce and others, 1984), showing major crustal
compositions and potential partial melting paths. Tick
marks show 5, 10, and 20% partial melting. UC, upper
crust; TC, total crust; LC, lower crust; NASC, North
American Shale Composite; PS, Pacific sediment; OS, ocea-
nic crust + 10% PS; see Table 2 for other abbreviations
Curve a - amphibolitic residue (see Table A6); curve g-
granulitic residue; curve gz - granulitic residue with zir-
con; curve gw - graywacke model of Conrad and others
(1989), showing range of compositions due to varying D;
curve e - eclogitic residue.
fields on Rb-Y + Nb (Figure 9) and Nb-Y (Figure
11) diagrams but lies in the within-plate granite
(WPG) field on Rb-Hf-Ta (Figure 10). Partial
melts of LC would thus be classified either as VAG,
COLG, or WPG, depending on which diagram was
selected. Lower crust may, in fact, contribute to
magmatism in all of these environments, and no
unique relationship exists between continental
source compositions and apparent tectonic environ-
ment.
Ocean-ridge sources (i.e., DM, OS, or average
MORB) plot outside the ORG fields on Rb-Y +Nb
(Figure 9) and Nb-Y (Figure 11) and could produce
granitic partial melts with ORG (amphibolitic
residue or eclogitic residue containing phlogopite) or
VAG (eclogitic residue) compositions. Addition of
small amounts of sediment to MORB (OS composi-
tions) does not significantly alter the compositions
of model granitoid melts, which remain in the VAG
field for most garnet-bearing residue compositions.
The composite axes (Y +Nb, Zr+Nb+Ce+ Y)
used for some discrimination diagrams involve ele-
ments that are sensitive to the presence of accessory
Tax3
Figure 10. Rb-Hf-Ta discrimination diagram for granitic
rocks (Harris and others, 1986). Curve gi - granulite
residue with magnetite and ilmenite; other curves labeled
as in Figure 9.
phases (e.g., allanite or zircon) and therefore may
behave in opposite ways during petrogenetic pro-
cesses. Petrogenetic interpretation of diagrams
with composite axes is extremely difficult.
DISCUSSION
Discriminant fields represent the approximate
ranges of compositional variation in igneous rocks.
Petrogenetic models should be able to reproduce
that range and to indicate what sources and
processes contribute to it. In most instances, model
compositions plot at least partly within the
discriminant fields they were intended to cover.
Most model compositions, however, also overlap
discriminant fields adjacent to them, which indi-
cates that specific models are not unique to the
discriminant fields they were designed to represent.
By extension, magmas formed in specific tectonic
environments are also expected to show a range of
compositions that is not restricted to the discrimin-
ant fields. Overlap is expected both in model com-
positions and in real rocks, and interpretations of
discrimination diagrams will necessarily involve
misidentification of tectonic environments for some
rocks. Models suggest that the extent of overlap is
Petrogenetic evaluation of trace element discrimination diagrams 115
Figure 11. Nb-Y discrimination diagram for granitic
rocks (Pearce and others, 1984). Curves labeled as in
Figure 9. Extreme depletion ofY in eclogite models (curve
e) is due to high Dy for garnet.
much greater than the"" 10% failure rate quoted for
several of the discrimination diagrams; for MORB
compositions, the extent of overlap is almost com-
plete.
Under the best circumstances, geochemical
data alone rarely support a unique petrogenetic or
petrologic interpretation. Discrimination diagrams
add another level of interpretation to that involved
in identifying magma sources and evolution pro-
cesses, and add another level of uncertainty as well.
Tectonic interpretation of geochemical variations
may be possible if the sources and processes in-
volved in petrogenesis are identified first. If, as for
many altered rocks in tectonically complex settings,
geochemical variation provides the primary evi-
dence for tectonic interpretation, the possibilities of
misinterpretation and the risk of misclassification
are increased.
Petrogenetic evaluation of discrimination dia-
grams for, mafic rocks at least, suggests that the
diagrams may distinguish magma series. It may be
possible to differentiate alkalic from tholeiitic and
calcalkalic suites. These distinctions depend on the
mineralogy of the residue in the source after melt-
ing, and some diagrams (especially those involving
Y, which is compatible in garnet) appear success-
fully to distinguish garnet-bearing and garnet-
absent melting histories.
The models presented above are not exhaustive
and do not adequately reflect the multiplicity of
processes and sources involved in igneous petro-
genesis. To what extent might real conditions differ
from those assumed in the calculations and allow a
better discrimination than the models predict? If
source composition, source mineralogy, and D for
individual elements (i.e., factors to which models
were most sensitive) in natural systems were
different from model values, natural magmas might
fit discriminant fields more closely than model
results do. Additionally, if processes other than
those considered in the models are operative, a dif-
ferent range of variability might result.
Mantle source compositions are almost certain-
ly more variable than those selected for modeling
and, in part, account for areas on discrimination
diagrams that are not covered by model results. A
broader range of sources, in all tectonic environ-
ments, would lead to increased variations in magma
compositions rather than to the more restricted
range that is required to minimize compositional
overlaps.
Source mineralogies are undoubtedly more
variable than those modeled. If residues after
melting include accessory phases in which trace
elements are compatible, a much larger range of
magma compositions could result. Accessory min-
erals have been invoked in explaining HFSE
depletion of arc magmas, but experimental data
suggest that Ti-rich minerals are unlikely to be
stable under conditions appropriate to generation of
arc magmas (Ryerson and Watson, 1987). Other ac-
cessory phases (carbonates, phlogopite, amphibole)
do not, for the most part, have D large enough to
have significant impact on bulk D during melting,
and thus they probably have minimal effects on
trace element concentrations in mantle melts (apa-
tite, with high D for REE and Th, is a notable
exception). For strongly alkalic magmas, where
unusual source compositions, source mineralogies,
and low degrees of partial melting might be in-
volved, a larger range of melt compositions exists
than is indicated by the WPB models.
Simple models were used in the petrogenetic
calculations. More complex models of partial melt-
ing and differentiation processes have been devised
and applied in specific case studies. Thus, for
example, multi-cycle melting (Langmuir and others,
1977), melt retention within the source region
(Langmuir and others, 1977), zone refining (Harris,
1974), in situ crystallization (Langmuir, 1989), cyc-
lic fractionation, tapping and replenishment of
magma chambers (RTF magma chambers; O'Hara
116 F. 6. Dudas
and Matthews, 1981), and assimilation coupled with
fractionation (AFC; DePaolo, 1981; Taylor, 1980;
Nielsen, 1989) have all been modeled. These models
require more constraints than can be applied to the
general cases treated here. In some of these models,
the range of compositional variation is extended
compared to simple partial melting or differen-
tiation (e.g.,in situ crystallization and RTF models),
whereas in others the range of variation is smaller
(e.g., AFC compared with simple bulk mixing).
Unless a complex model consistently applies to mag-
ma generation in a certain tectonic environment,
simple models provide an adequate estimate of
compositional variations.
In within-plate environments, enrichment or
metasomatism of mantle sources seems to be
commonplace (Menzies and Hawkesworth, 1987).
Enriched mantle is usually heterogeneous on a
small scale and commonly also contains accessory
minerals. The enriching agent is thought to be
silicic or carbonatitic melt (Eggler, 1987), possibly
derived from recycling of crust into the mantle
(Cohen and O'Nions, 1982; Hofmann and White,
1982; McKenzie and O'Nions, 1983). In some cases,
the enriched source has HFSE depletion and LILE
enrichment (Weaver and others, 1986), the diag-
nostic signature of arc magmatism. When such a
source is melted in a within-plate environment,
resulting magmas are HFSE depleted and LILE
enriched, they plot within the arc fields of dis-
crimination diagrams (Dudas, 1991), and they are
misclassified.
Long-term mixing of major geochemical reser-
voirs - DM, EM, LC, UC - is a necessary conse-
quence of mantle convection and crustal cycling in
subduction zones. Because recycled crust may pene-
trate at least to 700 km within the mantle (Silver
and others, 1988), the time lag between subduction
and the participation of subducted material in
subsequent magmatism may exceed a billion years.
Sources formed by these mixing processes may be
tapped by partial melting at times and places that
are remote from the subduction zones themselves.
Thus, some element patterns in MORB require a
recycled, old crustal component (Shirey and others,
1987; Ito and others, 1987) and some element abun-
dances in within-plate, ocean island rocks have a
"subduction-related" component (Shimizu, 1987;
Weaver and others, 1986), but in neither case is
continental crust or subduction directly involved.
HFSE depletion and LILE enrichment also occur in
some continental basalts (Weaver and Tarney, 1981;
Pegram, 1983), suggesting that the factors control-
ling this arc-like geochemical signature are not
unique to the arc environment.
No unique relationship exists between sources
and tectonic environments. The MORB source
contributes magmas to ocean ridge volcanism but is
found in within-plate (Fitton and Dunlop, 1985;
Allegre and others, 1981; Carlson, 1984) and arc
settings (Hole and others, 1984; Perfit and others,
1980; Rogers and others, 1985; Ryerson and Watson,
1987) as well. Enriched sources characterize
within-plate magmatism but also occur in arcs
(Rogers and others, 1985; Wheller and others, 1987)
and ocean ridge environments (Cohen and O'Nions,
1982; Ito and others, 1987).
The ability to decipher tectonic environments
from geochemical patterns retained by altered,
ancient rocks is an alluring target. Lack of direct
correlation between chemistry and tectonic setting
and absence of a conceptual basis for such a cor-
relation suggest that geochemical discrimination is
not possible without substantial risk of error. The
ease of using diagrams proposed in the literature is
seductive, especially if results fit with our prejudices
or if independent geological indicators support a
compatible tectonic interpretation. Alone, geo-
chemical discrimination diagrams convey empirical
interpretations but do not deal with geological facts.
CONCLUSIONS
1) Analytical uncertainty, "soft boundaries" of
discriminant fields, and geometric limitations of
individual discrimination diagrams conspire to re-
quire analyses of multiple samples and use of
multiple diagrams if tectonic classification by trace
element abundances is to be successfully attempted.
2) Tectonic discrimination diagrams can reliably
be used if five conditions are satisfied. In increasing
order of importance, these are (1) analytical un-
certainty must be minimal, compared to the dis-
criminant fields, (2) element mobility subsequent to
crystallization must be minimal, (3) initial con-
struction of discriminant diagrams must utilize data
from tectonic environments that are appropriate
analogues to those sought in data to be classified,
(4) petrogenetic processes must be primarily asso-
ciated with specific tectonic environments, and
(5) source compositions must be primarily associ-
ated with specific tectonic environments. If condi-
tions 3, 4, and 5 cannot be met, discrimination
Petrogenetic evaluation of trace element discrimination diagrams 117
diagrams are unlikely to be reliable. Of the dia-
grams examined, the spidergram approach of Holm
(1985) seems most promising.
3) Model results for MORB, V AB, and WPB usual-
ly cover portions of the discriminant fields they were
designed to reproduce but also show extensive over-
lap of melt compositions into adjacent discriminant
fields. The extent of overlap exceeds that expected
from success rates quoted for individual diagrams
and is most pronounced for MORB. Unless unusual
source compositions or distribution coefficients are
invoked, some portions of discriminant fields are not
accessible to model results. Petrogenetic processes
subsequent to partial melting have little influence
on trace element proportions and have little effect
on how model results plot on discrimination dia-
grams.
4) Model results for basaltic rocks are sensitive to
source composition, source mineralogy, extent of
partial melting, and variation of D. Model para-
meters for these factors are conservative in that
expected variations in natural environments exceed
those used in the models. Natural systems are
expected to have larger variations in melt com-
positions than those shown by the models and,
consequently, are also expected to have greater
overlap with other discriminant fields.
5) Model granitoid melts usually remain in the
discriminant field occupied by their source composi-
tion. The source dependence of tectonic classifica-
tion for granitoid melts implies that granites are
likely to inherit apparent tectonic characteristics
from their sources, regardless of the actual con-
ditions of magma generation.
6) Source compositions are not uniquely related to
tectonic environments. Cycling of materials
through the mantle as a consequence of subduction
implies long-term mixing of major geochemical
reservoirs and a lack of unique correspondence
between source characteristics and tectonic settings
of magmatism. Inheritance of geochemical com-
ponents from subducted materials can produce arc-
like features in magmas from any environment,
which consequently leads to misclassifications.
7) Geochemical data alone are not adequate for
tectonic classification.
ACKNOWLEDGMENTS
Discussions with D. H. Eggler, J. K. Meen, and
W. P. Leeman sharpened the focus of ideas presen-
ted in this paper. Reviews by A. N. LeCheminant,
R. L. Lustwerk, J. C. Roddick, R. Stern, and two ano-
nymous referees contributed significantly to the
final product. Though they may disagree with my
ultimate conclusions, I wish to thank them for their
patience and help.
REFERENCES
ALLEGRE, C. J., B. DUPRE, B. LAMB RET, and P. RICHARD,
1981, The subcontinental versus suboceanic debate, I:
Lead-neodymium-strontium isotopes in primary alkali
basalts from a shield area: The Ahaggar volcanic suite:
Earth and Planetary Science Letters, v. 52, p. 85-92.
ARCULUS, R. J., 1987, The significance of source versus
process in the tectonic controls of magma genesis: Journal
of Volcanology and Geothermal Research, v. 32, p. 1-12.
__ , and R. POWELL, 1986, Source component mixing in
the regions of arc magma generation: Journal of Geophysi-
cal Research, v. 91B, p. 5913-5926.
ARTH, J. G., 1976, Behavior of trace elements during
magmatic processes - A summary of theoretical models
and their applications: U.S. Geological Survey Journal of
Research, v. 4, p. 41-47.
BAILEY, J. C., 1981, Geochemical criteria for a refined
tectonic discrimination of orogenic andesites: Chemical
Geology, v. 32, p. 139-154.
BECCALUVA, L., D. OHNENSTETTER, and M. OHNEN-
STETTER, 1979, Geochemical discrimination between
ocean-floor and island-arc tholeiites - Application to some
ophiolites: Canadian Journal of Earth Science, v. 16, p.
1874-1882.
CARLSON, R. W., 1984, Isotopic constraints on Columbia
River flood basalt genesis and the nature of the sub-
continental mantle: Geochimica et Cosmochimica Acta, v.
48, p. 2357-2372.
CLEMENS, J. D., and D. VIELZEUF, 1987, Constraints on
melting and magma production in the crust: Earth and
Planetary Science Letters, v. 86, p. 287-306.
COHEN, R. S., and R. K. O'NIONS, 1982, Identification of
recycled continental material in the mantle from Sr, Nd
and Pb isotope investigations: Earth and Planetary Sci-
ence Letters, v. 61, p. 73-84.
CONDIE, K. C., 1989, Geochemical changes in basalts and
andesites across the Archean-Proterozoic boundary: Iden-
tification and significance: Lithos, v. 23, p. 1-18.
CONRAD, W. K., 1. A. NICHOLLS, and V. J. WALL, 1989,
Water-saturated and -undersaturated melting of meta-
luminous and peraluminous crustal compositions at 10 kb:
Evidence for the origin of silicic magmas in the Taupo
Volcanic Zone, New Zealand, and other occurrences:
Journal of Petrology, v. 29, p. 765-803.
118
F. o. Dudas
DAUTRIA, J. M., J. M. LIOTARD, N. CABANES, M. GIROD,
and L. BRIQUEU, 1987, Amphibole.rich xenoliths and host
alkali basalts: Petrogenetic constraints and implications
on the recent evolution of the upper mantle beneath Ahag
gar (Central Sahara, southern Algeria): Contributions to
Mineralogy and Petrology, v. 95, p. 133144.
DEPAOLO, D. J., 1981, Trace element and isotopic effects of
combined wallrock assimilation and fractional crystalliza
tion: Earth and Planetary Science Letters, v. 53, p. 189202.
DIETRICH, V. J.,W. FRANK, and K. HONEGGER, 1983, A
JurassicCretaceous island arc in the LadakhHimalayas:
Journal of Volcanology and Geothermal Research, v. 18, p.
405433.
DUDAs, F. 0.,1991, Geochemical features of igneous rocks
from the Crazy Mountains, Montana, and tectonic models
for the Montana alkalic province: Journal of Geophysical
Research, v. 96B, p.1326113277.
__ ,1. H. CAMPBELL, and M. P. GORTON, 1983, Geo
chemistry of igneous rocks in the Hokuroku District,
northern Japan; pp. 115133 in H. Ohmoto and B. J. Skin
ner (eds.), The Kuroko and Related Volcanogenic Mas
sive Sulfide Deposits: Economic Geology Monograph 5,
604p.
DUNCAN, A. R., 1987, The Karroo igneous province - A
problem for inferring tectonic setting from basalt geo
chemistry: Journal of Volcanology and Geothermal Re
search, v. 32, p. 1334.
DUNN, T., 1987, Partitioning ofHf, Lu, Ti and Mn between
olivine, clinopyroxene and basaltic liquid: Contributions to
Mineralogy and Petrology, v. 96, p. 476484.
__ , and 1. S. McCALLUM, 1982, The partitioning of Zr
and Nb between diopside and melts in the system diopside
albiteanorthite: Geochimica et Cosmochimica Acta, v. 46,
p.623629.
EGGLER, D. H., 1987, Solubility of major and trace ele
ments in mantle metasomatic fluids: Experimental con
straints; pp. 2141 in M. A. Menzies and C. J. Hawkes
worth (eds.), Mantle Metasomatism: Academic Press,
London, England, UK, 472 p.
ELLAM, R. M., and C. J. HAWKESWORTH, 1988, Elemental
and isotopic variations in subduction related basalts: Evi
dence for a three component model: Contributions to Min
eralogy and Petrology, v. 98, p. 7280.
FITTON, J. G., and H. M. DUNLOP, 1985, The Cameroon
Line, West Africa, and its bearing on the origin of oceanic
and continental alkali basalt: Earth and Planetary Science
Letters, v. 72, p. 2338.
FLOYD, P. A., and J. A. WINCHESTER, 1975, Magma type
and tectonic setting discrimination using immobile ele
ments: Earth and Planetary Science Letters, v. 27, p. 211
218.
FRANCALANCI, L., 1989, Trace element partition coeffi
cients for minerals in shoshonitic and calc alkaline rocks
from Stromboli Island (Aeolian Arc): Neues Jahrbuch fur
Mineralogie, Abhandlungen, v. 160, p. 229247.
FREY, F. A., D. H. GREEN, and S. D. ROY, 1978, Integrated
models of basalt petrogenesis: A study of quartz tholeiites
to olivine melilitites from southeastern Australia utilizing
geochemical and experimental petrological data: Journal
of Petrology, v. 19, p. 463513.
GANAPATHY, R., and E. ANDERS, 1974, Bulk compositions
of the moon and earth, estimated from meteorites [Pro
ceedings of the Fifth Lunar Conference, Supplement 5J:
Geochimica et Cosmochimica Acta, v. 2, p. 11811206.
GAST, P. W., 1968, Trace element fractionation and the
origin of tholeiitic and alkaline magma types: Geochimica
et Cosmochimica Acta, v. 32, p. 10571086.
GLASSLEY, W., 1974, Geochemistry and tectonics of the
Crescent volcanic rocks, Olympic Peninsula, Washington:
Geological Society of America Bulletin, v. 85, p. 785794.
GREEN, D. H., and E. RINGWOOD, A. 1967, The genesis of
basaltic magmas: Contributions to Mineralogy and Petro
logy, v. 15, p. 103190.
GREEN, T. H., and N. J. PEARSON, 1987, An experimental
study of Nb and Ta partitioning between Tirich minerals
and silicate liquids at high pressure and temperature:
Geochimica et CosmochimicaActa, v. 51, p. 5562.
__ , SIE, S. H., C. G. RYAN, and D. R. COUSENS, 1989,
Proton microprobedetermined partitioning of Nb, Ta, Zr,
Sr and Y between garnet, clinopyroxene and basaltic mag
ma at high pressure and temperature: Chemical Geology,
v. 74, p. 201216.
GROMET, L. P., R. F. DYMEK, L. A. HASKIN, and R. L.
KOROTEV, 1984, The "North American Shale Composite":
Its compilation, major and trace element characteristics:
Geochimica et Cosmochimica Acta, v. 48, p. 24692482.
HANSON, G. N., 1978, The application of trace elements to
the petrogenesis of igneous rocks of granitic composition:
Earth and Planetary Science Letters, v. 38, p. 2643.
HARKER, A., 1897, The natural history of igneous rocks:
Science Progress, v. 6, p. 1233.
HARRIS, N. B. W., J. A. PEARCE, and A. G. TINDLE, 1986,
Geochemical characteristics of collisionzone magmatism;
pp. 6781 in M. P. Coward and A. C. Ries (eds.J, CoUision
Tectonics: Geological Society of London, Special Publica
tion 19,415 p.
HARRIS, P. G., 1974, Origin of alkaline magmas as a result
of anatexis: a. Anatexis and other processes within the
mantle; pp. 427436 in H. (ed.), The Alkaline
Rocks: Wiley, London, England, UK, 622 p.
HART, S. R., and C. BROOKS, 1974, Clinopyroxenematrix
partitioning ofK, Rb, Cs, Sr and Ba: Geochimica et Cosmo
chimica Acta, v. 38, p. 17991806.
HAWKINS, J. W., Jr., 1980, Petrology of backarc basins
and island arcs: Their possible role in the origin of
Petrogenetic evaluation of trace element discrimination diagrams 119
ophiolites; p. 244-254 in A. Panayiotou (ed.), Ophiolites:
Proceedings of the International Ophiolite Sympos-
ium: Cypriot Ministry of Agriculture and Natural Re-
sources, Nicosia, Cyprus, 781 p.
HILDRETH, W., 1981, Gradients in silicic magma cham-
bers: Implications for lithospheric magmatism: Journal of
Geophysical Research, v. 86B, p. 10153-10192.
HOFMANN, A. W., and W. M. WHITE, 1982, Mantle plumes
from ancient oceanic crust: Earth and Planetary Science
Letters, v. 57, p. 421-436.
HOLE, M. J., A. D. SAUNDERS, G. F. MARRINER, and J.
TARNEY, 1984, Subduction of pelagic sediments: Implica-
tions for the origin of Ce-anomalous basalts from the
Mariana Islands: Journal of the Geological Society of Lon-
don,v.141,p.453-472.
HOLM, P. E., 1982, Non-recognition of continental tho-
leiites using the Ti-Y -Zr diagram: Contributions to Min-
eralogy and Petrology, v. 79, p. 308-310.
__ , 1985, The geochemical fingerprints of different
tectonomagmatic environments using hygromagmatophile
element abundances of tholeiitic basalts and basaltic an-
desites: Chemical Geology, v. 51, p. 303-323.
HUMPHRIS, S. E., and G. THOMPSON, 1978, Trace element
mobility during hydrothermal alteration of oceanic ba-
salts: Geochimica et Cosmochimica Acta, v. 42, p. 127-136.
IKEDA, Y., 1990, CeN/SrN/SmN: A trace element discri-
minant for basaltic rocks from different tectonomagmatic
environments: Neues Jahrbuch fur Mineralogie und
Paldontologie, Monatshefte, v. 17, p. 145-158.
IRVING, A. J., 1978, A review of experimental studies of
crystallliquid trace element partitioning: Geochimica et
CosmochimicaActa, v. 42, p. 743-770.
__ , and R. C. PRICE, 1981, Geochemistry and evolution
of lherzolite-bearing phonolitic lavas from Nigeria, East
Germany and New Zealand: Geochimica et Cosmochimica
Acta, v. 45, p. 1309-1320.
ITO, E., W. M. WHITE, and C. GOPEL, 1987, The 0, Sr, Nd
and Pb isotopic geochemistry ofMORB: Chemical Geology,
v. 62, p. 157-176.
JAGOUTZ, E., H. PALME, H. BADDENHAUSEN, K. BLUM,
M. CENDALES, G. DREillUS, B. SPETTEL, V. LORENZ, and
H. WANKE, 1979, The abundances of major, minor and
trace elements in the earth's mantle as derived from
primitive ultramafic nodules: Proceedings of the Tenth
Lunar and Planetary Science Conference, p. 2031-2050
JENSEN, L. S., 1976, A New Cation Plot for Classifying
SubalkaUne Volcanic Rocks: Ontario Ministry of Natur-
al Resources, Miscellaneous Paper No. 66,22 p.
KAY, R. W., 1980, Volcanic arc magmas: Implications ofa
melting-mixing model for element recycling in the crust-
upper mantle system: Journal of Geology, v. 88, p. 497-
522.
KENNEDY, W. Q., 1933, Trends of differentiation in basal-
tic magmas: American Journal of Science, v. 25, p. 239-
256.
LANGMUIR, C. H., 1989, Geochemical consequences of in
situ crystallization: Nature, v. 340, p. 199-205.
__ , J. F. BENDER, A. E. BENCE, G. N. HANSON, and S.
R. TAYLOR, 1977, Petrogenesis of basalts from the
FAMOUS area: Mid-Atlantic Ridge: Earth and Planetary
Science Letters, v. 36, p. 133-156.
LEROEX, A. P., 1987, Source regions of mid-ocean ridge
basalts: Evidence for enrichment processes; pp. 389-422 in
M. A. Menzies and C. J. Hawkesworth (eds.), Mantle
Metasomatism: Academic Press, London, England, UK,
472p.
LETERRIER, J., J. C. MAURY, P. THONON, D. GIRARD,
and M. MARCHAL, 1982, Clinopyroxene composition as a
method of identification of the magmatic affinities of
paleo-volcanic series: Earth and Planetary Science Letters,
v. 59, p. 139-154.
LIDIAK, E. G., and V. M. CECI, 1991, Authigenic K-
feldspar in the Precambrian basement of Ohio and its
effect on tectonic discrimination of the granitic rocks:
Canadian Journal of Earth Sciences, v. 28, p. 1626-1634.
LUDDEN, J. N., and G. THOMPSON, 1978, Behavior of rare
earth elements during submarine weathering of tholeiitic
basalt: Nature, v. 274, p. 147-149.
LUHR, J. F., and 1. S. E. CARMICHAEL, 1980, The Colima
volcanic complex, Mexico, I: Post-caldera andesites from
Volcan Colima: Contributions to Mineralogy and Petro-
logy, v. 71, p. 343-372.
__ ,1. S. E. CARMICHAEL, and J. C. VAREKAMP, 1984,
The 1982 eruptions of El Chi chOn volcano, Chiapas, Mexi-
co: Mineralogy and petrology of the anhydrite-bearing
pumices: Journal of Volcanology and Geothermal Re-
search, v. 23, p. 69-108.
McCALLUM, 1. S., and M. P. CHARETTE, 1978, Zr and Nb
partition coefficients: Implications for the genesis of mare
basalts, KREEP, and sea floor basalts: Geochimica et
Cosmochimica Acta, v. 42, p. 859-869.
McKENZIE, D., and R. K. O'NIONS, 1983, Mantle reser-
voirs and ocean island basalts: Nature, v. 301, p. 229-231.
MAHOOD, G., and W. HILDRETH, 1983, Large partition
coefficients for trace elements in high-silica rhyolites:
Geochimicaet CosmochimicaActa, v. 47, p. 11-30.
MANIAR, P. D., and P. M. PICCOLI, 1989, Tectonic dis-
crimination of granitoids: Geological Society of America
Bulletin, v. 101, p. 635-643.
MARSH, J. S., 1987, Basalt geochemistry and tectonic
discrimination within continental flood basalt provinces:
Journal of Volcanology and Geothermal Research, v. 32, p.
35-49.
120 F. o. Dudas
MEEN, J. K., 1987, Formation of shoshonites from calc-
alkaline basalt magmas: Geochemical and experimental
constraints from the type locality: Contributions to
Mineralogy and Petrology, v. 97, p. 333-35l.
__ , D. H. EGGLER, and J. C. AYERS, 1989, Experi-
mental evidence for very low solubility of rare-earth
elements in CO
2
-rich fluids at mantle conditions: Nature,
v. 340, p. 301-303.
MEIJER, A., 1976, Pb and Sr isotopic data bearing on the
origin of volcanic rocks from the Mariana island-arc
system: Geological Society of America Bulletin, v. 87, p.
1358-1369.
MENZIES, M. A., and C. J. HAWKESWORTH, 1987, mantle
metasomatism: Academic Press, London, England, UK,
472p.
MENZIES, M., SEYFRIED, W. E., Jr., and BLANCHARD,
D., 1979, Experimental evidence ofrare earth element im-
mobility in greenstones: Nature, v. 282, p. 398-399.
MESCHEDE, M., 1986, A method of discriminating between
different types of mid-ocean ridge basalts and continental
tholeiites with the Nb-Zr-Y diagram: Chemical Geology,
v. 56, p. 207-218.
MORRIS, P. A., 1988, Volcanic arc reconstruction using
discriminant function analysis of detrital clinopyroxene
and amphibole from the New England fold belt, eastern
Australia: Journal of Geology, v. 96, p. 299-311.
MULLEN, E. D., 1983, MnO/Ti0
2
/P
2
0
5
: A minor element
discriminant for basaltic rocks of oceanic environments
and its implications for petrogenesis: Earth and Planetary
Science Letters, v. 62, p. 53-62.
MURPHY, J. B., and A. J. HYNES, 1986, Contrasting
secondary mobility of Ti, P, Zr, Nb and Y in two meta-
basaltic suites in the Appalachians: Canadian Journal of
Earth Science, v. 23, p. 1138-1144.
MYERS, R. E., and J. H. BREITKOPF, 1989, Basalt
geochemistry and tectonic settings: A new approach to
relate tectonic and magmatic processes: Lithos, v. 23, p.
53-62.
NAGASAWA, H., and C. C. SCHNETZLER, 1971, Parti-
tioning of rare earth, alkali and alkaline earth elements
between phenocrysts and acid igneous magma: Geochimi-
ca et CosmochimicaActa, v. 35, p. 953-968.
NASH, W. P., and H. R. CRECRAFT, 1985, Partition
coefficients for trace elements in silicic magmas: Geo-
chimica et Cosmochimica Acta, v. 49, p. 2309-2322.
NEUMAN, H., J. MEAD, and C. J. VITALIANO, 1954, Trace
element variation during fractional crystallization as cal-
culated from the distribution law: Geochimica et Cosmo-
chimica Acta, v. 6, p. 90-100.
NIELSEN, R. L., 1989, Phase equilibria constraints on
liquid lines of descent generated by paired assimilation
and fractional crystallization: Trace elements and Sr and
Nb isotopes: Journal of Geophysical Research, v. 94B, p.
787-794.
NIGGLI, P., 1923, Gesteins- undMineralprovinzen: Ber-
lin, Germany, 602 p.
NISBET, E. G., and J. A. PEARCE, 1977, Clinopyroxene
composition in mafic lavas from different tectonic settings:
Contributions to Mineralogy and Petrology, v. 63, p. 149-
160.
O'HARA, M. J., and R. E. MATHEWS, 1981, Geochemical
evolution in an advancing, periodically replenished, peri-
odically tapped, continuously fractionated magma cham-
ber: Journal of the Geological Society of London, v. 138, p.
237-277.
OLAFSSON, M., and D. H. EGGLER, 1983, Phase relations
of amphibole, amphibole-carbonate, and phlogopite-car-
bonate peridotite: Petrologic constraints on the astheno-
sphere: Earth and Planetary Science Letters, v. 64, p. 305-
315.
PEACOCK, M. A., 1931, Classification of igneous rock
series: Journal of Geology, v. 39, p. 54-67.
PEARCE, J. A., 1975, Basalt geochemistry used to inves-
tigate past tectonic environments on Cyprus: Tectono-
physics, v. 25, p. 41-67.
__ , 1976, Statistical analysis of major element patterns
in basalts: Journal of Petrology, v. 17, p. 15-43.
1982 Trace element characteristics of lavas from
boundaries; pp. 525-548 in R. S. Thorpe
(ed.), Andesites: Orogenic Andesites and Related
Rocks: John Wiley and Sons, New York, New York, USA,
724p.
__ , 1983, Role of the subcontinental lithosphere in
magma genesis at active continental margins; pp. 230-249
in C. J. Hawkesworth and M. J. Norry (eds.), Continental
Basalts and Mantle Xenoliths: Shiva Publishing, Nant-
wich, Cheshire, England, UK, 272 p.
__ , 1987, An expert system for the tectonic charac-
terization of ancient volcanic rocks: Journal of Volcano-
logy and Geothermal Research, v. 32, p. 51-65.
__ , and J. R. CANN, 1971, Ophiolite origin investigated
by discriminant analysis using Ti, Zr and Y: Earth and
Planetary Science Letters, v. 12, p. 339-349.
__ , and J. R. CANN, 1973, Tectonic setting of basic
volcanic rocks determined using trace element analysis:
Earth and Planetary Science Letters, v. 19, p. 290-300.
___ , and G. H. GALE, 1977, Identification of ore-
deposition environment from trace-element geochemistry
of associated igneous host rocks; pp. 14-23 in M. J. Jones
(ed.), Volcanic Processes in Ore Genesis: Institution for
Mining and Metallurgy/Geological Society of London,
Special Publication 7, 188 p.
__ , and M. J. NORRY, 1979, Petrogenetic implications of
Petrogenetic evaluation oftrace element discrimination diagrams 121
Ti, Zr, Y and Nb variations in volcanic rocks: Contri-
butions to Mineralogy and Petrology, v. 69, p. 33-47.
__ , T. ALABASTER, A. W. SHELTON, andM. P. SEARLE,
1981, The Oman ophiolite as a Cretaceous arc-basin com-
plex: Evidence and implications: Philosophical Trans-
actions of the Royal Society of London, Series A, v. 300, p.
299-317.
__ , N. B. W. HARRIS, and A. G. TINDLE, 1984, Trace
element discrimination diagrams for the tectonic inter-
pretation of granitic rocks: Journal of Petrology, v. 25, p.
956-983.
PEARCE, T. H., B. E. GORMAN, and T. C. BIRKETT, 1975,
The Ti0z-K20-P205 diagram: A method of discriminating
between oceanic and non-oceanic basalts: Earth and
Planetary Science Letters, v. 24, p. 419-426.
__ , B. E. GORMAN, and T. C. BIRKETT, 1977, The
relationship between major element chemistry and tec-
tonic environment of basic and intermediate volcanic
rocks: Earth and Planetary Science Letters, v. 36, p. 121-
132.
PEGRAM, W. J., 1983, Geochemical Processes in the
Subcontinental Mantle and the Nature of the Crust-
Mantle Interaction - Evidence from the Mesozoic
Appalachian Tholeiite Province: PhD thesis, Massa-
chusetts Institute of Technology, Cambridge, Massachu-
setts, USA.
PERFIT, M. R, D. A. GUST, A. E. BENCE, R J. ARCULUS,
and S. R. TAYLOR, 1980, Chemical characteristics of
island-arc basalts: Implications for mantle sources:
Chemical Geology, v. 30, p. 227-256.
PLANK, T., and C. H. LANGMUIR, 1988, An evaluation of
the global variations in the major element chemistry of arc
basalts: Earth and Planetary Science Letters, v. 90, p. 349-
370.
PRESTVIK, T., 1982, Basic volcanic rocks and tectonic
setting: A discussion of the Zr-Ti-Y diagram and its
suitability for classification purposes: Lithos, v. 15, p. 241-
247.
RITTMANN, A., 1962, Volcanoes and Their Activity:
John Wiley and Sons, New York, New York, USA, 305 p.
RODEN, M. F., F. A. FREY, and D. M. FRANCIS, 1983, An
example of consequent mantle metasomatism in peridotite
inclusions from Nunivak Island, Alaska: Journal of
Petrology, v. 25, p. 546-577.
ROGERS, N. W., C. J. HAWKESWORTH, R J. PARKER, and
J. S. MARSH, 1985, The geochemistry of potassic lavas
from Vulsini, central Italy, and implications for mantle
enrichment processes beneath the Roman region: Contri-
butions to Mineralogy and Petrology, v. 90, p. 244-257.
RYERSON, F. J., and E. B. WATSON, 1987, Rutile satura-
tion in magmas: Implications for Ti-Nb-Ta depletion in
island-arc basalts: Earth and Planetary Science Letters, v.
86, p. 225-239.
SCHILLING, J.-G., and J. W. WINCHESTER, 1967, Rare-
earth fractionation and magmatic processes; pp. 267-283 in
S. K. Runcorn (ed.), Mantles of the Earth and Terres-
trial Planets: Interscience Publishers, New York, New
York, USA, 584p.
SCHNEIDER, M. E., and D. H. EGGLER, 1986, Fluids in
equilibrium with peridotite minerals: Implications for
mantle metasomatism: Geochimica et Cosmochimica Acta,
v. 50, p. 711-724.
SCHNETZLER, C. C., and J. A. PHILPOTTS, 1968, Partition
coefficients of rare-earth elements and barium between
igneous matrix material and rock-forming-mineral pheno-
crysts, I; pp. 929-938 in L. H. Ahrens (ed.), Origin and
Distribution of the Elements: Pergamon Press, Oxford,
England, UK, 1178 p.
SHAW, D. M., 1970, Trace element fractionation during
anatexis: Geochimica et Cosmochimica Acta, v. 34, p. 237-
243.
SHERV AIS, J. W., 1982, Ti-V plots and the petrogenesis of
modern and ophiolitic lavas: Earth and Planetary Science
Letters, v. 59, p. 101-118.
SHIMIZU, N., 1987, Trace element abundance patterns of
pyroxenes in spinel lherzolite nodules from Salt Lake
Crater, Oahu [abstract]: EOS, Transactions of the Ameri-
can Geophysical Union, v. 68, p. 446.
SHIREY, S. B., and G. N. HANSON, 1986, Mantle hetero-
geneity and crustal recycling in Archean granite-green-
stone belts: Evidence from Nd isotopes and trace elements
in the Rainy Lake area, Superior Province, Ontario,
Canada: Geochimica et Cosmochimica Acta, v. 50, p. 2631-
265l.
__ , J. F. BENDER, and C. H. LANGMUIR, 1987, Three-
component isotopic heterogeneity near the Oceanographer
Transform, Mid-Atlantic Ridge: Nature, v. 325, p. 217-223.
SILVER, P., R. W. CARLSON, and P. OLSON, 1988, Deep
slabs, geochemical heterogeneity, and the large-scale
structure of mantle convection: Investigation of an en-
during paradox: Annual Review of Earth and Planetary
Sciences, v. 16, p. 477-54l.
SYLVESTER, P. J., 1989, Post-collisional alkaline granites:
Journal of Geology, v. 97, p. 261-280.
TAYLOR, H. P., Jr., 1980, The effects of assimilation of
country rocks by magmas on 180 /
16
0 and 87Sr/86Sr syste-
matics in igneous rocks: Earth and Planetary Science
Letters, v. 47, p. 243-254.
TAYLOR, S. R, and S. M. McLENNAN, 1981, The com-
position and evolution of the continental crust: Rare earth
element evidence from sedimentary rocks: Philosophical
Transactions of the Royal Society of London, v. 301A, p.
381-399.
__ , and S. M. McLENNAN, 1985, The Continental
Crust: Its Composition and Evolution: Blackwell Sci-
entific Publishers, Oxford, England, UK, 312 p.
122
F. O. Dudas
TERA, F., L. BROWN, J. MORRIS, 1. S. SACKS, J. KLEIN,
and R. MIDDLETON, 1986, Sediment incorporation in
island-arc magmas: Inferences from lOBe: Geochimica et
CosmochimicaActa, v. 50, p. 535-550.
THIEBLEMONT, D., and B. CABAN IS, 1990, Utilisation
d'un diagramme (RbIlOOl-Tb-Ta pour la discrimination
geochimique et l'etude petrogenetique des roches magma-
tiques acides: Bulletin de la Societe Geologique de France,
v. 8, p. 23-35.
THOMPSON, R. N., 1982, Magmatism of the British
Tertiary Volcanic Province: Scottish Journal of Geology, v.
18, p. 49-107.
THOMPSON, R. N., 1983, Thermal aspects of the origin of
Hebridean Tertiary acid magmas, II: Experimental melt-
ing behaviour of the granites at 1 kbar P
H20
: Minera-
logical Magazine, v. 47, p. 111-121.
__ , M. A. MORRISON, D. P. MATTEY, A. P. DICKIN, and
S. MOORBATH, 1980, An assessment of the Th-Hf-Ta
diagram as a discriminant for tectonomagmatic classifica-
tions and in the detection of crustal contamination of
magmas: Earth and Planetary Science Letters, v. 50, p. 1-
10.
VIERECK, L. G.,M. F. J. FLOWER, J. HERTOGEN, H.-U.
SCHMINCKE, and G. A. JENNER, 1989, The genesis and
significance ofN-MORB sub-types: Contributions to Min-
eralogy and Petrology, v. 102, p. 112-126.
VILLEMANT, B., and C. FLEHOC, 1989, U-Th fractionation
by fluids in K-rich magma genesis: The Vico volcano,
central Italy: Earth and Planetary Science Letters, v. 91, p.
312-326.
WAGER, L. R., and R. L. MITCHELL, 1951, The distribution
of trace elements during strong fractionation of a basic
magma - A further study of the Skaergaard intrusion,
East Greenland: Geochimica et Cosmochimica Acta, v. 1, p.
129-208.
WALL, V. J., J. D. CLEMENS, and D. B. CLARKE, 1987,
Models for granitoid evolution and source compositions:
Journal of Geology, v. 95, p. 731-749.
WATSON, E. B., and F. J. RYERSON, 1986, Partitioning of
zirconium between clinopyroxene and magmatic liquids of
intermediate composition: Geochimica et Cosmochimica
Acta, v. 50, p. 2523-2526.
WEAVER, B. L., and J. TARNEY, 1981, The Scourie dyke
suite: Petrogenesis and geochemical nature of the Protero-
zoic sub-continental mantle: Contributions to Mineralogy
and Petrology, v. 78, p. 175-188.
__ , D. A. WOOD, J. TARNEY, and J.-L. JORON, 1986,
Role of subducted sediment in the genesis of ocean-island
basalts: Geochemical evidence from South Atlantic Ocean
islands: Geology, v. 14, p. 275-278.
WHALEN, J. B., K. L. CURRIE, and B. W. CHAPPEL, 1987,
A-type granites: Geochemical characteristics, discrimina-
tion and petrogenesis: Contributions to Mineralogy and
Petrology, v. 95, p. 407-419.
WHELLER, G. E., R. VARNE, J. D. FODEN, and M. J.
ABBOTT, 1987, Geochemistry of Quaternary volcanism in
the Sunda-Banda arc, Indonesia, and three-component
genesis of island-arc basaltic magmas: Journal of Vol-
canology and Geothermal Research, v. 32, p. 137-160.
WHITE, W. M., and B. DUPRE, 1986, Sediment subduction
and magma genesis in the Lesser Antilles: Isotopic and
trace element constraints: Journal of Geophysical Re-
search, v. 918, p. 5927-5941.
WILKINSON, J. F. G., and R. W. LEMAITRE, 1987, Upper
mantle amphiboles and micas and Ti0
2
, K
2
0 and P
2
0
5
abundances and 100 Mg/(Mg+Fe
2
+l ratios of common ba-
salts and andesites: Implications for modal mantle meta-
somatism and undepleted mantle compositions: Journal of
Petrology, v. 28, p. 37-73.
WOOD, D. A., 1980, The application of a Th-Hf-Ta
diagram to problems of tectonomagmatic classification and
to establishing the nature of crustal contamination of
basaltic lavas of the British Tertiary Volcanic Province:
Earth and Planetary Science Letters, v. 50, p. 11-30.
__ , J.-L. JORON, M. TREUIL, M. NORRY, and J.
TARNEY, 1979a, Elemental and Sr isotope variations in
basic lavas from Iceland and the surrounding ocean floor:
Contributions to Mineralogy and Petrology, v. 70, p. 319-
339.
__ , J.-L. JORON, and M. TREUIL, 1979b, Are-appraisal
of the use of trace elements to classify and discriminate
between magma series erupted in different tectonic set-
tings: Earth and Planetary Science Letters, v. 45, p. 325-
336.
WYLLIE, P. J.,M. R. CARROLL, A. D. JOHNSTON, M. J.
RUTTER, T. SEKINE, and S. R. VAN DER LAAN, 1989,
Interactions among magmas and rocks in subduction zone
regions: Experimental studies from slab to mantle to
crust: European Journal of Mineralogy, v. 1, p. 165-179.
YEATS, R. S., S. R. HART, and others, 1976: Initial
Reports of the Deep Sea Drilling Project, Volume 34:
U.S. Government Printing Office, Washington, DC, USA,
814p.
YODER, H. S., Jr., and C. E. TILLEY, 1962, Origin of
basaltic magmas: An experimental study of natural and
synthetic rock systems: Journal of Petrology, v. 3, p. 342-
532.
ZECK, H. P., and J. R. MORTHORST, 1982, Continental
tholeiites in the Ti-Zr-Y discrimination diagram: Neues
Jahrbuch fur Mineralogie, Monatshefte, v. 5, p. 193-200.
Petrogenetic evaluation oftrace element discrimination diagrams 123
APPENDIX
CALCULATIONS AND MODEL PARAMETERS
The equation for calculating concentration of an ele-
ment, C!> in a partial melt is (Shaw, 1970):
C] = Co (lI(D + F(l-P)))
where Co is concentration in the source material, D is the
bulk distribution coefficient, F is fraction of partial melt-
ing, and P is the distribution coefficient for phases par-
ticipating in melting. Values of these parameters have to
be estimated or assumed.
Source Compositions: Co
Source compositions (Table AI) were assumed to be
those of major geochemical reservoirs, or mixtures of
components from these reservoirs. Mantle compositions
included depleted mantle (DM), enriched mantle (EM), and
primitive mantle (PM; calculations for PM not shown);
crustal compositions included upper crust (UC), lower
crust (LC), total crust (TC), oceanic crust (OC), the North
American Shale Composite (NASC), and Pacific Authi-
genic Weighted Mean Sediment (PS). Mantle compositions
were used for MORB and WPB models; mantle-sediment
mixtures were used in V AB models.
There are few published estimates of trace element
compositions of mantle sources, and not all studies present
estimates of all elements of interest here. Missing values
in source concentration data were estimated by assuming
constant, approximately chondri tic element ratios in
mantle sources. For example, Hf was estimated from Zr,
given that ZrlHf"" 40. Notes to Table Al indicate which
values were missing and how they were estimated.
Partial Melting: F
The fraction of partial melting is constrained in
several models by major element modeling (Frey and
others, 1978; Viereck and others, 1989; Conrad and others,
1989). Like trace element models, major element models
have numerous assumptions, but experimentally deter-
mined phase relations of peridotite-basalt systems re-
present powerful constraints. In this study, extent of
melting for MORB and V AB was restricted to 10-20%
(Viereck and others, 1989; Plank and Langmuir, 1988) - a
range that is sufficiently small that changes in residual
phase assemblages or D are unlikely. This is not to suggest
that smaller percentage partial melts are absent from
natural environments. For WPB, the melting interval was
considered to be 1-20% - a range that is indicated by trace
element variations in several WPB suites (e.g., Frey and
others, 1978).
Distribution Coefficients: D andP
Bulk distribution coefficients are calculated for each
element from values for individual minerals: D = EfmD
m
,
where fm is the fraction that a mineral phase represents in
the residue after melting and Dm is the distribution
coefficient for that mineral. Dm is Cm/C!> the ratio of
element concentration in a mineral to that in coexisting
melt. The limiting value of D is 0, when an element is
completely excluded from the solid assemblage. Elements
with D SO.l are considered incompatible; concentrations of
incompatible elements in partial melts increase as fraction
of melting decreases. Elements with D 2: 1 are referred to
as compatible; concentrations of compatible elements
decrease as extent of partial melting decreases. Changes of
element proportions during processes involving mineral-
melt equilibria depend on ratios of D. For example, if
Dz!Dy = 1, there will be no change in Zr/Y as a function
of melting; if, however, Dz!Dy = 0.1, ZrIY will change
rapidly during partial melting, being very high in low
percentage melts, and approaching ZrIY of the source at
high degrees of melting.
Minerals in a source usually do not participate in
melting reactions in proportion to their abundances in that
source: melting is usually non-modal. For non-modal
melting, P = EPmDm, where Pm is the fraction that a min-
eral contributes to the melt. For modal melting, Pm = fm
and P = D. When insufficient data are available to specify
Pm' modal melting is used as an approximation.
Values of Dm are shown in Tables A2 and A4. Be-
cause Dm are a function of temperature, pressure and bulk
composition of coexisting melt, Dm for basaltic (Table A2)
and granitic (Table A4) magmas can be significantly dif-
ferent, and Dm for silica-rich granitic melts can vary
rapidly with small changes in bulk melt composition. As
with source concentration data, some Dm have been esti-
mated or arbitrarily set (see notes to tables). The range of
Dm represents absolute limits for D: bulk D can neither
exceed the highest nor be less than the lowest Dm' Note
that unless its Dm is very large, an accessory mineral has
little impact on D: D is detemined by the predominant
minerals in the residue and, consequently, has a narrow
range for basaltic melts extracted from peridotitic mantle,
in which olivine and orthopyroxene predominate.
MORBModel
All calculations were based on equilibrium non-modal
melting. Modell used melting proportions and plagioclase
lherzolite residue composition of model R2 of Viereck and
others (1989), whereas model 2, based on plagioclase-spinel
lherzolite, was identical with model R9 of Viereck and
others (1989). Details are presented in Table A3. Because
plagioclase andlor spinel were exhausted in model sources
by about 20% partial melting, melting calculations were
not extended beyond 20%.
Depleted mantle compositions are shown in Table
AI. Modell and model 2 partial melts were calculated for
each source composition. Fractional crystallization calcu-
lations were completed on 20% partial melts from modell,
assuming Rayleigh fractionation (Gast, 1968):
C] = C
2o
F(D-ll
where C] is element concentration in fractionated melt, C
20
is element concentration in unfractionated, 20% partial
melt, F is the fraction of melt remaining, and D is the bulk
distribution coefficient. Rayleigh fractionation maximizes
concentration changes during crystallization. Fractiona-
tion calculations covered a range up to 30% crystallization
for four different, assumed crystal assemblages: (1) oli-
T
a
b
l
e

A
I
.

M
a
n
t
l
e

s
o
u
r
c
e

c
o
m
p
o
s
i
t
i
o
n
s
.

D
e
p
l
e
t
e
d

M
a
n
t
l
e

P
r
i
m
i
t
i
v
e

M
a
n
t
l
e

E
n
r
i
c
h
e
d

M
a
n
t
l
e

1

2

3

4

5

6

7

8

9

1
0

T
i

1
1
6
6

1
1
7
7

8
3
9

1
6
3
2

5
0
0
f

1
3
0
0

1
5
2
7

9
6
0

2
0
0
0
h

1
0
4
4

Z
r

9
.
0
2

1
1
.
4

9
.
3

1
3
.
5

4
.
3

1
1

1
1

8
.
3

1
5
.
5

1
1
.
3

Y

4
.
9
9

4
.
1

3
.
2

5
.
1

2

4
.
6

4
.
8
7

3
.
4

5
.
2
5

2
.
7

S
r

1
2
.
4

1
3
.
2

1
6
.
6

1
9
.
4

6
.
3
7

2
8

2
3

1
7
.
8

2
7
.
5

1
7
.
9

H
f

0
.
2
4

0
.
2
8
5
a

0
.
3
3
8
a

O
.
l
l
a

0
.
3
5

0
.
3
5

0
.
2
7

0
.
3
9

0
.
2
6

T
a

0
.
0
1
5

0
.
0
2
2

0
.
0
3
4
e

0
.
0
0
3

0
.
0
4

0
.
0
4
3

0
.
0
4

0
.
0
3
e

0
.
0
6
2

T
h

0
.
0
1
3

0
.
0
2

0
.
0
3
d

0
.
0
1

0
.
0
9
4

0
.
0
9
6

0
.
0
6
4

0
.
1
5

0
.
0
6

C
e

0
.
7
9

0
.
9
5

1
.
4

0
.
4

1
.
6
3

1
.
9

1
.
4
3
6

4
.
7

1
.
6
8

Y
b

0
.
5
3

0
.
3
9
b

0
.
5
8

e

0
.
2
6
9

0
.
4
2

0
.
4
3
g

0
.
3
7
2

0
.
5
2

0
.
2
7
i

N
b

0
.
2
6

0
.
3
1

0
.
2
5

0
.
4
8

0
.
0
5
e

0
.
5
6
e

0
.
6
2

0
.
5
6
e

0
.
4
2
i

0
.
7
2

C
o
l
u
m
n
s

(
s
o
u
r
c
e
s
)
:

1
,

V
i
e
r
e
c
k

a
n
d

o
t
h
e
r
s

(
1
9
8
9
)

f
o
r

N
1
-
M
O
R
B
;

2
,

W
o
o
d

a
n
d

o
t
h
e
r
s

(
1
9
7
9
a
)
;

3
,

L
e
R
o
e
x

(
1
9
8
7
)
;

4
,

F
i
t
t
o
n

a
n
d

D
u
n
l
o
p

(
1
9
8
5
)
;

5
,

H
o
l
e

a
n
d

o
t
h
e
r
s

(
1
9
8
4
)
;

6
,

J
a
g
o
u
t
z

a
n
d

o
t
h
e
r
s

(
1
9
7
9
)
;

7
,

W
o
o
d

a
n
d

o
t
h
e
r
s

(
1
9
7
9
a
)
;

8
,

T
a
y
l
o
r

a
n
d

M
c
L
e
n
n
a
n

(
1
9
8
5
)
;

9
,

F
r
e
y

a
n
d

o
t
h
e
r
s

(
1
9
7
8
)
;

1
0
,

W
o
o
d

a
n
d

o
t
h
e
r
s

(
1
9
7
9
a
)
.

N
o
t
e
s
:

a
)

f
r
o
m

Z
r
/
H
f
=
=
4
0
;

b
)

0
.
9

t
i
m
e
s

p
r
i
m
i
t
i
v
e

m
a
n
t
l
e
;

c
)

f
r
o
m

T
b
l
T
a

=
=

1
4
;

d
)

c
a
l
c
u
l
a
t
e
d

a
s
s
u
m
i
n
g

F

=
=
0
.
1
,

D

=
=
0
.
0
2
3
4
;

e
)

c
a
l
c
u
l
a
t
e
d

f
r
o
m

a
v
g

N
-
M
O
R
B
,

F

=
=
0
.
1
5
,

D

=
=
0
.
0
2
3
4
;

f
j

f
r
o
m

T
i
l
Z
r

o
f

N
-
M
O
R
B
,

u
n
c
e
r
t
a
i
n
t
y

2
5
0
;

g
)

1
.
4
8
1
5

t
i
m
e
s

c
h
o
n
d
r
i
t
e

o
f

G
a
n
a
p
a
t
h
y

a
n
d

A
n
d
e
r
s

(
1
9
7
4
)
;

h
)

u
n
c
e
r
t
a
i
n
t
y

o
f


5
0
0
;

i
)

f
r
o
m

Z
r
l
N
b
=
=
3
7
;
j
)

f
r
o
m

Y
/
Y
b
=
=

1
0

i
n

p
r
i
m
i
t
i
v
e

m
a
n
t
l
e
.

T
a
b
l
e

A
2
.

D
i
s
t
r
i
b
u
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
s

f
o
r

b
a
s
a
l
t
i
c

m
e
l
t
s
.

O
l
i
v
i
n
e

O
r
t
h
o
p
y
r
o
x
e
n
e

C
l
i
n
o
p
y
r
o
x
e
n
e

G
a
r
n
e
t

P
l
a
g
i
o
c
l
a
s
e

S
p
i
n
e
l

A
m
p
h
i
b
o
l
e

T
i

0
.
0
5

0
.
1

0
.
7

0
.
3

0
.
0
1

e

0
.
1

1
.
2
5

d

Z
r

0
.
0
1

0
.
0
3
-
0
.
1
8

0
.
1
2

0
.
3
-
0
.
7

0
.
0
1

e

0
.
1

0
.
3
9

e

Y

0
.
0
0
2
-
0
.
2

0
.
0
0
9
-
0
.
2

0
.
2
-
0
.
9

1
.
4
-
9

0
.
0
2
-
0
.
2

0
.
0
2
-
0
.
2

0
.
5
7
f

S
r

0
.
0
1
6

0
.
0
1
6
-
0
.
0
4

0
.
0
6
-
0
.
1
6
5

0
.
0
1
4

2

0
.
0
0
1

e

0
.
6
1

H
f

0
.
0
1

a

0
.
0
3
a

0
.
3
6

0
.
5

0
.
0
1

e

0
.
1

0
.
3
9

T
a

0
.
0
1

b

0
.
1
5

b

0
.
0
1
3
-
0
.
4

0
.
0
6
-
0
.
3

0
.
0
1

e

0
.
0
0
1

e

0
.
4
1

T
h

0
.
0
0
1

c

0
.
0
0
1

e

0
.
0
0
5

0
.
0
0
1

e

0
.
0
1

c

0
.
0
0
1

e

0
.
0
3
8

C
e

0
.
0
0
0
8

0
.
0
0
0
9

0
.
0
4

0
.
0
2
1

0
.
0
8

0
.
0
0
1

e

0
.
2
5

Y
b

0
.
0
0
4

0
.
0
2
8
6

0
.
2

4

0
.
0
5

0
.
0
0
1

e

0
.
5
7

N
b

0
.
0
1

0
.
1
5

0
.
0
0
5
-
0
.
0
2

0
.
0
2
-
0
.
1

0
.
0
1

e

0
.
0
0
1

e

0
.
4
1
g

S
o
u
r
c
e
s
:

D
u
n
n

(
1
9
8
7
)
,

D
u
n
n

a
n
d

M
c
C
a
l
l
u
m

(
1
9
8
2
)
,

F
r
e
y

a
n
d

o
t
h
e
r
s

(
1
9
7
8
)
,

G
r
e
e
n

a
n
d

P
e
a
r
s
o
n

(
1
9
8
7
)
,

G
r
e
e
n

a
n
d

o
t
h
e
r
s

(
1
9
8
9
)
,

H
a
r
t

a
n
d

B
r
o
o
k
s

(
1
9
7
4
)
,

I
r
v
i
n
g

(
1
9
7
8
)
,

M
c
C
a
l
l
u
m

a
n
d

C
h
a
r
e
t
t
e

(
1
9
7
8
)
,

P
e
a
r
c
e

a
n
d

N
o
r
r
y

(
1
9
7
9
)
,

R
y
e
r
s
o
n

a
n
d

W
a
t
s
o
n

(
1
9
8
7
)
.

N
o
t
e
s
:

a
)

v
a
l
u
e

f
o
r

Z
r
;

b
)

v
a
l
u
e

f
o
r

N
b
;

c
)

a
r
b
i
t
r
a
r
y

v
a
l
u
e
;

d
)

f
o
r

k
a
e
r
s
u
t
i
t
e
;

c
a
l
c
u
l
a
t
e
d

f
r
o
m

d
a
t
a

i
n

D
a
u
t
r
i
a

a
n
d

o
t
h
e
r
s

(
1
9
8
7
)
;

e
)

v
a
l
u
e

f
o
r

H
f
;

f
j

v
a
l
u
e

f
o
r

V
b
;

g
)

v
a
l
u
e

f
o
r

T
a
.

.
.
.
.
.


o
j>
.

>
-
r
j

0
'

t
:
:
I


p
.
.
.


[
J
J

O
l
i
v
i
n
e

O
r
t
h
o
p
y
r
o
x
e
n
e

C
l
i
n
o
p
y
r
o
x
e
n
e

G
a
r
n
e
t

P
l
a
g
i
o
c
l
a
s
e

S
p
i
n
e
l

A
m
p
h
i
b
o
l
e

Q
u
a
r
t
z

T
i

0
.
0
2

Z
r

0
.
0
2
a

Y

0
.
0
0
1

b

S
r

0
.
0
1

c

H
f

0
.
0
2

T
a

0
.
0
1

T
h

0
.
0
0
8

C
e

0
.
0
1

Y
b

0
.
0
1
5

N
b

0
.
0
0
1

d

R
b

0
.
0
0
2
-
0
.
0
1
6

T
a
b
l
e

A
3
.

M
e
l
t
i
n
g

m
o
d
e
l

p
a
r
a
m
e
t
e
r
s

f
o
r

b
a
s
a
l
t
.

M
O
R
B

a
n
d

V
A
B

M
e
l
t
i
n
g

M
o
d
e
l
s

W
P
B

M
e
l
t
i
n
g

M
o
d
e
l
s

M
o
d
e
l
l

M
o
d
e
l

2

G
T
L
H
E
R
Z
I

S
P
L
H
E
R
Z
I

S
P
L
H
E
R
Z
2

R
e
s
l

M
e
l
t

R
e
s
l

M
e
l
t

R
e
s
2

R
e
s
2

R
e
s
3

6
5
.
1
9

1
0

5
7
.
7
8

1
2

5
8

6
5

7
6

2
3
.
4
7

1
5

2
8
.
0
7

1
3

2
4

2
4

1
7

5
.
7
7

2
6

9
.
5
7

2
1

1
1

1
0

6

7

5
.
5
7

4
9

2
.
2
1

3
9

2
.
3
6

1
5

1

1

R
e
f
e
r
e
n
c
e
s
:

1
)

V
i
e
r
e
c
k

a
n
d

o
t
h
e
r
s
,

1
9
8
9
;

2
)

F
r
e
y

a
n
d

o
t
h
e
r
s
,

1
9
7
8
;

3
)

R
o
d
e
n

a
n
d

o
t
h
e
r
s
,

1
9
8
3
.

T
a
b
l
e

A
4
.

D
i
s
t
r
i
b
u
t
i
o
n

c
o
e
f
f
i
c
i
e
n
t
s

f
o
r

g
r
a
n
i
t
i
c

m
e
l
t
s
.

P
l
a
g
i
o
-
O
r
t
h
o
-
C
I
i
n
o
-
c
l
a
s
e

K
-
F
e
l
d
s
p
a
r

p
y
r
o
x
e
n
e

p
y
r
o
x
e
n
e

G
a
r
n
e
t

A
m
p
h
i
b
o
l
e

B
i
o
t
i
t
e

0
.
1

0
.
4

1

I
e

7
f

2
0
f

0
.
1

0
.
1

a

0
.
3

0
.
3
a

0
.
5
a

1

0
.
1

0
.
0
0
5
-
0
.
0
4

b

0
.
5

1
.
5
b

3
5

3

0
.
0
5
-
1

4
.
0
-
1
0
.
0

0
.
0
2

0
.
5

0
.
0
2

0
.
5

0
.
3

0
.
0
5

0
.
0
0
5
-
0
.
1
3

0
.
1

0
.
3

0
.
5

0
.
5

0
.
3
-
1

0
.
0
2

0
.
0
0
1
-
0
.
0
8

0
.
2

0
.
1
5

0
.
3

0
.
5

1

0
.
0
1

0
.
1

0
.
1

0
.
1

e

0
.
0
7

0
.
2
-
1
.
5

0
.
1
5

0
.
2

0
.
4
-
1
0

0
.
2

0
.
4
-
1
.
5

0
.
3
-
2

0
.
0
3

1

1
.
5
-
5

3
8

1
.
5
-
5

0
.
4
-
1

0
.
0
3

0
.
0
0
1
-
0
.
0
4

d

0
.
5

0
.
1
5
d

0
.
3

2

2

0
.
0
1
6
-
0
.
1
9

0
.
3
1
-
1
.
8

0
.
0
0
3

0
.
0
3

0
.
0
0
8
5

0
.
0
0
8
-
0
.
0
1
4

2
.
0
-
5
.
3

A
M
L
H
E
R
Z

R
e
s
3

7
0

1
5

7

1

7

Z
i
r
c
o
n

6
0
-
2
2
0

2
6
0
0
-
3
7
5
0

4
0
-
5
5

4
0

0
.
0
0
1

c

S
o
u
r
c
e
s
:

A
r
t
h

(
1
9
7
6
)
,

F
r
a
n
c
a
l
a
n
c
i

(
1
9
8
9
)
,

G
r
e
e
n

a
n
d

P
e
a
r
s
o
n

(
1
9
8
7
)
,

H
a
n
s
o
n

(
1
9
7
8
)
,

I
r
v
i
n
g

(
1
9
7
8
)
,

I
r
v
i
n
g

a
n
d

P
r
i
c
e

(
1
9
8
1
)
,

L
u
h
r

a
n
d

C
a
r
m
i
c
h
a
e
l

(
1
9
8
0
)
,

L
u
h
r

a
n
d

o
t
h
e
r
s

(
1
9
8
4
)
,

M
a
h
o
o
d

a
n
d

H
i
l
d
r
e
t
h

(
1
9
8
3
)
,

N
a
g
a
s
a
w
a

a
n
d

S
c
h
n
e
t
z
l
e
r

(
1
9
7
1
)
,

N
a
s
h

a
n
d

C
r
e
c
r
a
f
t

(
1
9
8
5
)
,

P
e
a
r
c
e

a
n
d

N
o
r
r
y

(
1
9
7
9
)
,

V
i
l
l
e
m
a
n
t

a
n
d

F
l
e
h
o
c

(
1
9
8
9
)
,

W
a
t
s
o
n

a
n
d

R
y
e
r
s
o
n

(
1
9
8
6
)
.

N
o
t
e
s
:

a
)

v
a
l
u
e

f
o
r

H
f
;

b
)

v
a
l
u
e

f
o
r

Y
b
;

c
)

a
r
b
i
t
r
a
r
y

v
a
l
u
e
;

d
)

v
a
l
u
e

f
o
r

T
a
;

e
)

v
a
l
u
e
s

f
o
r

c
p
x
;

f
)

u
n
r
e
l
i
a
b
l
e
,

T
i

i
s

a

m
a
j
o
r

e
l
e
m
e
n
t
;

v
a
l
u
e
s

e
s
t
i
m
a
t
e
d

f
r
o
m

d
a
t
a

i
n

L
u
h
r

a
n
d

C
a
r
m
i
c
h
a
e
l

(
1
9
8
0
)
,

L
u
h
r

a
n
d

o
t
h
e
r
s

(
1
9
8
4
)
,

a
n
d

M
a
h
o
o
d

a
n
d

H
i
l
d
r
e
t
h

(
1
9
8
3
)
.

'
l
:
i

C
1
>

.
,
.
.
.

.
.
.
.

o

a
q

C
1
>

:
J

C
1
>

:
:
:
;
:

(
':
)

C
1
>

<

I
I
I


:
J

o

.
.
.
.
,

.
,
.
.
.

.
.
.
.

I
I
I

(
':
)

C
1
>

C
1
>


:
3

C
1
>


e
:

[
J
)

(
':
)

.
.
.
.

.

[

o


:
J

0
-
S
;
'

a
q

.
.
.
.

I
I
I

:
3

[
J
)

I
-
'

N

0
1

126
F. O. Dudas
vine, only; (2) 50% olivine and 50% clinopyroxene; (3) 50%
olivine and 50% plagioclase; and (4) 50% olivine, 25%
clinopyroxene, and 25% plagioclase. Similarity of D for
olivine and orthopyroxene suggests that olivine can sub-
stitute for orthopyroxene in these simplistic models.
The base case for both models was calculated using
low values of D from Table A2. High D cases were cal-
culated using either the high value indicated in Table A2
or an arbitrary value that ranged from 2 to 15 times the
low value of Table A2. Note, for example, that Dy for
orthopyroxene has a maximum value of 0.2; this derives
from rounding a measured, experimental value of 0.18
(Green and others, 1989) and a suggested value of 0.2
(Pearce and Norry, 1979). A corresponding high value for
olivine has not been determined. For high Dy calculations,
Dy for olivine was assumed equal to that for orthopyro-
xene. Models illustrating effects of variable D were cal-
culated both for partial melting (see Figure 5) and for
fractional crystallization (see Figure 4); bulk D are re-
corded in figure captions.
VABModel
The first model assumes source compositions formed
by bulk mixing between DM and NASC (Gromet and
others, 1984) and DM and PS (the Pacific Authigenic
Weighted Mean Sediment, or PA WMS of Hole and others,
1984). The second model assumes that hybridization of
DM with partial melts of subducted crust (including sedi-
ments, i.e., slab-derived melts; equivalent to the three-
component source of Ellam and Hawkesworth, 1988) yields
the range of relevant source compositions. In all cases, the
amount of admixed material was restricted to 5%. Isotopic
data indicate that the extent of sediment involvement in
arc magma genesis is limited to a few percent (Meijer,
1976; Tera and others, 1986; White and Dupre, 1986). This
is the basis for restricting source compositions to mixtures
containing s; 5% PS, NASC, or hybridizing melt.
These models represent gross simplifications. Bulk
mixing of DM and sediment, for example, is not a realistic
model for modification of the arc-source mantle. The pre-
ponderance of experimental evidence indicates that silicic
melts are primarily responsible for mass transport under
mantle conditions, and thus the second model is more
realistic. Mass transport and mixing involving volatiles
was not modeled, apart from models for hybridization with
melts, because the assumption that hydrous or carbonic
fluids can effectively or selectively transport components
from subducted crust to overlying mantle appears unten-
able (Olafsson and Eggler, 1983; Schneider and Eggler,
1986; Eggler, 1987; Meen and others, 1989).
Uncertainties in the DM hybridization model include
bulk composition of the melted oceanic crust, residue com-
position, extent of melting, and relative melting propor-
tions. Experiments suggest that partial melts of oceanic
crust would be dacitic to rhyolitic (Wyllie and others,
1989;, consequently, D appropriate to felsic melts were
used (Table A4). The melted, subducted crust was
assumed to be either the model 2 source composition of
Hole and others (1984) or 90% OC (Taylor and McLennan,
1985) plus 10% PS (Hole and others, 1984; this mixture is
shown on figures as +S or OS). Three different residue
mineralogies were used, all corresponding to eclogitic
sources: (1) 50% clinopyroxene and 50% garnet; (2) 60%
clinopyroxene, 30% garnet, and 10% amphibole; and
(1) 60% clinopyroxene, 30% garnet, and 10% phlogopite.
For all subducted crust melting models, the extent of
melting was assumed to be 20%, and modal melting was
assumed, in the absence of data on melting proportions in
composite sources. As for other models, a range of D was
used in the calculations. Mixtures of 5% of these partial
melts with DM (Table AI) were used as sources for V AB
models. Despite the range of variables used, the V AB
sources cover a relatively small area on discrimination
diagrams.
No models included crustal contamination of arc mag-
mas or mixing between magmas from different sources.
MORB melting models (i.e., plagioclase and plagio-
clase-spinel lherzolite) were used with V AB sources be-
cause there is permissive evidence that similar conditions
might prevail, at least in oceanic arcs (Ewart and Hawkes-
worth, 1987; Plank and Langmuir, 1988). Though H
2
0-
rich sources might contain amphibole lherzolite, and
hybridization of DM with silicic partial melts could
produce garnet lherzolite or garnet clinopyroxenite sources
(Wyllie and others, 1989), there is no clear indication that
amphibole or garnet are residual in arc basalt sources. In
particular, the absence of Yb-depleted REE patterns
among arc lavas indicates that garnet cannot playa pro-
minent role in their petrogenesis.
WPBModel
The WPB model used EM compositions (Table AI)
and garnet lherzolite (Frey and others, 1978), spinel
lherzolite (Frey and others, 1978; Roden and others, 1983),
and amphibole lherzolite (Roden and others, 1983) residue
mineralogies (Table A3). The EM source of Frey and
others (1978) was formulated for continental basaltic rocks
that ranged from tholeiitic to alkalic compositions; it
represents a modified pyrolite in which Ti content is poor-
ly constrained. Consequently, model calculations have
included a range of Ti contents in this EM source. The
other EM source (Wood and others, 1979a) was based on
data from Iceland and should be representative of OIB and
E-MORB sources.
High D and low D models were calculated for both EM
sources using all four residue mineralogies. For the most
part, amphibole lherzolite melts are very similar to spinel
lherzolite melts but usually have lower overall trace
element concentrations because amphibole has higher D
for most incompatible elements. A number of calculations
were done with arbitrarily assumed D, in order to extend
model coverage of WPB fields on the discrimination dia-
grams. These calculations suggest the range of EM source
compositions has to be much greater than that represented
in Table AI, because assumed D for tested sources were, in
some cases, unreasonable. The large range of source com-
Petrogenetic evaluation oftrace element discrimination diagrams 127
positions was also indicated by attempts to calculate
sources from inversion of extreme compositions on dis-
criminant diagrams.
Granite Models
Three sources were used in granite melt models
(Table A5): NASC, OS (the oceanic crust plus 10% sedi-
ment composition used in hybrid V AB models), and an
arbitrary LC-like composition. NASC models used the
graywacke melting model of Conrad and others (1989),
assuming residue proportions appropriate for 20% melting
(Table A6), and a range of D (Table A4). Both ilmenite-
present and ilmenite-absent models were computed; on
most diagrams; the impact of ilmenite was minor. The LC
composition was used with amphibolite and granulite
residues (Table A6); for the granulite residue, models
were also calculated with 1 % each of magnetite and zircon,
and 1% each of magnetite and ilmenite. The OS composi-
tion was used with eclogite residues, as described above for
mixing components in the V AB source.
Table A5. Possible crustal source compositions.
Upper Total Lower Oceanic Assumed
Crust Crust Crust Crust NASCa PSa O+Sa LC
Ti 3000 5400 6000 9000 4736 340
d
8134
Zr 190 100 70 80 200 21.57 74.16
Y 22x 22
X
22
X
32 22 b
46.31 33.43 22
Sr 350 260 230 130 142 1144 231.4
Hf 5.8 3 2.1 2.5 6.3 0.54e 2.3 3
Ta 2.2 1 0.6 0.3 1.12 0.05 0.28 1
Th 10.7 3.5 1.06 0.22 12.3 0.23 0.22
Ce 64 33 23 11.5 67 9.6 11.3
Yb 2.2 2.2 2.2 5.1 3.1 5.55 5.15
Nb 25 11 4
X
2.2 16
C
1.25 2.11 4
Rb 110 x 42
X
8
X
2.2 125 3.6 2.34 15
Sources: Taylor and McLennan (1985) for DC, TC, LC, and OC; Gromet and others (1984) for NASC; Hole and
others (1984) for PS. Notes: a) mixing components in VAB; b) assumed = DC; c) from Nb/Ta=14; d) calculated
by methods of Hole and others (1984) from data in Yeats and others (1976); e) from ZrlHf=40; x) values from
Taylor and McLennon (1981).
Table A6. Partial melting parameters for granite models.
Amphi- Granulite Graywacke (gw) Eclogite (e)
bolite
Quartz 10
Plagioclase 60
K-feldspar 10
Orthopyroxene
Clinopyroxene
Garnet
Biotite 10
Amphibole 10
Magnetite
Ilmenite
Zircon
g gz gi Res Melt Res Melt
10 10 10 25 30 25 30
70 70 70 44 50 43.5 50
10 9 9 5 1 5 1
10 9 9 12.5 4 12.5 3.5
13.5 15 13.5 15
1 1
1 0.5 0.5
1
Graywacke melting models based on Conrad and others (1989).
1
50
50
2
60
30
10
3
60
30
10
ABSTRACT
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Metallogeny at Precambrian and Mesozoic
continental margins
JOHN M. GUILBERT
Department of Geosciences, University of Arizona, Tucson, Arizona 85721, USA
(received August 29,1988; revision accepted June 19, 1989)
Modelling of Precambrian continental margin tectonic settings involves plate-tectonic-metallogenic
analyses, back-projection of modern lithotectonic settings and ore deposit classifications, and consideration of
changes in ore deposit genesis and characteristics through geologic time. In this paper, two approaches are
developed - namely, the judicious projection back to the Precambrian of Mesozoic-Cenozoic metallogenic
characteristics and interpretations, and the direct analysis of Archean-Proterozoic ore deposit settings and
their implications with respect to plate reconstructions and dynamics.
Metallogeny has taken four directions in the last 20 years. They are to produce: (1) mapped inventories of
occurrences and deposits and analysis of terranes therefrom; (2) mechanistic studies of how subduction
influences epicrustal metal distribution; (3) lithotectonic classification of ores and rocks leading to improved
understanding of both; and (4) theoretical analysis of economic- and trace-element behaviors in petrology and
tectonics. All have yielded constraints on continental margin systematics, especially for consuming margins.
Consuming margins are shown to exhibit characteristic polarized patterns of Fe-Cu-Mo-Zn-Pb-Ag-Au
occurrence and distributions of whole-rock major elements that can be related to subduction mechanisms.
Complications in specific margins arise from accretion and from variations in subduction geometry. Accreted
terranes involve pre-, syn-, and post-accretion components, and variations in (1) the amount and composition
of subducted material, (2) the thickness and composition of ocean crust subducted, (3) the amount of water,
(4) the array of isotherms, and (5) the angle and rate of subduction. All these variables must be factored into
the analysis of each accretionary event of a complex margin.
Analysis of ancient terranes must also recognize departures from actualistic metallogeny. We interpret
from terrane and ore-deposit fabrics and habitats that Archean plates and margins were more mobile than
they were later, and that Proterozoic craton stabilization fostered modified subduction styles and, with
atmospheric oxidation, new styles of metallogeny. Modern margin systematics (Windley, 1984) coupled with
firm compilations of the evolution of ore genesis through Earth history indicate that subduction was rapid,
short-range, and vigorous in the Archean, and that subduction similar but not identical to modern analogues
(with scales to thousands of kilometers) was established by 1.8 Ga, perhaps locally even earlier.
(abstract continued overleaf)
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 129-141. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
129
130 J. M. Guilbert
Abstract (continued)
Direct study of Archean crust in Fennoscandia, North America, Africa, and Australia shows global
patterns, with interpretable deviations, in volcano-sedimentary greenschist-facies constructions and their
intrinsic ore deposits. In the Proterozoic, a Yavapaian-Transhudsonian-Svecokarelian belt of volcanogenic
massive sulfide deposits suggests proto-subduction at Andean scale for the first time, as does the first arrival
of differentiated calcalkaline rocks and Cordilleran-style porphyry occurrences at 1.8 Ga. Anorthosite
emplacement-age constraints and the restriction to this time period of Broken-Hill-type Cu-Pb-Zn ores
suggest relatively thin, locally rifted cratons with steep geothermal gradients and high mantle-volatile fluxes.
Finally, detailed mapping, metallogenic analysis, and geochemistry in the Yavapai belt of Arizona indicate
that both intraoceanic island-arc and continental margin-arc subduction mechanisms operated there that
were similar but not identical to modern analogues (Anderson, 1986a,b).
Certainly due caution need be taken in such retrospection, but indisputably significant plate tectonic
information is recorded in Precambrian metallogeny. In combination with backward extrapolation of
Mesozoic-Cenozoic relationships, it shows that basically uniformitarian processes, with developmental
mileposts along the way, have marked continental margin dynamics from the Archean onward.
INTRODUCTION
This topic leads directly to the crossroads of
three major subjects: (1) plate tectonics-basement
tectonics and metallogeny; (2) lithotectonic settings
and classifications of mineral deposits, of their
crustal signatures, and of rocks that enclose them;
and (3) evolution of the fabric of ore deposits and
their genesis through time. This latter subject has
occupied many of us for the last two decades.
Among the few who have spoken and published
powerfully on the subject is Charles Meyer, first in
the 75th Anniversary Volume of Economic Geology
(Meyer, 1981) and then in the Annual Reviews of
Earth and Planetary Letters (Meyer, 1988), his last
paper completed before sudden death felled him on
November 15, 1987. It is especially appropriate that
his thoughts on the broad theme of plate tectonics
and metallogeny through time be focused here in
Butte, Montana - one of the many scenes of his
greatness. Having met him in Butte, having
worked with him in Butte, and having learned from
him here and elsewhere, I am pleased to dedicate
this paper to his memory.
We use two approaches to the comparison of
metallogeny at Precambrian and Mesozoic plate
boundaries. One is to project what we know of
metallogeny of Mesozoic-Cenozoic continental mar-
gins back to the Precambrian to draw inferences
about past mechanisms from their modern counter-
parts, and the second is to report metallogenic
studies of specific Precambrian terranes themselves.
No systematic distinction is made between the
Mesozoic and the Present.
New concepts and definitions in the broad sub-
ject area of metallogeny have literally sprung into
place during the last 20 years of perceptions brought
on by plate tectonics. To examine metallogeny we
need first to define the term. The roots of the word
are obvious, implying enquiry about the genesis and
distribution of metals in the Earth's crust, and by
extension metallic mineral deposits, although the
Greek root word "metallon" can include nonmetal-
lics. The term formerly related to the perception of
unifying characteristics of ore deposits or ore deposit
families, and thereby the definition of swaths of
terrane or periods of time or both, that contain
consanguineous, fundamentally similar deposits.
The concept has grown during the last two decades
to include greater consideration of trace elements,
geophysics, and other signatures than was formerly
the case. Nonmetallics and energy minerals are
now also included, and the focus is on the nature of,
and the reasons for, element distributions in re-
sponse to lithotectonics. Metallogeny, then, is the
study of the genesis of mineral deposits, with
emphasis on their relationships with geochemical,
petrologic, and tectonic features of the Earth's crust.
A metallogenic province is an area characterized by
abundance of an element or combination of elements
(Figure 1), by a particular assemblage of mineral
deposits, or by one or more characteristic types of
mineralization. A metallogenic belt is an elongate
metallogenic province. A metallogenic epoch is an
interval of geologic time that was favorable for
mineralization of an element, a set of elements, or a
characteristic type or types of deposit in a metal-
logenic province. Strong emphasis on lithotectonics
Metallogeny at Precambrian and Mesozoic continental margins 131
has come into metallogenic studies with the rela-
tively recent perception that mineral deposits are
much more an integral part of the rocks that contain
them than was thought to be the case even 25 years
ago. The modern economic geologist is almost
inescapeably applying metallogenic principles in
considering the distribution of epithermal hot-
spring or Carlin-type deposits, and the subject is of
increasing interest because what we have learned in
the theater of economic geology of metallogenic
provinces and epochs can certainly help decipher the
geologic record. It is, I think, more widely appre-
ciated than ever before that we can and should turn
"the logic" around: we need not use structural geo-
logy and tectonics in a one-way sense to find ore
deposits, but rather we now know so much about the
P-T-X details of ore-forming processes that we can
use ore-deposit characteristics to unravel many
plate-tectonic analytic details.
LITHOTECTONIC CLASSIFICATION
If it be assumed - (l)that plate tectonic mech-
anisms are really shaping the Earth's crust, (2) that
those mechanisms control the geochemistry and
petrology of the lithosphere, (3) that mineral de-
posits generally are integral parts of the rocks that
host them, (4) that there are metallogenic belts in
which families of ore deposits are predictably asso-
ciated with various lithotypes, and (5) that there are
trace element-trace mineral accessories to those
mineral concentrations - then the identification of
lithotectonic niches for ore deposits is feasible. Also,
the consideration of trace element variations at the
same scale as metallogenic belts - namely, the
scale of continental margins - is legitimate. Such
studies are not very old; one of the first quantitative
ones was C. Wayne Burnham's dissertation (Bur-
ham, 1959). He studied concentrations of trace
elements in sulfides in a broad distribution of occur-
rences in the western United States and found that
there were belts or zones of copper in sphalerite and
of zinc, tin, and silver in pyrite and chalcopyrite that
transcend individual districts, or in another sense,
link them together. These cryptic trace-element
anomalies and their ratios form belts roughly
parallel to the continental margin.
In the seventies and early eighties, metallogeny
went in four directions. One was to continue to
inventory mineral occurrences and deposits on
maps, to draw outlines around similar symbols to
42'
o
I
300 Mil ES
i KllOM TAts
EXPlANATlON
G VtJns. p pes
.. ..
Figure 1. An example of a metallogenic map. Con-
centrations of copper in the crust of the western United
States are seen to form swaths generally parallel to the
continental margin. Deposits of several genetic types are
distinguished but plotted together; no relative age infor-
mation is coded. The dashed, solid, and dot-dash lines are,
respectively, the west limit of the Paleozoic miogeocline,
the west limit of Srj>O.706, and the platform boundary.
The wide-ruled patterns denote geologic but subeconomic
concentrations. Figure 3 (Sillitoe, 1983) is an example of
the more common form of a metallogenic map. From
Albers (1983).
delineate metallogenic zones or belts, and to
attempt to analyze the nature of the terranes thus
defined (Nobel, 1970; Albers, 1981; Tooker, 1981).
Another was more mechanistic in that several
scholars tried to interpret the consequences of sub-
duction and continent collision in its many forms,
with sometimes elaborate and fanciful assignment
of epicrustal ore deposit signatures and geochemis-
tries to particular styles, rates, or geometries of
subduction. Such attempts were both descriptive
132
J . M. Guilbert
and predictive, and they have become useful to
explorationists and tectonicists alike (Mitchell and
Garson, 1972; Sawkins, 1972; Sillitoe, 1972; Hut-
chinson, 1980). The third approach was to dissect
metallogenic maps and their belts and epochs,
incorporate aspects of tectonic mechanisms and
lithotectonics, and produce plate-tectonic/litho-
tectonic classifications of ore deposits by assigning a
habitat to each deposit type. The technique was
obviously more successful with syngenetic than
with epigenetic occurrences, but it also enabled
some predictive determinations of ore deposit-host
rock relationships (Mitchell and Bell, 1973; Guil-
bert, 1981; Mitchell and Garson, 1981; Sawkins,
1984; Guilbert and Park, 1986). Lastly, several
theoretical analyses of economic-element and trace-
element partitioning in mantle, crust, and mineral
deposit systems have been instructive, most re-
cently by Brimhall (1987).
Consuming Margin Systematics
The cumulative, integrated result of these
efforts has provided a clearer but by no means com-
plete view of metallogeny at modern plate margins.
Vigorous debate continues not only by some indivi-
duals on the efficacy and degree of involvement of
plate tectonic mechanisms on metallogeny but also
on the sources of metals, and their containing
lithotypes, with respect to the subduction system.
The importance of lithospheric melting or anatexis
and near-surface derivation of components, as by
lateral secretions and source-bed hypotheses, are
being evaluated. A strong voice for primarily mag-
matic control is S. B. Keith, who has marshalled
thousands of whole-rock and trace-element analyses
of rocks and ores, mainly from consuming-margin
environments, and constructed a descriptive and
predictive model of metallogenic zones with respect
to those margins that is based on the major-element
and petrographic characteristics of ore-deposit host
rocks (Keith, 1978; Westra and Keith, 1981; Keith,
1986). Keith's conclusions must remain somewhat
disputed until his proprietary use of his data is
supplanted by its full publication, but his talks,
limited publications, and private use of those data
has won him a school of international adherents.
His principal conclusions applicable here are sum-
marized in Figure 2, which shows that the habitats
of several deposit types, which are metal-compo-
sition variations on the porphyry theme, can be
eu-Mo-
CaIC_" 1-"""1
..",. ()ornax..",.
I
A

EI"CSIa.IC back-arC
300
AI
Figure 2. A model describing interaction between sub-
duction, igneous rock chemistry, and ore deposits assem-
bled for a section from present-day California through
Nevada and Utah into Colorado. Note the element ratio
and absolute element variation shown in the upper por-
tion. From Westra and Keith (1981).
directly related to depth of melting along a sub-
ducting slab. Keith subscribe to the K
2
0-d argu-
ments and major alkali-alkali earth petrologic
classification that are expressed and implied in
Figure 2. Isotopic, geochemical, and geophysical
arguments in Westra and Keith (1981) support the
model, and Keith has gone on to state that melting
of the lithosphere above the subducting slab appears
to contribute little to the trace-element profiles of
either the calcalkaline-stem rocks or the ore depo-
sits associated, and presumably cognate, with them.
Keith and his coworkers, then, see the distributions
of elements recognized for decades as Cordilleran-
style metallogenic belts and epochs (see Figure 1) to
be visceral to subduction geometries and systema-
tics; they see the source of the metals of cogenetic
orebodies to be the subducted slab and closely
related material itself. More is said of the alter-
native positions in succeeding paragraphs, but it
can be noted here that Keith insists that he and his
group can sensitively and predictively forcast the
trace- and economic-element profiles of ore systems
by knowing associated magma-chemical details, and
vice versa.
Metallogeny at Precambrian and Mesozoic continental margins 133
.'.
I
00'
,
I
I
I
,
,
,
\
Y,
+,
.... ,
V'\

,
I
;: 10"
, ,
,
COPPJ> - IGOlO-MOI.,aC(NUM I
coPPER - LEAO-Zlm; -SILvER
TIN. (JUNGSTEN
A.nt lplono - Pur-a
RtC.,,1 VOIconot's
BRAZ IL
I
ijC'w
km
Figure 3. Metallogenic belts parallel to the coastline,
continental margin, and the Peru-Chile trench of South
America. They are thought to be subduction related and
the "normal" metallogenic array.
It would be expected, if subduction is a
generally regular process and Keith and his group
are correct, that there might exist a normal and
reproducible zonal pattern of rocks and ore deposit
types that are symmetrical to, and polarized from,
the "normal" consuming margin. Such a pattern
has been recognized for many years by workers too
numerous to mention individually. A simplified
summary (Figure 3) shows relatively clear-cut
bands essentially parallel to the continental margin
and showing successive belts that are enriched in
iron, copper, copper-molybdenum, copper-Iead-zinc-
silver, and then tin-wolfram-silver inboard from the
trench. A copper-gold belt is now recognized in
eastern Chile and western Argentina east of the
zones in Figure 3. The degree to which the zones
might be broadened or narrowed presumably de-
pends upon subduction angle (Coney and Reynolds,
1977), and the degree to which they might overlap
could be ascribed to constancy of that angle through
time. Mitchell and Garson (1972) made early
reference to the presence of high fluorine-tin asso-
ciations on the inboard side of the Anadaman arc -
a relation carried into Westra and Keith (1981) (see
Figure 2) and ascribed by several scientists, most
thoroughly Sillitoe (1976), to subduction manifesta-
tions rather than to deeper partial melting of litho-
sphere or actual continental basement inboard from
the margin. New isotopic data cloud this issue, with
neodymium-samarium and initial strontium models
(Farmer and DePaolo, 1983; Anthony and Titley,
1988) requiring melting and assimilation of sialic
crustal material that should skew simple subduc-
tion-driven metallogenic distributions by variable
assimilation of lithophile elements into calcalkaline
melts and resultant ores. It appears valid, nonethe-
less, cautiously to apply the simple geometry and
relative disposition of elements of Figure 2 to the
Mesozoic-Cenozoic North American Cordilleran
margin in general, and to similar subducting arcs
elsewhere.
Not only can economic element distributions be
construed to define belts, but also deposit types are
characteristically and predictively arrayed. These
arrays are the basis of lithotectonic classifications
as indicated earlier (Guilbert, 1981; Sawkins, 1984)
and have been described with relation to Cordiller-
an margins and island arcs by Sawkins (1972) and
Sillitoe (1981). The latter separated six subsets of
ore-deposit types embracing variations on the
porphyry, skarn, limestone repacement, massive
sulfide, and volcanic-hosted themes as functions of
associated granitoid types (I, S, A, R, oxidation
state), tectonic-subduction relations, host-rock
types, and so forth. The deposit types occupy belts
(like those in Figures 1 and 3) even in Mexico's
post-accretion terrane. Sillitoe presents arguments
that the inboard occurrence of the lithophile ele-
ments Sn-W-Mo-Be is subduction-driven and pro-
duces what he calls Bolivian-type back-arc settings,
and is not the result of continental crust hy-
bridization. These lithotectonic classifications can
134 J. M. Guilbert
reliably be projected back in Earth history and can
thus help identify tectonic fabrics and settings -
overlays of modern ore-deposit characteristics on
Paleozoic and Mesozoic plate boundary lithotypes fit
well (Guilbert, 1972).
Accreted Terranes
If we assume that simple subductive consuming
margins display the metallogenic array of Figures
2 and 3, we can first enquire of the perturbations
that prevent all consuming margins from conform-
ing to the model, and then begin to consider what to
conclude about geologically older analogues. Three
readily identifiable complications are accreted or
"suspect" terrane development, short time-scale
variations of subduction mechanisms, and changes
of ore- and rock-forming processes through geologic
time.
Concepts involving accreted terranes have
developed more recently than many depictions of
continental margin metallogenic belts, and it is not
surprising that areas such as Alaska-British Colum-
bia were not early candidates for selection as "type
areas" for zoning analyses. If accreted terranes do
indeed involve microplate masses of exotic continen-
tal crust of various provenances, slices of oceanic
crust, island arcs, and rafts of seamounts, then
metallogenic belts or epochs should be similarly
chaotic. Metallogenic analysis of accreted margins
must incorporate consideration of two major aspects:
first, what was the metallogenic nature of each
accreted microplate at the time of its mooring at the
accreting margin; and second, what have subduction
mechanisms, rifting, or other aspects of plate tec-
tonics done to the assembly since docking.
We can therefore consider pre-accretion, syn-
accretion, and post-accretion effects. Pre-accretion
effects are those aspects of metallogeny locked into
an island arc (porphyry copper-gold, epithermal
precious metal, volcanogenic sulfide sequences), a
slice of continental crust (porphyry copper-moly,
Cordilleran vein, epithermal precious metal), or a
wedge of oceanic crust (a Cyprus-type Cu-Zn vol-
canogenic massive sulfide, a scarf of tectonized peri-
dotite chromitite) that are welded, intact, onto the
widening margin. The accreted fragments will also,
of course, carry prisms of sedimentary and volcanic
rocks, granitoids, and metamorphic assemblages
appropriate to their origins that may become subject
to later events. It appears that several of the many
porphyry base-metal systems of British Columbia
were in fact delivered to the accreting margin as
though on a raft.
Some microplates may have been delivered in
passive tectonic media such as along a strike-slip or
medial fault, but most were probably carried less
obliquely toward trenches and Benioff zones to their
collisions and dockings, and were thus subject
almost immediately to another metallogenic event
or influence from a subjacent subduction system.
Syn-accretion metallogenic effects would include
every aspect from blueschist metamorphism to full-
blown magmatic arc mechanisms, with the oppor-
tunity of development of a second set of consuming
margin manifestations superimposed on the pre-
accretion microplate ones. Finally, as the margin
widens as the subjacent subduction zone evolves to
varying degrees of complexity or as the result of
much later orogeny, post-accretion overprinting on
the already heterogeneous suspect terrane would
further complicate the final composite metallogenic
inheritance.
Griffiths and Godwin (1983) examined the
metallogeny of British Columbia before suspect
terrane analysis was generally appreciated. Their
analysis, coupled with more recent isotopic studies
aimed at root understanding of microplate origins,
gives us a look at accretion metallogeny, particu-
larly at what are interpretable as post-accretion
effects. They indicate that the tholeiitic and cal-
calkaline melts which moved up from the partially
melting slab into the volcanic arc implanted copper-
gold porphyry systems at the outboard edge, and
that inboard portions of the margin (now inter-
preted as microplates) were affected by upwarped
isotherms and melting curves, thereby producing
mineral deposits of varying Cu-Au-Mo-Sn-W ratios
(Figure 4). It can now be appreciated, especially
since Griffiths' and Godwin's ideas run counter to
Keith's, that the metallogeny of accreted terranes
can be exceedingly complex, although detailed study
with the awareness of the complications of over-
prints of pre-accretion effects by syn- and post-
accretion ones can be effective, as shown by Albers
(1983) for the western United States. We now
recognize that much of the western margin from
Alaska to Panama is in fact suspect, so metallo-
geneticists must incorporate that knowledge. Study
of each crustal fragment with careful isotopic and
geochemical "time-slice" or "still-frame" analysis is
proving succesful (Keppie and Dallmeyer 1989;
Keppie, this volume).
Metallogeny at Precambrian and Mesozoic continental margins 135
Subduction Variation
A second cause of variation in metallogenic
patterns is short-term variation in subduction
mechanisms. If regular, smooth subduction (see
Figure 2) produces regular metallogenic belts (see
Figure 3), then one aspect of imperfection in belts
can be attributed to variations on the subduction
theme. That variability can include variation in the
amount and composition of arc-trench sediment
being deposited and subducted, the thickness and
composition of the oceanic crust being consumed
(does it carry Cyprus-type Cu-Zn massive sulfides
with it?), the amount of water going down, the array
of isotherms around the slab and in the lithosphere
above it (see Figures 2 and 4), and the angle and
rate of subduction.
The density of radiometric age determinations
and petrologic identifications in the western United
States has revealed the well-known pattern that has
been interpreted to reflect plate angle variation
through the Cenozoic with concomitant migration of
arc magmatism (Coney and Reynolds, 1977). Trans-
gression-regression superimposed metallogenic
zones as it superimposed lithologic ones. Keith
VOl..C""'lC AlAe.
of------- COHTIHEMTAI. MARGI N
(1986) and Keith and Wilt (1986) meticulously
described four major serial lithogenic-metallogenic
assemblages that swept across Arizona and adjacent
areas and back during 85-43 Ma. Each lithotectonic
assemblage is now studded with mines and
prospects with different production histories and
major- and trace-element profiles.
Other variations in subduction geometry in-
clude scissors-disruption, with adjacent sections
plunging at different angles and rates, different
penetration depths, and hence different K
2
0-d
(Dickinson and Hatherton, 1967) and CI-F (Mitchell
and Garson, 1972), dismembered or double slabs
(Lipman Prostka, and Christiansen, 1971), and the
more complex results of complicated consuming
margins. The interpretive metallogenic problems
also become formidable, with overlaps, omissions,
and distortions of patterns to be expected.
Metallogeny and Earth History
A final aspect of metallogenic variability con-
sidered here is the effects of changes in metallogenic
processes through time. Regular relationships
TAE- HeM I TYPE Cu- Au I ' UP! CUoMo 5' lYP! SnoW
Figure 4. A model by Griffiths and
Godwin (1983) constructed to show
probable post-accretion metallogenic
variability and perhaps explain some
of it.
o
lone 01 pallogl meu1r.g
ond of
11.e and
(,0 - 0 cd- t'r'IC9"O'
200

GRANI TOIDS
l' (.l\ ....... ..-
'0
1
1500
5UUAC, IPU - !R0510 )
YOUHGr----------------------------,
R!lAIIV! AG!
Of 5EOIM!NI5
t,N() VQi.CANIC5
136 J. M. Guilbert
described above reach back at least to the earliest
Cenozoic, and they include distributions gelling
today: EI Teniente, one of the giant porphyry copper
deposits in Chile, is only 4.5 Ma old, Panguna in the
Solomons is 1.5 Ma, and the epithermal bonanza
gold at Lihir Island is forming now, as may be a
porphyry copper deposit under Mount Saint Helens.
But as we look further back into time in an attempt
to draw comparisons with Proterozoic and even
Archean margins - as we do in the next section of
this paper - these temporal trends seem to have
been significant.
Charles Meyer and Richard Hutchinson stand
out among the authors who have been concerned
with the topic. Both have attempted to sort out
waxings and wanings, even extinctions of ore-
deposit lineages through time. Meyer (1981) pre-
sented an elaborate table with geologic age plotted
against ore-deposit types in major lithologies, with
individual district names in the body of the table. It
is too large to be reproduced here, but the serious
scholar should see the foldout as published. Many
aspects of the table are included in Hutchinson
(1980 and 1981), summarized by Windley (1984),
and discussed by Meyer (1988).
Metallogenic mechanisms have clearly evolved
through time in essential lockstep with changes in
atmosphere, hydrosphere, lithosphere and, as is be-
coming increasingly more evident, biosphere. Oxy-
genation of the atmosphere at "" 2 Ga had obvious
effects on sedimentary iron formation composition,
Figure 5. Evolution of metallo-
genesis and ore-deposit speciation
of major mineral deposit types.
From Hutchinson (1981).
essentially ending as it did the Superior-type BIF
accumulation. Higher oxygen activity also modified
placer-deposit mineralogies by eliminating sulfides.
But by far the greater changes in metallogenesis
can be related to profound changes in what we now
perceive as plate tectonics and Earth history:
(1) massive submarine mafic platform growth to
felsic volcanism, sedimentation, and greenstone-
belt tectonics in the Archean; (2) cratonization and
the preservation of little-deformed supracrustal
sedimentary basins with lesser volcanism in the
Proterozoic; and (3) development ,of global-scale
miogeoclines and calcalkaline magmatic arcs in the
Proterozoic and Phanerozoic. Somewhere along the
way, global-scale subduction as we know it today
came to pass. As we will see, the ore-deposit-
metallogeny procession suggests the Proterozoic as
that key time.
Figure 5 (Hutchinson, 1981) summarizes the
ebb and flow of various metallogenic processes. The
earliest ore-forming environments were almost
certainly in and on the flanks of vast Archean
submarine volcanic edifices, mafic at the bases
progressing to felsic and sediment-dominated at the
tops, in turn built upon thin Archean protomantle or
oceanic crust, perhaps above rifts in those platforms.
Sedimentary textures that are indistinguishable
from modern ones require nearly S-T-P conditions at
the Archean surface, so thermal gradients must
have been steep. Among the earliest ore deposits
were reduced iron formations with or without gold,
Au Au --_______ <::::::::::: Au lod'!s, pluC'rs
lodes U. undSlone type
FomUltion
Au
Udeposils::::::
disc. lode';; Unconformit'1 -------,;;. = bluk sh.lC's
-....... lJ deps. M,uine blolCk
','-...... ,...U V'genic Ni -- f Ni
71'--
Uhnm.fic
vok.nism
KENORAN
OROGENY
I'4i-Cu,. c,. -::::==:::::::===:::=;:=:::::::::>
""ilie C1'\
1 ____ ARCHEAN __ P1I0TIROZOIC ---oi-- PHANEROZOIC
3500 2500 1500
500
Metallogeny at Precambrian and Mesozoic continental margins 137
volcanogenic massive Cu-Zn-S and Cu-Ni-S bodies,
and chromite and asbestos occurrences. The pat-
terns of distribution of these deposits and their
greenstone matrices suggest rapid overturn, pro-
bably short-range, local, rapid, subduction-style
zones, with scales of heat-loss mechanisms on the
order of tens or hundreds of kilometers.
Metallogenesis changed dramatically with
stabilization of continental cratons, which many
equate with the onset of the Proterozoic. Great
thicknesses of epicrustal sediments include paleo-
placers of the Witwatersrand-Jacobina-Tarkwa-
Blind River type, copper- and manganese-rich sedi-
ments, and Superior-type iron formations in tectoni-
cally stable basins. Among the definitive metallo-
genic relationships is that of porphyry coppers.
Most economic geologists accept the Lowell and
Guilbert (1970) definition of "PCDs." According to
that definition, the oldest unequivocal Cordilleran-
style porphyry system is the Haib occurrence in the
Richtersveld province of southern Namibia. The
province is one of the oldest systemic calcalkaline
belts in the world, and it is also host to the oldest
PCD. A cluster of subeconomic porphyry copper-
moly occurrences in Finland is associated with
calcalkaline rocks ofthe same 1.8-1.9 Ga age (Gaal
and Isohanni, 1979).
Because the calcalkaline belt-porphyry copper
association is well established in Phanerozoic
geologic history, and because calcalkaline belts are
now firmly linked to global-scale subduction, it is
inferred that continental-scale subduction mech-
anisms were probably established by 2.0 Ga. It is
tempting to suggest that cratonization permitted
both preservation of supracrustal sedimentary
basins and development of subduction systems
measured in thousands of kilometers. During the
Proterozoic, rifting and proto-rifting of newly com-
petent cratonic masses permitted development of
layered mafic complexes, some massive sulfide
deposits, the Broken-Hill-type (BHD) deposits men-
tioned below, and carbonatites and kimberlites.
The Phanerozoic has seen an increased speci-
fication of ore deposit types, largely integral to the
development of passive continental margins and
miogeoclines. Mississippi-Valley type, Cordilleran
vein limestone replacement deposits, and true
sedimentary exhalative black shale-hosted base
metal deposits are characteristically Paleozoic, as
are evaporite sequences in failed rifts or aulacogens.
The continent-scale calcalkaline belts, with
their contained porphyry and related ore systems,
grew in importance through the Paleozoic, pre-
servation potential and erosional destruction effects
notwithstanding. PCD occurrences, known in Asia,
Appalachia, and British Columbia, are of Paleozoic
and Mesozoic age, but the metallogenesis of por-
phyry systems blossomed in the Mesozoic with
scores of occurrences peaking in the Laramide.
Predictably, calcalkaline volcanism and porphyry
copper emplacement ceased in the western United
States in the mid-Cenozoic when the North
American plate overrode the East Pacific Rise and
the subduction zone between them ceased. Epi-
thermal precious metal mineralization has long
been thought to have been a mid-Cenozoic pheno-
menon, but epithermal districts such as the Silbak-
Premier gold-silver system in British Columbia, of
Jurassic age, indicate that subaerial, volcanic-
hosted, epithermal mineralization is also related to
calcalkaline belts in general and can be used as a
paleotectonic indicator.
Volcanogenic massive sulfide deposits first
developed in Archean greenstone belts and have
persisted in eugeoclinal environments until the
present. Archean deposits in the Abitibi belt of
Ontario are similar to, but distinguishable from, the
1.8 Ga occurrence at Jerome, Arizona, in the Yava-
pai greenschist belt; Cambrian deposits in the Mt.
Reed Volcanics of Tasmania are almost indis-
tinguishable from Miocene Kuroko deposits in
Japan. As Hutchinson (1981) reiterated, the Pb
content of volcanogenic massive sulfides has in-
creased through time.
Clearly, analyses of metallogenesis and Earth
history reveal patterns that can be interpreted in
terms of evolution of consuming margin tectonics.
Windley (1984) summarized the consanguinity of
tectonics and metallogeny in much greater detail
and from the viewpoint of economic geologists like
Meyer and Hutchinson. Insights into the relation-
ship continue to grow through application and deve-
lopment of lithotectonic classifications of ore de-
posits and analyses of accreted terranes.
PROTEROZOIC TERRANES
Detailed studies of metallogeny in Proterozoic
terranes are frustratingly few. Rickard (1987)
described volcanogenic mineralization styles in the
Proterozoic; Colley and Westra (1986) considered
Early Proterozoic mineralization in Finland; Plimer
(1986) differentiated the tectonic settings and geo-
138
J. M. Guilbert
chemistries of the Broken Hill- and Mt. Isa-type
volcanogenic sediments; Herz (1976) treated titan-
ium deposits in mid-Proterozoic anorthosite massifs;
and Anderson (1986a,b) scrutinized lithotectonics
and metallogeny in the Yavapai greenschist belt of
Arizona.
Rickard (1987) examined overall metallogeny of
the Proterozoic, comparing it with Archean and
Phanerozoic phenomena. He noted that two end-
members among the several ore-deposit styles in the
Proterozoic stand out, the volcanogenic massive
sulfides and the Broken Hill-type established by
Plimer (1986). Rickard discerned the fact that
strong lithostratigraphic similarities link the mas-
sive sulfides of Finland (Outokumpu, Orijarvi, and
others), Sweden (the Skellefte district and Falun),
the circum-Ungava and central Canadian deposits
(Flin Flon, La Ronge), several in Wisconsin, and the
cluster of deposits in the Jerome district, Arizona.
Plotted on reconstructions of the Proterozoic super-
continent, they all fall in a metallogenic belt "'" 500
X 8000 km on the "east" side of the continental
mass. Rickard concluded that this (Yavapaian)-
Transhudsonian-Svecokarelian belt "may trace the
scar of an early Proterozoic Wilson cycle." Major,
global-scale subduction appears to have been estab-
lished from both porphyry base-metal and volcano-
genic massive sulfide deposit metallogenic records.
Broken Hill-type deposits appear to be uniquely
mid-Proterozoic, "'" 1.9 Ga. Plimer (1986) concluded
that they formed as a result of successful rifting of
thin crust, extrusion of mafic volcanics, and ascent
of abundant altering, mineral-component-rich,
CO
2
-, F -, and P-rich mantle fluids. Subsequent
underplating of that thin crust, cratonization, and
development of thicker supracrustal sequences may
have jointly ended the BHD era.
Anderson (1986a) recently compiled new field,
geochemical, and geochronological data on the 1.9-
1.65 Ga Proterozoic of Arizona based on several
man-years of mapping and analysis. His compre-
hensive study was aimed at establishing the nature
of Proterozoic plate tectonic mechanisms and metal-
logeny (Anderson and Guilbert, 1979), among other
things. The massive sulfide deposits proved to be
contemporaneous with the maximum expression of
felsic volcanism atop tholeiitic to calcalkaline dif-
ferentiation sequences (see also Meyer, 1988) in the
five sub-belts of the Yavapai. These belts, and the
ores in them, are not ensialic but are either intra-
oceanic island arcs on Proterozoic oceanic crust or
true consuming continental-margin arcs. Rickard's
deposit description and tectonic model are strikingly
similar, especially with regard to the second of
Anderson's options. The Proterozoic massive sul-
fides and their settings are similar to the Kuroko
analogues in Japan, so the conclusion can be drawn
(Anderson, 1986b) that "broadly similar processes of
lithospheric subduction generated the magmas and
ultimately the massive sulfide deposits." Ander-
son's summary diagram (Anderson, 1986b, his Fig.
6) shows his interpretation of plate tectonic mech-
anisms that operated in generation of the Yavapai
greenschist belt, including both an intraoceanic
island arc at 1.8 Ga and a continental consuming
margin 80 million years later.
Anderson (1986a) concluded that (1) the Yava-
pai belts resulted from Proterozoic island arcs that
are analogous to modern arcs, with expectable
differences in size, composition, and other aspects;
(2) the lack of ophiolites and ultramafic volcanic
sequences suggests that Proterozoic oceanic crust
was quite different from either Archean or Phanero-
zoic material, and is commensurately difficult to
identify; (3) the lack of blueschists and classic
melange in the Yavapai imply that the P-T or rate
regime of subduction were different in detail from
modern subduction, but that broad similarity
existed; (4) intense penetrative deformation with
steep lineation and strong vertical extension mark
the Yavapai rocks - Archean deformation in
general being variable in plunge and strain state,
and Phanerozoic time showing more horizontal
transport and lesser penetrative effects; and (5) sev-
eral Archean and Phanerozoic lithotypes, such as
komatiite, graywacke, and molasse, are rare in the
Proterozoic, and many volcanic rocks differ sys-
temically in chemical detail. "Such contrasts imply
that tectonic processes broadly similar to modern
ones operated during the Proterozoic era, but
produced tectonic features that are importantly
different to modern ones in many essential details,
thus constituting a style of plate tectonics that was
unique to Proterozoic convergent margins ... [and] in
many respects transitional in its tectonic features."
(Anderson, 1986a).
CONCLUSIONS
1) Metallogenic belts and epochs, in terms of both
mineralization occurrences and trace element distri-
bution, reflect crustal development mechanisms
through time.
Metallogeny at Precambrian and Mesozoic continental margins 139
2) Lithotectonic classifications of mineral deposits
illuminate both metallogenesis and lithogenesis in
ancient and modern settings.
3) Subduction generates consistently polarized
metallogenic belts characterized by progressive
inboard Fe-Cu-Mo-Zn-Pb-Ag-Au peaks that vary
with the alkali-alkali earth profiles of their cognate
igneous host-rocks.
4) Variation of subduction angle, rate, uniformity,
and ingestion of sediment, water, and oceanic crust
affect epicrustal geometry and nature of metallo-
genic belts in ways that are becoming more pre-
dictable and manageable.
5) Accretion of margins can be interpreted by de-
tailed, time-slice methods designed to recognize
intrinsic pre-accretion microplate metallogeny and
the effects of syn-accretion docking tectonics and
post-accretion subduction, if and when it occurs.
6) Evolution of ore-deposit types and genetic
frameworks through Earth history can be sym-
biotically interpreted in terms of continental
margin processes and growth mechanisms, thereby
illuminating and extending both.
7) Archean, Proterozoic, and Phanerozoic metallo-
geny confirm gross aspects of those periods deduced
otherwise, but provide insights not otherwise
perceived: (l)A belt of volcanogenic massive sulfide
deposits from Finland to Arizona requires an 8000-
km-Iong subducting margin along one side of Gond-
wanaland at 1.8 Ga. (2) The appearance of fully
developed, "typical" porphyry base metal deposits in
calcalkaline belts at 1.8 Ga similarly indicates sub-
duction processes, but their global dispersion
suggests that several subducting margins existed,
including one that produced the Richtersveld pro-
vince in Namibia. (3) Anorthosite and Broken Hill-
type settings imply relatively thin Proterozoic
cratons that were rifted, in the case of BHDs
permitting effusion of mafic volcanics and abundant
mantle-derived volatiles with steep thermal gra-
dients. Neither of these deposit-set environments
recurred during the rest of Earth history. (4) Epi-
crustal placer gold-uraninite-pyrite deposits inter-
leaved in thick sections of siliciclastic little-
deformed sediments bespeak reducing atmosphere
and stabilized craton, probably marking the shift to
increased cratonization at the Archean-Proterozoic
boundary.
8) Metallogeny in the Yavapai greenschist belt
(Anderson, 1986a,b) shows operation of Proterozoic
intraoceanic island arc and then continental margin
subduction analogous but not identical to today's
scenarios. The absence of ophiolites and ultramafic
volcanic rocks is distinctive, and the lack of blue-
schists and melanges indicates higher P-T gra-
dients. Massive sulfides are "Proterozoic specific,"
with felsic aqua gene basalts and tuffs below and
blister domes and ashflow turbidites above, suggest-
ing an Andean-type subducting margin. Intense
penetrative deformation, with strong vertical linea-
tion and extension, contrasts with more variably
directed Archean modes of deformation and lesser
deformation coupled with greater horizontal trans-
port in the Phanerozoic.
(9) Metallogenic assessment of plate tectonic mech-
anisms and terranes can powerfully extend tradi-
tional methods of continental margin analysis.
REFERENCES
ALBERS, J. P., 1981a, Distribution of mineral deposits in
accreted terranes and cratonal rocks of western United
States: Canadian Journal of Earth Science, v. 20, p. 1019-
1029.
__ , 1981b, A lithologic-tectonic framework for the
metallogenic provinces of California: Economy Geology, v.
76, p. 765-790.
ANDERSON, P., 1986a, The Proterozoic Tectonic Evolu-
tion of Arizona: PhD dissertion, University of Arizona,
Tucson, Arizona, USA, 416 p.
__ , 1986b, Summary of the Proterozoic plate tectonic
evolution of Arizona from 1900 to 1600 Ma; pp. 5-11 in B.
Beatty and P. A. K. Wilkinson (eds.), Frontiers in Geo-
logy and Ore Deposits of Arizona and the Southwest:
Arizona Geological Society Digest, v. 16, 594 p.
__ , and J. M. GUILBERT, 1979, The Precambrian mas-
sive sulfide deposits of Arizona - A distinct metallogenic
epoch and province; pp. 39-48 in Proceedings of the Fifth
Quadrennial IAGOD Symposium (Salt Lake City,
Utah, 1978): Nevada Bureau Mines and Geology, Report
33.
ANTHONY, E. Y., and S. R. TITLEY, 1986, Geochemical
evidence for crustal melting at the Sierrita porphyry
copper deposit, southeastern Arizona [abstract]; in Ab-
stracts of the Seventh Quadrennial IAGOD Sym-
posium (Lulea, Sweden): Terra Cognita, v. 6, p. 552.
BRIMHALL, G. H., Jr., 1987, Preliminary fractionation
patterns of ore metals through Earth history: Chemical
Geology., v. 64, p 1-16.
BURNHAM, C. W., 1959, Metallogenic provinces of the
southwestern United States and northern Mexico: New
140 J. M. Guilbert
Mexico Bureau of Mines and Mineral Resources Bulletin, v.
65,p.1-76.
CAMPA, M. F., and P. J. CONEY, 1981, Tectono-stragraphic
terranes and mineral resource distributions in.Mexico:
Canadian Journal of Earth Science, v. 20, p. 1040-1051.
COLLEY, R., and L. WESTRA, 1987, The volcano-tectonic
setting and mineralization of the Early Proterozoic Kemio-
Orijarvi-Lohja belt, SW Finland; pp. 95-112 in T. C.
Pharaoh, R. D. Beckinsale, and D. Rickard (eds.), Geo-
chemistry and Mineralization of Proterozoic Volcanic
Suites: Blackwell Scientific Publishers, London, England,
UK,575p.
CONEY, P. J., and S. J. REYNOLDS, 1977, Cordilleran
Benioff zones: Nature, v. 270, p. 403-406.
DICKINSON, W. R., and T. HATHERTON, 1967, Andesitic
volcanism and seismicity around the Pacific: Science, v.
157, p. 801-803.
FARMER, G. L., and D. J. DEPAOLO, 1983, Origin of
Mesozoic and Tertiary granite in the western United
States and implications for pre-Mesozoic crustal structure,
I: N d and Sr isotopic studies in the geocline of the north-
ern Great Basin: Journal of Geophysical Research, v. 88, p.
3379-3401.
GAAL, G., and M. ISOHANNI, 1979, Characteristics of
igneous intrusions and various wall rocks in some Precam-
brian porphyry copper-molybdenum deposits in Pohjan-
maa, Finland: Economic Geology, v. 74, p. 1198-1210.
GRIFFITHS, J. R., and C. 1. GODWIN, 1983, Metallogeny
and tectonics of porphyry copper-molybdenum deposits in
British Columbia: Canadian Journal of Earth Science, v.
20, p. 1000-1018.
GUILBERT, J. M., 1972, Emplacement of mineralized in-
trusive rocks with regard to regional structures and plate
tectonics: Colorado Mining Association, Mining Yearbook,
p.44-51.
___ , 1981, A plate-tectonic-lithotectonic classification of
ore deposits; pp. 1-10 in W. R. Dickinson and W. D. Payne
(eds.), Relations of Tectonics to Ore Deposits in the
Southern Cordillera: Arizonia Geological Society Digest,
v. 14, p. 288.
__ , and C. F. PARK, Jr., 1986, The Geology of Ore
Deposits: W. H. Freeman, New York, New York, USA,
986p.
HERZ, N. L., 1976, Geology and Resources of Titanium:
U.S. Geological Survey, Professional Paper 959A-F, 49 p.
HUTCHINSON, R. W., 1980, Massive base metal sulphide
deposits as guides to tectonic evolution; pp. 659-684 in D.
W. Strangway (ed.), The Continental Crust and its Min-
eral Deposits: Geological Association Canada, Special
Paper 20,804 p.
__ , 1981, Mineral deposits as guides to supracrustal
evolution; pp. 120-140 in R. J. O'Connell, and W. S. Fyfe
(eds.), Evolution of the Earth: American Geophysical
Union/Geological Society of America, Geodynamics Series
3,282p.
KEITH, S. B., 1978, Paleosubduction geometries inferred
from Cretaceous and Tertiary magmatic patterns in south-
western North America: Geology, v. 6, p. 516-521.
1986 Petrochemical variations in Laramide mag-
and their relationship to Laramide tectonic and
metallogenic evolution in Arizona and adjacent regions;
pp. 89-101 in B. Beatty and P. A. K. Wilkinson (eds.),
Frontiers in Geology and Ore Deposits of Arizona and
the Southwest: Arizona Geological Society Digest, v. 16,
594p.
__ , and J. C. WILT, 1986, Laramide orogeny in Arizona
and adjacent regions: A strato-tectonic synthesis; pp. 502-
554 in B. Beatty and P. A. K. Wilkinson (eds.), Frontiers
in Geology and Ore Deposits of Arisona and the
Southwest: Arizona Geological Society Digest, v. 16, 594
p.
KEPPlE, J. D., and R. D. DALLMEYER (comp.), 1989,
Tectonic Map of the Pre-Mesozoic Terranes in the
Circum-Atlantic Phanerozoic Orogens (1:5,000,000):
Wolfville, Nova Scotia, Canada.
LAMBERT, 1. B., and D. 1. GROVES, 1981, Early Earth
evolution and metallogeny; pp. 339-447 in K. H. Wolf (ed.),
Handbook of Strata-Bound and Stratiform Ore Depo-
sits: Elsevier, New York, New York, USA, 592 p.
LIPMAN, P. W., H. J. PROSTKA and R. L. CHRISTIAN-
SEN, 1971, Evolving subduction zones in the western
United States as interpreted from igneous rocks: Science,
v. 174, p. 821-825.
LOWELL, J. D., and J. M. GUILBERT, 1970, Lateral and
vertical alteration-mineralization zoning in porphyry ore
deposits: Economic Geology, v. 65, p. 373-408.
MITCHELL, A. H. G., and J. D. BELL, 1973, Island-arc
evolution and related mineral deposits: Journal of Geo-
logy, v. 81, p. 381-405.
__ , and M. S. GARSON, 1972, Relationship of porphyry
copper and circum-Pacific tin deposits to palaeo-Benioff
zones: Transactions of the Institute of Mining nad Metal-
lurgy, v. 81, p. B10-B25.
MEYER, C., 1981, Ore-forming processes in geologic his-
tory; pp. 6-41 in B. J. Skinner (ed.), Seventy-Fifth Anni-
versary Volume: Society for Economic Geology, Econo-
mic Geological Publishing Company, 964 p.
__ , 1988, Ore deposits as guides to geologic history of
the Earth: Annual Reviews of Earth and Planetary Sci-
ence, v. 16, p. 47-171.
Metallogeny at Precambrian and Mesozoic continental margins 141
NOBLE, J. A., 1970, Metal provinces of the western United
States: Geological Society of America Bulletin, v. 81, p.
1607-1624.
PLIMER, 1. R, 1986, Sediment-hosted exhalative Pb-Zn
deposits - Products of contrasting ensialic rifting: Trans-
actions of the Geological Society of South Africa, v. 89, p.
57-73.
RICKARD, D., 1987, Proterozoic volcanogenic mineraliza-
tion styles; pp. 23-35 in T. C. Pharaoh, R D. Beckinsale,
and D. Rickard (eds.), Geochemistry and Mineralization
of Proterozoic Volcanic Suites: Blackwell Scientific
Publishers, London, England, UK, 575 p.
SAWKINS, F. J., 1972, Sulfide ore deposits in relation to
plate tectonics: Journal of Geology, v. 80, p. 377-397.
__ , 1984, Metal Deposits in Relation to Plate Tec-
tonics: Springer-Verlag, New York, New York, USA, 325
p.
SILLITOE, R H., 1972, A plate tectonic model for the origin
of porphyry copper deposits: Economic Geology, v. 67, p.
184-197.
__ , 1981, Ore deposits in Cordilleran and island arc
settings; pp. 49-69 in B. Beatty and P. A. K. Wilkinson
(eds.), Frontiers in Geology and Ore Deposits of Ari-
zona and the Southwest: Arizona Geological Society
Digest, v.16, 594p.
TARLING, D. H. (ed.), 1981, Economic Geology and Geo-
tectonics: Halsted Press/John Wiley, New York, New
York, USA, 213 p.
TOOKER, E. W., 1981, Correlation of metal occurrence and
terrane attributes in the northwestern conterminous
United States: Canadian Journal of Earth Science, v. 20,
p.1030-1040.
WESTRA, G., and S. B. KEITH, 1981, Classification and
genesis of stockwork molybdenum deposits: Economic
Geology, v. 76, p. 844-873.
WINDLEY, B. F., 1984., The Evolving Continents: John
Wiley, New York, New York, USA, 399 p.
CHAPTER 2
CORDILLERAN MESOZOIC MARGIN
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Post-Archean crustal evolution, terrane accretion, and
metamorphism of the Western Cordillera, with emphasis on
northern and central California *
W.G.ERNST
School of Earth Sciences, Stanford University, Stanford, California 94305-2210, USA
*UCLA Institute of Geophysics and Planetary Physics Publication No. 3217
(received August 16, 1988; revision accepted December 27,1988)
ABSTRACT
In the conterminous United States, narrow Phanerozoic high-P blueschist + serpentinized peridotite
zones lie outboard from coeval high-T metamorphosed calcalkaline magmatic belts bordering the Pacific
Ocean; both lie parallel to the North American margin. They represent recrystallized products of mid-
Paleozoic and younger subduction zones, Andean margins and exotic arcs. These belts contain allochthonous
oceanic units but, judging from provenance and isotope data, voluminous terrigineous debris and parental
magmatic zones appear to be native to the evolving North American margin. Far-traveled sialic
micro continental scraps are rare. Metamorphic parageneses farther within the continent are associated with
chiefly mesozonal granitic plutons. Glaucophane schists, eclogites, and large tracts of ophiolitic ultramafic
rocks are absent, partly due to overprinting of accretionary margin assemblages by subsequent orogenies,
high-T recrystallization and anatexis. For example, at depth within the California Coast Ranges, high-P
phases are currently being destroyed. Inland, a paucity of petrotectonic assemblages of oceanic affinities
probably reflects differential sinking of these dense crustal units during tectonism. The late Mesozoic
Franciscan trench complex/Great Valley forearc basin/Sierra Nevada magmatic arc constitutes the most
completely preserved accretionary segment along the western conterminous U.S. continental margin, and it
serves as a type example ofparautochthonous crustal evolution.
Precambrian Pb, Nd, and Sr isotope ratios and igneous/metamorphic formation ages, in comparison with
Phanerozoic recrystallization and metamorphic-facies distributions, suggest a gradual zonal enlargement of
the continent southward from the Wyoming craton during Early and Middle Proterozoic time, followed by
Late Proterozoic-Phanerozoic westward development. Sialic growth in the western U.S. Cordillera appears to
have been due chiefly to partial fusion and magma ascent above continentward subducting paleo-Pacific
lithospheric plates, combined with metamorphic and sedimentary reworking in the forearc + trench + back-
arc regions, rather than resulting from the amalgamation of variably old and young geologically unrelated,
far-traveled, exotic crustal fragments.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 145-168. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
145
146 W. G.Ernst
INTRODUCTION
Long-sustained plate-tectonic motions affect
the thermal evolution of the continental crust,
which in turn is attested to by preserved meta-
morphic assemblages (Miyashiro, 1973). Construc-
tional stages are typified by head-on and oblique
convergent plate motions, resulting in the develop-
ment of: (1) broad inboard, low PIT, high heat-flow
regimes, the addition of primary calcalkaline
igneous arc rocks, the anatectic, metamorphic, and
sedimentary reworking of pre-existing materials;
and (2) narrow outboard belts of high PIT, low heat-
flow subduction complexes, the suturing of far-
travelled oceanic assemblages island arcs and
allochthonous microcontinental fragments, and a
lack of coeval calcalkaline igneous activity. Such
paired belts provide a characteristic pattern which
can be used to document the nature of continental
accretion. Episodic rifting and strike-slip faulting,
which attend divergent and transform plate
motions, respectively, modify and obscure the record
of real sialic growth. Remnants of these meta-
morphic belts, therefore, are scattered piecemeal
and annealed/recrystallized by later thermotectonic
events. Rifting and transform motions thus re-
arrange previously produced geologic terranes but
do not result in an increase in the aggregate mass of
the Earth's sialic crust, hence these processes
cannot account for continental growth. (They do
result in the enlargement of acceptor continental
assemblies at the expense of donor continents, of
course.)
Experimental phase relations for synthetic and
natural rock and mineral systems, taken together
with
18
0/
16
0 and DIH isotopic geothermometers and
diverse mineralogic thermobarometers, allow the
erection of a metamorphic-facies grid. Facies illus-
trated schematically in Figure 1 are appropriate for
rocks of roughly basaltic composition, at oxygen
fugacities within the magnetite f02-T stability field
and - except for granulitic lithologies where low
f
H2
0 prevails at a relatively high fco
2
- at high
fHj!O' Also projected onto this diagram are phase
relations for Al
2
Si0
5
polymorphs that would be
stable in spatially-associated aluminous para-
schists.
The thermal structures of divergent and con-
vergent plate boundaries have been modeled
numerically by many workers (e.g., Oxburgh and
Turcotte, 1971; Goto and others, 1987; van den
Beukel and Wortel, 1987). Employing computed P-

/ / srable
/ 00> / ' .... ...'Plate
t rench complex / ,,<J / .' interior
y (7V .I
/ (6) /.1 volcanic/
/ / / plutonic ,I
/ / . arc /
k!:J / / \/
(3)/ / / /
/ / /
/ (5) / / '0 /
/ - / i /
prehnite- 1 / -a/ (al
" ,,/ - E
pumpel Ylte ., / /0 granulite
I (2) Z'/ .:;: /
I 0 11
I /0. /' l si
I / / / ... '. .21
/ //<:,""
/ (I) / // "
8

/ / /v \ ...... "\
2 / oS? "" ...... '. spreading
/ / ./'/' ,,/ ' . center
I / ./ / ,,'/ d ' .
I /././ / // an ,
/ ....... ._..... hornfels
\.
o 200 400 600 800
Temperature in C
Figure 1. Petrogenetic grid for basaltic rock types at
intermediate oxidation states, with P-T assignments of
metamorphic facies, after Ernst (1976). An aqueous fluid
is assumed to be present except for granulite facies
metamorphic environments where fluid may be absent or,
if present, apparently is rich in CO
2
and very low in H
2
0.
Kyanite, sillimanite and andalusite fields of stability
(Holdaway, 1971) in associated aluminous pelites are also
indicated. Typical examples of metamorphic P-T trajec-
tories or field gradients (Richardson, 1970; Spear and
others, 1984) are illustrated.
T arrangements, the geologic disposition of meta-
morphic mineral assemblages indicated in the facies
grid of Figure 1 are shown for simplified divergent
and convergent lithospheric plate junctions in
Figures 2 and 3, respectively. The very high heat-
flow regime characterized by an oceanic spreading
system accounts for a telescoped metamorphic
zonation and relatively shallow-level development
of high-rank hornfels beneath the ridge (Figure 2).
In contrast to this simple, bilaterally symmetrical
pattern, the more complex convergent plate-tectonic
setting is characterized by a broad, relatively high
heat-flow environment in the magmatic arc, re-
flected by a general symmetric upwarp of the
metamorphic-facies assemblages + anatexis of
basal portions of thick, "juicy" continental crust,
The post-Archean Western Cordillera, California
o 10 2Okm.
"""""""'" ,...........,
oceanic
ridge sea level
YOOOOt::;; 0&
(I) zeol ite ()
(5) greenschist
lithosphere
:'. """ (J' .
, '?'\ "":' {j ' ..
: \'" '0 '. -:.'
.. U":':', ., ....
. asthenosphere
'.. . .
147
Figure 2. Schematic distribution of metamorphic facies in the high heat-flow regime of a simple divergent plate boundary,
after Ernst (1976) . Individual facies are numbered as in Figure 1.
....
trench
complex forearc basin
(5)
(11 zeolite (laumontite)
(2)p'ehnite-pumpellyile
(3lblueschist (lawsonite)
(4leclogile
(5) 9'eenschlst (chlorite)
(6)low,onk amphibolite
(biotilel
(7)high,ank amphibolite
(8)Qronuhle. hornf'els
asthenosphere
magmatic arc
Figure 3. Schematic distribution of metamorphic facies in the low heat-flow trench environment and high heat-flow
magmatic-arc environment appropriate to a simple convergent plate boundary, after Ernst (1976). Individual facies are
numbered as in Figure 1. The broad expanse of high-T assemblages in the arc and the downward projection of high-P
assemblages in the subducting slab are consequences of magmatic heat transport and cool lithospheric plate descent,
respectively.
and a spectacular downward, asymmetric projection
of relatively low-T, high-P phase compatibilities in
the subduction zone (Figure 3).
Crustal metamorphic environments and geo-
chemical signatures (major elements, REE, Pb, Nd,
Sr) of igneous rocks derived from the mantle and/or
148
W. G. Ernst
__ __ ''_''_' '_- -- - -
- .. -
o Z-PP: zeolite + prehnite-pumpellyite
o B-E: blueschist + eclogite
o G: greenschist
A: amphibolite
f] G-H: granulite + pyroxene hornfels

r
J.'

. \
"
..
Figure 4. Generalized areal disposition of metamorphic belts of the western conterminous U.S., based on regional summary
contributions in Ernst (1988); serpentinized peridotites shown in black. The dominant regional metamorphic facies
assemblages are presented irrespective of age. Physical conditions inferred for these metamorphic facies are illustrated in the
poT diagram of Figure 1. Facies assemblages of Figure 1 are combined, however, as shown in the inset [(1) + (2) = Z-PP; (3)
+ (4) = B-E; (6) + (7) = AJ.
()
6
N
o
Z
w
u
a

u
6
N
o
CIJ
w
::E
w

c(
-l
..
The post-Archean Western Cordillera, California
- ._ .. _ .. _ .. _ .. _ .. _ .. -
o Late Meso-Mid Cenozoic
El Mid Paleo-Early Mesozoic
ca Proterozoic
r::J Archean
(
1-'
,
. \
149
Figure 5. Approximate ages of principal recrystallization events characteristic of the dominant metamorphic assemblages
(see Figure 4), based on regional summary contributions in Ernst (1988); serpentinized perioditites shown in black. Ages of
metamorphism are combined as shown in the inset.
150
W. G.Ernst
the deep crust, combined with regional structural
and lithologic/tectonic age relationships, provide
constraints regarding the complex interplay of
processes attending continental accretion. As
reviewed below, isotopic data suggest that for the
western U.S. Cordillera, the bulk of sialic material
was added northwest and, especially, south of the
Archean shield during Early and Middle Proterozoic
time, whereas Late Proterozoic-Phanerozoic growth
occurred predominantly along the western margin
of the North American craton. Continental en-
largement was accompanied by the development of
successively younger igneous + metamorphic belts.
Geologic relations of the Franciscan/Great Valley/
Sierra Nevada lithotectonic triad is a well-preserved
Mesozoic example of continental growth and pro-
vides a possible analogue of some older lithotectonic
complexes.
TECTONOMETAMORPHIC BELTS OF THE
WESTERN U.S.
Schematic maps summarizing metamorphism
in the Western Cordillera are presented as Figures
4 and 5; compilations are based chiefly on relation-
ships described by many workers (e.g., Ernst, 1988).
Although the Mesozoic history of the continental
margin is reasonably well known, younger, relat-
ively intact metamorphic tracts are still largely
buried. In contrast, pre-Mesozoic belts have been
deeply eroded, or overprinted by later dynamo-
thermal and tectonic processes, and thus are
generally preserved in fragmentary fashion at best.
For this reason, Mesozoic belts seem to represent the
optimum stage of preservation/exposure for de-
ciphering the petrogenetic evolution. Inboard from
the Pacific margin, the direct relationship between
lithotectonic assemblages and inferred plate tec-
tonic motions, and a relatively complete rock record,
are progressively obliterated, making study of the
Phanerozoic continent and its Precambrian base-
ment more difficult.
Nevertheless, several conclusions may be
drawn from the compilation maps (Figures 4 and
5). Archean metamorphism, of which chiefly high-
T, low- to moderate-P amphibolite and granulite
facies assemblages are preserved, is confined to
terranes of Wyoming and adjacent parts of Idaho,
Utah, Montana, and South Dakota. Cratonization
was completed here by late Archean time, and much
of the ancient crust was not influenced significantly
by later dynamothermal metamorphism. Weak re-
crystallization characterizes the Upper Proterozoic
passive-margin section northwest of the Wyoming
nucleus. Parageneses developed in Middle Protero-
zoic accretionary sialic crust to the south in Colo-
rado, Utah, and New Mexico are dominantly of
intermediate-P amphibolite type. A regional transi-
tion of uncertain origin exists from greenschists in
southeastern Arizona to low-P amphibolites and
low-rank granulites in western Arizona. In the
Mojave Desert and eastern Californian Transverse
Ranges, high-rank, low-P amphibolites of Protero-
zoic age are scattered along the modern San
Andreas right-lateral transform system. In Cali-
fornia, east-central Oregon, and northwestern
Washington, fragmentary scraps as well as coherent
metamorphic belts, reflecting Ordovician through
Cretaceous recrystallization events, are preserved
as relics of paleo-Pacific convergent margins; sut-
ures are marked by tectonized meta-ophiolites,
mantle fragments, and rare allochthonous micro-
continental scraps. Oceanward, subduction-zone,
high-P, non-volcanic metamorphic belts are invari-
ably juxtaposed against landward, high-T, contin-
ental-margin, calcalkaline batholithic realms and
associated metamorphic wall rocks.
Westward from the North American cratonal
hinge line, these belts have been partly to thorough-
ly overprinted by Late Jurassic-Cretaceous and
early to mid-Cenozoic intracontinental polymeta-
morphism accompanying widespread igneous acti-
vity. Products of this event include the mobilization
of inboard, high-rank core complexes (Coney, 1980;
Armstrong, 1982; Anderson, 1988) and broad-scale
development of greenschist- and low-P amphibolite-
facies assemblages in rocks of the Basin and Range,
the Mojave-Sonoran Desert, and the Sierra Nevada-
Peninsular Ranges. Greater depths of emplacement
are recorded in country rocks surrounding and north
of the Idaho batholith, where moderately high-P
amphibiIites of Late Cretaceous metamorphic age
are exposed. Outboard from the Late Triassic North
American margin in western California and north-
western Washington, accretionary prisms contain-
ing abundant oceanic as well as terrigenous
materials have been subjected to zeolite- to blue-
schist-facies metamorphism during the suturing of
these ensimatic belts against the continent.
Metamorphic terranes are successively younger
from the Wyoming continental nucleus toward the
Pacific Basin and the Gulf of Mexico. Where de-
tailed geologic relationships are available within
The post-Archean Western Cordillera, California 151
anyone lithotectonic province (e.g., Cheyenne belt,
Colorado-New Mexico, northwestern Washington,
Klamaths, Sierra Nevada, California Coast
Ranges), the ages of original rock assemblages and
times of recrystallization generally decrease sea-
ward. The observed temporal sequence offormation
is compatible with parautochthonous growth but
would not be expected if random exotic slices of older
and younger oceanic and microcontinental terranes
had been stranded haphazardly at the accreting
western margin of North America.
Paired metamorphic belts are preserved only in
Phanerozoic sections of the Cordillera directly
bordering the Pacific Ocean. Inland, recrystallized
terranes of diverse ages exhibit lithologic assem-
blages produced mainly inboard in continental-
margin and island-arc settings. Conspicuous by
their absence are outboard blueschist belts and
extensive tracts of peridotite; if such associations
were formed in early Phanerozoic and older plate-
tectonic environments of the western U.S., they
were either selectively destroyed by subsequent
diastrophism or transported away along margin-
parallel shear zones.
CASE HISTORY OF A CONSTRUCTIONAL
CONTINENTAL MARGIN IN CALIFORNIA
Evidence concerning the largely Mesozoic ac-
cretion of sialic crust in northern and central
California is preserved in metamorphosed litho-
tectonic belts of the Sierra Nevada, Great Valley,
and Coast Range provinces. These contrasting
assemblages display geologic affinities with the
Klamaths on the north and the Transverse Ranges
+ Mojave Desert + Peninsular Ranges to the south
and southeast. Terranes and areal distribution of
metamorphic belts are presented in generalized
form in Figures 6 and 7, respectively.
Sierra Nevada
The Sierra Nevada appears to represent the
southeastward continuation of the Klamath pro-
vince (Davis, 1969; Hietanen, 1981; Day and others,
1988). However, as presently exposed, the Kla-
maths consist mainly of metamorphosed country
rocks intruded by discrete, pre-Cretaceous calc-
alkaline plutons, whereas the Sierra Nevada con-
sists principally of coalescing, Jura-Cretaceous
batholithic units, separated imperfectly by thin
metamorphic septa. The northwest foothills belt is
the only major segment of Sierran wall rocks
preserved adjacent to the plutonic series, although
isolated roof pendants occur scattered throughout
the range. Compared to the Klamaths, the Sierra
Nevada may represent a deeper level of crustal
exposure, especially at its southern extremity where
high-rank amphiboliticllow-rank granulitic meta-
igneous rocks beneath the batholith are exposed
(Sams and Saleeby, 1988).
Sutures bounding individual Sierran metamor-
phic belts are marked by serpentinized peridotites,
but the amount of ultramafic material decreases to
the southeast. Different lithotectonic units (Sharp,
1988) from east to west include: the early to mid-
Paleozoic Shoo Fly terrane; the Feather River
peridotite and blueschists of the Melones fault zone;
the late Paleozoic-early Mesozoic Calaveras-
Melones complex; and the western Jurassic belt.
These lithotectonic units display a steeply east-
dipping imbrication; isoclinally folded sections have
been recognized, but where primary flow tops and
sedimentary laminations are preserved, sections
typically face east (Clark, 1964,1976). In general,
age of formation, metamorphic grade, and struc-
tural complexity all increase eastward (Schweickert
and others, 1988).
The Shoo Fly assemblage, apparently of local
provenance (Hannah and Moores, 1986), constitutes
some of the most ancient rocks exposed in the Sierra
Nevada. It includes: (1) metamorphosed fine-
grained siliceous sediments, quartzites, and inter-
calated phyllites suggestive of continental shelf,
slope, and rise sedimentation; (2)cherts; (3) a green-
stone + limestone sequence; and (4) a serpentinite-
bearing melange. The complex, at least partly
Silurian in age, is overlain unconformably by Upper
Devonian-Permian arc rocks, which are covered
eastward by stratified Mesozoic calcalkaline vol-
canogenic units as young as Jurassic (Schweickert,
1981; Varga and Moores, 1981). The Shoo Fly is
tectonically juxtaposed and underlain on the west
by a major, steeply east-dipping ultramafic slab, the
== 325-387 Ma old Feather River peridotite (Ehren-
berg, 1975; Standlee, 1978; Saleeby and Moores,
1979). The northern portion of the intervening fault
zone encloses the peridotite and places Shoo Fly and
Feather River mafic-ultramafic units against the
upper Paleozoic-lower Mesozoic Calaveras melange
complex on the west. Farther south, the Calaveras
belt lies between the Melones fault on the west and
152
Figure 6.
Lithotectonic
belts of northern
and central Califor-
nia, based on the
geologic map of Cali -
fornia (Jennings, 1977)
as interpreted by Ernst
(1983) from numerous
literature sources. The
eastern, mid-Mesozoic
sedimentary and volcanic
stratified rocks of the eastern
Klamath and northern Sierra
are depicted with a map pattern
somewhat similar to that of the
western Jurassic belt because,
although not necessarily related,
these section.s were deposited nearly
contemporaneously. For simplicity, the
"Sur series" of the Salin ian block is shown
with a map pattern not greatly different from
tha t employed for eastern Klamath and Shoo Fly
terranes, but again no genetic relationship is implied.
w. G. Ernst
U Quaternary alluvium
D Coastal belt Froociscan
D Froociscon melange bell
Eastern Franciscon bell
MJ:;@ GrtOt Volley Group
o Western Jurossic bell
., Western Poleozliic IJld iiossic bell, CoIaYeros COOlIIIel
L "':;l Centrol metalTlOfphic bell
D [ostein Klomotl1 and SOOo Fly terrooes, 'sur series'
o Plutonic igneous rocks, chiefly granitoids
IC!iI Ultramafic bodies
/ Fault
/" Thrust faull
/ Geologic contact
a 31 <0 filKm
The post-Archean Western Cordillera, California
Figure 7.
Metamorphic
zonation deve-
loped in litho-
tectonic belts of
northern and central
California, based on the
geologic map of California
(Jennings, 1977) as interpreted
by Ernst (1983) from numerous
literature sources. Although high-
grade metamorphic rocks are present,
especially in Sierran roof pendants, only
the high-rank amphibolitic and granulitic
zones of the Klamath central metamorphic
belt and the Salin ian terrane, respectively,
have been distinguished from other, less
intensely recrystallized rocks (biotitic green-
schists, amphibolites, etc,) in this compilation.
' 2O" W
Quaternary atluvi\JTI
Metamorphic Zones
o Zeciile (\cMJrnof1tile -be<l'r.g <n1 diogerIeticl
_ Prehnite-pumpell yite
Blueschist (mostly lawsonite-bearing)
Chlorite greenschist
.. Biotitic cxnpIIibolite <n1 h9le' grade
L:;:':-;::::I PkJtooic igneous rocks, chiefly granitoids
/ FauH
/" Thrust fouR (barbs on upper pIoIe)
/ Geologic contact
153
154 W.G.Ernst
the Shoo Fly terrane on the east, from which it is
separated by the foothills suture, an east-dipping
thrust fault (Schweickert, 1981; Saleeby, 1982).
Greenschist and prehnite-pumpellyite facies
assemblages characterize much of the northern
Shoo Fly (Day and others, 1988; Harwood, 1988),
but rocks reach amphibolite or higher metamorphic
grade east of the foothills suture in the southern
portion of this terrane and in roof pendants
distributed along the length of the batholith (e.g.,
Ernst, 1974). Polymetamorphism probably was
roughly coeval with Permo-Triassic docking of the
outboard Calaveras-Melones complex against this
older assemblage (Schweickert, 1981); accretion of
the former was completed prior to emplacement of
the 170-172 Ma Standard pluton (Snoke and others,
1982), which crosscuts the Shoo Fly-Calaveras
boundary in the southern Sierra Nevada foothills.
Directly south of the Feather River peridotite,
blueschist facies rocks occur in slices within the
Melones fault zone (Hietanen, 1981). Because of the
likelihood of argon loss, the K-Ar radiometric ap-
parent age of 174 Ma (Schweickert and others, 1980)
provides a minimum date of metamorphism for
these high-P schists.
The somewhat chaotic Calaveras-Melones com-
plex consists of several different lithotectonic enti-
ties of cor.trasting origins, brought together along
the anastomosing shear system of the western
Sierra Nevada, probably through oblique conver-
gence. Each segment is rich in meta chert, meta-
argillite, phyllite, fine-grained marble, and meta-
volcanics. Ophiolitic lenses and carbonate blocks in
oceanic melange, some of which contain cosmo-
politan late Paleozoic Tethyan faunas (Douglass,
1967), are evidently exotic, in contrast to associated
terrigenous materials. The main Calaveras belt lies
east of the southern part of the Melones fault zone
and west of the foothills suture. In the northern
Sierran foothills, a much less extensive belt of ap-
parent Pennsylvanian depositional age occurs
directly west of the Feather River peridotite. To the
west, it is juxtaposed against the upper Paleozoic-
Triassic Franklin Canyon Formation (Hietanen,
1981). This unit consists principally of meta-
morphosed basalt, andesite, dacite and rhyodacite,
and minor lenses of phyllite. The westernmost
Calaveras-Melones belt in the northern foothills is
the Horseshoe Bend Formation, an imbricate
melange (Hietanen, 1981); its western boundary is
the Big Bend fault. The Horseshoe Bend consists of
calcalkaline metavolcanics, metasedimentary rocks,
and mafics + ultramafic units of Permo-Triassic
age. To the south, the Bear Mountains and Big
Bend faults merge, constituting the western boun-
dary of the terrane. In this southern region, the
Penon Blanco volcanics and mafic + ultramafic plu-
tonic complex constitute a major portion of the unit.
Tectonic blocks of various metamorphic grades, but
including crossite-bearing metabasaltic materials,
are associated with this ophiolitic slice (Morgan,
1976; Saleeby, 1982); the 196-200 Ma minimum
radiometric apparent age of the latter suggests the
possibility of correlation with the Feather River
blueschists. Other than this isolated occurrence of
high-P tectonic blocks, the Calaveras-Melones com-
plex displays widespread development of biotite ( +
locally garnet) in metapelitic units, and both
epidote-amphibolite and garnet-amphibolite meta-
morphic facies in mafic schists, especially in the
northern part of the belt.
Farther to the southeast, both Melones and
Bear Mountains fault systems are obliterated by
intrusives of the Sierra Nevada batholith. The 200
Ma Kings meta-ophiolite complex, and the == 300 +
Ma Bear Mountains and Kaweah meta-ophiolite
melanges, remnants of oceanic crust + uppermost
mantle, crop out along extensions of these shear
zones (Behrman, 1978; Saleeby, 1982). They are
probably correlative with the Penon Blanco mafic
+ ultramafic body and the Feather River peridotite,
respectively. Judging from the equatorial fauna
contained in associated Permian limestone blocks,
these southernmost pre-Mesozoic mafic-ultramafic
complexes evidently represent far-traveled oceanic
units swept against the accreting North American
continent in Triassic time (Saleeby, 1981). Amphi-
bolite-facies metamorphism is widespread in this
zone and may have been produced during docking of
these oceanic complexes along the developing
margin of California. However, in spite of abundant
ophiolite material, much of the lithotectonic belt
lying west of the Melones fault zone is calcalkaline-
arc debris which apparently accumulated nearly in
place relative to the magmatic belt (Behrman,
1978).
The outboard western Jurassic belt is, in gen-
eral, bounded on the east by the Big BendiBear
Mountains fault system and on the west by on-
lapping strata of the Great Valley Group. Locally
along the fault zone, a narrow belt of Jurassic slates
rests with angular unconformity on Calaveras-
Melones basement. Two lithotectonic entities make
up much of the western Jurassic belt: the Mariposa
The post-Archean Western Cordillera, California 155
Slate, including interlayered calcalkaline volcano-
genic metasedimentary and metavolcanic rocks; and
the intra-arc Smartville ophiolite, a mafic (+ ultra-
mafic) complex overlain by weakly metamorphosed
basalts, andesites, and dacites (Xenophontos and
Bond, 1978; Beard and Day, 1987). The smaller
Pine Hill pluton + dike swarm occurs farther south
in a similar geologic setting. Sedimentary proto-
liths were laid down during later Jurassic time but
are no younger than Oxfordian-Kimmeridgian; the
ophiolite complexes have crystallization ages of
"'" 160 Ma (Saleeby, 1982). The entire belt was per-
vasively but weakly chloritized, with metamorphic
grade confined to the prehnite-pumpellyite and
lower greenschist facies. According to Behrman and
Parkison (1978), it formed through the accumula-
tion of detritus from nearby calcalkaline volcanism
on pre-existing units of the more easterly Calaveras
and northern Sierra Nevada and southern Klamath
provinces.
Northern and Central Franciscan-Great Valley
Complex
Franciscan and Great Valley units are are ally
associated from southern Oregon to west-central
Baja California. Belts consist dominantly of clastic
sedimentary rocks, recrystallized to contrasting
extents, and each has a latest Jurassic-Paleogene
depositional range. The great ensimatic, tectoni-
cally imbricated prism of partly chaotic (Hsl1, 1968,
1971), chiefly east-dipping and east-facing Francis-
can strata, lies closest to the Pacific Ocean and is
confined to the Coast Range province; it is separated
from the broadly contemporaneous Klamath/
Sierran igneous arc by the well-bedded Great Valley
Group. The two terrigenous sedimentary belts have
been regarded as trench and forearc-basin deposits,
respectively (Ernst, 1970; Dickinson, 1972,1976).
However, paleomagnetic and fossil data docu-
ment the far-traveled nature of contained oceanic
units, especially deep-sea cherts and pillow basalts
(e.g., Blake (ed.), 1984; Beck, 1986), mainly of Late
Jurassic and younger age. These relationships
evoked the hypothesis of juxtaposition of unrelated
Franciscan and Great Valley terranes (Blake and
Jones, 1978, 1981; Blake and others, 1988), and in
general, the accretion of exotic terranes throughout
the Western Cordillera (Coney and others, 1980;
Jones and others, 1983). The present intimate spa-
tial association of the two metasedimentary assem-
blages in the California Coast Ranges reflects
relative westward thrusting of Great Valley strata
and underlying oceanic basement (the Coast Range
ophiolite) over the higher pressure but coeval Fran-
ciscan Complex (Bailey and others, 1964, 1970),
followed by, or concurrent with, some sort of exten-
sional return flow for the Franciscan (Suppe, 1972;
Cloos, 1982; Platt, 1986). Seismic reflection and
refraction profiles indicate that the Franciscan may
have behaved as a tectonic indentor, wedging east-
ward between overlying Great Valley strata and the
Sierran basement (Wentworth and others, 1984);
but inasmuch as Quaternary units are involved, this
structure, if real, may represent chiefly Neogene
deformation.
Recognizing that most pillow lavas and cherts
of oceanic character are clearly exotic, geologic
relations nevertheless suggest that the volumetri-
cally dominant clastic rocks of western California
are autochthonous or parautochthonous in relation
to the evolving continental margin. Studies of
paleocurrent vectors and sandstone + conglomerate
petrofacies demonstrate a common Klamath/Sier-
ran provenance for Franciscan and Great Valley
detritus (Jacobson, 1978; Dickinson and others,
1982; Ingersoll, 1983; Seiders and Blome, 1984,
1988). Similarities of quartz, feldspar, and lithic-
fragment proportions within the forearc basin and
trench complex are illustrated in Figure 8. Slight
but systematic contrasts in quartz contents may be a
reflection of sedimentary concentration during wea-
thering and transport of the terrigenous debris.
Figure 9 shows inferred paleogeography and
sediment-distribution trajectories (Dickinson and
others, 1982). Spatial contiguity seems to be re-
quired during deposition of these units. Moreover,
Upper Cretaceous Great Valley trench-slope units
locally rest with angular unconformity on the
underlying Franciscan accretionary prism (Max-
well, 1974; Smith and others, 1979). Evidently
forearc debris was in intimate proximity to the
rising mass of decoupled trench melange during
deposition of the former.
Four major lithotectonic belts (divided by some
workers into more numerous tectonostratigraphic
terranes) crop out in the northern Coast Ranges
and, on a smaller scale, in the southern Coast
Ranges (Blake and others, 1984, 1985, 1988). From
east to west, these are the Great Valley Group, the
eastern Franciscan belt, the central Franciscan
melange belt, and the coastal belt of Franciscan.
156 W.G.Ernst
quartz
Great Valley clast averages
(e) uppennost K
(d) Upper K
(c) basal Upper K
(b) Lower K
(a) Upper Ju-Iowermost K
Figure 8. Modal proportions of clastic quartz, feldspars,
and lithic fragments from 232 Great Valley and 203 Fran-
ciscan sandstones from northern and central California,
compared with typical magmatic-arc derived sandstones,
after Dickinson and others (1982). Evolution in average
Great Valley modes during unroofing ofvolcanic/metamor-
phic cover and exposure ofK-feldspar-bearing granitoids is
illustrated by the following petro stratigraphic intervals:
(a) Upper Jurassic-lowermost Cretaceous; (b) Lower Creta-
ceous; (c) basal Upper Cretaceous; (d) Upper Cretaceous;
and (e) uppermost Cretaceous.
Great VaUey Group. The well-bedded Great
Valley clastic prism floors both Sacramento and San
Joaquin Valleys, constituting an asymmetric syn-
clinorium with a near-vertical western limb and
gently west-dipping eastern limb (Hackel, 1966).
The assemblage is transgressive onto the North
American continent; thus, although uppermost
Jurassic deep-water turbidites on the west lie
conformably on oceanic crust (the Coast Range
ophiolite), Upper Cretaceous shallow-marine clastic
rocks on the east side of the Great Valley rest un-
conformably on the granitoid-invaded western
Sierran foothills (Ingersoll, 1978, 1979). The clastic
section is feebly metamorphosed to zeolite-facies
assemblages near the base of its thickest, westerly
sections (Dickinson and others, 1969; Bailey and
Jones, 1973).
The Coast Range ophiolite, which crops out
along the west side of the Great Valley, includes
variably serpentinized peridotite tectonite, over-
lying mafic + ultramafic plutonic rocks, minor pla-
t22I Franciscan
Complex
Modoc
Plateau
Figure 9. Paleogeography and sediment distribution
paths in northern and central California during Cre-
taceous time, after Dickinson and others (1982) (see also
Seiders, 1988). The major suture zone shown as a thrust
contact (barbs on upper plate) is the so-called Coast Range
thrust, which experienced compound movement, including
earlier subduction (underflow and compression) and later
uplift (return flow and/or extension) as documented by
Bailey and others (1970), Ernst (1970), Cloos (1982), Platt
(1986), and Jayko and others (1987).
giogranite, sheeted mafic dikes and sills, and a sur-
mounting group of massive and pillowed tholeiitic
basalts + breccias (Bailey and others, 1970; Bailey
and Blake, 1974). It probably includes several
rather different oceanic units. Some were generated
"" 165 Ma at an oceanic-spreading center (Hopson
and others, 1981), while others locally are associ-
ated with slightly younger ("" 153 Ma) calcalkaline
arc rocks (Evarts, 1977). Most segments of the
ophiolite, particularly in the southern part of the
Coast Ranges, are capped by Upper Jurassic deep-
sea radiolarian cherts that pass upward to rhyth-
mically layered, thin-bedded Great Valley distal
mudstones. Occurrences of the ophiolite inter-
layered with magmatic-arc rocks may have been
generated nearby (Sharp and Evarts, 1982), where-
as other segments, having remnant paleomagnetic
The post-Archean Western Cordillera, California 157
inclinations compatible with near-equatorial forma-
tion (Luyendyk, 1982), probably are exotic to the
Californian continental margin. Such paleotectonic
contrasts are also reflected in geochemical differ-
ences among geographically disparate mafic units of
the Coast Range ophiolite (Shervais and Kim-
brough,1985).
Eastern Franciscan Belt. The eastern Fran-
ciscan belt and its southern extension in the Diablo
Range consist of a relatively coherent series of
tectonically imbricated phyllitic schists, quartzo-
feldspathic metagraywackes, dark metashales,
lenses of greenstone + serpentinite, and widespread
but volumetrically minor chert layers. The complex
is bounded to the east by the Coast Range fault and
the structurally overlying ophiolite + Great Valley
Group and to the west by a fault juxtaposing the
tectonically lower central Franciscan melange belt.
Imbricate structures within the eastern belt, as well
as bounding faults, dip eastward (Suppe, 1973).
Depositional ages of turbiditic rocks of this belt
range from latest Jurassic to mid-Cretaceous.
Rocks of the eastern belt are characterized by
the growth of neoblastic lawsonite throughout.
Regional metamorphism has generated several dis-
tinctly different lithologic assemblages in portions
of this terrane:
Mafic tectonic blocks of high-grade garnet-
bearing blueschist, amphibolite and eclogite occur
as isolated, scattered fragments in a chaotic me-
lange. Actinolite + chlorite or talc rinds marginal
to the blocks indicate that the latter were not in
chemical equilibrium with the enclosing lower
metamorphic grade matrix of the olistostrome, tec-
tonic melange or serpentinite host. Recrystalliza-
tion ages of these high-grade blocks typically are on
the order of "" 160 Ma (Coleman and Lanphere,
1971; Mattinson, 1988) - older than the sparsely
fossiliferous matrix in which they are found.
Interlayered schistose metapelitic and meta-
basaltic blueschists occur in the northeastern part of
the Franciscan Complex directly beneath the Coast
Range fault (Blake and others, 1967, 1969; Irwin
and others, 1974; Brown and Ghent, 1983). This
entity, the South Fork Mountain Schist, is the
highest structural unit in a stack of imbricate, east-
dipping sheets (Worrall, 1981). The minimum
metamorphic K/Ar age of these schists is "" 120 Ma
(Lanphere and others, 1978).
Relatively undeformed metagraywacke + meta-
shale sequences, especially extensive along the east
side of the Diablo Range, contain widespread neo-
blastic aragonite and associated jadeitic clinopyro-
xene+quartz (McKee, 1962a,b; Ernst, 1971a). The
time of this very high-P, low-T recrystallization
seems to have been ""90-120 Ma (Ernst, 1971b;
Suppe and Armstrong, 1972; Bostick, 1974; Mattin-
son and Echeverria, 1980).
Central Franciscan Belt. The central Fran-
ciscan belt consists predominantly of chaotic me-
lange, with variable proportions of more coherent
strata. It also contains the most abundantly dis-
persed high-grade mafic tectonic blocks (eclogites,
garnet amphibolites, and glaucophane schists) and
serpentinized peridotite lenses within the Fran-
ciscan Complex. The sedimentary age of the shaly
melange matrix appears to be chiefly mid- and Late
Cretaceous, although included blocks and broken-
formation slabs range in age from Late Jurassic to
Late Cretaceous and are of both foreign and cognate
lithologies (Blake and Jones, 1981). Fractured and
attenuated blocks may be regarded as boudins,
whereas with lesser degrees of disruption, tectonic
melange passes gradually into semistratified bro-
ken formation (Hsii, 1968,1974). The lower the
sandstone/shale ratio, the more thoroughly disrup-
ted the sections appear to be. Laminar flow within
the Franciscan subduction zone was modeled by
Cowan and Silling (1978) and Cloos (1982). Foreign
blocks evidently were spalled off from different loci
along boundaries of the subduction-induced me-
lange circulation system; this phenomenon, coupled
with differential settling velocities of dense blocks
during flow, accounts for both the lithologic varia-
tion and the size distribution of the tectonic
fragments (Cloos, 1982,1986).
Metamorphism of the central Franciscan belt
produced pumpellyite and lawsonite but not jadeitic
pyroxene (Blake and others, 1967; Suppe, 1973;
Cloos, 1983). Moderate pressures and low tem-
peratures are indicated. The recrystallization age of
the matrix has not been determined but is probably
Late Cretaceous, judging by the contained mid-
Cretaceous fossils and pre-existing, high-rank tec-
tonic blocks.
Exotic pelagic limestone blocks occur in tec-
tonized melange near Laytonville, California, and
elsewhere along the western margin of the central
belt (Bailey and others, 1964; Gucwa, 1975).
Remanent-magnetic and abundant microfossil data
(Alvarez and others, 1980; Tarduno and others,
1986) demonstrate that some of these carbonates
158 W. G.Ernst
formed ==88-101 Ma ago at latitude 14-17S. The
origin in mid-Cretaceous time was probably as deep-
water biogenic limestones fringing seamounts. The
limestone blocks evidently were supplied to the
Franciscan lithotectonic environment by rapid
northeastward translation, which brought oceanic
crust and bathyal sediments to the Late Cretaceous
trench (Sliter and others, 1986; Debiche and others,
1987). The allochthonous nature of such pelagic
limestone scraps provides few, if any, constraints
regarding provenance of the voluminous Franciscan
clastic wedge, however.
Coastal Franciscan Belt. The coastal Fran-
ciscan belt is bounded on the east by the apparently
east-dipping coastal belt thrust, which separates it
from the structurally higher central Franciscan
melange belt. On the west, the coastal belt con-
stitutes the North American continental margin
and slope and is bounded by strands of the San
Andreas fault system. Stratal continuity is charac-
teristic of this rather arkosic lithotectonic unit.
Sediments were deposited during Late Cretaceous to
Miocene time (McLaughlin and others, 1982). Pre-
metamorphic rock types include deep-water strata
on the west, apparently associated with an altered
basaltic substrate and chert; to the east, this section
passes by degrees to mid-fan turbiditic, andesitic
graywacke and abundant quartzofeldspathic gray-
wacke (Bachman, 1978; Blake and Jones, 1981).
Greenstone and serpentinite lenses are scarce, and
high-grade tectonic blocks are extremely rare in this
belt (Blake and others, 1988).
Metamorphism in the coastal belt of the Fran-
ciscan is weak and has not been investigated in
detail. Laumontite seems to be widely distributed
as a vein mineral. A few metasedimentary para-
geneses include prehnite and/or pumpellyite, but
blueschists have not been reported (Bailey and
others, 1964; McLaughlin and others, 1982). Phase
assemblages are similar to those described from the
most deeply buried portions of the Great Valley
Group. This seaward belt of the Franciscan appar-
ently was not subjected to profound underflow prior
to decoupling from the subducting plate and under-
plating along the continental margin.
Continental Accretion and Metamorphism
As evident from the above description - and as
summarized in Figures 4 and 5 and especially Fig-
ures 6 and 7 - two major sequences of sialic
growth and metamorphic recrystallization are ex-
hibited by lithotectonic belts of northern and central
California:
1) Evidence for late Paleozoic-early Mesozoic ac-
cretion is preserved in the Sierran foothills, but
much of the earlier, progressively westward addi-
tion of new crust was overprinted and obscured by
the thermotectonic regime established accompany-
ing the Jura-Cretaceous batholithic invasion and
associated volcanism.
2) To the west lie the late Mesozoic Great Valley
and Franciscan terranes, themselves representing
reworked, transported, and recrystallized products
of the contemporaneous, more easterly magmatic
arc.
Allochthonous additions to both lithotectonic
sequences are minor in amount and chiefly oceanic
in nature. In contrast, the voluminous terrigenous
wedges represent the erosion, deposition, deforma-
tion, and recrystallization of intermediate and felsic
calcalkaline igneous contributions to the continen-
tal crust. Thus, the primary additions of material to
the North American crust were sited in the
volcanic/plutonic arc, with sedimentary/metamor-
phic reworking taking place in the outboard forearc
and trench environments.
Through both earlier and later constructional
events, the ages of deposition of the superjacent
volcanics + sediments, and intensities + ages of
metamorphism, generally decrease seaward (Ernst,
1983). This overall systematic, oceanward progres-
sion reflects growth at a convergent plate boundary,
but would not be readily explicable if geologically
unrelated, exotic, older and younger continental and
oceanic terranes had been stranded and recrys-
tallized in random fashion along the evolving sialic
margin.
JURASSIC AND YOUNGER SUBDUCTION,
TRANSFORM MOTION AND THE LITHO-
LOGIC RECORD OF THE WESTERNMOST
U.S. CORDILLERA
Post-Middle Jurassic sea-floor spreading is re-
corded in oceanic crust and overlying hemipelagic
sediments of the Pacific Basin (Pitman and others,
1974; Engebretson and others, 1985; Debiche and
others, 1987). Whereas parts of western limbs of
several oceanic plates are still extant, eastern limbs
of these paleo-Pacific spreading systems have been
The post-Archean Western Cordillera, California 159
overridden by the North American plate over the
past 165 m.y. (Hamilton, 1969). More than 10,000
km of eastward subduction must be accounted for
(Ernst, 1984; Engebretson and others, 1988), aver-
aging nearly 6 cm/year. Thus, in spite of several
thousand kilometers of northward drift of oceanic
crust-capped lithosphere relative to North America,
underflow must have been the dominant mechanism
responsible for production of paired metamorphic
belts and voluminous calcalkaline igneous activity,
reflecting Cordilleran continental growth during
mid-Mesozoic to mid-Cenozoic time. Furthermore,
true accretion (increase in the total mass of terres-
trial sial) requires the addition of new calcalkaline
igneous material to the continents, rather than the
rearrangement of pre-existing crustal fragments.
On the basis of measured high heat flow in the
blueschist-bearing California Coast Ranges and
reasonable mineralogic transformation rates, Cloos
and Dumitru (1987) calculated that rapid thermal
obliteration of high-P metamorphic mineral assem-
blages is currently in progress at depths greater
than 5-10 km. The current strike-slip regime of
western California, therefore, evidently has been
recently imposed on a chiefly convergent Mesozoic-
Cenozoic margin (Atwater, 1970), where the process
of subduction until recently has sustained the
nearly continuous refrigeration and preservation of
blueschist lithologies since their late Mesozoic
formation. The same general conclusion was drawn
by Peacock (1988) for retained inverted metamor-
phic gradients in the westernmost Cordillera. Later
terrane shuffling along the dextral shear system of
the California Coast Ranges, therefore, has partly
obscured the effects of the main constructional stage
of sialic growth accompanying oblique or head-on
lithospheric plate descent.
Both the eastward underflow of great tracts of
paleo-Pacific oceanic crust and the apparent recent
change in thermal structure of the westernmost con-
tinental crust argue for long-continued subduction
as the chief plate-tectonic process operating in the
Western Cordillera during mid-Mesozoic to mid-
Cenozoic time. Such a process explains the contem-
poraneity, spatial association, and contrasting P-T
histories of landward calcalkaline arcs, forearc-
basin deposits, and oceanward trench complexes -
not only in the western conterminous U.S. but
around the modern Pacific rim in general. The total
absence of old blueschist belts within the interior of
the U.S. Cordillera may be accounted for by thermal
overprinting. The scarcity of negatively buoyant
ophiolitic peridotites within the same region may
reflect systematic downward sagging of dense
mantle + mafic crustal material accompanying
thermal softening and crustal remobilization. Of
course, it is also possible that ancient plate-tectonic
regimes were sufficiently different from modern
analogs, especially in terms of crustal temperature,
structure and mantle chemistry, that blueschists
and ophiolites were produced less abundantly than
currently.
ISOTOPIC PROVINCES AND TRUE CRUSTAL
GROWTH OF THE WESTERN CORDILLERA
Tectonometamorphic trends in the western
United States, concentrating on the Phanerozoic
evolution of portions of northern and central Cali-
fornia, have been reviewed thus far. This section
presents a summary of isotopic data for basement
rocks of the continental interior. This information
helps to illuminate crustal growth, and the petro-
genesis of Precambrian metamorphic belts, pre-
served in a more fragmentary fashion in the
Cordillera than younger analogues.
Magmas, derived from partial fusion of deep-
crustal and upper-mantle source materials, provide
geochemical constraints on the nature of the base-
ment. Mesozoic and Cenozoic igneous rocks of the
western U.S. reflect a continental lithosphere typi-
fied by discrete Pb and Nd isotopic provinces
(Zartman, 1974; Farmer and DePaolo, 1983,1984;
Wooden and others, 1988); these igneous units
exhibit variable contributions from mantle and
crustal protoliths. The age of separation of the deep
continental crust from an evolving mantle reservoir,
as indicated by Precambrian Pb and Nd model and
crystallization ages, monotonically decreases from
the Archean Wyoming province through the Early
Proterozoic central Great Basin + Mojave Desert to
the Middle Proterozoic Sonoran Desert. Within the
Mesozoic calcalkaline belts, a systematic oceanward
decrease in the degree of continental involvement,
as expressed by decreasing 87Sr/86Sr initial ratios,
and increasing eNd values for volcanics and plu-
tonics, is well documented (Kistler and Peterman,
1978; DePaolo, 1980,1981; Farmer and DePaolo,
1984).
The Sr/Rb ages of arc-magma emplacement and
country-rock recrystallization provide evidence con-
cerning the time, or times, of stabilization of an
orogenic belt (Armstrong and others, 1977; Condie,
160 W.G.Ernst
1981, 1986; Bickford, 1988). Pre-existing thermal
events generally are imperfectly preserved in relict
assemblages and are indicated by disturbed isotopic
systematics.
the Paleozoic and especially the Mesozoic, when
the continent grew westward.
The age of separation from the mantle of materials
constituting the Precambrian basement gradually
decreases from the Wyoming Province (>2.6 Ga),
through the central Great Basin + Mojave Desert
(2.0-2.3 Ga) and the Colorado Plateau (1.8-2.0 Ga),
to the Sonoran Desert (1.7-1.8 Ga), and the mid-
continent granite-rhyolite belt 1.4 Ga). Chrono-
logic relationships are also mirrored by geochemical
trends. Lead isotope studies of Mesozoic and Ceno-
zoic igneous rocks (Zartman, 1974; Wooden and
others, 1988) reflect a continental lithosphere typi-
fied by distinctly different geochemical realms. Nd
isotopic data from crustally derived magmatic rocks
Accretion of the western U .8. continental crust
occurred principally during three major time inter-
vals, through the primary formation of calc alkaline
batholiths and related volcanogenic arcs (collec-
tively referred to as magmatic arcs):
the later Archean (2.5-3.3 Ga), when the con-
tinental nucleus was assembled;
the Early and Middle Proterozoic (1.4-2.3, but
chiefly 1.7-1.9 Ga), when a major part of the
sialic basement was generated progressively
southward; and
-.. - - : '- -', -,-'- _ .. - .. _. _ - .. - -_.- _ .. _ .. - ',- - "-
r \
'- I ! 20-230. I?)
...... I "
""... j "1 \ _ -- "
( t -_-
C!\
\. r --.- - ' -'-'-- -- '\ \
'.J_J . \
i I I
I >260. \.- L-._.--
I ;
I X _
I .t-r-
I
i 1 __ -./ __ .-1-.- '- ',
--,;,- - \
23 Cia : Sffl ...... 'f -
20- I Lok. __ -' ! \
I \
""-1 i \
/ i ,.1-.200. I _...\-
\ I
\ I ...-/- _._._,_1-
\ ____ - - - _-7.L- ---- t--
\ I /./ \
, // 1 i
". \ \ ,...,.../ 1 \1-180. \
'I 1./ \ \
r- \ I
I i
/ ' i i
0_ I ' I
{ I i _-1
;
u i '/ ___ .-J
$.1'100'1 s.. '- " - \ t--=- -
"- - ' __ .. _ .L ..... "
Figure 10. Isotopic provinces of the western conterminous United States: Nd-depleted mantle model ages (Farmer and
DePaolo, 1984; Farmer, 1988; Bennett and DePaolo, 1987), and 87Sr/86Sr (= 0.706 and 0.704) initial ratio limits in Mesozoic
and Cenozoic granitoids (Armstrong and others, 1977; Kistler and Peterman, 1978). Mid-Proterozoic and older basement lies
inboard of the 0.706 line.
The post-Archean Western Cordillera, California 161
of Precambrian to Tertiary age further define these
contrasting isotopic provinces (Bennett and De
Paolo, 1987).
Geochemical provinces are illustrated in Fig-
ures 10 and 11. These isotopic belts border the
Archean cratonal nucleus in Wyoming and exhibit
overall Proterozoic and Phanerozoic enlargement
towards the modern continental margins. Relation-
ships hold for model mantle separation ages as well
as times of crustal stabilization (igneous emplace-
ment and regional metamorphism). Lithotectonic
belts apparently developed asymmetrically and
episodically rather than concentrically and con-
tinuously, indicating sequential zonal growth
rather than gradual circumferential enlargement.
Major truncations of these belts also suggest the
probability of occasional rifting and removal of pre-
viously accreted segments of the North American
continental crust (Burchfiel and Davis, 1975).
Conclusions regarding Precambrian crustal
evolution must be regarded as tentative because of a
limited data base. However, broad trends are com-
patible with an overall growth of the Western
Cordillera mainly by the generation and telescoping
of new continental crust surmounting convergent
plate junctions, with the incorporation of variable
amounts of reworked old sialic materials, and by the
random accretion of volumetrically minor, exotic,
mainly oceanic terranes of unrelated geochemistry.
-T--- -f! -i- ---- --- --- ----C -.--
II , i
\ r 1,\ \
\ , \ "1 I
l __ ..... __ -----.-\ ( \-._._-'-'-
\; I
, r - -.- - _._._.- - -
;" \ "\
S \ J_J I
I \, \ /
I it\ 10 I > 25 Go c:
-'--'1- l __ J _ -'3
1
_ __ _ _ I 4 -'-
I r .. -.... --.- .... ..
I I ,/ I <\. :;1 ./.. __ ...1 -'-'-'1
III I i r;: '{.. "\j - i
I i II I' ,\ - _. '-... /... .. -..1/. t"'-
I i of i / : ..... 1 i
I ! I: V . i
'" ::, /"-'- I \
/_.... I. ,,,8Ga i
1;-- ---1 ... ' 1/ Ib I i
\ . "., . L.
\ ........... < .. 2G-23G'>I?/ ' '. _____ ._._ -'- -;::=._
p, ,
i t ,I I
I
ii " r' ,,' 1
,i' ' ./'. i
i!........... 1 : ./' \
-0"" -...... I', , .,/'" ,
...... .( .r-- ' \
II . .,/" '311 CIa
', "'1, / I
- '-..., / ' - _.- -.- _._--'
-_ .. J 1 ,. - - '-
.......... ___ L--.... \", '-
Figure n. Isotopic provinces of the western conterminous United States: Pb isotope provinces (Zartman, 1974; Wooden and
others, 1988) and areas characterized by radiometric, chiefly Rb/Sr and UlPb, bulk-rock ages of magmatic crystallization,
modified after Condie (1981, 1986) and Bickford (1988); regions I, II and Ill, separated by dotted lines, are typified by lead from
post-Paleozoic igneous rocks occurring in regions principally of Precambrian basement, derivative sedimentary strata, and
young, eugeoclinal mixed provenance, respectively (Zartman, 1974).
162 W. G. Ernst
Because of the observed monotonic decrease in
ages and the coherence of isotopic data for the belts
proceeding outward from the >2.6 Ga Wyoming
craton, it is evident that addition of far-traveled
microcontinental terranes consisting of sialic base-
ment detectably older than the developing native
continental margin did not characterize growth of
the western conterminous U.S. Probable exceptions
to this generalization include: (1) the Salin ian con-
tinental salient west of the San Andreas fault (Fig-
ure 10), although the magnitude of northward drift
since rifting of this granitic fragment is uncertain
(Hill and Dibblee, 1953; Page, 1982); (2) eastern-
most central California, where a westward step in
the 87Sr/86Sr initial ratio, offsetting the late Precam-
brian continental margin (Figure 10) may mark a
faulted terrane boundary (Kistler and Peterman,
1978); and (3) the Yellow Aster Complex (Misch,
1966) of the northwestern Cascades, Washington
(see Figure 4), an outboard fragment of old sialic
crust, the volume of which is, however, relatively
small.
SUMMARY
True continental growth resulting from the
separation of silicic, LIL, volatile, and alkali-rich
material from the mantle, in contrast to re-
arrangment of pieces of already extant, old sialic
crust, seems to require an important component of
convergent plate motion (Ernst, 1984). Phanerozoic
metamorphism (see Figures 4-7), the Precambrian
isotopic provinces and rock record (Figures 10 and
11), and the gradual oceanward decrease in initial
87Sr/86Sr ratios and increase in eNd values in con-
tinental igneous rocks (e.g., DePaolo, 1980, 1981),
are all compatible with a process involving crustal
growth and metamorphism in the Western Cordille-
ra dominantly by proximal accretion. Chiefly near
the Pacific margin of the U.S., allochthonous ophio-
litic debris is abundant and exotic Phanerozoic
microcontinental fragments of uncertain source are
present but rare. However, accretion primarily in-
volved the sweeping back into the North American
margin of native terrigenous debris and previously
metamorphosed sialic fragments of local or regional
provenance. Development of the late Mesozoic tri-
ad, Franciscan trench/Great Valley forearc/Sierran
calcalkaline arc, provides a relatively well pre-
served example of the growth and reworking pro-
cess; the parautochthonous setting is demonstrated
by provenance and transport vectors as well as by
petrochemical similarities. Lateral tectonic/meta-
morphic rearrangement of parautochthonous slices
of the margin as well as the episodic removal of
continental-margin/island-arc sections and docking
at distant sites are important complications which
cannot be ignored. Nevertheless, subduction-
related arc processes (and recycling in forearc,
backarc, and trench environments) evidently have
been responsible for most of the real enlargement of
continental crust documented for the Western Cor-
dillera.
ACKNOWLEDGMENTS
The author acknowledges the scientific help of
participants at a UCLA Rubey Colloquium "Meta-
morphism and Crustal Evolution of the Western
United States" which took place in January 1986.
That symposium, and the subsequent publication
(Ernst, 1988), provided much of the framework for
the present paper. Support was provided by UCLA
and by the U.S. Department of Energy, through
grant FG03-87ER13806. R. V. Ingersoll and V. L.
Hansen reviewed the first-draft manuscript; their
comments helped to ensure a better final version of
this paper.
REFERENCES
ALVAREZ, W., D. V. KENT, 1. P. SILVA, R. A. SCHWEICK-
ERT, and R. A. LARSON, 1980, Franciscan Complex lime-
stone deposited at 17' South paleolatitude: Geological
Society of America Bulletin, v. 91, p. 476-484.
ANDERSON, J. L., 1988, Core complexes of the Mojave-
Sonoran Desert: Conditions of plutonism, mylonitization,
and decompression; pp. 502-525 in W. G. Ernst (ed.), Meta-
morphism and Crustal Evolution of the Western Uni-
ted States (Rubey Volume 7): Prentice-Hall, Inc., Engle-
wood Cliffs, New Jersey, USA, 1153 p.
ARMSTRONG, R. L., 1982, Cordilleran metamorphic core
complexes - from Arizona to southern Canada: Annual
Review of Earth and Planetary Sciences, v. 10, p. 129-154.
__ , W. H. TAUBENECK, and P. O. HALES, 1977, Rb-Sr
and K-Ar geochronometry of Mesozoic granitic rocks and
their Sr isotopic composition, Oregon, Washington, and
Idaho: Geological Society of America Bulletin, v. 88, p. 397-
411.
The post-Archean Western Cordillera, California 163
ATWATER, T., 1970, Implications of plate tectonics for the
Cenozoic tectonic evolution of western North America:
Geological Society of America Bulletin, v. 81, p. 3513-3536.
BACHMAN, S. B., 1978, A Cretaceous and early Tertiary
subduction complex, Mendocino coast, northern California;
pp. 419-430 in D. G. Howell and K. A. McDougall (eds.),
Mesozoic Paleogeography of the Western United
States: Society of Economic Paleontologists and Minera-
logists, Pacific Coast Paleogeography Symposium 2, 573 p.
BAILEY, E. H., and D. L. JONES, 1973, Metamorphic facies
indicated by vein minerals in basal beds of the Great
Valley Sequence, northern California: U.S. Geological
Survey Journal of Research, v. 1, p. 383-385.
__ , W. P. IRWIN, and D. L. JONES, 1964, Franciscan
and Related Rocks, and Their Signifzeance in the
Geology of Western California: California Division of
Mines and Geology, Bulletin 183, 171 p.
__ , and M. C. BLAKE, Jr., 1974, Major chemical char-
acteristic of Mesozoic Coast Range ophiolite in California:
U.S. Geological Survey Journal of Research, v. 2, p. 637-
656.
__ , M. C. BLAKE, Jr., and D. L. JONES, 1970, On-land
Mesozoic oceanic crust in California Coast Ranges: U.S.
Geological Survey, Professional Paper 700-C, p. 70-81.
BEARD, J. S., and H. W. DAY, 1987, The Smartville in-
trusive complex, Sierra Nevada, California: The core of a
rifted volcanic arc: Geological Society of America Bulletin,
v. 99, p. 779-791.
BECK, M. E., 1986, Model for late Mesozoic-early Tertiary
tectonics of coastal California and western Mexico and
speculation on the origin of San Andreas fault: Tectonics,
v. 5, p. 49-64.
BEHRMAN, P. G., 1978, Pre-Callovian rocks west of the
Melones fault zone, central Sierra Nevada foothills; pp.
337-348 in D. G. Howell and K. A. McDougall (eds.),
Mesozoic Paleogeography of the Western United
States: Society of Economic Paleontologists and Minera-
logists, Pacific Coast Paleogeography Symposium 2, 573 p.
__ , and G. A. PARKISON, 1978, Paleogeographic sig-
nificance of the Callovian to Kimmeridgian strata central
Sierra Nevada foothills, California; pp. 349-360 in D. G.
Howell and K. A. McDougall (eds.), Mesozoic Paleogeo-
graphy of the Western United States: Society of Econo-
mic Paleontologists and Mineralogists, Pacific Coast Paleo-
geography Symposium 2, 573 p.
BENNETT, V. C,. and D. J. DEPAOLO, 1987, Proterozoic
crustal history of the western United Stastes as deter-
mined by Neodymium isotopic mapping: Geological Soci-
ety of America Bulletin, v. 99, p. 674-685.
BICKFORD, M. E., 1988, The accretion of Proterozoic crust
in Colorado: Igneous sedimentary, deformational, and
metamorphic history; pp. 411-430 in W. G. Ernst (ed.),
Metamorphism and Crustal Evolution of the Western
United States (Rubey Volume 7): Prentice-Hall, Inc.,
Englewood, Cliffs, New Jersey, USA, 1153 p.
BLAKE, M. C., Jr. (ed.), 1984, Franciscan Geology of
Northern CaUfornia: Pacific Section, Society of Economic
Paleontologists and Mineralogists, v. 43, 254 p.
__ , and D. L. JONES, 1978, Allochthonous terranes in
northern California? - A reinterpretation; pp. 397-400 in
D. G. Howell and K. A. McDougall (eds.), Mesozoic Paleo-
geography of the Western United States: Society of Eco-
nomic Paleontologists and Mineralogists, Pacific Coast
Paleogeography Symposium 2, 573 p.
__ , and D. L. JONES, 1981, The Franciscan assemblage
and related rocks in northern California: A reinterpreta-
tion; pp. 307-328 in W. G. Ernst (ed.), The Geotectonic
Development of California (Rubey Volume 1): Pren-
tice-Hall, Inc., Englewood Cliffs, New Jersey, USA, 706 p.
__ , D. G. HOWELL, and A. S. JAYKO, 1984, Tectono-
stratigraphic terranes of the San Francisco Bay region; pp.
5-22 in M. C. Blake, Jr. (ed.), Franciscan Geology of
Northern California: Society of Economic Paleontologists
and Mineralogists (Pacific Section), v. 43, 254 p.
__ , W. P. IRWIN, and R. G. COLEMAN, 1967, Upside-
down metamorphic zonation, blueschist facies, along a
regional thrust in California and Oregon: U.S. Geological
Survey, Professional Paper 575-C, p. 1-9.
__ , W. P. IRWIN, and R. G. COLEMAN, 1969, Blueschist-
facies metamorphism related to regional thrust faulting:
Tectonophysics, v. 8, p. 237-246.
__ , A. S. JAYKO, and R. J. McLAUGHLIN, 1985, Tec-
tonostratigraphic terranes of the northern Coast Ranges,
California; pp. 159-171 in D. G. Howell (ed.), Tectono-
stratigraphic Terranes of the Circum-Pacific Region:
Circum-Pacific Council on Energy and Mineral Resources,
Earth Science Series 1, 581 p.
A. S. JAYKO, R. J. McLAUGHLIN, and M. B. UNDER-
1988, Metamorphic and tectonic evolution of the
Franciscan Complex, northern California; pp. 1035-1060 in
W. G. Ernst (ed.), Metamorphism and Crustal Evolution
of the Western United States (Rubey Volume 7): Pren-
tice-Hall, Inc., Englewood Cliffs, New Jersey, USA, 1153 p.
BOSTICK, N. H., 1974, Phytoclasts as indicators of thermal
metamorphism, Franciscan assemblage and Great Valley
sequence (Upper Mesozoic), California; pp. 1-17 in R. R.
Dutcher, P. A. Hacquebard, J. M. Schops, and J. A. Simon
(eds.), Carbonaceous Materials as Indicators of Meta-
morphism: Geological Society of America, Special Paper
153,108p.
164 w. G. Ernst
BROWN, E. H., and E. D. GHENT, 1983, Mineralogy and
phase relations in the blueschist facies of the Black Butte
and Ball Rock areas, northern California Coast Ranges:
American Mineralogist, v. 658, p. 365-372.
BURCHFIEL, B. C., and G. A. DAVIS, 1975, Nature and
controls of Cordilleran orogenesis, western United States:
Extensions of an earlier synthesis: American Journal of
Sciences, v. 275-A, p. 363-396.
CLARK, L. D., 1964, Stratigraphy and Structure of Part
of the Western Sierra Nevada Metamorphic Belt, Cali-
fornia: u.S. Geological Survey, Professional Paper 410,70
p.
__ ,1976, Stratigraphy of the North Halfofthe West-
ern Sierra Nevada Metamorphic Belt, California: U.S.
Geological Survey, Professional Paper 923, 25 p.
CLOOS, M., 1982, Flow melanges: Numerical modeling of
geological constraints on their origin in the Franciscan
subduction complex, California: Geological Society of
America Bulletin, v. 93, p. 330-345.
__ , 1983, Comparative study of melange matrix and
metashales from the Franciscan subduction complex with
the basal Great Valley sequence, California: Journal of
Geology, v. 91, p. 291-306.
__ , 1986, Blueschists in the Franciscan Complex of
California: Petrotectonic constraints on uplift mechan-
isms; pp. 77-93 in B. W. Evans and E. H. Brown (eds.),
Blueschists and Eclogites: Geological Society of Ameri-
ca, Memoir 164, 432 p.
__ , and T. DUMITRU, 1987, Blueschist terranes in the
Franciscan complex of California: Their future character
and implications for past plate interactions (abstract):
Geological Society of America Abstracts with Programs, v.
19, p. 366.
COLEMAN, R. G., and M. A. LANPHERE, 1971, Distribu-
tion and age of high-grade blueschists, associated eclogites,
and amphibolites from Oregon and California: Geological
Society of America Bulletin, v. 82, p. 2397-2412.
CONDIE, K. C., 1981, Precambrian Rocks of the South-
western United States and Adjacent Areas of Mexico:
New Mexico Bureau of Mines Mineral Resources, Resource
Map 13, scale 1:1,000,000.
__ , 1986, Geochemistry and tectonic setting of early
Proterozoic supracrustal rocks in the southwestern United
States: Journal of Geology, v. 94, p. 845-864.
CONEY, P. J., 1980, Cordilleran metamorphic core com-
plexes: An overview; pp. 7-31 in M. D. Crittenden, Jr., P. J.
Coney, and G. H. Davis (eds.), Cordilleran Metamorphic
Core Complexes: Geological Society of America, Memoir
153,490 p.
__ , D. L. JONES, and J. W. H. MONGER, 1980, Cor-
dilleran suspect terranes: Nature, v. 288, p. 329-333.
COWAN, D. S., and R. M. SILLING, 1978, A dynamic, scaled
model of accretion at trenches and its implications for the
tectonic evolution of subduction complexes: Journal of
Geophysical Research, v. 83, p. 5389-5396.
DAVIS, G. A., 1969, Tectonic correlations, Klamath Moun-
tains and western Sierra Nevada, California: Geological
Society of America Bulletin, v. 80, p. 1095-1108.
DA Y, H. W., P. SCHIFFMAN, and E. M. MOORES, 1988,
Metamorphism and tectonics of the northern Sierra
Nevada; pp. 737-763 in W. G. Ernst (ed.), Metamorphism
and Crustal Evolution of the Western United States
(Rubey Volume 7): Prentice-Hall, Inc., Englewood Cliffs,
New Jersey, USA, 1153 p.
DEBICHE, M. G., A. COX, and D. ENGEBRETSON (eds.),
1987, The Motion of Allochthonous Terranes Across
the North Pacific Basin: Geological Society of America,
Special Paper 207,49 p.
DEPAOLO, D. J., 1980, Sources of continental crust:
Neodymium isotope evidence for the Sierra Nevada and
Peninsular Ranges: Science, v. 209, p. 684-687.
__ , 1981, A neodymium and strontium isotopic study of
the Mesozoic calc-alkaline granite batholiths of the Sierra
Nevada, and Peninsular Ranges, California: Journal of
Geophysical Research, v. 86, p.l0,470-10,488.
DICKINSON, W. R., 1972, Evidence for plate-tectonic re-
gimes in the rock record: American Journal of Science, v.
272, p. 551-576.
__ , 1976, Sedimentary basins developed during evolu-
tion of Mesozoic-Cenozoic arc-trench systems in western
North America: Canadian Journal of Earth Sciences, v.
13, p. 1268-1289.
__ , R. V. INGERSOLL, D. S. COWAN, K. P. HELMOLD,
and C. A. SUCZEK, 1982, Provenance of Franciscan gray-
wackes in coastal California: Geological Society of America
Bulletin, v. 93, p. 95-107.
__ , R. W. OJAKANGAS, and R. J. STEWART, 1969,
Burial metamorphism of the late Mesozoic Great Valley
sequence, Cache Creek, California: Geological Society of
America Bulletin, v. 80, p. 519-526.
DOUGLASS, R. C., 1967, Permian Tethyan fusilinids from
California: U.S. Geological Survey, Professional Paper
583-A, p. 7-43.
EHRENBERG, S. M., 1975, Feather River ultramafic body,
northern Sierra Nevada, California: Geological Society of
America Bulletin, v. 86, p. 1235-1243.
The post-Archean Western Cordillera, California 165
ENGEBRETSON, D. C., A. COX, and DEBICHE, M., 1988,
Estimates for the age, geometry, and amount of oceanic
lithosphere subducted along western North America since
the Jurassic [abstract]: Geological Society of America
Abstracts with Programs, v. 20, p. 158.
__ , A. COX, and R. G. GORDON (eds.), 1985, Relative
Motions Between Oceanic and Continental Plates in
the Pacific Basin: Geological Society of America, Special
Paper 206, 59 p.
ERNST, W. G., 1970, Tectonic contact between the Fran-
ciscan melange and the Great Valley sequence, crustal
expression of a late Mesozoic Benioff zone: Journal of
Geophysical Research, v. 75, p. 886-901.
__ , 1971a, Petrologic reconnaissance of Franciscan
metagraywackes from the Diablo Range, central Cali-
fornia Coast Ranges: Journal of Petrology, v. 12, p. 413-
437.
__ , 1971b, Do mineral parageneses reflect unusually
high-pressure conditions of Franciscan metamorphism?:
American Journal of Science, v. 270, p. 81-108.
__ , 1974, Metamorphism and ancient convergent con-
tinental margins; pp. 907-919 in C. A. Burk and C. L.
Drake (eds.), The Geology of Continental Margins:
Springer-Verlag, New York, New York, USA, 1009 p.
__ , 1976, Petrologic Phase Equilibria: W. H. Free-
man Co., San Francisco, California, USA, 333 p.
__ , 1983, Phanerozoic continental accretion and the
metamorphic evolution of northern and central California:
Tectonophysics, v. 100, p. 287-320.
__ , 1984, Californian blueschists, subduction, and the
significance of tectonostratigraphic terranes: Geology, v.
12, p. 436-440.
__ , (ed.), 1988, Metamorphism and Crustal Evolu-
tion of the Western United States (Rubey Volume 7):
Prentice-Hall, Inc., Englewood Cliffs, New Jersey, USA,
1153 p.
EVARTS, R. C., 1977, The geology and petrology of the Del
Puerto ophiolite, Diablo Range, central California Coast
Ranges; pp. 121-139 in R. G. Coleman and W. P. Irwin
(eds.), North American Ophiolites: Oregon Department
of Geology and Mineral Industries, Bulletin 95, 183 p.
FARMER, G. L., 1988, Isotope geochemistry of Mesozoic and
Tertiary igneous rocks in the western United States and
implications for the structure and composition of the deep
continental lithosphere; pp. 87-109 in W. G. Ernst (ed.),
Metamorphism and Crustal Evolution of the Western
United States (Rubey Volume 7): Prentice-Hall, Inc.,
Englewood Cliffs, New Jersey, USA, 1153 p.
__ , and D. J. DEPAOLO, 1983, Origin of Mesozoic and
Tertiary granite in the western United States and implica-
tions for pre-Mesozoic crustal structure, 1: Nd and Sr
isotopic studies in the geocline of the northern Great
Basin: Journal of Geophysical Research, v. 88, p. 3379-
3401.
___ , and D. J. DEPAOLO, 1984, Origin of Mesozoic and
Tertiary granite in the western United States and implica-
tions for pre-Mesozoic structure, 2: Nd and Sr isotopic
studies of unmineralized and Cu- and Mo-mineralized
granite in the Precambrian craton: Journal of Geophysical
Research, v. 89, p. 10,141-10,160.
GOTO, K., Z. SUZUKI, and H. HAMAGUCHI, 1987, Stress
distribution due to olivine-spinel phase transition in des-
cending plate and deep focus earthquakes: Journal of Geo-
physical Research, v.92, p. 13,811-13,820.
GUCWA, P. R., 1975, Middle to Late Cretaceous sedimen-
tary melange, Franciscan complex, northern California:
Geology, v. 3, p. 105-108.
HACKEL, 0., 1966, Summary of the geology of the Great
Valley; pp. 217-238 in E. H. Bailey (ed.), Geology of
Northern California: California Division of Mines and
Geology, Bulletin 190, 508 p.
HAMILTON, W., 1969, Mesozoic California and the under-
flow of the Pacific mantle: Geological Society of America
Bulletin, v. 80, p. 2409-2430.
HANNAH, J. L., and E. M. MOORES, 1986, Age relationship
and depositional environments of Paleozoic strata, north-
ern Sierra Nevada, California: Geological Society of Am-
erica Bulletin, v. 97, p. 787-797.
HARWOOD, D. S., 1988, Tectonism and metamorphism in
the northern Sierra terrane, northern California; pp. 764-
788 in W. G. Ernst (ed.), Metamorphism and Crustal
Evolution of the Western United States (Rubey Vol-
ume 7): Prentice-Hall, Inc., Englewood Cliffs, New Jersey,
USA,1153p.
HIETANEN, A., 1981, Petrologic and Structural Studies
in the Northwestern Sierra Nevada, California: U.S.
Geological Survey, Professional Paper 1226A-C, 59 p.
HILL, M. L., and T. W. DmBLEE, Jr., 1953, San Andreas,
Garlock, and Big Pine faults, California: Geological Soci-
ety of America Bulletin, v. 64, p. 443-458.
HOLDAWAY, M. J., 1971, Stability ofandalusite and silli-
manite and the aluminum silicate phase diagram: Am-
erican Journal of Sciences, v. 271, p. 97-131.
HOPSON, C. A., J. M. MATTINSON, and E. A. PESSAGNO,
Jr., 1981, Coast Range ophiolite, western California; pp.
418-510 in W. G. Ernst (ed.), The Geotectonic Develop-
ment of California (Rubey Volume 1): Prentice-Hall,
Inc., Englewood Cliffs, New Jersey, USA, 706 p.
166 w. G. Ernst
HSO, K. J., 1968, Principles of melanges and their bearing
on the Franciscan-Knoxville paradox: Geological Society of
America Bulletin, v. 79, p. 1063-1074.
__ , 1971, Franciscan melanges as a model for eugeosyn-
clinal sedimentation and underthrusting tectonics: Jour-
nal of Geophysical Research, v. 76, p. 1162-1170.
__ , 1974, Melanges and their distinction from olisto-
stromes; pp. 321-333 in R. H. Dott, Jr., and R. H. Shaver
(eds.), Modern and Ancient Geosynclinal Sedimenta-
tion: Society of Economic Paleontologists and Mineralo-
gists, Special Publication 19, 380 p.
INGERSOLL, R. V., 1978, Paleogeography and paleo-tec-
tonics of the late Mesozoic forearc basin of northern and
central California; pp. 471-482 in D. G. Howell and K. A.
McDougall (eds.), Mesozoic Paleogeography of the
Western United States: Society of Economic Paleonto-
logists and Mineralogists, Pacific Coast Paleogeography
Symposium 2, 573 p.
__ , 1979, Evolution of the Late Cretaceous forearc
basin, northern and central California: Geological Society
of America Bulletin, Part I, v. 90, p. 813-826.
__ , 1983, Petrofacies and provenance of late Mesozoic
forearc basin, northern and central California: American
Association of Petroleum Geological Bulletin, v. 67, p. 1125-
1142.
IRWIN, W. P., E. W. WOLFE, M. C. BLAKE, Jr., and G. C.
CUNNINGHAM, 1974, Geologic Map of the Pickett Peak
Quadrangle, Trinity County, California: U.S. Geologi-
cal Survey, Geological Quadrangle Map GQ-llll , scale
1:62,500.
JACOBSON, M. 1., 1978, Petrologic variations in Francis-
can sandstone from the Diablo Range, California; p.p 401-
417 in D. G. Howell and K. A McDougall (eds.), Mesozoic
Paleogeography of the Western United States: Society
of Economic Paleontologists and Mineralogists, Pacific
Coast Paleogeography Symposium 2,573 p.
JAYKO, A. S., M. C. BLAKE, Jr., and T. HARMS, 1987, At-
tenuation of the Coast Range ophiolite by extensional
faulting, and nature of the Coast Range "thrust," Cali-
fornia: Tectonics, v. 6, p. 475-488.
JENNINGS, C. W., 1977, Geologic Map of California:
California Division of Mines and Geology, Scale 1:750,000.
JONES, D. L., D. G. HOWELL, P. J. CONEY, and J. W. H.
MONGER, 1983, Recognition, character, and analysis of
tectonostratigraphic terranes in western North America;
pp. 21-35 in M. Hashimoto and S. Uyeda (eds.), Accretion
Tectonics in the Circum-Pacific Region: Terra Science
Publishing Co., Tokyo, Japan, 358 p.
KISTLER, R. W., and Z. E. PETERMAN, 1978, A Study of
Regional Variations of Initial Strontium Isotopic Com-
position of Mesozoic Granitic Rocks in California: u.S.
Geological Survey, Professional Paper 1071,17 p.
LANPHERE, M. A., M. C. BLAKE, Jr., and W. P. IRWIN,
1978, Early Cretaceous metamorphic age of the South Fork
Mountain schist in the northern Coast Ranges of Cali-
fornia: American Journal of Science, v. 278, p. 798-815.
LUYENDYK, B. P., 1982, Paleolatitude of the Point Sal
ophiolite [abstract): Geological Society of America Ab-
stracts with Programs, v. 14, p. 182.
McKEE, B., 1962a, Widespread occurrence of jadeite, law-
sonite, and glaucophane in central California: American
Journal of Science, v. 260, p. 596-610.
__ , 1962b, Aragonite in the Franciscan rocks on the
Pacheco Pass area, California: American Mineralogist, v.
47, p. 379-387.
McLAUGHLIN, R. J., S. A. KLING, R. Z. POORE, K.
McDOUGALL, and E. C. BEUTNER, 1982, Post-middle
Miocene accretion of Franciscan rocks, northwestern
California: Geological Society of America Bulletin, v. 93, p.
595-605.
MA'ITINSON, J. M., 1988, Constraints on the timing of
Franciscan metamorphism: Geochronological approaches
and their limitations; pp. 1023-1034 in W. G. Ernst (ed.),
Metamorphism and Crustal Evolution of the Western
United States (Rubey Volume 7): Prentice-Hall, Inc.,
Englewood Cliffs, New Jersey, USA, 1153 p.
__ , and ECHEVERRIA, L. M., 1980, Ortigalita Peak
gabbro, Franciscan complex: U-Pb dates of intrusion and
high-pressure-low-temperature metamorphism: Geology,
v. 8, p. 589-593.
MAXWELL, J. C., 1974, Anatomy of an orogen: Geological
Society of America Bulletin, v. 85, p. 1195-1204.
MISCH, P., 1966, Tectonic evolution of the northern Cas-
cades of Washington State; pp. 108-148 in H. C. Gunning
(ed.), Tectonic History and Mineral Deposits of the
Western Cordillera: Canadian Institute of Mining and
Metallurgy, Special Volume 8, 353 p.
MIYASHIRO, A., 1973, Metamorphism and Metamor-
phic Belts: George Allen & Unwin Ltd., London, England,
UK,492p.
MORGAN, B. A., 1976, Geology of the Chinese Camp
and Moccasin Quadrangles, Toulumne County, Cali-
fornia: u.S. Geological Survey, Miscellaneous Field Stu-
dies Map MF-840, Scale 1:24,000.
The post-Archean Western Cordillera, California 167
OXBURGH, E. R, and D. L. TURCOTTE, 1971, Origin of
paired metamorphic belts and crustal dilation in island arc
regions: Journal of Geophysical Research, v. 76, p. 1315-
1327.
PAGE, B. M., 1982, Migration of Salinian composite block,
California, and disappearance of fragments: American
Journal of Sciences, v. 282, p. 1694-1734.
PEACOCK, S. M., 1988, Inverted metamorphic gradients in
the westernmost Cordillera; pp. 953-975 in W. G. Ernst
(ed.), Metamorphism and Crustal Evolution of the
Western United States (Rubey Volume 7): Prentice-
Hall, Inc., Englewood Cliffs, New Jersey, USA, 1153 p.
PITMAN, W. C., R L. LARSON, and E. M. HERRON, 1974,
The Age of the Ocean Basins: Geological Society of Am-
erica Map Series, Horizontal Scale 1:40,000,000.
PLATT, J. P., 1986, Dynamics of orogenic wedges and the
uplift of high-pressure metamorphic rocks: Geological
Society of America Bulletin, v. 97, p. 1037-1053.
RICHARDSON, S. W., 1970, The relation between a petro-
gnetic grid, facies series and the geothermal gradient in
metamorphism: Fortschritte fur Mineralogie, v. 47, p. 65-
76.
SALEEBY, J. B., 1981, Ocean floor accretion and volcano-
plutonic arc evolution of the Mesozoic Sierra Nevada; pp.
132-181 in W. G. Ernst (ed.), The Geotectonic Develop-
ment of California (Rubey Volume 1): Prentice-Hall,
Inc., Englewood Cliffs, New Jersey, USA, 706 p.
__ , 1982, Polygenetic ophiolite belt of the California
Sierra Nevada: Geochronological and tectonostratigraphic
development: Journal of Geophysical Research, v. 87, p.
1803-1824.
__ , and E. M. MOORES, 1979, Zircon ages on northern
Sierra Nevada ophiolite remnants and some possible
regional correlations [abstract): Geological Society of Am-
ericaAbstracts with Programs, v. 11, p. 125.
SAMS, D. B., and J. B. SALEEBY, 1988, Geology and
petrotectonic significance of crystalline rocks of the
southernmost Sierra Nevada, California; pp. 865-893 in W.
G. Ernst (ed.), Metamoprhism and Crustal Evolution of
the Western United States (Rubey Volume 7): Prentice-
Hall, Inc., Englewood Cliffs, New Jersey, USA, 1153 p.
SCHWEICKERT, R A., 1981, Tectonic evolution of the
Sierra Nevada Range; pp. 87-131 in W. G. Ernst (ed.), The
Geotectonic Development of California (Rubey Vol-
ume 1): Prentice-Hall, Inc., Englewood Cliffs, New Jersey,
USA,706p.
__ , R. L. ARMSTRONG, and J. E. HARAKEL, 1980, Law-
sonite blueschist in the northern Sierra Nevada, Cali-
fornia: Geology, v. 8, p. 27-31.
__ , C. MERQUERIAN, and N. L. BOGEN, 1988,
Deformational and metamorphic history of Paleozoic and
Mesozoic basement terranes in the western Sierra Nevada
metamorphic belt; pp. 789-822 in W. G. Ernst (ed.), Meta-
morphism and Crustal Evolution of the Western
United States (Rubey Volume 7): Prentice-Hall, Inc.,
Englewood Cliffs, New Jersey, USA, 1153 p.
SEIDERS, V. M., 1988, Origin of conglomerate stratigraphy
in the Franciscan assemblage and Great Valley sequence,
northern California: Geology, v. 16, p. 783-787.
__ , and C. D. BLOME, 1984, Clast compositions of upper
Mesozoic conglomerates of the California Coast Ranges
and their tectonic significance; pp. 135-148 in M C. Blake,
Jr. (ed.), Franciscan Geology of Northern California:
Society of Economic Paleontologists and Mineralogists
(Pacific Section), v. 43, 254 p.
__ , and M. C. BLAKE, Jr., 1988, Implications of upper
Mesozoic conglomerate for suspect terrane in western Cali-
fornia and adjacent areas: Geological Society of America
Bulletin, v. 100, p. 374-391.
SHARP, W. D., 1988, Pre-Cretaceous crustal evolution in
the Sierra Nevada region, California; pp. 823-864 in W. G.
Ernst (ed.), Metamorphism and Crustal Evolution of
the Western United States (Rubey Volume 7): Prentice-
Hall, Inc., Englewood Cliffs, New Jersey, USA, 1153 p.
__ , and R. C. EVARTS, 1982, New constraints on the
environment of formation of the Coast Range ophiolite at
Del Puerto Canyon, California [abstract): Geological Soci-
ety of America Abstracts with Programs, v. 14, p. 233.
SHERVAIS, J. W., and D. L. KIMBROUGH, 1985, Geo-
chemical evidence for the tectonic setting of the Coast
Range ophiolite: A composite island arc-oceanic crust
terrane in western California: Geology, v. 13, p. 35-38.
SLITER, W. V., R J. McLAUGHLIN, G. KELLER, and W. R.
EVITT, 1986, Paleogene accretion of Upper Cretaceous
oceanic limestone in northern California: Geology, v. 14, p.
350-353.
SMITH, G. W., D. HOWELL, and R. V,. INGERSOLL, 1979,
Late Cretaceous trench-slope basins of central California:
Geology, v. 7, p. 303-306.
SNOKE, A. W., W. D. SHARP, J. E. WRIGHT, and J. B.
SALEEBY, 1982, Significance of mid-Mesozoic peridotitic to
dioritic intrusive complexes, Klamath Mountains-western
Sierra Nevada, California: Geology, v. 10, p. 160-166.
SPEAR, F. S., J. SELVERSTONE, D. HICKMONT, P. CROW-
LEY, and K. V. HODGES, 1984, P-T paths from garnet
zoning: A new technique for deciphering tectonic processes
in crystalline terranes: Geology, v. 12, p. 87-90.
168 W.G.Ernst
STANDLEE, L. A., 1978, Middle Paleozoic ophiolite in the
Melones fault zone, northern Sierra Nevada, California
[abstract]: Geological Society of America Abstracts with
Programs, v. 10, p. 148.
SUPPE, J., 1972, Interrelationships of high-pressure meta-
morphism, deformation, and sedimentation in Franciscan
tectonics, USA: Reports of 24th International Geology
Congregation (Montreal, Canada), Section 3, p. 552-559.
__ , 1973, Geology of the Leech Lake Mountain-Ball
Mountain region, California: University of California
Publications: Geological Sciences, v. 107, p. 1-82.
__ , and R. L. ARMSTRONG, 1972, Potassium-argon dat-
ing of Franciscan metamorphic rocks: American Journal
of Science, v. 272, p. 217-233.
TARDUNO, J. A., M. McWILLIAMS, W. V. SLITER, H. E.
COOK, M. C. BLAKE, Jr., and I. PREMOLI-SILVA, 1986,
Southern hemispheric origin of the Cretaceous Laytonville
Limestone of California: Science, v. 231, p. 1425-1428.
VAN DEN BEUKEL, J. and R. WORTEL, 1987, Tempera-
tures and shear stresses in the upper part of a subduction
zone: Geophysical Research Letters, v. 14, p. 1057-1060.
VARGA, R. J., and E. M. MOORES, 1981, Age, origin, and
significance of an unconformity that predates island-arc
volcanism in the northern Sierra Nevada: Geology, v. 9, p.
512-518.
WENTWORTH, D. M., M. C. BLAKE, Jr., D. L. JONES, A.
W. WALTER, and M. D. ZOBACK, 1984, Tectonic wedging
associated with emplacement of the Franciscan assem-
blage, California Coast Ranges; pp. 163-173 in M. C.
Blake, Jr. (ed.), Franciscan Geology of Northern Cali-
fornia: Society of Economic Paleontologists and Min-
eralogists (Pacific Section), v. 43, 254p.
WOODEN, J. L., J. S. STACEY, B. R. DOE, K. A. HOWARD,
and D. M. MILLER, 1988, Pb isotopic evidence for the
formation of Proterozoic crust in the southwestern United
States; pp. 69-86 in W. G. Ernst (ed.), Metamorphism and
Crustal Evolution of the Western United States (Ru-
bey Volume 7): Prentice-Hall, Inc., Englewood Cliffs,
New Jersey, USA, 1153 p.
WORRALL, D. M., 1981, Imbricate low angle faulting in
uppermost Franciscan rocks, South YoUa Bolly area,
northern California: Geological Society of America Bulle-
tin, v. 92, p. 703-709.
XENOPHONTOS, C., and G. C. BOND, 1978, Petrology, sedi-
mentation, and paleogeography of the Smartville terrane
(Jurassic) - Bearing on the genesis of the Smartville
ophiolite; pp. 291-302 in D. G. Howell and K. A. McDougall
(eds.), Mesozoic Paleogeography of the Western United
States: Society of Economic Paleontologists and Minera-
logists, Pacific Coast Paleogeography Symposium 2, 573 p.
ZARTMAN, R. E., 1974, Lead isotope provinces in the Cor-
dillera of the western United States and their geologic
significance: Economic Geology, v. 69, p. 792-805.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Mesozoic and Cenozoic intrusions and batholiths of the circum-
Pacific region as analogues of pre-Phanerozoic batholiths:
A summary
DAVID A. BREW
Department ofInterior, U.S. Geological Survey, Menlo Park, California 94025, USA
(received November 21,1988; revision accepted April 7, 1989)
INTRODUCTION
This report is a summary of part of a review
presented at the Eighth International Conference
on Basement Tectonics. The review had four main
components: (1) systematic analysis of the Mesozoic
and Cenozoic batholiths and intrusions that ring the
Pacific Basin - their modal compositions, chemical
characteristics, internal and external structural
features, and tectonic settings; (2) application of the
results of the above analysis to determine which
characteristics are common to most of the batho-
liths, which ones are uncommon, and which are
potentially diagnostic as to tectonic setting; (3) clas-
sification of the characteristics identified during the
above examination to establish a framework or
scheme appropriate for both summarizing the essen-
tial information on the circum-Pacific batholiths
and applying that information to other batholiths in
the world; and (4) actual application of the classi-
fication scheme thus developed to a few pre-
Phanerozoic batholiths to test whether it would aid
in deducing their original tectonic environments.
This summary emphasizes the last two components.
The tectonic settings of the Mesozoic and Ceno-
zoic batholiths were emphasized mainly because the
circum-Pacific regions contain a variety of batho-
liths in a variety of recognized tectonic settings;
thus any linkages between settings and batholithic
characteristics have direct transfer value to other
batholiths of other ages elsewhere in the world.
The circum-Pacific region constitutes the
Earth's greatest mobile belt and contains the largest
and most continuous belts of batholiths and intru-
sions. Figure 1 shows the distribution of groups of
intrusions and batholiths of Mesozoic and Cenozoic
ages in the circum-Pacific regions. The analysis was
facilitated by dividing the study area into eight
large segments: (1) western South America (north-
ern and southern Andes); (2) northwestern Mexico
and southwestern USA. (Peninsular Ranges);
(3) western Canada and southeastern Alaska (West-
ern Cordillera); (4) central Alaska (Alaska Range)
and the Aleutian Islands; (5) Kurile arc and north-
eastern Siberia (Russia); (6) Japan, Korea, and
eastern China; (7) southeastern Asia (Malaysia,
Philippines, and the Indonesian region in general);
and (8) New Zealand. Pertinent literature for each
segment was examined and compositional, struc-
tural, and tectonic-setting data were evaluated for
the batholiths and intrusions. Recent syntheses
(Roddick, 1983; Pitcher, 1982; Armstrong, 1988:
Bateman, 1988; Brew, 1988), together with selected
other papers and the unpublished results of my
studies in Alaska, provided the primary information
base.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 169-177. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
169
170 D.A.Brew
CONCLUSIONS
This summary presents in slightly elaborated
form the study's two main conclusions, one minor
conclusion, and a selected bibliography. These four
items are the most important outcomes to date of the
investigation.
The first conclusion is that the circum-Pacific
batholiths and intrusions share many common fea-
tures (Table 1) and that, collectively, these features
constitute a set of recognition criteria that may be
applicable to batholiths throughout the world,
including those of pre-Phanerozoic age.
Many variations exist in age relations, modal
and chemical composition, and other features of the
circum-Pacific intrusive bodies but several impor-
tant commonalities link them:
the bodies form great, linear, composite en-
tities;
they were generally intruded over tens of mil-
lions of years;
they are associated in most cases with very
long, and in some cases enigmatic, lineaments,
faults, or sutures;
some (but not all) of the lineaments, faults, and
sutures are major plate boundaries, tectono-
stratigraphic, or tectonomagmatic boundaries;
the loci of intrusion migrate, generally craton-
ward, with time;
the major-, minor-, trace-, rare-earth-elements
and the modal and isotopic compositions vary
systematically in space and time;
distinctive abundances and types of volcanic
products are associated with different types of
plutons;
compositional segmentation is common along
the length of belts of one class of batholiths
(referred to below as the continental-margin-lip
type); and
many batholiths are superimposed on signifi-
cantly older tectonomagmatic and metamorphic
belts.
The second conclusion of the study is that a
modified version of Pitcher's (1982) classification of
granitic rock types and batholiths is appropriate for
categorizing almost all the circum-Pacific batholiths
and probably other batholiths of the world. An
updated and modified version of Pitcher's (1982)
classification of batholithic types is therefore pro-
posed. The batholiths are classified in the modified
Pitcher scheme as oceanic island (On, continental-
margin-lip (CL), continental-margin-uplift (CD),
continental-collision (CC), or anorogenic/post-oro-
genic (AP) types. These terms are explained below
Table 1. Major common general features of circum-Pacific Mesozoic and Cenozoic batholiths and intrusions.
1. Great, linear, multiple, composite entities
2. Generally intruded over tens of millions of years
3. Association with very large, in some cases enigmatic, lineaments, faults, or sutures
4. Association in most cases with major plate boundaries or contrasting tectonomagmatic regime boundaries
5. Migration ofloci ofintrusion, generally continentward, with time
6. Variation of modal, minor-trace-element-chemical and isotope compositions systematically with time and
place
7. Distinctive abundances and types of volcanic products associated with different types of batholiths
8. Longitudinal compositional segmentation common in continental-margin batholiths
9. Many batholiths superimposed on significantly older tectonomagmatic belts
10. Continentward progression from oceanic-island arc granitic rocks (01) to continential-collision-related (CC)
and/or continental-margin granitic rocks (CL), to continental-uplift-related granitic rocks (CU) with, in some
cases, superimposed anorogenic or post-orogenic granitic rocks (AP) (See text and Table 2 for explanation of
these terms)
'
0

'
0

'
0
0

"

.
.
.
.
.
,

"
"
-

.
.

:


.
.
.
.
.
.


.
.


'

'
,
'


,
.

'
'
'
'
'

0
"
-
;
'


'
"
,

/
"
/
F
'

-
-

t
.
;

,
,
:
:;
r
:
i
-
-
-
:
,

'
O
<
!
&


"


.
/
{
)

j
"

.
..
.
.

_
.
.

,
.
'
'
4
0

'
.


"
1
'1

'
8
0

;

,


.
.
.

..
.
;

.
;
-
-


\
'
"

.
J
<
"
'.

,
\

:
J
.
.
r
1
>
!
I
"
9
0
.
!
.
.
"
,


.

t
.


\
:
.

J

i
'1
'f
J
J
i
f
'

'

'
'
'
'

.
.

'
.
,
.

"
.
"
,
.


"
"
"

.
.
.
.
,
.

<
'

"

.
.

"

.
J
"

"
0
0


"
"
,
.

,
.

J
o

,


,
J
9
d
t

"
'
P
J
I
l
,
.


.
.


I
I

I

i

.
.
o
n


"


.
:

;
;
.

i
!


A
g
e


;
;
.

5
-


"

O
S

N
e
o
g
e
n
e

N
:
J

N
s
f

.
.
.
.

"
'
"

'
"
"

.
.
.
.


P
a
l
e
o
g
e
n
o

P
g

"
"

I
'J
1

''9
:
1

.
"
.
,


"
"

C
l
f
l
l
a
c
C
O
U
5

K
Q

K
s
y

I
(
g

I
C
g
:
I

"
'
"

1
(
0

u

0

.
,
U

.a
s
s
c


.
.
.
,

J
g

J
g
O

.
.
.
.

j
J


T
n
.
J
S
5
i
;

.
g

1
.
.
.
.

T
I
/
'
T
i
l
"

T
"
"

'
0


-
-
-
U
-
e
S
O
I
(
H
C

M
I
g

"
"
'
Y

.
.
.
'V
'
"
"
I
>
J

I
.
"
"
,

"
'
"


"
0

.
J


:
;

N
(
p


"
I
#
J

"
'
"


'
0
0

'
-
;
.

;
-
;
"
'

',
'r


... .
i
f


Il'i
Q
c
!)

"
"
,


l
o
a


'9
<
h
"


J
4
.
J
9
I11
,lo
:
liiI
p
P
.

'
"

l
"
'
"

:
.
'

F
i
g
u
r
e

1
.

D
i
s
t
r
i
b
u
t
i
o
n
,

a
g
e
,

a
n
d

c
o
m
p
o
s
i
t
i
o
n

o
f

c
i
r
c
u
m
-
P
a
c
i
f
i
c

M
e
s
o
z
o
i
c

a
n
d

C
e
n
o
z
o
i
c

p
l
u
t
o
n
i
c

r
o
c
k
s

(
a
f
t
e
r

B
a
t
e
m
a
n

a
n
d

o
t
h
e
r
s
,

1
9
8
2
)
.


'
"

9
.

"
"
l

n


(
!
)

U
l

0

N

S
.

n

p
o

t
:
i

p
.
.

0

(
!
)

5

N

0

"
'
.

5
'

M
-
"
"
l


U
l

o


t
:
i

U
l

p
o

t
:
i

p
.
.

0
"

p
o

.
.
.
.

:
o
r


M
-
:
o
r

U
l

.
.
.
.
.
.

.
.
.
,

.
.
.
.
.
.

172 D.A.Brew
Table 2. Modified Pitcher (1982) classification of granitic batholiths and intrusions:
Their character and geological environments.
Continental-
Oceanic Island Margin,Lip
(01) (CL)
Distribution
Small, quartz Great multiple,
diorite gabbro linear batholiths;
composite plu- arrays of com-
tons; sometimes posite and caul-
zoned drons at higher
levels, migmati-
tic roots at lower
levels
Lithologies
Plagiogranite Tonalite domin-
subordinate to ant, but broad
gabbro compositional
spectum - dio-
rite to monzo-
granite - with
wide Si0
2
-range;
major association
with gabbro
Varietal and
Hornblende and Honrblende and
accessory
biotite; pyroxene biotite; magne-
minerals
tite, sphene
Inclusions and
Mafic igneous Dioritic; may be
screens
restitic, cognate,
or from separate
magma
Chemical
Tholeiitic, calc- Calcalkalic
classification
alkalic
(Irvine and
Barager, 1971)
Chemical
Subaluminous to Metaluminous
classification
metaluminous
(Shand, 1959)
and in Table 2. All five types of batholiths have
characteristic distributional, lithologic, minera-
logic, chemical, isotopic, tectonic, and other features
that are of potential use in the study of pre-
Phanerozoic batholiths. None of the types can be
Continental- Continental- Anorogenicl
Margin, Uplift Collision Post-Orogenic
(CU) (CC) (AP)
Dispersed, Multiple batho- Multiple, cen-
isolated com- liths, plutons and tered, cauldron-
plexes ofmul- sheets, less vol- complexes of
tiple plutons and uminous and relatively small
sheets; some more commonly volume
multiple linear diapiric than CL
batholiths andCUtypes
Granodiorite- Granites with Biotite granite in
granite in con- high but narrow evolving series
trasted associa- range of Si0
2
; with alkalic
tion with minor leucocratic mon- granite and
bodies of horn- zogranites pre- syenite; highly
blende, diorite dominate but contrasted felsic-
and gabbro granitoids with mafic relation-
high biotite con- ship
tent locally im-
portant
Biotite pre- Muscovite and Green biotite;
donimates; il- red biotite; il- alkali amphi-
meniteand menite, mona- boles and
magnetite zite, garnet, cor- pyroxenes in
dierite alkalic granites
Mixed dioritic Metasedimen- Cognate; also
and meta- tary rocks mafic from sepa-
sedimentary dominant rate magma
Calcalkalic Calcalkalic Calcalkalic to
alkalic to per-
alkalic
Metaluminous Peraluminous to Peraluminous to
and peralumin- strongly pera- strongly pera-
ous luminous luminous
classified with a single criterion. Some batholiths
and belts of plutons are not readily classified.
The third and minor conclusion is that although
pre-Phanerozoic batholiths and intrusions can be
classified according to the modified Pitcher scheme
Circum-Pacific Mesozoic and Cenozoic intrusions and batholiths 173
Table 2
(continued)
Continental- Continental- Continental- Anorogenicl
Oceanic Island Margin, Lip Margin, Uplift Collision Post-Orogenic
(01) (CL) (CU) (CC) (AP)
Typical initial
<0.704 <0.706 >0.705- <0.709 >0.708 0.703-0.712
87Srj86Sr ratios
Duration and
Short, sustained; Very long dura- Short sustained; Sustained, of Short-lived
kinematic set-
prekinematic tion episodic; postkinematic moderate dura- stable environ-
ting
synkinematic tion; syn- and ments
and late kine- post- kinematic
matic
Associated
Island-arc basalt Great volumes of Sometimes with Lacking in Caldera-centered
volcanism
and andesite andesite and basalt-andesite voluminous alkalic lavas
dacite plateaus volcanic equi-
valents; may
have silicic lavas
Associated
Open folding; Vertical move- Vertical and Isoclinal folding; Doming, rifting;
deformation;
burial ments, open to transcurrent regional low P, lowP& T
metamorphism
isoclinal folding; faulting; low P intermediate T,
burial to inter- in discordant and lowP & T
mediateP& T aureoles, inter-
mediate T, also
retrograde
Associated
Au, porphyry Cu Porphyry Cu, Rarely strongly Sn and W skarn, Sn, F, Nb, U, Th
metaUic min-
porphyryMo mineralized vein, and
eraUzation
replacement
Tectonic set-
Oceanic island Andean type Caledonian type Hercynotype Post-orogenic
ting given by
arc marginal con- post-closure up- continental col- and anorogenic
Pitcher (1982,
tin ental arc lift, tensional lision. Also en- situations
Table 1, Fig. 7)
tectonics; major cratonic ductile
faulting shear belts
Associated
Volcaniclastic Volcanic fillings Rapid erosion Filling of Uplift may pro-
sedimentation
aprons offault-bounded results in extensional duce molasse (?)
basins and molasse basins
shallow troughs
Selected
Bougainville, Coastal batho- Sierra Nevada White Moun- Began and New
examples
New Ireland, lith (Peru); great (California, tains (California, England batho-
(Papua New tonalite sill (SE USA); Idaho USA); Llotse and Ii ths (Australia)
Guinea); Cor- Alaska, USA and batholith (Idaho- Manaslu batho-
dillera Occi- Canada) Montana, USA) liths (Nepal)
dental (Colom-
bia)
174 D. A. Brew
and meaningful tectonic inferences made, exactly
analogous batholithic types and commonalities
probably did not develop in pre-Phanerozoic time
because of the smaller crustal plates, probably
faster spreading rates, hotter oceanic crust, and
steeper geothermal gradients. For the pre-Phanero-
zoic batholiths, I infer that:
magmatic arcs (and their batholithic roots)
formed closer to trenches,
all batholith and batholithic-component dimen-
sions were smaller,
chemical and modal segmentation were not as
prevalent,
associated metamorphism was predominantly
low pressure and high temperature, and
there was less "contamination" by mantle and
lower crustal material.
COMMON GENERAL FEATURES OF CIRCUM-
PACIFIC CENOZOIC AND MESOZOIC
BATHOLITHS
This study focused on the common features
shared by the circum-Pacific batholiths and smaller
intrusions. These common features are those cons-
idered most likely to have persisted in convergent-
plate-margin situations throughout Earth history
and thus are most likely to be valid recognition
criteria for pre-Phanerozoic batholiths. This em-
phasis necessarily understates the diversity of
features that exist in circum-Pacific regions and
downplays real and important differences between
these regions. Also, to a large extent, this overview
consists of "first approximations" of actual situa-
tions, and many known details and exceptions are
purposefully ignored.
This review of circum-Pacific Mesozoic and
Cenozoic batholiths uses my own data from the
western Cordillera region, several papers contained
in the final report of the IGCP Circum-Pacific
Plutonism Project (Roddick, 1983), and numerous
papers published since the end of that project.
Especially useful among the latter are Mukasa
(1986), Bateman (1988), and Armstrong (1988).
The study was directed at determining if signi-
ficant common features exist in the circum-Pacific
batholiths, or if differences are so great that com-
monalities are relatively unimportant. 1 found that
there are both significant commonalities (see Table
1) and significant differences. The significant differ-
ences are of three main types:
The absolute timing of pluton emplacement
varies significantly from region to region, and
the relative timing of different classes of plu-
tons varies within regions; this is interpreted to
be due to variation in subduction convergence
rates and other factors.
Chemical and modal compositions of different
classes of plutons vary both between and within
regions; this is because of the reasons given
above (variations in the subducted material and
subduction zone steepness), and because of
variations in the crust overlying the subduction
zones.
The overall complexity of the batholiths and
smaller plutons varies from region to region;
greater complexity is interpreted to be caused
mainly by two factors: the presence of accreted
terranes and the subduction of triple junctions.
CLASSIFICATION OF BATHOLITHS
It was apparent early in this study that classify-
ing batholiths of the circum-Pacific region into a
relatively small number of categories would be a
useful way of grouping important features that
could be used as recognition criteria in pre-Phanero-
zoic terranes. Several classification schemes for
batholiths, as opposed to the chemical classification
of individual rock samples, have been proposed. All
of them are simplistic, as each mobile belt is to some
degree unique (Pitcher, 1979, p. 652). The three
schemes that have received the widest attention,
and have thus been most thoroughly tested, are the
comprehensive, multifactor approach of Pitcher
(1982), the geochemical "1-" and "S-granite" ap-
proach of Chappell and White (1974) and White
(1979), and the geochemical "ilmenite-" and "mag-
netite-series" granitoid scheme of Takahashi and
others (1980). Another recently proposed scheme is
that of Ague and Brimhall (1988a,b).
The classification devised by Pitcher (1982) was
found to be useful for this study, mainly because of
the breadth of factors it includes. It is modified here
(see Table 2) to accomplish three purposes: (1) to
change the system from a geochemical orientation
to a tectonic orientation; (2) to incorporate addi-
tional factors used by Pitcher and other authors in
describing granitoids and their environments; and
(3) to clarify the names applied to different groups of
batholiths and plutons by eliminating both provin-
cial terms ("Cordilleran," "Caledonian," and "Her-
Circum-Pacific Mesozoic and Cenozoic intrusions and batholiths 175
cynian") and mixed geochemical and tectonic abbre-
viations ("M," "I," "S," and "A") (see Pitcher, 1982,
Table 1), then replacing them with a consistent set
of tectonic-setting terms.
As noted above, several factors influenced the
modification of the Pitcher (1982) classification.
Changing the names applied to the different cate-
gories of batholiths and intrusions requires the most
explanation. The names and designations Pitcher
(1982) used were a mixture of chemical abbrevia-
tions and of provincial terms that were subject to
varying interpretations, depending on the bias and
experience of the reader.
As part of the reorientation to a tectonic classi-
fication, I decided to use descriptive tectonic terms
for exactly the same groups of batholiths and
intrusions classified by Pitcher (1982). The main
thrust of the present study was tectonic, and I felt
that a consistent set of descriptive, yet general,
tectonic terms would be more understandable and
more widely applicable than the original terms
used. The terms selected are defined briefly below.
Oceanic-island (abbreviated 01) is used to de-
signate batholiths and intrusions that occur in
oceanic island arcs as the roots of volcanic arcs.
Pitcher (1982) called these "M-type."
Continental-margin lip (CL) is used to desig-
nate batholiths and groups of intrusions that
are related to the subduction of oceanic crust
and that occur close to the oceanward margins
of the overriding continental plates. They may
be emplaced in continental crust or, less com-
monly, in accreted subduction-complex rocks.
Pitcher (1982) referred to these as "1- (Cor-
dilleran) type."
Continental-margin uplift (CU) is used for
batholiths and groups of intrusions that are like
"continental-margin lip bodies," are oceanic-
crust-subduction-related, and are close to the
boundary between the subducted material and
the continent itself. The difference between
these and the CL type is that these bodies are
emplaced after subduction has ceased and the
thickened crust has been uplifted appreciably.
Pitcher (1982) called these "1- (Caledonian)
type."
Continental collision (CC) is used for batholiths
and intrusions that are continent-contine nt-
collision-related and occur within the thickened
crust of the collision zone. Pitcher (1982) re-
ferred to these as "S-type."
Anorogenic/post-orogenic (AP) is ued to desig-
nate batholiths, groups of intrusions that occur
in diverse settings wihtin cratons and along
continental margins. They are apparently not
related to subduction of oceanic crust, nor to
continent-continent collisions, and may instead
be localized by deep crustal processes and
structures. This category was termed "A-type"
by Pitcher (1982).
As shown in Table 2, the modified Pitcher
classification emphasizes tectonic setting. The sup-
porting descriptive information for each category of
batholith covers a variety of features, including geo-
chemical information of the type that is emphasized
in other classifications. Thus, the modified Pitcher
classification can be used in several ways.
A composite transect across a hypothetical
"complete" segment of a circum-Pacific batholithic
region would include (from the ocean to the con-
tinent): an accreted oceanic-island-arc (01) batho-
lith; a suture; a continental-margin-lip (CL) batho-
lith; and a continental-margin-uplift (CU) batho-
lith; a continental-collision (CC) batholith might be
present further continentward; and an anorogenic/
post-orogenic (AP) batholith might be present
anywhere in the transect. No single circum-Pacific
transect contains all these elements. The first four
of these classes (01, CL, CU, and CC) would have a
rough temporal and spatial relation to each other,
with 01 bodies occurring outboard of the others and
being the oldest. The other bodies would be pro-
gressively younger continentward. The anorogenic
bodies could be of almost any relative age and occur
both as cratonic plutons with very obscure controls
or in continental margin areas. They quite clearly
are not part of the idealized subduction- and col-
lision-related sequence of events.
ACKNOLWEDGMENTS
This study was prompted by D. L Hyndman's
conviction that the Mesozoic and Cenozoic batho-
liths and intrusions of the circum-Pacific regions
provide valuable keys to understanding older mag-
matic belts. This summary benefited greatly from
reviews by Hyndman, M. J. Bartholomew, P. C.
Bateman, and especially Bela Csejtey, Jr., and from
editing by M. L. Callas.
176 D.A.Brew
SELECTED BIBLIOGRAPHY
AGUE, J. J., and G. H. BRIMHALL, 1988a, Magmatic arc
asymmetry and distribution of anomalous plutonic belts in
the batholiths of California: Effects of assimilation, crus-
tal thickness, and depth of crystallization: Geological
Society of America Bulletin, v. 100, p. 912-927.
__ , 1988b, Regional variations in bulk chemistry, min-
eralogy, and the composition of mafic and accessory
minerals in the batholiths of Califormia: Geological
Society of America Bulletin, v. 100, p. 891-911.
ARMSTRONG, R L., 1988, Mesozoic and early Cenozoic
magmatic evoluntion of the Candian Cordillera; pp. 55-91
in S. P. Clark, Jr., B. C. Burchfiel, and J. Suppe (eds.),
Processes in Continental Lithospheric Deformation:
Geological Society of America, Special Paper 218,212 p.
ATHERTON, M. P., W. J. McCONT, L. M. SANDERSON, and
W. P. TAYLOR, 1979, Origin of granite batholiths:
Geochemical evidence; pp. 45-64 in M. P. Atherton and J.
Tarney (eds.), Origin of Granite Batholiths [Based on a
Meeting of the Geochemistry Group of the Minera-
logical Society of Great Britian and Northern Ire-
land]: Shiva Publishing Ltd, Nantwich, Cheshire, Eng-
land, UK, 133 p.
BATEMAN, P. C., 1983, A summary of critical relations in
the central part ofthe Sierra Nevade batholith, California,
U.S.A.; pp. 241-252 in J. A. Roddick (ed.), Circum-Pacific
Plutonic Terranes: Geological Society of America, Mem-
oir 159,316 p.
__ , 1988, Geologic and geophysical constraints on
models for the origin of the Sierra Nevada batholith,
California; pp. 72-86 in W. G. Ernst (ed.), Metamorphism
and Crustal Evolution of the Western United States
(Rubey Volume 7): Prentice-Hall, Inc., Edgewood Cliffs,
New Jersey, USA, 1153 p.
__ , N. A. CHILO, Y. M. PUSHCHAROVSKY, A. M.
FIRSOV, L. AGUIRRE, D. A. BREW, D. HYNDMAN, G.
GASTIL, W. PITCHER, J. COBBIN, H. DAVIES, T. NOZAWA,
N. MURAKAMI, N. YAMADA, A. WHITE, R PRICE, J.
RODDICK, O. KIM, D. LEE, S. SINGH, 1982, and L. 1.
KRASNY, A. P. MILOV, Y. A. IVANOV, and Y. B.
BENNENOV (eds.), Map of Circum-Pacific Magmatism:
North-East Interdisciplinary Scientific Research Institute
and the Far-East Scientific Centre of the USSR Academy
of Sciences, scale 1:10,000,000 with scale 1:1,500,000 in-
sertmaps.
BREW, D. A., 1988, Latest Mesozoic and Cenozoic
Igneous Rocks of Southeastern Alaska - A Synopsis:
U.S. Geological Survey, Open-File Report 88-405,29 p.
CHAPPELL, B. W., and A. J. R WHITE, 1974, Two con-
trasting granite types: Pacific Geology, v. 8, p. 63-65.
COBBING, E. J., 1974, The tectonic framework of Peru as a
setting for batholithic emplacement: Pacific Geology, v. 8,
p.63-65.
__ , W. S. PITCHER, and W. P. TAYLOR, 1977, Segments
and super units in the coastal batholith of Peru: Journal of
Geology, v. 85, p. 625-631.
DICKINSON, W. R, 1973, Widths of modern arc-travel gaps
proportional to past duration of igneous activity in asso-
ciated magmatic arcs: Journal of Geophysical Research, v.
78, no. 17, p. 3376-3389.
__ , 1975, Potash-depth (K-h) relations in continental
margin, and intra-oceanic magmatic arcs: Geology, v. 3,
no. 2, p. 53-56.
EMERY, K. P., 1977, Stratigraphy and structure of pull-
apart margins; pp. B1-B20 in Geology of Continental
Margins: American Association of Petroleum Geologists,
Research Committee and Continuing Education Com-
mittee, Short Course, 121 p.
GRIFFIN, T. J., 1983, Granitoids of the Tertiary continent-
island arc collision zone, Papua, New Guinea; pp. 61-76 in
J. A. Roddick (ed.), Circum-Pacific Plutonic Terranes:
Geological Society of America, Memoir 159,316 p.
HYNDMAN, D. L., 1983, The Idaho batholith and associa-
ted plutons, Idaho and western Montana; pp. 213-240 in J.
A. Roddick (ed.), Circum-PacirI.C Plutonic Terranes:
Geological Society of America, Memoir 159, 316 p.
IRVINE, T. N., and W. R BARAGER, 1971, A guide to the
chemical classification of the common volcanic rocks:
Canadian Journal of Earth Sciences, v. 8, no. 5, p. 523-548.
MARSHAK, R S., and D. E. KARIG, 1977, Triple junctions
as a cause for anomalously near-trench igneous activity
between the trench and volcanic arc: Geology, v. 5, p. 233-
236.
MUKASA, S. B., 1986, Zircon U-Pb ages of super units in
the Coastal Batholith, Peru: Implications for magmatic
and tectonic processes: Geological Society of America
Bulletin, v. 92, no. 2, p. 241-254.
NOZAWA, T., 1983, Felsic plutonism in Japan; pp. 105-122
in J. A. Roddick (ed.), Circum-Pacific Plutonic Ter-
ranes: Geological Society of America, Memoir 159, 316 p.
PITCHER, W. S., 1974, The Mesozoic and Cenozoic batho-
liths of Peru: Pacific Geology, v. 8, p. 51-62.
__ , 1979, The nature, ascent and emplacement of grani-
tic magmas: Journal of the Geological Society of London, v.
136, p. 628-662.
__ , 1982, Granite type and tectonic environment; pp.
19-40 in K. J. Hsu (ed.), Mountain Building Processes:
Academic Press, London, England, UK, 263 p.
Circum-Pacific Mesozoic and Cenozoic intrusions and batholiths 177
RODDICK, J. A., 1983, Geophysical review and composition
of the Coast Plutonic Complex, south of latitude 55N; pp.
195-211 in J. A. Roddick (ed.), Circum-Pacific Plutonic
Terranes: Geological Society of America, Memoir 159, 316
p.
SHAND, S. J., 1951, Eruptive Rocks: J. Wiley and Sons,
New York, N ew York, USA, 488 p.
TAKAHASHI, M., S. ARAMAKI, and S. ISHIHARA, 1980,
Magnetite-series/ilmenite series vs. I-type/S-type grani-
toids; pp. 13-28 in S. Ishihara and S. Takenouchi (eds.),
Granitic Magmatism and Related Mineralization:
Mining Geology of Japan, Special Issue 8,247 p.
TULLOCH, A. J., 1983, Granitoid rocks of New Zealand-
A brief review; pp. 5-20 in J. A. Roddick (ed.), Circum-
PaciflC Plutonic Terranes: Geological Society of Ameri-
ca, Memoir 159, 316 p.
WHITE, A. J. R., 1979, Sources of granite magmas [ab-
stract]: Geological Society of America Abstracts with Pro-
gram, v.11, no. 7, p. 539.
YAMADA, N., T. NOZAWA, Y. HAYAMA, and T. YAMADA,
1977, Mesozoic felsic igneous activity and related meta-
morphism in central Japan - From Nagoya to Toyama;
pp. 1-88 in Guidebook for Excursion 4: Geological Sur-
vey of Japan, Kawasaki, Japan, 103 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Island-arc evolution and fracture-zone tectonics in the
Mesozoic Sierra Nevada, California, and implications for
transform offset of the Sierran/Klamath convergent margins
YILDIRIM DILEK
Department of Geology and Geography, Vassar College, Poughkeepsie, NY 12601, USA
ELDRIDGE M. MOORES
Department of Geology, University of California, Davis, California 95616, USA
(received August 12, 1988; revision accepted March 1, 1989)
ABSTRACT
In the foothills of the northern Sierra Nevada of California, an early Mesozoic island-arc terrane,
composed of Jurassic plutonic and volcanic rocks of the Smartville, Slate Creek, and Lake Combie complexes,
tectonically overlies a late Paleozoic to early Mesozoic melange with a continental margin affinity.
Fragments of this N/S-trending arc terrane are bounded to the east by E-directed thrust faults that are in turn
cut by E-dipping folds and reverse faults, and to the north by the EIW-trending Big Bend fault zone. Triassic
to lower Jurassic amphibolite-grade mafic-ultramafic rocks within this fault zone form discontinuous and
dismembered exposures of an ophiolite complex that demonstrates an origin within an oceanic fracture zone.
This fracture zone is interpreted as part of a tectonic boundary at the northern edge of the early Mesozoic
island-arc terrane. Eastward thrusting and accretion of the arc complex onto the continental margin may
have occurred around Middle Jurassic time, as suggested by a 1652 Ma stitching pluton. The presence of
Upper Jurassic dikes and plutons that intrude the Smartville complex indicates a post-accretion intrusive
event and in situ rifting within the arc. This Late Jurassic intra-arc rifting and magmatism in the
northwestern Sierra Nevada was coeval with development of the Josephine back-arc basin and its ophiolite in
the western Klamath Mountains, suggesting a period of regional extension and associated magmatism in the
western United States. The regional extension was followed by a contractional episode, known as the Late
Jurassic Nevadan orogeny, during which the Upper Jurassic and older rocks in both regions were deformed.
The nature and chronology. of early Mesozoic tectonic and magmatic events in the northern Sierra Nevada are
similar to those documented in the Klamath Mountains, indicating a common tectonic history for both
regions. We suggest that a trench/trench/transform-fault plate boundary may have separated the early
Mesozoic arc-trench systems of the northern Sierran and western Klamath provinces.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 179-196. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
179
180
Y. Dilek and E. M. Moores
INTRODUCTION
Early Mesozoic volcanic arc complexes and
ophiolitic rocks constitute a significant part of the
northern Sierra Neveda and the Klamath Moun-
tains in California (Figure 1). The tectonomagma-
tic evolution of these complexes and their accretion
to the continental margin of the western United
States are controversial among Cordilleran geolo-
gists. Lithological, mineralogical, and geochemical
characteristics consistently indicate a volcanic arc
origin for these complexes (e.g., Beard and Day,
1987; Harper and Wright, 1984; Menzies and others,
1980; Saleeby and others, 1982), suggesting a
subduction-zone tectonic environment. Accretion of
these subduction-related tectonostratigraphic units
to North America was attributed to the Late
Jurassic Nevadan orogeny, long accepted as the
principal Jurassic deformation in the Sierra Nevada
and the Klamath Mountains (e.g., Day and others,
1985; Harper and Wright, 1984; Schweickert, 1981).
Models of the Nevadan orogeny were based mainly
on inferred polarity of volcanic arc complexes and
related subduction zones. Some workers proposed a
collisional model for the Late Jurassic Nevadan
orogeny, implying allochthonous origin of arc
complexes (e.g., Moores, 1972; Schweickert and
Cowan, 1975; Schweickert, 1981; Moores and Day,
1984). Others supported a non-collisional intra-
plate deformation, suggesting autochthonous deve-
lopment of arc complexes over a single and long-
lived E-dipping subduction zone (e.g., Burchfiel and
Davis, 1981; Harper and Wright, 1984; Saleeby,
1981; Wright and Fahan, 1988).
Recent field and geochronological studies in the
northern Sierra Nevada and the Klamath Moun-
tains show that some volcanic arc complexes, pre-
viously thought to be Late Jurassic in age, are as old
as Early Jurassic and that they were deformed and
accreted to the continent during a Middle Jurassic
shortening event (Bickford and Day, 1988; Dilek,
1989; Dilek and others, 1988b; Harper and Wright,
1984; Saleeby and others, 1989; Wright and Fahan,
1988). These accreted Jurassic arc complexes were
rifted apart and intruded by Late Jurassic dikes and
plutons, which were subsequently deformed and
metamorphosed during the Late Jurassic Nevadan
orogeny. These relations imply that a major sutur-
ing event in both regions may have occurred during
Middle Jurassic time, long before the alleged Neva-
dan characteristics of the Jurassic arc terrane and
its tectonic basement in the northern Sierra
Nevada, orogeny. In this paper, we describe the
areal extent and the structural and petrological
California. We then discuss the geochronologic
order of the tectonic and magmatic events in this
region. We suggest that early Mesozoic arc com-
plexes and ophiolitic rocks in both the northern
Sierra Nevada and western Klamath Mountains
may represent ensimatic arc-trench systems sepa-
rated by an oceanic fracture zone. This oceanic frac-
ture zone is interpreted to be a fossil trench/
trench/transform-fault plate boundary, the nature
and evolution of which are defined by the geometry
and polarity of the bounding arc-trench systems. In
the last section of the paper, we examine several
alternative models for the tectonic evolution of these
arc systems and the intervening fracture zone.
REGIONAL GEOLOGIC SETTING
The Sierra Nevada foothills in northern Cali-
fornia consist of, from east to west, the Paleozoic to
Mesozoic Eastern Belt, the Paleozoic Feather River
peridotite, the late Paleozoic to early Mesozoic
Central Belt, and the Jurassic Western Belt (Figure
1). The Eastern Belt includes lower Paleozoic meta-
sedimentary rocks of the Shoo Fly Complex suc-
cessively overlain by Devonian-Mississippian and
Permo-Triassic volcanic arc assemblages that collec-
tively form the basement of a Jurassic arc sequence
(D'Allura and others, 1977; Hannah and Moores,
1986; Schweickert, 1981). Predominant structures
in this composite basement include a series of E-
vergent isoclinal folds (Figure 2) and associated W-
dipping axial-plane cleavage that are cut by E-
directed thrust faults (Moores and Day, 1984).
Adjacent to the Eastern Belt is the Feather River
peridotite, which includes Paleozoic mafic, ultra-
mafic, and sedimentary rocks together with amphi-
bolite-facies metamorphic rocks and blocks of
blueschist (Day and others, 1985; Edelman, 1987;
Hietanen, 1981). The Feather River peridotite is
separated from the Central Belt by a steeply E-
dipping fault (Day and others, 1985). The Central
Belt consists of a melange terrane tectonically
intercalated with fragments of Jurassic and older
ophiolitic (e.g., Jarbo Gap ophiolite; Dilek, 1989)
and volcanoplutonic complexes (e.g., Slate Creek
and Lake Combie complexes; Day and others, 1985;
Edelman, 1987). The melange terrane includes
blocks of upper Paleozoic to lower Mesozoic sedi-
mentary rocks and mafic to ultramafic rocks in a
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada 181
42* -
39"N _
124'
I
o
o
o
m
1>
'Z
123'
I
122' W
I o 50
t
,
KM
o GRANITIC BATHOUTHS (Mltlt1I.Junu/c and youngef)
o FLVSCH DEPOSfTS IU'. Jut8nIC}
OPHIOUTE COMPI..EJeES (u,. Jurns/c)
o ISLAND ARC COMPLEXES (EBIy,O Mlddl. Jur.uk)
.',. CONTINENTAL (1) ARC COMPLEXES
G:J OCEANIC BASEMiENT TERRANES (Trl.ule?)
D ACCRETIONARY MELANGE lu. PaJHuolc , ' . IrMJOZlc}
...
,..0' \30' IH'
I
Figure 1. Distribution of major tectonostratigraphic units in the Klamath Mountains and northern Sierra Nevada provinces
of California. Data from Cady (1975), Day and others (1985), Dilek (1989), Harper and Wright (1984), Hietanen (1981, Irwin
(1981), Saleeby and others (1982), and Schweickert (1981). Thick dashed line south of Redding depicts inferred structural
discontinuity between Sierran and Klamath provinces. Stippled pattern and dash-dot line in the Great Valley indicate
gravity 1000 y) and magnetic anomalies (50 mgals), respectively. A-A' and B-B' are geologic cross-section lines in Figure 2.
Key to numbers: 1, Josephine ophiolite; 2, Galice Formation; 3, Rattlesnake Creek terrane; 4, western Hayfork terrane;
5, Eastern Hayfork terrane; 6, Stuart Fork-Fort Jones terrane; 7, North Fork-Salmon River terrane; 8, central metamorphic
belt; 9, Condrey Mountain Schist; 10, Yreka-Callahan terrane (Paleozoic); 11, Trinity Complex; 12, South Fork Mountain
Schist (Upper Jurassic-Lower Cretaceous(?) blueschist-greenschist terranes); 13, Eastern Klamr.th arc; 14, Coast Range
ophiolite; 15, Franciscan Complex; 16, Smartville Complex; 17, Slate Creek Complex; 18, Lake G.mbie Complex; 19, Jarbo
Gap ophiolite (oceanic basement); 20, Argillite matrix melange; 21- Feather River Peridotite; 22, Shoo Fly complex;
23, Paleozoic volcanic arc complexes; 24, Jurassic volcanic arc complexes. In the northern Sierra Nevada, 22-24 together
constitute the Eastern Belt; 17-20 the Central Belt; and 16 the Western Belt. Inset map shows major tectonic boundaries in
the North American Cordillera. KMP, Klamath Mountains province, SNP, Sierra Nevada provinc,).
fine-grained argillite matrix (Day and others, 1985;
Dilek, 1989; Hietanen, 1981; Vaitl, 1980). The
Western Belt consists of volcanic, volcaniclastic,
plutonic, and hypabyssal rocks of a Jurassic island-
arc terrane known as the Smartville complex (Beard
and Day, 1987; Day and others, 1985; Menzies and
others,1980). Volcanic and hypabyssal rocks of the
Smartville complex tectonically overlie metamor-
182
Y. Dilek and E. M. Moores
t o
--
--
D Batholithic Rocks
[;: ..... <i Franciscan complex
Blueschist-amphibolite
_ Jurassic ophiolite
t- : >: :1 Jurassic arc complex

D Oceanic basemenVarc rool (?)
D Melange (Sierra foothills)
KLAMATH MOUNTAINS
Great Valley Sequence
Swedes Flat
Pluton
I
Yuba Rivers
Pluton
I
o km 30
I , I
No vertical exaggeration
-- ............

NORTHERN SIERRA NEVADA
Figure 2. Interpretive geologic cross-sectons of Klamath Mountains (A-A') and northern Sierra Nevada CB-B'), California,
showing relationships between different tectonostratigraphic units. Symbols are as in Figure 1. EKT, Eastern Klamath
terranes, including Stuart Fork-Fort Jones terrane, central metamorphic belt, and Yreka-Callahan terrane.
phosed and deformed igneous and sedimentary
rocks of the melange in the Central Belt along
gently to steeply W-dipping faults (Day and others,
1985; Moores and Day, 1984; Ricci and others,
1985). Volcanic and volcaniclastic rocks of the
Smartville complex are overlain depositionally on
the west by clastic sediments of the Late Jurassic to
Cretaceous Great Valley Sequence (Figure 2).
Magnetic and gravity anomalies (Cady, 1975) and
seismic reflection data (Wentworth and others,
1984) suggest, however, that mafic hypabyssal and
volcanic rocks of the Smartville complex continue as
a W-dipping slab beneath the Great Valley Se-
quence (Figure 2).
Rocks and associated structures within the
Eastern Belt and the Feather River peridotite
display evidence for nearly continuous deposition,
recurrent magmatism, and episodic deformation
occurring near and/or at the active continental
margin of western North America during Permo-
Triassic through Late Jurassic time (Burchfiel and
Davis, 1981; D'Allura and others, 1977; Hannah and
Moores, 1986; Schweickert, 1981). Although the
paleogeography of Paleozoic island-arc assemblages
and their accretion to western North America are
still controversial, it is commonly accepted that the
Jurassic volcanic arc assemblage of the Eastern Belt
was constructed on the previously accreted terranes
and over an E-dipping subduction zone (see Burch-
fiel and Davis, 1981; Schweickert, 1981; and re-
ferences therein). Some major outstanding ques-
tions regarding the early Mesozoic tectonic evolu-
tion of the northern Sierra Nevada foothills
involved paleogeography of the Smartville complex
in the Western Belt and its pre-tectonic rela-
tionships with the tectonostratigraphic units in the
region. These questions constitute one emphasis of
this paper, and thus the Smartville complex in the
Western Belt and the tectonostratigraphic units in
the Central Belt are discussed in some depth in the
next section.
WESTERN BELT: SMARTVILLE COMPLEX
The Smartville complex crops out as a N/S-
trending tectonostratigraphic unit in the northern
Sierra Nevada foothills (Figure 1) and forms an E-
verging antiform composed of Jurassic volcanic, vol-
caniclastic, plutonic, and hypabyssal rocks. Rocks of
the Smartville complex continue in a narrow belt
south of latitude 39N as the Jurassic Foothills
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada 183
sequence in the central Sierra Nevada (Saleeby,
1981). The stratigraphic sequence of the Smartville
complex includes, from bottom to top: (1) massive
diabase and microgabbro; (2) pillowed and massive
lava flows; and (3) intermediate to felsic volcanic
and volcaniclastic rocks (Day and others, 1985;
Menzies and others, 1980). Intermediate to felsic
volcanic and volcaniclastic rocks of the upper
Smartville include dacite, rhyolite, andesite, basal-
tic andesite, and volcanic tuff, and contacts between
these rocks are generally well preserved. Whereas
the trace-element geochemistry of these interme-
diate to felsic volcanic rocks mainly shows calc-
alkaline trends (Menzies and others, 1980), whole-
40' -
39' -
I
121 ' 30' 121'
o
,
KM
rock trace-element concentrations and compositions
of pyroxenes from pillow and massive lavas show
both calcalkaline and tholeiitic trends (Menzies and
others, 1980; Xenophontos, 1984).
V olcanic and volcaniclastic rocks of the Smart-
ville complex are overlain by a flysch sequence (Fig-
ure 3), consisting of turbiditic sandstone, conglo-
merate, shale, pebbly mudstone, and olistostromal
blocks of aphanitic mafic volcanic rocks (Vaitl, 1980;
Dilek, 1989). The occurrence of Buchia concentrica
in a tuffaceous sandstone in the turbiditic sequence
suggests a middle Oxfordian to late Kimmeridgian
age for the flysch (Marlette and others, 1979; Vaitl,
1980).
25
.
E X P L A A T I 0
w ERN BELT
SMAR1VIU.E COMI'lfX
D rd Aysch deposiu
Q uvu Upper ,,'olcank: unit
Iillll!lIlI od- Sheeted dikes
r. II>' TonaJit1c plutons
I: : : : :1 zp- Zoned gobbro-diori.e plutofU
D mg Meugabll<o
Ivu- L.o\A,'C,f volcanic unit
md Massive diabase
CE 1TRAL BELT
VQLCA:o.;OPLUTOZ'Io'1C COMPLEXES
V ... !,1 LC Lal;e Combic. SeC Sioic ClOck
OCEANIC CRUST
IE IGO- larbo G.p ophiolite
MELANGE
D M- Argilli.e maui. tntiMl,.
FEATHER RI VER PERIDOTITE
O
- FRP Uhr1/lUfic and mafic roek.s.
amphibolite, and blueschist
MAJOR FA LT
BBF- Big Bend Fault
DPF Dog"'OOd Poll Fault
GCF Croek F.ull
GHF- Gillis Hill Faull
WCF WolfCreek F.ull
M J OR PL ONS
IILP Bucks LaI;. Pluton
BRP Billd Rock Pluton
CP- Cascade Pluton
GP Grizzly Pluton

SFP Swedes A .. Pluton
YRP- Yuba Rivers Pluton
Folded tluuSl rault
Figure 3. Generalized tectonic map of northern Sierra Nevada, California. Data from Beard and Day (1987), Day and others
(1985), Dilek (1989), Edelman (1987), Ricci and others (1985), and Vaitl (1980).
184
Y. Dilek and E. M. Moores
Volcanic and volcaniclastic rocks of the Smart-
ville complex are intruded by 163-159 Ma mafic to
felsic dikes and plutons (U-Pb zircon ages; Saleeby,
1981; Saleeby and others, 1989), which range in
composition from gabbro, gabbronorite to diorite,
monzodiorite, and tonalite (Beard and Day, 1987;
Xenophontos, 1984). Sheeted and un sheeted dikes
range in composition from diabase to basaltic ande-
site, andesite, dacite, and rhyolite, and they display
a geochemical trend encompassing both tholeiltic
and calc-alkaline compositions (Xenophontos, 1984).
Generally N/S-trending elongated plutons in the
center of the complex (Figure 3) are compositionally
similar to NINW -striking dikes. Ages and cross-
cutting relationships suggest that dike injection and
plutonism were mostly coeval (Beard and Day, 1987;
Saleeby and others, 1989).
V olcanic rocks of the Smartville complex are
lithologically, mineralogically, and geochemically
similar to those observed in modern island arcs in
the western Pacific region (e.g., Gill, 1981), and the
Smartville complex has been interpreted as an ensi-
matic arc terrane (Menzies and others, 1980; Beard
and Day, 1987). Sheeted dikes and coeval plutons
intrude all stratigraphic levels in the complex and
are thus younger than the Smartville volcanic
rocks. This interpretation is consistent with a Pb-U
zircon date of 2055 Ma (Bickford and Day, 1988;
Saleeby and others, 1989) from a tonalitic pluton
associated with the massive diabase unit, which
suggests that the core of the Smartville complex is
at least Early Jurassic in age. A SmlNd isochron
age of 178 21 Ma from the massive diabase and
pillow basalt units in the complex (Saleeby and
others, 1989) can be interpreted to support the Early
Jurassic age ofthe Smartville complex.
Hypabyssal and volcanic rocks of the Smartville
complex are separated from various rock types of
the Central Belt by complex fault systems in the
north and east. The northern margin of the Smart-
ville complex is bounded by the EIW-trending Big
Bend fault zone, which is characterized by 204-202
Ma and older mafic-ultramafic rocks and related
EIW-trending structures (Figure 3; Dilek, 1989;
Dilek and Moores, 1986). Southwest of the Big Bend
fault zone, volcanic rocks of the Smartville complex
structurally overlie argillite matrix melange of the
Central Belt along an E-directed subhorizontal
thrust fault that was subsequently deformed by E-
dipping high-angle reverse faults (Ricci and others,
1985). Similarly, at the southeastern margin of the
Smartville complex near Higgins Corner (Figure 3),
hypabyssal and volcanic rocks tectonically overlie
melange in the Central Belt along an E-vergent
subhorizontal fault (Day and others, 1985).
CENTRAL BELT
The Central Belt extends in a 25-35-km-wide
belt north and east of the Smartville complex
(Figure 3). It includes upper Paleozoic to lower
Mesozoic sedimentary rocks in an argillite matrix
melange, as well as fragments of Jurassic and older
ophiolitic and volcanoplutonic complexes.
ArgiUite Matrix Melange
The melange includes blocks of fossiliferous
Permian limestone, chert, graywacke, turbiditic
sandstone, volcaniclastic rocks, conglomerate, peb-
bly mudstone, gabbro, and serpentinite in a fine-
grained argillite matrix (Vaitl, 1980; Day and
others, 1985; Dilek, 1989). Some volcanic clasts and
blocks in the matrix display their primary shape
and igneous texture (i.e., well-preserved vesicles),
implying the existence of nearby volcanic centers.
The presence, in the matrix, of metamorphic two-
mica-schist clasts of possible continental origin
suggests that the depositional environment of ar-
gillaceous rocks was proximal to the continent
(Dilek, 1989). The depositional age of the argillite
matrix melange is not well-constrained because of
intense deformation and tectonic mixing; the
existence of Permian limestone blocks that are
equivalent to the McCloud limestone of the eastern
Klamath Mountains (Watkins and others, 1987) and
chert blocks with Triassic to Lower Jurassic radio-
larians (Hietanen, 1981; Irwin and others, 1978)
suggests a late Paleozoic to early Mesozoic age for
the melange. Lithologic characteristics of the
melange are similar to those of subduction-accretion
melanges widespread in the Mesozoic Cordillera of
the western United States (Cowan, 1985).
Volcanoplutonic Complexes
V olcanoplutonic complexes in the Central Belt
include the Lake Combie complex to the south and
the Slate Creek complex to the north (Figure 3) and
consist mainly of plutonic units (ranging in com-
position from gabbro to diorite and tonalite), hyp-
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada 185
abyssal diabase, massive to pillowed lava flows, and
overlying volcanic, volcaniclastic, and epiclastic
rocks (Day and others, 1985; Edelman, 1987). These
lithologies are similar to those in the Smartville
complex and characterize an island arc environment
of origin. U-Pb zircon dates from a tonalite pluton
in the Slate Creek complex indicate 205 5 Ma
emplacement ages (Bickford and Day, 1988; Saleeby
and others, 1989), suggesting that the Slate Creek
complex may represent a fragment of an Early
Jurassic arc terrane. There are no radiometric dates
available yet from the Lake Combie complex;
similarities in stratigraphic sequences between the
Smartville and Lake Combie complexes suggest,
however, that these two arc sequences may be gene-
tically related (Day and others, 1985).
Ophiolite Complexes
Ophiolitic rocks in the Central Belt include
serpentinized peridotites, gabbros, and volcanic and
sedimentary rocks, all of which are metamorphosed
up to amphibolite facies. These rocks represent a
dismembered ophiolite sequence which is deformed
locally as a serpentinite matrix melange. A leu co-
diorite lens within the metagabbro unit in the Jarbo
Gap ophiolite north of the Smartville complex
(Figure 3) yielded concordant U-Pb zircon dates of
204-202 Ma, whereas intermediate to felsic dikes
intruding a serpentinized peridotite unit gave con-
cordant U-Pb zircon dates of 1964 Ma (Saleeby
and others, 1989). These age relationships suggest
that mafic-ultramafic rocks of the Jarbo Gap ophio-
lite are coeval with or older than the'" 200 Ma Slate
Creek and Smartville complexes. Dikes of 196+4
Ma age intruding mantle rocks of the Jarbo Gap
ophiolite have compositions of basaltic andesites
formed from depleted mantle wedge above a sub-
ducting slab in an incipient island-arc setting (Thy
and Dilek, 1987). These intrusions may therefore
represent the inception ofthe Early Jurassic island-
arc magmatism in the northern Sierra Nevada
foothills.
INFERRED TECTONIC AND MAGMATIC
EVENTS IN THE NORTHERN SIERRA NEVADA
Stratigraphic, structural, and geochronologic
relationships suggest that the Smartville, Slate
Creek, and Lake Combie complexes may be frag-
ments of an Early Jurassic island-arc terrane. We
interpret dismembered and deformed ophiolitic
rocks in the Central Belt as remnants of the oceanic
basement of this Early Jurassic arc (Dilek, 1989;
Dilek and others, 1987). We summarize here the
inferred tectonic evolution of the Early Jurassic
island-arc terrane, which is depicted on a recon-
structed stratigraphic/structural section (Figure 4).
Radiometric dates from the Kings-Kaweah
ophiolite belt in the southern Sierra Nevada suggest
a period of ocean floor genesis and ophiolite deve-
lopment west of the North American continent
186
Y. Dilek and E. M. Moores
ONAG
NQR"lll cMN SIEKkA i'I:.V AD,\
WESUki\' KL'\'l"lll MOt 'HA1:\:S NORTlIF.R"\ COAST R,\XGES
lXJlfu. 11 1Uf'I (),:olonnallon
.' 13SIT1:lu...m
1),:0110"1001'1 [)('fo rrnatlOO

(v. r. Op) (f'<>,) (IT, '1. Or) (V,P.Op) (170") (H, M. Or ) (V. P.Op) (Fo,) (Tf.M,Or)

- 12S
,.
ti
..J


<

P
"'
r "" -
lJ,

l=
I
p

I
<
I
Fos

I
Fos. ---:,-=0<
I
I
F",
I
M. Tf ..J
I
P.
'\ -=--
Op
-
163
v, P Op I

'"
':l

I
P

'"
is
v
..-:;-
<
"
I

ex i1
v
:0
-
f- 187
>-
..J v. r
I
'" < P
"'
I
- I- 208
Fos
'"
I
I

.... I
< I
Op
..J I
<
I


2)0
-
;;
240
Figure 5. Magmatism, deposition, and deformation in northern Sierra Nevada, western Klamath Mountains, and northern
Coast Ranges. V, volcanism; P, plutonism; Op, ophiolite genesis; Fos, fossil ages; Tf, thrust faulting; M, metamorphism;
Op, orogenic phase. Stippled boxes denote maximum duration of Nevadan orogeny; ruled areas are maximum duration of
Middle Jurassic contraction. Deformation data from the northern Coast Ranges include information from amphibolite-facies
blocks in the Franciscan Complex, which may indicate initiation of Franciscan subduction zone. See Figure 4 and Table 1 for
data sources. Additional data from Ingersoll and Schweickert (1986) and Schweickert (1981).
during Middle to Late Triassic time (Saleeby, 1981).
Triassic to Early Jurassic dismembered ophiolitic
rocks and associated serpentinite-matrix melange
in the northwestern Sierra Nevada may also have
originated at this time. The existence of 196-204
Ma dikes and plutons intruding this ophiolitic
basement indicates active plutonism during Early
Jurassic time, fonning the core of the arc terrane
(Figures 4 and 5). Continued magmatic activity
during Early (to Middle?) Jurassic time resulted in
active volcanism and construction of a volcanic-arc
edifice, primarily and possibly entirely in a submar-
ine environment. This submarine arc was accreted
to continental margin along E-vergent thrusts
around Middle Jurassic time, as suggested by the
1652 Ma Scales pluton cutting across the thrust
fault beneath the Slate Creek complex. This accre-
tion may have been synchronous with regional
defonnation and metamorphism of all juxtaposed
terranes and tectonostratigraphic units in the
northern Sierra Nevada. The accreted arc terrane
was intruded by 163-159 Ma dikes and plutons,
recording arc-related magmatism during early Late
Jurassic time (Figures 3 and 5). The extensive
sheeted dike complex in the Smartville complex
suggests in situ rifting of the arc during this period.
Flysch sediments of 160-155 Ma age overlying the
Smartville complex indicate tectonic uplifting, rapid
erosion, and sedimentation, followed by culmination
ofthe Nevadan orogeny ("" 155 Ma) that resulted in
development of the E-dipping Foothills fault system.
The existence of late- and/or post-Nevadan plutons
(Figures 3 and 5) throughout the northern Sierra
Nevada indicates continued magmatic activity
during and after the main Nevadan defonnation.
CORRELATIONS WITH THE WESTERN
KLAMATH MOUNTAINS
The Klamath Mountains province consists of
Paleozoic and Mesozoic igneous and sedimentary
rocks with both oceanic and island-arc affinities
that fonn an imbricate stack of tectonostratigraphic
units bounded by generally E-dipping regional
faults (Figures 1 and 2; Burchfiel and Davis, 1981;
Irwin, 1981). These tectonostratigraphic units be-
come progressively younger westward, and thus the
relatively youngest rocks (i.e., Franciscan Complex)
occupy the lowest structural position within this
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada 187
collage of imbricate thrust packages. Although the
possible existence of E-directed thrust systems in
the province - particularly in the western Klamath
Mountains (see for example Cannat and Boudier,
1985; Roure, 1983) - cannot be ruled out, the re-
gional NE-dipping structural grain and the strati-
graphic and intrusive relationships have led many
workers to suggest an oceanward-migrating single
arc-trench system beneath the western edge of
North America (e.g., Burchfiel and Davis, 1981;
Harper and Wright, 1984; Wright and Fahan, 1988).
The intervening Great Valley Sequence over-
lying early Mesozoic and older basement rocks and
the"" 100 km across-strike offset of the Sierran and
Klamath provinces make direct correlations oflitho-
tectonic units and related structures in both regions
difficult. However, recent studies in the western
Klamath Mountains (e.g., Harper and Wright, 1984;
Saleeby and others, 1982; Wright and Fahan, 1988)
show remarkable similarities between the two
regions, as proposed earlier by Davis (1969). These
studies suggest that an early Middle Jurassic arc
terrane (western Hayfork terrane) characterized by
mafic to felsic volcanic and volcaniclastic rocks, and
contemporaneous intrusive complexes was construc-
ted through and across a structurally complex
ophiolitic terrane (Rattlesnake Creek terrane) of
Late Triassic to Early Jurassic age (Figures 4 and
5; Wright and Fahan, 1988). These documented
stratigraphic relationships in the western Klamath
Mountains are analogous to those between the
Jurassic arc and its ophiolitic basement in the
northern Sierra Nevada (Figure 4). Radiometric
dates show that the western Klamath island-arc
terrane and its ophiolitic basement underwent a
major contractional deformation during Middle
Jurassictime (= 168 Ma; Wright and Fahan, 1988),
causing the arc terrane to thrust westward over its
ophiolitic basement (Figure 5). Following this
regional Middle Jurassic deformation, rocks of the
arc-ophiolite complex were intruded by 163-159 Ma
mafic dikes of the Preston Peak complex (Figures 4
and 5), which are interpreted to represent the onset
of early Late Jurassic intra-arc rifting in the west-
ern Jurassic belt (Saleeby and others, 1982).
Continued rifting resulted in the opening of a back-
arc basin in which the 163-157 Ma Josephine ophio-
lite developed (Figures 4 and 5, Table 1). Flysch
sediments of the overlying Galice Formation and
supracrustal rocks of the Josephine ophiolite were
deformed and imbricated along W-directed thrusts
during the Late Jurassic Nevadan orogeny (Table
1; Saleeby and others, 1982). It thus appears that
metamorphic belts in both the northwestern Sierra
Nevada and the western Klamath Mountains
experienced nearly similar and synchronous tec-
tonomagmatic events during early Mesozoic time,
suggesting a tectonic history probably common to
both regions.
FRACTURE ZONE TECTONICS AND
TRANSFORM OFFSET OF THE SIERRAN-
KLAMATH PROVINCES
The nature of the apparent offset and deflection
between coeval tectonostratigraphic units and
structures in the northern Sierra Nevada and
Klamath Mountains is enigmatic. Davis (1969)
pointed out that many lithologic belts correlate
between these two regions, and he further proposed
that the apparent deflection may be due to flat-
tening of steep W-directed reverse faults in the
northern Sierra Nevada to shallow-dipping W-
directed thrust faults in the Klamath province.
Apparently opposite Middle Jurassic(?) transport
directions of the arc terranes in both regions are not
consistent with this interpretation, however.
Furthermore, the N/S-trending gravity and magne-
tic anomalies between the two provinces, which are
interpreted to signify oceanic rocks beneath Ceno-
zoic volcanic cover and the Great Valley Sequence
(Cady, 1975; A. Griscom, oral commun., 1988), do
not continue northward beyond a presumed EIW-
trending structural discontinuity south of the Kla-
math Mountains region (Figure 1). This discont-
inuity suggests significant lateral displacement
between the two regions.
An alternative interpretation of the apparent
offset suggests that differential movement may
have occurred along a system of left-lateral tear
faults within a WNW -trending Cretaceous fault
zone (Jones and Irwin, 1971). This NW- to SE-
trending fault zone offsets an Early Cretaceous
shoreline for > 100 km in the southern Klamath
Mountains and juxtaposes basement rocks of dif-
ferent character (Jones and Irwin, 1971). The basal
Jurassic sedimentary rocks rest on oceanic crust
within and south of the fault zone, whereas basal
Hauterivian (Early Cretaceous) strata rest on crust
of the Klamath Mountains north of it (Jones and
Irwin, 1971). These features support existence of a
major tectonic boundary along this zone of apparent
offset. A Cenozoic origin of the offset was proposed
188 Y. Dilek and E. M. Moores
Table 1. Geochronological constraints pertinent to tectonic and magmatic events in northwestern Sierra Nevada,
western Klamath Mountains, and northern Coast Ranges.
Sample Interpreted Significance' Data Source
1381 PbIU Hornblende-biotite Igneous age of post-Nevadan Saleeby and others, 1989
Zircon tonalite Grizzly pluton; intrudes rocks
of Central Belt
1582PbIU Tonalitic Igneous age ofsyn-Nevadan Yuba Saleeby and others, 1989
Zircon Rivers pluton
160-155 Buchia concentrica in a Depositional age of flysch overlying Marlette and others, 1979;
Fossil tuffaceous sandstone Smartville volcanic rocks Vaitl,1980
1625PbIU Plagiogranitic bodies and Igneous age of Smartville sheeted Bickford and Day, 1988;
Zircon a leucotonalite dikes and associated intrusions Saleeby and others, 1989
2055PbIU Tonalitic Igneous age of a pluton within Bickford and Day, 1988
Zircon massive diabase unit; age of core
of Smartville complex
168-165 PbIU Tonalitic Igneous age of a pluton intruding Bickford and Day, 1988;
Zircon east-vergent thrust fault beneath Saleeby and others, 1989
Slate Creek complex
2055PbIU Dioritic and tonalitic Igneous age of plutonic sequence Bickford and Day, 1988;
Zircon in Slate Creek complex Saleeby and others, 1989
1944PbIU Silicic tuff (quartz- Eruptive age ofvo1canic rocks Saleeby and others, 1989
Zircon keratophyre) from overlying Jarbo Gap ophiolite in
south of J arbo Gap Central Belt
1964PbIU Quartz diorite cross- Minimum igneous age of dike Saleeby and others, 1989
Zircon cutting intermediate swarms intruding J arbo Gap
to felsic dikes ophiolite in Central Belt
2042PbIU Leucodiorite lens in a Igneous age of a gabbroic intrusion Saleeby and others, 1989
Zircon flaser gabbro north of in Jarbo Gap ophiolite
JarboGap
205 43 SmIN d Suite of dioritic, its host Isotopic igneous age of ophiolitic Saleeby and others, 1989
gabbro, pyroxenite, and intrusive rocks in Jarbo Gap
flaser gabbro north of ophiolite
JarboGap
157-150 PbIU Cross-cutting plutons Maximum and minimum ages of References in Wright and
Zircon intruding Galice, J ose- Nevadan deformation; plutons cut Fahan,1988
phine, and Rattlesnake across Nevadan thrust faults
Creek terranes between terranes
160-155 Buchia concentrica near Depositional age of flysch overlying Saleeby and others, 1982
Fossil base of Galice Fm Josephine ophiolite
157PbIU Plagiogranite as a screen Igneouis age of Josephine ophiolite Saleeby and others, 1982
Zircon rock in sheeted dike
complex
159PbIU Quartz diorite from Initial rifting age of Josephine Saleeby and others 1982
Zircon Preston Peak mafic dike backarc basin
complex
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada
Sample
1683.4K/Ar Garnet-bearing
Hornblende amphibolite
168-170 PbJU Ironside Mtn batholith;
Zircon cross-cutting pluton
1772.8K/Ar Crystal tuff in western
Hornblende Hayfork arc terrane
170-169 PbJU Gabbroic-dioritic intru-
Zircon sives in western Hayfork
arc terrane
194-212 PbJU Gabbro-diorite plutons
Zircon intruding ultramafic
rocks of Rattlesnake
Creek terrane
160 Ar/Ar Amphibolite block in
Hornblende Franciscan complex
163-153 Radiolarian cherts and
Fossil tuffaceous cherts
165-155 Pb/U Felsic dikes and intru-
Zircon sions in Coast Range
ophiolite
Table 1
(continued)
Interpreted Significance
Metamorphism and deformation
age of western Hayfork arc terrane
Thrusting age of western Hayfork
arc terrane onto Rattlesnake Creek
terrane
Eruptive and/or depositional age of
volcaniclastic strata in arc terrane
Igneous age of plutonic complexes in
western Hayfork arc terrane
Igneous age of mafic plutons
intruding ophiolitic basement in
Rattlesnake Creek terrane
Development of a metamorphic sole
beneath Coast Range ophiolite; ini-
tiation of Franciscan subduction
zone (?)
Depositional age of sedimentary
rocks overlying Coast Range
ophiolite
Igneous age of Coast Range
ophiolite
Data Source
Wright and Fahan, 1988
Wright and Fahan, 1988
Wright and Fahan, 1988
Wright and Fahan, 1988
Wright and Fahan, 1988
Ross and Sharp, 1986
Hopson and others, 1981
Hopson and others, 1981
189
by Hamilton (1978), who suggested that the offset
may have developed due to a combination of rifting,
rotation, and strike-slip faulting resulting from
Basin and Range extension in the western U.S.A.
Paleomagnetic studies (e.g., Magill and Cox, 1981)
show, however, that during rotation of the Cascades,
Coast Range, Klamath, and Sierra Nevada crustal
blocks due to extension within the Basin and Range
province, all these blocks moved westward together,
retaining their E-W-spacing. These data suggest
that the offset between the Sierran and Klamath
blocks existed prior to Basin and Range extension.
systems of the northern Sierra Nevada and the
western Klamth Mountains prior to their accretion
to the continent.
As an alternative model we propose the exis-
tence of an oceanic paleo-transform-fault boundary
between the Sierran and Klamath provinces (Dilek
and Moores, 1986). In this model, the EIW-trending
Big Bend fault zone north of the Smartville complex
(Figures 1 and 3) is interpreted to represent a seg-
ment of this transform boundary, which once sepa-
rated the intraoceanic Jurassic island arc-trench
Existence of a fracture zone in the northwestern
Sierra Nevada is suggested for two major reasons:
(1) the abrupt bend of N/S-trending structures and
tectonic boundaries to a conspicuous E-W orienta-
tion immediately north of the Smartville arc, and
termination of the Smartville complex and arc-
related rocks against the Big Bend fault zone; and
(2) oceanic fracture zone characteristics of mafic-
ultramafic rocks and overlying supracrustal rocks
locally embedded in a schistose and detrital ser-
pentinite matrix within this fault zone (Dilek and
Moores, 1986). Contacts between ophiolitic sub-
units (include tectonized harzburgite, dunite, wehr-
lite, flaser gabbro, amphibolite, and metamorphosed
volcanic and sedimentary rocks) exhibit complex
intrusive and tectonic relationships that show both
along- and across-strike variations. Reworked and
190 Y. Dilek and E. M. Moores
sedimentary serpentinites, ophicalcites, mafic talus
breccias, and serpentinite diapirs indicate that
oceanic crust composed of these rock types was
exposed on a sea floor with high local relief and
active tectonism (Dilek and Moores, 1986).
Andesitic and basaltic-andesitic dikes of boninitic
origin intrude serpentinized peridotites (Thy and
Dilek, 1987), overlain by fault-bounded basinal
sedimentary sequences. These features are reminis-
cent of oceanic fracture zones and support an oceanic
fracture-zone origin of these ophiolitic fragments
and associated serpentinite-matrix melange within
the Big Bend fault zone. Based on structural,
intrusive, and geochronological relationships, Dilek
and others (1987) suggested that this oceanic frac-
ture zone may constitute the northern termination
and the oceanic basement of the Smartville arc
complex.
Fracture-zone-related oceanic rocks exist along
the same latitude farther west in the northern Coast
Range ophiolite (see Figure 1). Ophiolitic talus
breccias and interbedded pillow-lavas overlying
plutonic and ultramafic rocks within fault-bounded
blocks in the northern Coast Range ophiolite (see
Figure 4) are interpreted as having formed in an
oceanic fracture zone (Harper and others, 1985).
These features suggest that the arc-related southern
Coast Range ophiolite (Shervais and Kimbrough,
1985) is bounded in the north by a fracture zone
ophiolite, similar to the northern boundary of the
Smartville arc terrane. However, ophiolitic rocks in
the northern Coast Ranges are Late Jurassic in age
(see Figure 5) and thus younger than those in the
northwestern Sierra Nevada. These age relation-
ships may imply westward propagation of the frac-
ture zone, assuming that both the northwestern
Sierra and the northern Coast Range fracture zones
are continuous. Alternatively, they may imply
development of a continental-margin fracture zone
at the site of the already-accreted oceanic fracture
zone.
The inferred EIW -trending oceanic fracture
zone between the northern Sierra Nevada and the
Klamath Mountains represents a transform fault
boundary separating the early Mesozoic arc-trench
systems of these regions. Such a transform fault
requires the plate boundary to be a trench/trench/
transform fault. The type (i.e., shortening, length-
ening, or constant length) and evolution of this
transform fault depend mainly on inferred geometry
and polarity of subduction zones and upper plates in
both regions.
MODELS FOR EARLY MESOZOIC TECTONICS OF
THE WESTERN U.S. CONTINENTAL MARGIN
Plate Tectonic Constraints
Computed plate motions of western North Am-
erica and adjacent oceanic plates, and paleomagne-
tic data from both accreted terranes and the craton,
place constraints on tectonic models for Mesozoic
western United States. Reconstructions by Page
and Engebretson (1984) suggest a large sinistral
oblique component of convergence ("" 6.0 cm/yr)
between the Farallon and North American plates
"" 160-120 Ma. This plate motion model is consis-
tent with accreted oceanic terranes in the north-
western Sierra Nevada arriving from the northwest
relative to present-day North American coordinates.
There are no direct constraints on orientation of the
North American continental margin during and
before accretion of these terranes; however, the
trend of the Jurassic Sierran plutonic belt suggests
that the active margin of North America had a
N400W orientation (Page and Engebretson, 1984).
These reconstructions and computed plate motions
do not provide any constraints, nevertheless, on
travel distance of accreted terranes.
Paleomagnetic studies of several Upper Cre-
taceous plutons in the central Sierra Nevada yield a
mean 100-90 Ma paleopole of 71.5N, 174.9E for the
Sierra Nevada (Frei, 1986). When compared with
the mean cratonal pole of the same age (66.4N,
187.4E), =::600 km of poleward displacement of the
Sierra Nevada relative to the craton is required
(Frei, 1986). This post-Late Cretaceous northward
displacement of the Sierra Nevada could have
resulted from large-scale dextral shear along its
eastern boundary that preceded Basin and Range
extension (Dilek and others, 1988a; Frei, 1986).
Paleomagnetic studies of a number of Upper Juras-
sic plutons that intrude the Middle Jurassic arc and
its oceanic basement in the western Klamath Moun-
tains indicate clockwise rotation of "" 60, which is
comparable to that recorded by Jurassic strata in
the eastern Klamath Mountains (Fagin and Gose,
1983; Irwin, 1985; Mankinen and others, 1988).
Paleomagnetic data from both the western and
eastern Klamath terranes suggest a pre-Late Juras-
sic amalgamation of these terranes, possibly around
Middle Jurassic time, in agreement with both the
geologic and geochronologic record. These paleo-
magnetic results from the Klamath Mountains
imply a more westerly trend for the terranes and the
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada 191
continental margin prior to the onset of clockwise
rotation. Irwin (1985) suggested that clockwise
rotation of the Klamath terranes should have ceased
by Early Cretaceous time, synchronous with depo-
sition of Lower Cretaceous Great Valley sediments
on Klamath basement.
Tectonic Models
On the basis of the regional geology and plate-
tectonics constraints, we present here several
hypothetical evolutionary models for the early
Mesozioc continental margin of the western United
States. Each model attempts to reconcile as much
data and different opinions on the Sierran and
Klamath geology as possible.
WestFacing.Arcs Model. This model depicts
a continuous and long-lived subduction zone be-
neath the leading edge of the North American
continent during much of the early Mesozoic and
thus defines an oceanward migrating arc-trench
system (Figure 6). W-facing arc systems in both
regions represent oceanward continuation of the
North American continental arc. A modern ana-
logue for this kind of arc system is the Japanese
Islands fringing continental Asia. The transform-
fault boundary separating the fringing Sierran and
Klamath arcs may represent, in this case, a left-
lateral fracture zone. Modern examples for this type
of transform fault include the West Aleutians-
Kommandorski Islands plate boundary in the north-
western Pacific region.
& previously accreted terranes
o Island-arc terrane
o Oceanic crustophloh e
o SubduCllon melange (FranCiscan complex)
Transform faufl
...... AChve subduction zone
.-- ThruSf lault (barbS on upper plale)
.......... Suture zone
i( Oceanic spreading
Figure 6. Schematic diagrams showing tectonic evolution of W-facing arcs in Sierran/Klamath regions. Oceanward-
migrating fringing arc-trench systems of Sierran/Klamath regions separated by left-offset transform fault. Subduction and
clockwise rotation of Klamath Mountains continue throughout early Mesozoic. Both arc systems undergo regional thrusting
and metamorphism and are accreted to continent during Middle Jurassic time ("=175 Ma) . Continuous subduction and
magmatism in Late Jurassic time result in rifting of accreted arc terranes and opening of Josephine and Coast Range ophiolite
basins. Rifting event and genesis of ophiolites nearly coincide with initiation of Franciscan subduction zone (see Figure 5).
192
Y. Dilek and E. M. Moores
East-Facing Intraoceanic Arcs ModeL This
model invokes intraoceanic, and thus exotic, arc
systems in the Sierran and Klamath regions (Fig-
ure 7). An intraoceanic nature of arc systems is
suggested by the oceanic character of the arc
substratum and early eastward transport directions
in the northern Sierra Nevada and possibly in the
Klamath Mountains (see Roure, 1983; Cannat and
Boudier,1985). The inferred cusp between the sub-
duction zones results in clockwise rotation of the
Klamath arc and initiation and propagation of a
left-lateral transform fault between the Sierran and
Klamath arc-trench systems. A modern analogue
for such a cuspate arc-trench system is the Pulau
Seram-Timor volcanic islands developed over SW-
and W-dipping subduction zones in the Banda Sea
region that are connected by a left-lateral
transform-fault plate boundary. The presence of
late Paleozoic and early Mesozoic continental arc
rocks in the Eastern Belt and the eastern Klamath
Mountains (see Figure 1) indicates that an E-
dipping subduction zone was present beneath the
continental margin of North America during evo-
lution of the intraoceanic Sierran and Klamath arcs
(Figure 7). This argument implies that the intra-
oceanic Sierran and Klamath arcs may have deve-
loped above different subduction zone(s) presumably
dipping away from the continent. This scenario de-
picts oppositely dipping subduction zones west of the
continent in both the Sierran and Klamath regions.
A modern example of this kind of paleogeography is
the Molucca Sea region, where two island arcs
(Halmahera and Sangihe), facing each other over
oppositely dipping subduction zones, have been col-
liding since mid-Tertiary time (Moore and Silver,
1983).
Intraoceanic Arcs with Opposite Polarities
Model. This model depicts oppositely dipping intra-
oceanic subduction zones and related arc systems
separated by a left-lateral transform fault (Figure
8). The northern Sierran arc represents an oceanic
island arc developed over a NW-dipping subduction
zone (in present coordinates), whereas the western
Klamath arc of the same age delineates an oceanic
arc over a NE-dipping subduction zone. These arc-
trench systems are separated by a left-lateral trans-
form fault that lengthens by an amount equal to the
mutual rate of subduction, assuming rigid plates.
The best modern analogue for this kind of tectonic
setting exists in the southwestern Pacific, where the
E-facing Tonga-Kermadec and the W-facing New
Hebrides volcanic islands are offset by a sinistral
fault zone that includes the Fiji and Hunter fracture
zones. The existence of an E-dipping subduction
zone at the leading edge of the continent, as dis-
cussed in the previous model, implies, in this case, a
collision of two oppositely dipping subduction zones
in the northern Sierra Nevada, and docking and
amalgamation of continental and oceanic arcs in the
Klamath Mountains.
DISCUSSION
The geology of the northern Sierra Nevada and
the Klamath Mountains is highly complex, and
models presented in this paper portray oversimpli-
fied approaches for a better understanding of this
part of the North American Cordillera. Although
tectonic scenarios in each of the models are in-
ternally consistent, they do not necessarily fit all
known geological constraints. In the first model of
W-facing arcs, for example, there is no convincing
explanation for the cause of contractional deforma-
tion during Middle and Late Jurassic times when
subduction was still continuing. Documented trans-
port directions of the arc terranes in both provinces
are opposite: eastward in the northern Sierra Neva-
da, but westward in the western Klamath Moun-
tains. Furthermore, it is difficult to explain east-
ward thrusting and emplacement of the arc in the
northern Sierra Nevada in a dominantly E-dipping
convergent system.
The model of E-facing intraoceanic arcs strong-
ly suggests a collision orogen for Middle Jurassic
deformation, implying that arc terranes in both
provinces are exotic to North America. These im-
plications are incompatible with current interpre-
tations of tectonics of the Klamath Mountains (e.g.,
Harper and Wright, 1984; Wright and Fahan, 1988).
Furthermore, generally W-directed thrust faults in
the upper plate are antithetic to postulated sub-
duction polarity in the western Klamaths.
Documented transport directions of the Jurassic
arc terranes in both provinces support the plate
geometry presented in the model of intraoceanic
arcs with opposite polarities. The presumption here
is that different transport directions of mantle rocks
and arc complexes may point to fundamental dif-
ferences in the geometry of plate boundaries and a
change in polarity of the subduction zones. The W-
dipping seismic reflector beneath the oceanic slab
underneath the Great Valley Sequence (Wentworth
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada
Figure 7. Schematic diagrams showing
tectonic evolution of E-facing intraoceanic
arcs in Sierran/Klamath regions. Symbols
as in Figure 6. Arc-trench systems become
separated by a left-lateral transform fault
as Intraoceanic Sierran/Klamath Moun-
tains rotate clockwise. Intervening oceanic
crust between arc-trench systems and ac-
tive continental margin of North America
consumed at oppositely-dipping subduction
zones. Intraoceanic arc-trench systems
collide with continental margin 175 Ma and
become accreted to continent. Following
collision event, E-dipping Franciscan sub-
duction zone develops beneath newly-
assembled continental margin. Josephine
and Coast Range ophiolite basins open up
above this subduction zone during early
Late Jurassic time.
Figure 8. Schematic diagrams showing
tectonic evolution of introceanic arcs with
opposite polarities in Sierran/Klamath
regions. Symbols as in Figure 6. Oppositely
facing Klamath and Sierran oceanic arcs
separated by transform fault that length-
ens with time as subduction and clockwise
rotation of Klamath Mountains continue.
Collision of Sierran arc and docking of
Klamath arc with active continental mar-
gin around Middle Jurassic time Ma)
result in regional deformation. Continued
convergence between North America and
oceanic plates causes initiation and south-
ward propagation of E-dipping Franciscan
subduction zone beneath continent. Initia-
tion and development of this subduction
zone nearly coincides with of Late Jurassic
Nevadan orogeny in both regions (see
Figure 5).
Ma
Ma
Ma
193
194 Y. Dilek and E. M. Moores
and others, 1984) may be interpreted as a W-dipping
mantle thrust consistent with a W -dipping sub-
duction-zone model. Such a subduction zone may
have facilitated eastward displacement and trans-
port of the island arc and its oceanic basement over
the accretionary complex of the active North
American margin. However, an interpretation of
subduction polarity cannot be based solely on ver-
gence of a thrust fault that emplaced the arc terrane
onto the continental margin, and consequently an
adequate documentation of evidence indicating a W-
dipping subduction zone is necessary.
The models presented in this paper do not
consider the probable post-accretion dispersion of
accreted terranes as implied by paleomagnetic data
discussed earlier. These data suggest lateral dis-
placements parallel to the continental margin and
thus imply that the accreted terranes might not
have developed in their present relative positions.
Inferred post-Late Cretaceous northward transla-
tion of the Sierra Nevada apparently post-dates both
the Middle and Late Jurassic deformational events.
How and where this northward translation is ac-
commodated remains a major question in Cordille-
ran geology. Dilek and others (1988a) suggested
that the dextral Walker Lane shear zone east of the
Sierra Nevada batholith may be the accomodation
zone for such a large-scale ("" 600 km) migration
(inset map in Figure 1). Northward continuation of
the Walker Lane shear zone in northern California
and Oregon is not known, however, because of the
cover of Cenozoic volcanic rocks.
The likely northward translation of the Sierran
arc has a major bearing on the interpretation of the
transform offset between the Sierran and Klamath
arcs. The interpretation of this transform boundary
as a pre-Nevadan structure implies that the Kla-
math arc also migrated northward for an equivalent
amount with the Sierran arc. The Walker Lane
shear zone may continue northward, in this case,
east of the Klamath arc. More paleomagnetic data
from the Klamath Mountains province are essential
to test this hypothesis. A relatively parauthoch-
thonous position of the Klamath arc, if demon-
strated, signifies a post-Nevadan origin of the
transform boundary between the Sierran and Kla-
math arcs. This interpretation suggests that the
boundary may be part of the Walker Lane shear
zone extending south of the Klamath arc. Addi-
tional field, geochronologic, and paleomagnetic data
are crucial to test this hypothesis and to refine the
models presented here.
ACKNOWLEDGMENTS
Support for field studies in the northern Sierra
Nevada provided by R. C. Baker and Cordell Durrell
research grants (Department of Geology, University
of California) to Y. Dilek and by NSF grant EAR-80-
19697 to E. M. Moores and H. W. Day is gratefully
acknowledged. Our interpretations of Sierran geo-
logy have benefited from numerous discussions and
communications with H. W. Day, S. H. Edelman, J.
B. Saleeby, R. A. Schweickert, and R. K. Springer.
Critical and constructive reviews by J. W. Goodge,
G. D. Haeck, and R. J. Twiss improved the paper
substantially.
REFERENCES
BEARD, J. S., and H. W. DAY, 1987, The Smartville
intrusive complex, Sierra Nevada, California: The core of
a rifted volcanic arc: Geological Society of America
Bulletin, v. 99, p. 779-79l.
BICKFORD, M. E., and H. W. DAY, 1988, Jurassic ages of
arc-ophiolite complexes, northern Sierra Nevada: Implica-
tions for duration of the Nevadan "Orogeny" [abstract]:
Geological Society of America Abstracts with Programs, v.
20, p. A274.
BURCHFIEL, B. C., and G. A. DAVIS, 1981, Triassic and
Jurassic tectonic evolution of the Klamath Mountains-
Sierra Nevada geologic terrane; pp. 50-70 in W G. Ernst
(ed.), The Geotectonic Development ofCaUfornia (Ru-
bey Volume 1): Prentice-Hall, Englewood Cliffs, New
Jersey, USA, 706p.
CADY, J. W., 1975, Magnetic and Gravity Anomalies in
The Great Valley and Western Sierra Nevada Meta-
morphic Belt, California: Geological Society of America.
Special Paper 168,56 p.
CANNAT, M., and F. BOUDIER, 1985, Structural study of
intra-oceanic thrusting in the Klamath Mountains, north-
ern California: Implications on accretion geometry: Tec-
tonics, v. 4, p. 435-452.
COWAN, D. S., 1985, Structural styles in Mesozoic and
Cenozoic melanges in the western Cordillera of North
America: Geological Society of America Bulletin, v. 96, p.
451-462.
D'ALLURA, J. A., E. M. MOORES, and L. ROBINSON, 1977,
Paleozoic rocks of the northern Sierra Nevada: Their
structural and paleogeographic implications; pp. 395-408
in J. H. Stewart, C. H. Stevens and A. E. Fritsche (eds.),
Paleozoic Paleogeography of the Western United
States: Society of Economic Paleontologists and Mineralo-
gists, Pacific Coast Paleogeography Symposium 1, p ..
Island-arc evolution and fracture-zone tectonics in the Mesozoic Sierra Nevada 195
DA VIS, G. A., 1969, Tectonic correlations, Klamath Moun-
tains and western Sierra Nevada, California: Geological
Society of America Bulletin, v. 80, p. 1095-1108.
DAY, H. W., E. M. MOORES, and A. C. TUMINAS, 1985,
Structure and tectonics of the northern Sierra Nevada:
Geological Society of America Bulletin, v.96, p.436-450.
DILEK, Y., 1989, Tectonic signiflcance of post-accretion
rifting of a Mesozoic island-arc terrane in the northern
Sierra Nevada, California: Journal of Geology, v. 97, p.
503-518.
__ , and E. M. MOORES, 1986, A possible allochthonous
oceanic transform fault zone in the northwestern Sierra
Nevada, California [abstract): Geological Society of Am-
ericaAbstracts with Programs, v. 18, p. 101.
E. M. MOORES, and M. C. ERSKINE, 1988a, Ophio-
litic thrust nappes in western Nevada: Implications for the
Cordilleran orogen: Journal of the Geological Society of
London, v. 145, p. 969-975.
__ , E. M. MOORES, K. L. VEROSUB, and P. THY, 1988b,
Petrologic, paleomagnetic, and radiometric constraints on
the timing and nature of the accretion of a Mesozoic arc
ophiolite in the northern Sierra Nevada [abstract): Geo
logical Society of America Abstracts with Programs, v. 20,
p.155.
__ , P. THY, and E. M. MOORES, 1987, Fracture zone
basement of the Smartville complex, northwestern Sierra
Nevada, California [abstract): Geological Society of Ameli-
caAbstracts with Programs, v. 19, p. 372.
EDELMAN, S. H., 1987, Evidence for two Mesozoic arc-
continent collisions in the western Sierra Nevada meta-
morphic belt, California: Plate kinematic interpretation of
a terrane analysis [abstract): Geological Society of Ameri-
caAbstracts with Programs, v. 19, p. 651.
FAGIN, S. W., and W. A. GOSE, 1983, Paleomagnetic data
from the Redding section of the eastern Klamath belt,
northern California: Geology, v. 11, p. 505-508.
FREI, L. S., 1986, Additional paleomagnetic results from
the Sierra Nevada: Further constraints on Basin and
Range extension and northward displacement in the west-
ern United States: Geological Society of America Bulletin,
v. 97, p. 840-849.
GILL, J. B., 1981, Orogenic Andesites and Plate
Tectonics: Springer-Verlag, Heidelberg, Germany, 390 p.
HAMILTON, W., 1978, Mesozoic tectonics of the western
United States; pp. 33-70 in D. G. Howell and K. A. Mc
Dougall (eds.), Mesozoic Paleogeography of the West-
ern United States: Society of Economic Paleontologists
and Mineralogists, Pacific Coast Paleogeography Sympo-
sium 2, 573 p.
HANNAH, J. L., and E. M. MOORES, 1986, Age relation-
ships and depositional environments of Paleozoic strata,
northern Sierra Nevada, California: Geological Society of
America Bulletin, v. 97, p. 787-797.
HARPER, G. D., and J. E. WRIGHT, 1984, Middle to Late
Jurassic tectonic evolution of the Klamath Mountains,
California-Oregon: Tectonics, v. 3, p. 759-772.
J. B. SALEEBY, and E. A. S. NORMAN, 1985,
Geometry and tectonic setting of sea-floor spreading for the
Josephine ophiolite, and implications for Jurassic accre-
tionary events along the California margin; pp. 239-257 in
D. G. Howell (ed.), Tectonostratigraphic Terranes of
the Circum-PacifIc Region: Circum-Pacific Council for
Energy and Mineral Resources, Earth Science Series 1, 581
p.
HIETANEN, A., 1981, Petrologic and Structural Studies
in the Northwestern Sierra Nevada, California: U.S.
Geological Survey, Professional Paper 1226,59 p.
HOPSON, C. A., J. M. MATTINSON, and E. A. PESSAGNO,
1981, Coast Range ophiolite, western California; pp. 419-
510 in W. G. Ernst (ed.), The Geotectonic Development
of California (Rubey Volume 1): Prentice-Hall, Engle-
wood Cliffs, New Jersey, USA, 706 p.
INGERSOLL, R. V., and R. A. SCHWEICKERT, 1986, A
plate-tectonic model for Late Jurassic ophiolite genesis,
Nevadan orogeny and forearc initiation, northern Califor-
nia: Tectonics, v. 5, p. 901-912.
IRWIN, W. P., 1981, Tectonic accretion of the Klamath
Mountains; pp. 29-49 in W. G. Ernst (ed.), The Geotec-
tonic Development of California (Rubey Volume 1):
Prentice-Hall, Englewood Cliffs, New Jersey, USA, 706 p.
___ ,1985, Age and tectonics of plutonic belts in accreted
terranes of the Klamath Mountains, California and Ore-
gon; pp. 187-199 in D. G. Howell (ed.), Tectonostrati-
graphic Terranes of the Circum-PacifIc Region:
Circum-Pacific Council for Energy and Mineral Resources,
Earth Science Series 1, 581 p.
D. L. JONES, and T. A. KAPLAN, 1978, Radiolarians
from pre-Nevadan rocks of the Klamath Mountains,
California and Oregon; pp. 303-310 in D. G. Howell and K.
A. Mc Dougall (eds.), Mesozoic Paleogeography of the
Western United States: Society of Economic Paleonto-
logists and Mineralogists, Pacific Coast Paleogeography
Symposium 2, 573 p.
JONES, D. L., and W. P. IRWIN, 1971, Structural impli-
cation of an offset Early Cretaceous shoreline in northern
California: Geological Society of America Bulletin, v. 82, p.
815-822.
MAGILL, J., and A. COX, 1981, Post-Oligocene tectonic
rotation of the Oregon Western Cascade Range and the
Klamath Mountains: Geology, v. 9, p.127-131.
196 Y. Dilek and E. M. Moores
MANKINEN, E. A., W. P. IRWIN, and C. S. GROMME, 1988,
Paleomagnetic results from the Shasta Bally plutonic belt
in the Klamath Mountains province, northern California:
Geophysical Research Letters, v. 15, p. 56-59.
MARLETIE, J. W., R. 1. AKERS, K. A. COLE, and R. D.
McJUNKIN, 1978, Geologic investigations: The August 1,
1975 Oroville earthquake investigation: California De-
partment of Water Resources, Bulletin 204-78, p. 15-103.
MENZIES, M., D. BLANCHARD and C. XENOPHONTOS,
1980, Genesis ofthe Smartville Arc-Ophiolite, Sierra Nev-
ada Foothills, California: American Journal of Science, v.
280-A, p. 329-344.
MOORE, G. F. and E. A. SILVER, 1983, Collision processes
in the northern Molucca Sea; pp. 360-372 in D. E. Hayes
(ed.), The Tectonic and Geologic Evolution of South-
east Asian Seas and Islands, Part 2: American Geo-
physical Union Monograph 27,396 p.
MOORES, E. M., 1972, Model for Jurassic island arc-
continental margin collision in California [abstract]:
Geological Society of America Abstracts with Programs, v.
4, p. 202.
__ , and H. W. DAY, 1984, An overthrust model for the
Sierra Nevada: Geology, v. 12, p. 416-419.
PAGE, B. M., and D. C. ENGEBRETSON, 1984, Correlation
between the geologic record and computed plate motions
for central California: Tectonics, v. 3, p. 133-155.
RICCI, M. P., E. M. MOORES, K. L. VEROSUB, and J. S.
McCLAIN, 1985, Geologic and gravity evidence for thrust
emplacement of Smartville ophiolite: Tectonics, v. 4, p.
539-546.
ROSS, J. A., and W. D. SHARP, 1986, 4oAr/39Ar and SmlNd
dating of garnet amphibolite in the Coast Ranges [ab-
stract): EOS, Transactions of the American Geophysical
Union, v. 67, p. 1249.
ROURE, F., 1983, New data on vergence and tectonic
history in the Klamath Mountains [abstract): Geological
Society of America Abstracts with Programs, v. 15, p. 426.
SALEEBY, J. B., 1981, Ocean floor accretion and volcano-
plutonic arc evolution in the Mesozoic Sierra Nevada,
California; pp. 132-181 in W. G. Ernst (ed.), The Geotec-
tonic Development of California (Rubey Volume 1):
Prentice-Hall, Englewood Cliffs, New Jersey, USA, 706 p.
__ , G. D. HARPER, A. W. SNOKE, and W. D. SHARP,
1982, Time relations and structural-stratigraphic patterns
in ophiolite accretion, west central Klamath Mountains,
California: Journal of Geophysical Research, v. 87, p.
3831-3848.
__ , H. F. SHAW, S. NIEMEYER, E. M. MOORES, and S.
H. EDELMAN, 1989, U/Pb, SmINd, and Rb/Sr geochrono-
logical and isotopic study of northern Sierra Nevada ophio-
litic assemblages, California: Contributions to Mineralogy
and Petrology, v. 102, p. 205-220.
SCHWEICKERT, R. A., 1981, Tectonic evolution of the
Sierra Nevada Range; pp. 87-131 in W G. Ernst (ed.), The
Geotectonic Development of California (Rubey Vol-
ume 1): Prentice-Hall, Englewood Cliffs, New Jersey,
USA, 706p.
__ , and D. S. COWAN, 1975, Early Mesozoic tectonic
evolution of the western Sierra Nevada, California: Geo-
logical Society of America Bulletin, v. 86, p. 1329-1336.
SHERVAIS, J. W., and D. L. KIMBROUGH, 1985, Geo-
chemical evidence for the tectonic setting of the Coast
Range ophiolite: A composite island arc-oceanic crust ter-
rane in western California: Geology, v. 13, p. 35-38.
THY, P., and Y. DILEK, 1987, Boninitic dikes from a
fracture zone basement of the Smartville complex, NW
Sierra Nevada, California [abstract): EOS, Transactions of
the American Geophysical Union, v. 68, p. 1519.
VAITL, J., 1980, Geology and Structure of a Northern
Sierra Nevada Foothill Melange and Adjacent Areas,
Butte County, California: MS thesis, University of Cali-
fornia, Davis, California, USA, 93 p.
WATKINS, R., C. E. REINHEIMER, J. W. WALLACE, andM.
K. NESTELL, 1987, Paleogeographic significance of a Per-
mian sedimentary megamictite in the Central Belt of the
northern Sierra Nevada: Geological Society of America
Bulletin, v. 99, p. 771-778.
WENTWORTH, C. M., M. D. ZOBACK, and A. W. WALTER,
1984, Testing models of Franciscan and Nevadan obduc-
tion by scientific drilling in central California [abstract):
EOS, Transactions of the American Geophysical Union, v.
65,p.1102
WRIGHT, J. E., and M. R. FAHAN, 1988, An expanded view
of Jurassic orogenesis in the western U.S. Cordillera:
Middle Jurassic (pre-Nevadan) regional metamorphism
and thrust faulting within an active arc environment,
Klamath Mountains, California: Geological Society of
America Bulletin, v. 100, p. 859-876.
XENOPHONTOS, C., 1984, Geology, Petrology, and Geo-
chemistry of Part of the Smartville Complex, Northern
Sierra Nevada, California: PhD dissertation, University
of California, Davis, California, USA, 446 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Problems concerning collisional vs noncollisional deformation
at continental-margin orogens, with an example from the
Mesozoic Cordilleran orogen
STEVEN H. EDELMAN
Department of Geological Sciences. University of California, Santa Barbara, California 93106, USA
(received September 30, 1988; revision accepted April 20, 1989)
ABSTRACT
The relative importance of collisional vs noncollisional orogenesis at continental-margin orogens is an
unresolved problem in tectonics. The fundamental problem is that contraction of continental lithosphere in a
subduction hanging wall could be due to subcontinental subduction (noncollisional orogenesis) or to arc-
continent collision preceding subcontinental subduction initiated by subduction flip (collisional orogenesis). A
secondary problem arises in proving that sutures representing ocean-basin closure and arc collision exist in
the complex oceanic terranes that typically constitute the internal parts of continental-margin orogens. This
problem exists in part because "tectonic flaking" results in replacing the pre-collision subduction-zone fault
with a new fault during collision. Tectonic flaking and subsequent thrust imbrication, folding, normal
faulting, and strike-slip faulting make arc-continent sutures in collisional orogens difficult to distinguish
from deformed marginal arcs in noncollisional orogens. Phanerozoic continental-margin orogens contain
evidence for early orogenic arc accretion, and the Mesozoic cordilleran orogen of the western United States, a
classic example of a noncollisional, "Andean-style" orogen, fits a collisional model in detail. Tectonic models
for contractional deformation in Archean and Proterozoic continental-margin orogens should consider both
collisional and noncollisional processes.
INTRODUCTION
A fundamental ambiguity exists with inter-
preting plate-tectonic settings of orogenic belts
formed at continental margins (Dickinson, 1971). In
contrast to intracontinental orogens such as the
Alps, Himalayas, Urals, and the late Paleozoic
Appalachians, which are convincingly continent-
continent collisions, orogens at continental margins
may form by (1) noncollisional subcontinental sub-
duction, or (2) arc-continent collision. In the latter
case, subduction may flip after collision and dip
beneath the continent (Dewey and Bird, 1970;
Moores, 1970). For instances in which subcon-
tinental subduction is accompanied by orogenesis -
i.e., in an "Andean-style" orogen, no obvious way
exists to determine whether orogenic contraction of
the continental margin is noncollisional and related
to the flipped subcontinental subduction or is
collisional and related to an earlier arc collision.
This paper outlines this problem and cites an
example from the Mesozoic Cordilleran orogen of
the western United States. New constraints on the
existence, timing, and geometry of arc collisions in
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 197-204. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
197
198 S. H. Edelman
the internal part of this orogen are summarized in
order to outline the case for collisional deformation.
This deformation occurred along with subcontinen-
tal subduction and has generally been ascribed to
noncollisional processes (Dewey and Bird, 1970;
Coney, 1973; Burchfiel and Davis, 1975; Hamilton,
1985). The ambiguities of this relatively well-
studied and well- preserved orogen suggest that
plate-tectonics interpretations of contractional
deformation in Archean and Proterozoic Andean-
style orogens may be problematic or indeterminate.
CRITERIA FOR COLLISIONAL VS NONCOLLISIONAL
OROGENESIS
Collisional Orogens
Well-known collisional orogens show great
variability but display consistent geometrical and
chronological patterns. Progressing from the adja-
cent craton toward the orogenic belt, a collisional
orogen displays (1) an external fold and thrust belt
with dominant vergence toward a craton; (2) a
metamorphic core zone characterized by medium
PIT (Barrovian) facies series, including high grade
rocks that have had up to 25-40 km of overburden
removed by erosion and possibly normal faulting
(Platt, 1986); and (3) a suture zone marked by ophio-
lites, blueschists, pelagic sedimentary rocks, and
other evidence of ocean-basin closure.
Modern concepts of collisional orogenesis em-
phasize attempted subduction of a continental
margin and consequent imbrication of that margin
to form an orogenic "wedge" (Davis and others,
1983; Platt, 1986; Jamieson and Beaumont, 1988).
Maintenance of a "critical taper" results in migra-
tion of the front of deformation from the suture zone
toward the foreland with time as the wedge shortens
and thickens. Thickening of the crust may lead to
gravitational collapse manifested by normal fault-
ing within the thickened zone and possibly by late
thrust faulting at the margins of the thickened zone
(Dewey, 1988). Regardless of the mechanics, defor-
mation associated with collision typically lasts for
several tens of millions of years from the time of
initial collision to the cessation of external foreland
deformation (Edelman, 1991a).
Suture zones contain tectonic "flakes" (Ox-
burgh, 1972). Tectonic flakes were originally con-
ceived as being emplaced antithetic to subduction,
both for oceanic flakes involved in ophiolite em-
placement and for continental flakes (Figure 1a,c).
Most subsequent interpretations of suture zones
suggest that flakes are emplaced synthetic to sub-
duction (Figure 1b,d). For example, the type area
of flake tectonics, the S-dipping Austroalpine
nappes of the eastern Alps, is now widely inter-
preted as resulting from southward subduction
beneath Italy (Roeder and Bogel, 1978; Butler,
1986; many others), as in Figure 1b, rather than
northward subduction beneath southern Europe as
suggested by Oxburgh (1972; Figure 1a).
An important result of flake tectonics is re-
moval of the middle and lower crust beneath the
main overthrust mass so that the ancient subduc-
tion system is largely displaced to greater depth and
not exposed. In all models of flake tectonics (Figure
1), the upper crustal suture between collided blocks
is not the old subduction zone, but rather is a new
fault created during the collision process. If con-
vincing stratigraphic or paleogeographic evidence
for large (> 200 km) displacements is lacking, then
a flake thrust fault, marking the site of a cryptic
suture, may be difficult to distinguish from a thrust
fault within a single, imbricated terrane.
The flake-suture is the earliest major structure
associated with collision. Subsequent thrust imbri-
cation, folding, and normal and strike-slip faulting
deform the flake fault, which is already difficult to
recognize as a suture. For example, the flake-suture
in the Alps is folded (Alpine root zone), faulted
(Insubric line), and reworked by normal faulting
(Selverstone, 1988).
Special Problem of Identifying Arc-Continent
Collisions
The hanging wall of collisional subduction must
be an arc, but a well-developed arc may be lacking
due to insufficient duration of subduction before
collision as in the Alps, or to extensive rifting of the
subduction hanging wall as in ophiolites (Edelman,
1988a). The hanging-wall arc would be a continen-
tal-margin, Andean-style arc in a continent-contin-
ent collision, or an intra-oceanic arc in an arc-
continent collision (Dewey and Bird, 1970). When
subduction flips after collision between an oceanic
arc and a continent, ensuing deformation may result
from the collision or from non collisional processes
associated with the flipped subcontinental sub-
duction. Because collisional deformation may last
for tens of millions of years, contractional deforma-
Collisional us noncollisional deformation at continental-margin orogens 199
Continental flakes
Cal
Antithetic flake, lithospheric
wedge (Oxburgh 1972)
Synthetic flake,
lithospheric duplex
Oceanic flakes
............ . .
""ok_ ...... " '::::::::.::: :
........ ..
(c) ."

........ ....
Antithetic flake, ----
lithospheric wedge "'1.
(Coleman 1971)
Synthetic flake,
lithospheric duplex
Figure 1. Schematic cross sections of tectonic flaking processes. The upper crustal suture in each post-collision diagram is not
the subduction zone fault but rather a new fault in the position of the heavy dotted line in each pre-collision diagram. All
geometries result in juxtaposition of collided (sutured) masses along young, collision-related faults. Key to patterns: hachures
= continental crust at passive margin; crosses = continental crust at convergent Andean-style margin; stipple = oceanic
crust; dashed line = Moho.
tion above a subcontinental subduction zone can be
ascribed either to arc-continent collision or to the
subcontinental subduction for tens of millions of
years after initiation of subcontinental subduction.
Thus, the simple presence of a subcontinental
subduction zone does not prove that orogenic de-
formation of the subduction hanging wall is non-
collisional. The presence of previously accreted
oceanic arcs and/or ophiolites in the subduction
hanging wall allows a collisional interpretation. All
Phanerozoic orogenic belts cited by Dewey and Bird
(1970) as noncollisional ("Cordilleran-type"), includ-
ing the classic example of the Mesozoic Cordilleran
orogen of the western United States, display
evidence for early accreted arcs and ophiolites and
thus could be collisional (Edelman, 1991a).
In summary, orogenic belts at continental mar-
gins occur above subcontinental subduction zones,
in contrast to continent-continent collisional oro-
gens within assembled supercontinents. Contin-
ental-margin orogenesis may be due to arc-contin-
ent collision with ensuing reversed subduction
having no orogenic manifestations, to subcontinen-
tal subduction with arc accretion involved only in
initiating the subcontinental subduction, or to a
combination of the two. Phanerozoic continental-
margin orogens contain evidence for arcs and ophio-
lites, but it is difficult to prove that the arcs and
ophiolites are allochthonous, collided masses. This
difficulty arises from complexities in potential
suture zones; the critical evidence for ocean-basin
closure is typically destroyed by tectonic flaking as
well as by subsequent deformation. Therefore,
typical continental-margin orogens developed in
subduction hanging walls and containing fragments
of arc and ophiolitic rocks have inherent interpre-
tational ambiguities in determining the cause of
deformation and the paleogeography of possible
accreted material.
200
S. H. Edelman
Late
Jurassic-
early Tertiary
Figure 2. Generalized map of the
southwestern United States showing
major features of the Mesozoic oro-
gen (including early Tertiary defor-
mation in the Laramide belt). Oro-
genesis could be noncollisional and
related to subcontinental subduc-
tion, represented by the Franciscan
subduction complex and the Sierra
Nevada batholith (SNB), or could be
collisional, with accreted oceanic arc
and ophiolite terranes (black) in the
Klamath Mountains (KM) and Sier-
ra Nevada (SN), as interpreted in the
text.
" .
" " \ Middle Jurassic-
:Y,(Cretaceous
200 kre
REVERSED SUBDUCTION SYSTEM


Subduction zone Continental margin arc
COLLISIONAL OROGEN
':':'J.':;'
Oceanic arcs Metamorphic
and ophiolites core zone
v v v
Foreland fold
and thrust belt
MESOZOIC OF THE WESTERN UNITED STATES
Mesozoic orogenesis in the Cordillera of the
western United States is a frequently cited example
of a noncollisional orogen (Dewey and Bird, 1970;
Coney, 1973; Burchfiel and Davis, 1975; Hamilton,
1985), but it contains all the elements of an arc-
continent collision (Edelman and Moores, 1984).
The paired Sierra Nevada batholith (and its exten-
sions) and Franciscan subduction complex (Figure
2) are a classic example of an ancient continental-
margin arc-trench system (Hamilton, 1969). This
subcontinental subduction coincided with activity of
a continental-margin orogenic belt represented by
the Nevadan, Sevier, and Laramide belts (Figure
2).
Contractional deformation in the Nevadan belt
is Middle Jurassic to Cretaceous; the deformed rocks
include ophiolites, volcanic-plutonic arc complexes,
and oceanic sedimentary rocks. Contractional de-
formation in the Sevier and Laramide belts is Late
Jurassic-early Tertiary; the deformed rocks are
miogeoclinal strata in eastern Nevada and Utah
and deeper-water sedimentary rocks in western
Nevada (including the middle Paleozoic-Triassic
Roberts Mountain and Golconda allochthons).
Metamorphic core complexes in eastern Nevada and
adjacent areas in the "Sevier hinterland" contain
early Tertiary low-angle normal faults; the
metamorphic rocks in the footwalls of the normal
faults display both Barrovian and higher tempera-
ture facies series. The age of metamorphism is Late
Jurassic-Cretaceous (Labotka and Albee, 1988;
Miller and others, 1988; Snoke and others, 1982).
The Barrovian metamorphism, which is older than
regional pluton-related, high-temperature meta-
morphism, reflects Mesozoic crustal contraction and
thickening (Coney and Harms, 1984) similar to that
in collision zones.
Nevadan deformation in the Sierra Nevada
(Figure 2) was ascribed to arc-continent collision by
Moores (1970, 1972) and Schweickert and Cowan
(1975). These workers interpreted Middle-Late
Jurassic volcanic, intrusive, and ophiolitic rocks in
the western part of the Sierra Nevada mountains as
the accreted oceanic arc. In contrast, Davis and
others (1978) and Saleeby (1981) suggested that the
proposed accreted rocks formed as an autochthonous
Collisional us noncollisional deformation at continental-margin orogens 201
Figure 3. Comparative stratigraphic records of the
Early Jurassic continental margin and now-adjacent
oceanic arc fragments. Locations of arc fragments shown
in black in Figure 2. Note ' that the Early Jurassic
igneous event is absent in the Eastern Hayfork and
Calaveras terranes, and late Paleozoic deposition is ab-
sent in the Rattlesnake Creek and Tuolumne River
terranes (except in olistostrome clasts) where J urassic-
Triassic deposits rest directly on Paleozoic ophiolitic
basement . From Gray (1986), Edelman and Sharp
(1989), and references therein.
western facies of a wide continental-margin arc, and
Harper and Wright (1984) and Wright and Fahan
(1988) presented evidence that all basement ter-
ranes in the Klamath Mountains were amalga-
mated before deposition and intrusion of the Middle-
Late Jurassic arc rocks. Edelman and Sharp (1989)
outlined evidence for a similar interpretation for the
Sierra Nevada, thus supporting the view that this
arc is essentially autochthonous and contradicting
the Nevadan collisional models.
However, older basement terranes representing
Early Jurassic oceanic arcs and ophiolites that are
probably remnants of a collided arc system are
present (black areas in Figure 2). The Rattlesnake
Creek terrane in the Klamath Mountains and the
Tuolumne River terrane in the Sierra Nevada
consist of Early Jurassic (200 Ma) arc igneous rocks
that are intruded into and extruded upon disrupted
ophiolitic basement (Figure 3). The Early Jurassic
arc rocks overlie, intrude, and are intercalated with
Triassic-Early Jurassic chert, argillite, and olisto-
strome. The Slate Creek terrane in the northern
Sierra Nevada is a 200 Ma ophiolitic nappe without
known basement (Figure 3; Edelman 1991b).
The 200 Ma igneous event recorded in these
terranes is not recognized in the more easterly
Eastern Hayfork and Calaveras terranes, suggest-
ing that the Early Jurassic arcs did not form in their
present positions. In addition, the late Paleozoic
depositional record in the Eastern Hayfork and
Calaveras terranes is missing in the Rattlesnake
Creek and Tuolumne River terranes. These strati-
W Kla.math Mountains
and Sierra Nevada:
Sierra Nevada Ranlesnake Cleek
Slate Creek and Tuolumne River
terra e teuanes
Ea rl), .Iu r , <> ;c
OPHIOLITE
T ri:\'\i';.ic
Per III i ;t n
,\1 is si ssi l>pia n
E
Ktamalh
Mountains and
Sierra Nevada'
Eastern Haylorl(
and Calaveras
terranes
ARGILLITE,
LIMESTONE, VOLCANIC
ROCKS
--+

CONTINE TAL
MARGIN
EARLY
JURASSIC
-- OCEANIC
ARCS
graphic contrasts cannot be explained by facies
changes and differ from the Middle-Late Jurassic
arc rocks, which intrude or overlie all Klamath-
Sierra terranes (Davis and others, 1978; Snoke and
others, 1982). The stratigraphic evidence for sub-
stantial transportation of the Early Jurassic arc and
ophiolite terranes, and for their formation in intra-
oceanic environments, suggest they are remnants of
a collided oceanic arc complex. Juxtaposition of the
200 Ma arc complex with the Eastern Hayfork and
Calaveras terranes is constrained by cross-cutting
plutons to have occurred before = 170 Ma, in the
late Early or early Middle Jurassic (Gray, 1986;
Wright and Fahan, 1988; Edelman and Sharp,
1989).
Saleeby (1981) cited the absence of major sub-
duction complexes in the Sierra Nevada as evidence
against collision and in favor of a single, imbricated
arc. Edelman (1988b) outlined a tectonic flake in-
terpretation of Nevadan structure which explains
sharp fault contacts in the positions of sutures. The
cross sections in Figure 4 show the upper crustal
flake structure of the Klamath Mountains and the
Sierra Nevada. The "earlier faults" (Figure 4) may
be interpreted as imbrication of a single wide arc,
but the stratigraphic contrasts across some of these
faults (Figure 3), combined with the possibility that
collisional flaking could replace an old subduction
zone with a new, relatively unspectacular fault
(Figure 1), suggest a collisional interpretation.
The Sevier and Laramide deformations have
been widely ascribed to the subcontinental subduc-
202
S. H. Edelman
w

I
1,0 2,0 3p km
W
T
Nevadan (post-Kimmeridgian) fault
E
: :::: : EK:: :.: : : :: : :::
E
NS
Earlier fault
Figure 4. Generalized cross sections of two segments of the Nevadan suture zone. See Figure 2 for locations of Klamath
Mountains and Sierra Nevada. Earlier faults bound relatively thin thrust sheets or tectonic flakes. Middle-Late Jurassic,
post-amalgamation arc and ophiolite assemblages are represented by horizontal line pattern and include the Josephine
ophiolite, Galice Formation, and related units (JG), the Western Hayfork terrane (WH), and the Smartville Complex, Logtown
Ridge Formation, and related units (SL). Early Jurassic, possibly collided arcs and their basements are represented by the
Rattlesnake Creek (RC), Tuolumne River (TR), and Slate Creek (SC) terranes. Key to other terranes: C = Calaveras; CM =
Central metamorphic; EH = Eastern Hayfork; EK = Eastern Klamath; FR = Feather River; NF = North Fork; NS =
Northern Sierra; RA = Red Ant; SF = Stuart Fork; T = Trinity. Klamath section from Irwin (1977), with modifications
based on Gray (1986) and Wright and Fahan (1988); Sierra Nevada section modified from Edelman and others (1989) and
Edelman and Sharp (1989).
tion associated with the Sierra-Franciscan arc-
trench system (Dewey and Bird, 1970; Coney, 1973;
Burchfiel and Davis, 1975; Allmendinger and Jor-
dan, 1981; Hamilton, 1985). In contrast, Edelman
and Moores (1984) suggested that Sevier and Lara-
mide deformation may be associated with Nevadan
collisions. Their model encountered some chrono-
logical problems because the Late Jurassic Nevadan
collision involved in their model was too young to
explain some features. The evidence for earlier
collision (late Early-early Middle Jurassic) cited
above alleviates these problems.
The configuration of an Early-Middle Jurassic
arc-continent suture zone, a Late Jurassic-Creta-
ceous Sevier hinterland metamorphic core zone, and
a Late Jurassic-early Tertiary Sevier-Laramide
foreland fold and thrust belt (Figure 2) contains all
the major geometrical-chronological elements pre-
dicted by an arc-continent collisional model. The
Sierra-Franciscan arc-trench system would repre-
sent the post-collisional reversed subduction. A
problem with the collisional model is that deforma-
tion lasted for "" 120 million years ("" 180-60 Ma)
which is longer than better-known collisional oro-
gens (Edelman, 1989a). The later, Late Cretaceous-
early Tertiary stages of deformation may be related
to noncollisional processes such as shallow subduc-
tion (Coney and Reynolds, 1977), transpression (Ave
Lallement and Oldow, 1988), and extensional
collapse of the collision(?)-thickened lithosphere
(Dewey, 1988). It is also possible that the arc and
ophiolite collisions described above produced little
deformation and noncollisional deformation was
dominant throughout the history of the orogen, or
that the Cordilleran orogen is a very long-lived
Collisional us noncollisional deformation at continental-margin orogens 203
collisional orogen. No obvious method exists for
distinguishing these alternatives.
CONCLUSIONS
Unless evidence for accreted arcs and ophiolites
is demonstrably lacking at a continental-margin
orogen, a fundamental uncertainty exists regarding
whether orogenesis was due to arc-continent colli-
sion or subcontinental subduction. Tectonic flaking
during collisions and subsequent contractional,
extensional, and strike-slip disruption obscure the
distinctions between accreted and autochthonous
arcs in potential suture zones.
The Mesozoic orogen of the western United
States contains all the major geologic features of an
arc-continent collision, although orogenesis oc-
curred in the hanging wall of a subcontinental sub-
duction zone. This orogen and other examples of
Phanerozoic continental-margin orogens are per-
missively arc-continent collisions. Determination of
plate-tectonics origins of contractional deformation
at continental-margin orogens is a major problem,
and new approaches are needed to distinguish
collisional vs noncollisional models. Interpretations
of the plate-tectonics settings of Archean and
Proterozoic Andean-style orogens should consider
both collisional and noncollisional models.
ACKNOWLEDGMENTS
Discussions with Eldridge M. Moores and
James S. Beard helped formulate these ideas, and
discussions with George V. Albino, M. E. Bickford,
Howard W. Day, Gary G. Gray, Bradley R. Hacker,
Scott R. Paterson, Jason B. Saleeby, Richard A.
Schweickert, and Warren D. Sharp contributed to
my knowledge of Sierra-Klamath geology. Manu-
script preparation was supported by a grant from
the University of Tennessee to Robert D. Hatcher,
Jr. Reviews by David Alt and anonymous reviewers
were very helpful.
REFERENCES
ALLMENDINGER, R. W., and T. E. JORDAN, 1981, Mesozoic
evolution, hinterland of the Sevier orogenic belt: Geology,
v. 9, p. 308-313.
AVE LALLEMENT, H. G., and J. S. OLDOW, 1988, Early
Mesozoic southward migration of Cordilleran trans-
pressional terranes: Tectonics, v. 7, p. 1057-1075.
BURCHFIEL, B. C., and G. A. DAVIS, 1975, Nature and con-
trols of Cordilleran orogenesis, western United States:
Extensions of an earlier synthesis: American Journal of
Science, v. 275-A, p. 363-396.
BUTLER, R. W. H., 1986, Thrust tectonics, deep structure
and crustal subduction in the Alps and Himalayas: Jour-
nal of the Geological Society of London, v. 143, p. 857-873.
COLEMAN, R. G., 1971, Plate tectonic emplacement of upper
mantle peridotites along continental edges: Journal of
Geophysical Research, v. 76, 1212-1222.
CONEY, P. J., 1973, Non-collision tectogenesis in western
North America; pp. 713-727 in D. H. Turling and S. K.
Runcorn (eds.), Implications of Continental Drift to the
Earth Sciences: Academic Press, New York, New York,
USA, 1184p.
__ , and T. A. HARMS, 1984, Cordilleran metamorphic
core complexes: Cenozoic extensional relics of Mesozoic
compression: Geology, v. 12, p. 550-554.
__ , and S. J. REYNOLDS, 1977, Cordilleran Benioff zones:
lVature,v.270,p.403-406.
DAVIS, D., J. SUPPE, and F. A. DAHLEN, 1983, Mechanics of
fold-and-thrust belts and accretionary wedges: Journal of
Geophysical Research, v. 88, p. 1153-1172.
DAVIS, G. A., J. W. H. MONGER, and B. C. BURCHFIEL, 1978,
Mesozoic construction of the Cordilleran "collage", central
British Columbia to central California; pp. 1-32 in D. G.
Howell and K. A. McDougall (eds.), Mesozoic Paleogeo-
graphy of the Western United States: Society of Econo-
mic Paleontologists and Mineralogists, Pacific Coast Paleo-
geography Symposium 2, 573 p.
DEWEY, J. F., 1988, Extensional collapse of orogens: Tec-
tonics, v. 7, p. 1123-1139.
__ , and J. M. BIRD, 1970, Mountain belts and the new
global tectonics: Journal of Geophysical Research, v. 75, p.
2625-2647.
DICKINSON, W. R., 1971, Plate tectonic models for orogeny
at continental margins: lVature, v. 232, p. 41-42.
EDELMAN, S. H., 1988a, Ophiolite generation and emplace-
ment by rapid subduction hinge retreat on a continent-
bearing plate: Geology, v. 16, p. 311-313.
__ , 1988b, Continental growth by stacking, folding, and
reverse faulting of oceanic tectonic flakes, western Sierra
Nevada metamorphic belt, California [abstract]: Geologi-
cal Society of America Abstracts with Programs, v. 20, p.
184.
204
S. H. Edelman
___ , 1991a, A critical review of tectonic processes at
continental-margin orogens: Tectonophysics, v. 191, p.
199-212.
__ , 1991b, Relationships between kinematics of arc-
continent collision and kinematics of thrust faults, folds,
shear zones, and foliations in the Nevadan orogen, north-
ern Sierra Nevada, California: Tectonophysics, v. 191,
p.223-236.
__ , and E. M. MOORES, 1984, Late Mesozoic collision
orogen in the western USA [abstract]: Geological Society of
America Abstracts with Programs, v. 16, p. 498.
__ , and W. D. SHARP, 1989, Terranes, early faults, and
pre-Late Jurassic amalgamation of the western Sierra
Nevada metamorphic belt, California: Geological Society
of America Bulletin, v. 101, p. 1420-1433.
__ , H. W. DAY, E. M. MOORES, S. M. ZIGAN, T. P.
MURPHY, and B. R HACKER (eds.), 1989, Structure Across
a Mesozoic Ocean-Continent Suture Zone in the
Northern Sierra Nevada, California: Geological Society
of America, Special Paper 224, 56 p.
GRAY, G. G., 1986, Native terranes of the central Klamath
Mountains, California: Tectonics, v. 5, p. 1043-1054.
HAMILTON, W., 1969, Mesozoic California and the under-
flow of Pacific mantle: Geological Society of America
Bulletin, v. 80, p. 2409-2430.
__ , 1985, Subduction, magmatic arcs, and foreland de-
formation; pp. 259-262 in D. G. Howell (ed.), Tectono-
stratigraphic Terranes of the Circum-Pacific Region:
Circum-Pacific Council for Energy and Mineral Resources.
Earth Science Series 1, 581 p.
HARPER, G. D., and J. E. WRIGHT, 1984, Middle to Late
Jurassic tectonic evolution of the Klamath Mountains,
California-Oregon: Tectonics, v. 3, p. 759-772.
IRWIN, W. P., 1977, Review of Paleozoic rocks of the
Klamath Mountains; pp. 441-454 in J. H. Stewart, C. H.
Stevens and A. E. Fritsche (eds.), Paleozoic Paleogeo-
graphy of the Western United States: Society of Econo-
mic Paleontologists and Mineralogists, Pacific Coast
Paleogeography Symposium 1, 502 p.
JAMIESON, R. A., and C. BEAUMONT, 1988, Orogeny and
metamorphism: A model for deformation and pressure-
temperature-time paths with applications to the central
and southern Appalachians: Tectonics, v. 7, p. 417-445.
LABOTKA, T. C., and A. L. ALBEE, 1988, Metamorphism and
tectonics of the Death Valley region, California; pp. 715-
736 in W. G. Ernst (ed.), Metamorphism and Crustal
Evolution of the Western United States (Rubey Vol-
ume 7): Prentice-Hall, Englewood Cliffs, New Jersey,
USA, 1153 p.
MILLER, E. L., P. B. GANS, J. E. WRIGHT, and J. F. SUTTER,
1988, Metamorphic history of the east-central Basin and
Range province: Tectonic setting and relationship to
magmatism; pp. 650-682 in W. G. Ernst (ed.), Metamor-
phism and Crustal Evolution of the Western United
States (Rubey Volume 7): Prentice-Hall, Englewood
Cliffs, New Jersey, USA, 1153 p.
MOORES, E. M., 1970, Ultramafics and orogeny, with
models of the US Cordillera and the Tethys: Nature, v.
228, p. 837-842.
__ , 1972, Model for Jurassic island arc-continent
collision in California [abstract]: Geological Society of
America Abstracts with Programs, v. 4, p. 202.
OXBURGH, E. R, 1972, Flake tectonics and continental
collision: Nature, v. 239, p. 202-204.
PLATT, J. P., 1986, Dynamics of orogenic wedges and the
uplift of high-pressure metamorphic rocks: Geological
Society of America Bulletin, v. 97, p. 1037-1053.
ROEDER, D. H., and H. BOGEL, 1978, Geodynamic inter-
pretation of the Alps: Inter- Union Commission on Geo-
dynamics, Scientific Reports, no. 38, p. 191-212.
SALEEBY, J. B., 1981, Ocean floor accretion and volcano-
plutonic arc evolution of the Mesozoic Sierra Nevada; pp.
132-181 in W. G. Ernst (ed.), The Geotectonic Develop-
ment of California (Rubey Volume 1): Prentice-Hall,
Englewood Cliffs, New Jersey, USA, 706 p.
SCHWEICKERT, R A., and D. S. COWAN, 1975, Early Meso-
zoic tectonic evolution of the western Sierra Nevada,
California: Geological Society of America Bulletin, v. 86, p.
1329-1336.
SELVERSTONE, J., 1988, Evidence for east-west extension in
the eastern Alps: Implications for the unroofing history of
the Tauern window: Tectonics, v. 7, p. 87-105.
SNOKE, A. W., and D. M. MILLER, 1988, Metamorphic and
tectonic history of the northeastern Great Basin; pp. 607-
648 in W. G. Ernst (ed.), Metamorphism and Crustal
Evolution of the Western United States (Rubey Vol-
ume 7): Prentice-Hall, Englewood Cliffs, New Jersey,
USA,1153p.
__ , W. D. SHARP, J. E. WRIGHT, and J. B. SALEEBY, 1982,
Significance of mid-Mesozoic peridotitic to dioritic intru-
sive complexes, Klamath Mountains-western Sierra Neva-
da, California: Geology, v. 10, p. 160-166.
WRIGHT, J. E., and M. R FAHAN, 1988, An expanded view of
Jurassic orogenesis in the western United States Cordille-
ra: Middle Jurassic (pre-Nevadan) regional metamor-
phism and thrust faulting within an active arc environ-
ment, Klamath Mountains, California: Geological Society
of America Bulletin, v. 100, p. 859-876.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Plutonism across the Tujunga-North American
terrane boundary: A middle to upper crustal view of
two juxtaposed magmatic arcs
J.LAWFORDANDERSON\ ANDREWP.BARTH
2
, EDWARDD. YOUNG
3
,
E. ERIKBENDER\ MARKJ. DAVIS\ DANIEL L. FARBER
5
,
ELIZABETH M. HA YES\ AND KEITH A. JOHNSON
1
lDepartment of Geological Sciences, University of Southern California, Los Angeles, California 90089-0740, USA
2Department of Geology, Indiana University/Purdue University, Indianapolis, Indiana 46202, USA
3Geophysical Laboratory, Carnegie Institution of Washington, Washington, DC 20015-1305, USA
4Department of Geology and Geophysics, Yale University, PO Box 6666, New Haven, Connecticut 06511, USA
5Earth Sciences Board of Studies, U ni versity of California, Santa Cruz, California 95064, USA
(received October 17,1988; revision accepted August 10, 1989)
ABSTRACT
Opportunities to study middle crust magmatic arc construction are uncommon. Petrologic studies in
southern California have revealed middle to upper crustal portions of two juxtaposed Mesozoic magmatic arcs,
one being a native terrane of the Cordilleran orogen and the other the "suspect" Tujunga (or San Gabriel)
terrane. Crystallization thermobarometry of succes:o;i vely intruded and dated plutons provide "crustal nails"
that track the variable depth history of orogenic crust in both terranes. Deep crust is now exposed at the
surface by virtue of two contrasting tectonic settings - lower plates of extensional detachment faults in
Tertiary metamorphic core complexes and upper plates of Mesozoic basement-involved thrust faults.
The goal of this research has been to document the nature of Mesozoic arc development as a function of
crustal depth (from > 25 km to subvolcanic) across the inferred terrane boundary and to utilize this
information to constrain the accretionary history of this region. Timing of the terrane boundary has been
enigmatic and existing interpretations vary from the pre-Middle Jurassic Mojave-Sonoran megashear to
thrust faults of pre-Middle Jurassic to Late CretaceQus age. The Tujunga terrane is everywhere in fault
contact with adjacent units. Plutonism, thrust and detachment faults, and Neogene disruption by the San
Andreas fault have obscured the nature ofthe boundary and the origin ofthe Tujunga terrane.
Our data indicate that the Tujunga terrane has Cordilleran roots regardless of the magnitude of
displacement by bounding faults. A unique feature of the Tujunga terrane is its Middle Proterozoic history,
including the emplacement of a 1.2 Ga anorthosite-charnockite complex into 1.4 Ga granulites. Much of this
apparent "suspect" nature of the Tujunga terrane stems from its partial derivation from the middle crust.
Early Proterozoic, Triassic, Jurassic, and Cretaceous plutonism of the Tujunga terrane are all separately
distinct, yet each has close compositional affinity to, and most are indistinguishable from, that in the adjacent
Mojave Desert region of the Cordillera. Mesozoic and Cenozoic tectonic disruption are thus viewed as intra-
arc events and not the result of original suturing of a long-displaced terrane.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 205-230. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
205
206
.J. L. Anderson, A. P. Barth, E. D. Young, and others
INTRODUCTION
Through evaluation of source heritage, crustal
interaction, and emplacement conditions of Protero-
zoic and Mesozoic plutons, we have sought to con-
strain the timing and nature of the accretionary
history of the "suspect" Tujunga or San Gabriel
terrane relative to native North America. The
Tujunga terrane is exposed in Southern California
and adjacent southwestern Arizona and is one of
several displaced, possibly allochthonous crustal
blocks that occur in the westernmost portion of the
Cordillera. This is an ongoing project, and this
paper builds on the preliminary results presented in
Anderson and others (1990).
Terranes of Southern California
The Baldy and Tujunga/San Gabriel are two
prominent "suspect" terranes in southern Califor-
nia. The terranes are presently juxtaposed by one or
more major thrust faults and are composed of
oceanic (Mp unit, Figure 1) and continental crustal
(SGT unit, Figure 1) rock units, respectively. Their
separate origins remain a major enigma for the
tectonic evolution of this portion of the southern
Cordillera. Our study concerns the origin of the
Tujunga terrane. Several previous workers (Blake
and others, 1982; Howell and others, 1983) used the
term Tujunga terrane more broadly than we have
here (Table 1) and included it in both the San
Gabriel terrane of Silver (1982) and the Joshua Tree
terrane of Powell (1981) - the latter being a
Proterozoic gneiss complex that,structurally under-
lies the Tujunga terrane in the eastern Transverse
Ranges. Our work (Bender, 1989) indicates that the
Joshua Tree terrane has contiguity with the Mojave
Desert region of the Cordillera. Thus, we argue that
Tujunga should not include the Joshua Tree
terrane, leaving the Tujunga and San Gabriel
terranes to be equivalent. Notably, both were
separately named (see Blake and others, 1982;
Silver, 1982) for the same basement complex that
comprises much the San Gabriel Mountains plus
correlative units offset by the San Andreas fault.
Vedder and others (1983) and Champion and others
(1984) argued for continuity between Tujunga and
the Salinian terrane (exposed in the southern Coast
Ranges of California, north of the area depicted in
Figure 1) and included Tujunga as part of their
Santa Lucia-Orocopia allochthon. They concluded
that correlative rocks of the Tujunga terrane are
nonconformably overlain by Late Cretaceous or
Paleocene sediments, an interpretation critical to
their paleomagnetic results. Our data, however,
demonstrate (see below) that much of the Tujunga
crystalline assemblage was at a depth of 25 km
when these sediments were being deposited. Thus,
we see no reason to extend Tujunga into the Coast
Ranges.
Setting and Structural History of the Tujunga
Terrane
The Tujunga (or San Gabriel) terrane is com-
posed of a Proterozoic to Mesozoic igneous and
metamorphic complex presently exposed throughout
much of southern California owing to variable dis-
ruption by the Neogene San Andreas fault. It is
everywhere allochthonous, forming an upper plate
to three different thrust systems. In the San Gab-
riel, Orocopia, and Chocolate Mountains, the Tujun-
ga terrane structurally overlies a Mesozoic oceanic
metasedimentary unit, termed the Pelona/ Orocopia
schist (Baldy terrane of Coney and others, 1980), as
the upper plate to the Late Cretaceous/ early
Tertiary Vincent-Orocopia-Chocolate Mountains
thrust fault (Ehlig, 1981; Haxel and Dillon, 1978).
In the Chuckwalla, Eagle, and Pinto Mountains of
the eastern Transverse Ranges, the Tujunga terrane
occurs structurally above Proterozoic gneisses
comprising the Joshua Tree terrane along the Red
Cloud thrust fault of Powell (1981). Powell recog-
nized that the Red Cloud thrust had two stages of
movement, the younger being pre-latest Jurassic.
Davis and Farber (1989) further constrained the age
of the early Red Cloud thrust to be Proterozoic and
Figure 1. Geologic map of a portion of southern California and adjacent Arizona depicting pre-Tertiary units. Neogene
offset on the San Andreas fault has been restored. Unit designations: SGT, San Gabriel terrane Proterozoic crystalline
rocks; TR, Triassic plutons; J, Jurassic plutons; K, Cretaceous plutons; Mp, Pelona/Orogopia schists; Mm, McCoy
Formation and older Mesozoic strata. Unlabeled areas northeast of the Tujunga terrane are pre-Mesozoic rocks inferred to
be native to North America.
,
N

0
,
Plutonism across the Tujunga-North American terrane boundary
10
,
K
.
Lava
K
J J Marble
MIni .

Amboy ;'\
Bullion Vi
Turlle
Chemehuevi
MIni.
MIn.. t K
'-- ------:J;""
K Sh ...
eep "
K MIni. ,
,
"
\ K
.... CA'oIZ VALLEY r
BATHOLITH River.ide
Mlnl.11I .$Mln .
20 30 40 50Km.
,
I I I
207
208
J. L. Anderson, A. P. Barth, E. D. Young, and others
Table 1. Terrane Terminology of the Southwestern Cordillera.*
San Gabriel Joshua Tree Pelonal
Terrane Terrane Orocopia Schist
This paper SGT/TT JTT POS
Coney (1980) SGT SGT BT(OT)
Howell and others (1983) TT TT BT
Vedder and others (1983) SLOA ? SLOA
Jones and others (1983) TT TT BT
Champion and others (1984) TT TT BT
Siberling and others (1987) TT TT BT
Silver (1982) SGT FTT POS
Dibblee (1982) SGT ? POS
*SGT, San Gabriel terrane; JTT, Joshua Tree terrane; BT, Baldy terrane; OT, Orocopia terrane; POS, Pelona-
Orocopia schist; TT, Tujunga terrane; SLOA, Santa Lucia-Orocopia allochthon.
the late Red Cloud thrust to be Middle Jurassic. In
the Mule Mountains, Mesozoic portions of the Tu-
junga terrane occur above the Late Cretaceous Mule
Mountain thrust fault, overriding Cretaceous meta-
sedimentary rocks of the McCoy Mountains Forma-
tion (Tosdal, 1982, 1984).
The accretionary history of the Tujunga/San
Gabriel terrane has remained speculative in terms
of both timing and mechanism. A portion of the
boundary is overlain by the McCoy Formation,
which is exposed in a structural basin (Figure 1)
bounded on the north and south by overriding thrust
faults with opposed vergence. Harding and Coney
(1985) suggested that the original depositional
basin was the Jurassic(?) Mojave-Sonoran mega-
shear of Anderson and Silver (1979). Other possible
sutures occur south and west of the basin, speci-
fically the above-mentioned Vincent-Chocolate,
Mule Mountains, and late Red Cloud thrust faults.
Paleomagnetic studies led some to suggest that
several outboard terranes, including the Tujunga,
were amalgamated in the Cretaceous and then
transported 1500 km northward prior to early Ter-
tiary accretion to North America (Vedder and
others, 1983; Champion and others, 1984; Morris
and others, 1986). Tosdal and others (1987), how-
ever, concluded that such a magnitude of transport
is unlikely for the Tujunga terrane because of a lack
of significant lithologic contrast across the Mule
Mountains thrust - a finding that is in agreement
with results of our investigation.
Although the project is ongoing in nature, im-
portant findings include the following. (1) Several
Mesozoic and Tertiary crystalline sections in the
region were derived from the middle crust (depths of
origin >20 km) including (a) Whipple and Cheme-
huevi core complexes (Anderson and others, 1988b;
Anderson, 1988; John and Wooden, 1989); (b) Old
Woman and Granite II Mountains (Young and
Wooden, 1988); and (c) portions of the Tujunga ter-
rane, including that in the San Gabriel, Chuckwalla
(pre-Red Cloud thrust units), and Cargo Muchacho
Mountains (Barth and others, 1988; Barth, 1990;
Hayes, 1989; Farber and Davis, 1989; Anderson and
others, 1988a). Plutons in the Old Woman and the
Cargo Muchacho Mountains intrude metamor-
phosed Paleozoic and Mesozoic supra-crustal units,
respectively, implying pre-intrusion crustal thick-
ening on the scale of ;;:: 20 km. (2) The San Gabriel!
Tujunga terrane appears not to have an extra-
Cordilleran origin. Its unusual or "suspect" (Silver,
1982; Blake and others, 1982) features stem prin-
cipally from its partial derivation in the middle
crust. Our work demonstrates that Proterozoic
(Bender, 1989) and Mesozoic (Barth, 1989; Anderson
and others, 1990) plutonism in the San Gabriel
terrane has strong compositional affinities to arc
development native to North America. (3) Cretace-
ous underthrusting of the oceanic Pelona-Orocopia
schist along the Vincent-Chocolate Mountains
thrust and overthrusting of the terrane over the
McCoy Mountains Formation along the Mule
Mountains thrust are intra-arc events (Barth and
others, 1988) unrelated to any inferred accretion of a
"suspect" terrane. (4) Pre-Jurassic evidence of the
Mojave-Sonoran megashear in this region of the
Cordillera (Figure 1) is lacking - a conclusion
echoed by Hamilton (1988).
Plutonism across the Tujunga-North American terrane boundary 209
Details of these results are more fully developed
below. An exciting aspect of the work is the docu-
mentation of deep crustal exposures in both the
Tujunga terrane and native portions of the Cor-
dillera. Geologic sampling of deep crust is derived
from two tectonic environments: the lower plate of
extensional detachment faults of metamorphic core
complexes, and the upper plate of basement-cored
thrust faults. Exceptions to these generalizations
include the lower plate in the Sacramento Moun-
tains core complex and the crustal assemblages
above the Late Cretaceous Mule Mountains thrust
fault, which are both of upper crustal origin 10
km) (Anderson and others, 1988a; Barth and others,
1988). The question of "how deep" appears to be a
function of longevity and style of upward transport
along these zones of crustal shear, whether deve-
loped during regional extension (core complexes) or
compression (thrust faults).
The above conclusion is derived through ther-
mobarometry of pluton emplacement which, when
there are successive stages of intrusion, allows
tracking of ascent and/or descent of crust during
orogenic tectonism. The impetus for this type of
application stemmed from work in metamorphic
terranes where pressure-temperature-time (P-T-t)
paths were obtained from compositional zoning
profiles and prograde inclusions in porphyroblasts
(e.g., Spear and Selverstone, 1983). We are one of a
few research groups who have extended this method
in igneous suites (see also Zen, 1985; Hollister and
others, 1987). The success of the approach is strik-
ing, despite the uncertainties of some of the thermo-
barometers applicable to plutonic rocks.
There is much interest today in the physical
construction of the middle crust, including attempts
to seismically image and drill into deep crust. Our
field-based research has shown that such crust is
locally exposed at the Earth's surface, the study of
which provides a base of ground truth and should
accelerate the benefits derived from these other
projects.
PRE BATHOLITHIC UNITS
Proterozoic Basement: Tujunga Terrane and
Mojave Desert
The Early Proterozoic crystalline assemblages
of the Tujunga/San Gabriel terrane exhibit marked
similarities with respect to rocks of similar age
exposed throughout the Mojave-Sonoran Desert
regions of the southern Cordillera. We recently
completed a number of investigations dealing with
Proterozoic crustal history of the Mojave region
(Anderson, 1983; Anderson, 1987; Thomas and
others, 1988; Orrell and others, 1987; Bender and
others, 1988; Anderson and Bender, 1989; Young
and others, 1989). The oldest rocks in both the
Tujunga terrane (Ehlig, 1981) and the Mojave (An-
derson and others, 1990) include layered gneisses of
probable metasedimentary and metavolcanic origin
that were regionally metamorphosed at high grade
and intruded by synkinematic to late kinematic
foliated granitoids and augen gneisses at "" 1.7 Ga.
The chemistry of these "" 1.7 Ga plutons is in-
distinguishable (see below). In the Mojave Desert
region, these metamorphic rocks record a high am-
phibolite to granulite metamorphic event at low
pressure ("'" 3.5 kb) and subsequently were intruded
by high potassium (rapakivi) granites at 1.4 Ga and
swarms of younger diabase dikes as part of a trans-
continental "anorogenic" magmatic event (Ander-
son, 1983, 1987; Anderson and Bender, 1989).
A key feature of the Tujunga terrane is its
unique Middle Proterozoic history with no 1.4 Ga
granites or Proterozoic diabase swarms. A granulite
metamorphic event was defined by Silver and others
(1963) at 1.4 Ga based on U-Pb dating of zircon from
a pegmatite. Granulites of 1.4 Ga age have not been
recognised elsewhere in North America; however,
such rocks should be abundant at depth as a litho-
logic record of magma generation of the widespread
1.4 Ga granites. Subsequent intrusions include an-
orogenic massif-type anorthosite and charnockite at
1.22 Ga (Carter and Silver, 1972), the only such
occurrence of that age in the western United States.
The anorthosite-charnockite rocks are distinctive
and served as key lithologic elements in constrain-
ing the magnitude of offset by the San Andreas fault
(Crowell, 1962, 1981).
We are not yet able to estimate the depth of
emplacement of the anorthosite complex - notably,
the anorthosite occurs only in sections that were at
middle crustal levels during all or part of the Meso-
zoic (San Gabriel and Chuckwalla Mountains), im-
plying that the anorthositic portion ofthe complex is
deep seated. In contrast, the associated charnockitic
intrusives (jotunite and syenite) are more wide-
spread and occur in areas that were at both deep
(San Gabriel, Chuckwalla, and Little Chuckwalla
Mountains) and shallow (Eagle and Orocopia Moun-
tains) crustal levels by Mesozoic time.
210 J. L. Anderson, A. P. Barth, E. D. Young, and others
Paleozoic Supracrustal Rocks
Paleozoic rocks, transitional from cratonic to
miogeoclinal facies, locally overlie Proterozoic rocks
of the Mojave Desert region. The sequence, occur-
ring in Mesozoic fold and thrust nappes in the Big
Maria and Old Woman Mountains and in thrust
slices in the Little Harquahala and Granite Wash
Mountains, is commonly attenuated and/or dis-
rupted by late Mesozoic tectonic events (Stone and
others, 1983). In a few areas, the section is un-
metamorphosed, such as in the Marble and Provi-
dence Mountains. Elsewhere, metamorphic facies
range from middle greenschist to upper amphibolite
(Hoisch, 1985a,b; Hoisch, 1987; Hoisch and others,
1988) at pressures ranging from 3 kb (Big Maria
Mountains) to >5 kb (Old Woman Mountains). The
section is absent throughout much of the region
owing to removal by pre-Miocene erosion. In several
ranges, strata are preserved that are structurally
overlain by Proterozoic rocks (Spencer and others,
1987).
Although pelitic and carbonate metasedimen-
tary rocks occur in the southeastern San Gabriel
Mountains (Dibblee, 1982; Barth and May, 1987),
their correlation with native Paleozoic sections is
uncertain. Possible Paleozoic rocks have not been
identified elsewhere in the Tujunga terrane.
Mesozoic Supracrustal Rocks
Most stratified rocks of Mesozoic age occur in
the southern portion of the area depicted in Figure
1 and can be divided into four lithostratigraphic
units, as described most recently by Tosdal and
others, (1987). Three of these units occur primarily
in or near the structural McCoy Basin and are only
weakly metamorphosed, including, in ascending
order: (1) a Triassic to lower Jurassic clastic section,
(2) the dominately volcanic Dome Rock sequence of
Jurassic age, and (3) the Jurassic(?) to Cretaceous
McCoy Mountains Formation of Harding and Coney
(1985).
The fourth supracrustal unit is predominantly
an oceanic metasedimentary section now exposed in
an intracontinental position. The unit has been
termed, depending on locality, the Rand, Pelona, or
Orocopia schist and collectively has been referred to
as the Baldy terrane by Coney and others (1980) and
Blake and others (1982), based on exposures that
underlie Mount Baldy of the San Gabriel Moun-
tains. Its Neogene dispersal due to strike-slip fault-
ing in southern California is similar to that of the
structurally overlying Tujunga terrane. The princi-
pal lithology is metagreywacke schist derived from
continental detritus, with only a minor component
of magmatic arc material (Haxel and others, 1987).
Subordinate metamorphosed lithologies include tho-
leiitic (MORB) basalt, chert, marble, and serpen-
tinite. The metamorphism, coeval with develop-
ment of overlying thrust (presumably subduction-
related) faults, was of greenschist to rare blueschist
to lower amphibolite facies at pressures> 8 kb (Gra-
ham and England, 1976; Graham and Powell, 1984).
Thus, like the Tujunga terrane, the Pelona/Orocopia
schist resided at deep crustal levels during the Late
Cretaceous and underwent subsequent rapid up-
ward transport during the early Tertiary. Protolith
age for the schist is constrained to be Mesozoic
(Haxel and Tosdal, 1986; Haxel and others, 1987).
Tosdal and others (1987) suggested that the original
sedimentary basin of the schist protolith may have
originated within the Mojave-Sonoran megashear,
similar to the proposed model for deposition of the
McCoy formation by Harding and Coney (1985).
METHODOLOGY
Why Study Granites?
This effort is directed at the plutonic history of
the craton and adjacent Tujunga terrane. Addi-
tional goals are to evaluate the origin of magmatism
leading to crustal development and to constrain the
depth history (P-T-t) of crust during the orogenic
process. A provocative variation of crustal depth is
exposed on both sides of the terrane boundary, al-
lowing the exciting opportunity to evaluate mid- to
upper-crustal arc construction of two juxtaposed
sections of crust.
Granitoid magmas can undergo major modifica-
tion during ascent and emplacement. Determina-
tion of original magma composition requires "seeing
through" these changes. The problem is often work-
able and important, as the derived data provide the
only evidence available on the age and composition
of the source region that, due to tectonic separation,
may no longer underlie the magmatic arc.
Depths of emplacement are obtainable through
use of several new thermobarometers (see below)
applicable to granitic rocks. Contact metamorphic
rocks are also useful in this regard, but in multiply
Plutonism across the Tujunga-North American terrane boundary 211
<' /
WHIPPLE
MOUNTAINS
WHIPPLE /
I
-I
SANTA
CATALINA
MOUNTAINS
WASH /
33 4km )
(89
E
.x

l-
e...
w
c
30
20
10
AXTEL
29 1 km
(73 3 Ma)
AGE (Ma)
/
/
/ / LEATHERWOOD
I/'- 20.7 1.1 km
(68 !8 Ma)
15 :3.5 km
(47!3 Ma)
MYLONITIZATION
9.3 1.9 km
(-37 18 Ma)
50 100
Figure 2. Upward transport of the Whipple and Santa Catalina core complexes as a function of age. Depth calculated from an
assumed geobarometic gradient of 3.7 kmlkb (for average crustal densities) derived from barometry of plutonic emplacement
and deformational events. From Anderson and others (1988b).
intruded terranes, such as those in this study, such
applications are difficult owing to extensive over-
printing. In contrast, high-temperature mineral
phases of igneous rocks often retain primary com-
positions, despite the occurrence of subsequent plu-
tonism or deformation. Thermochronological data
are also often used to constrain depth history but are
hampered by large uncertainties in assumed ther-
mal gradients. Our approach is more direct and
utilizes dated plutons as crustal nails to constrain
the variable ascent and/or descent of the crustal
section during its tectonic history. An example of
this application is shown in Figure 2, which depicts
the depth-time path for lower plate assemblages of
two metamorphic core complexes - those of the
Whipple and Santa Catalina Mountains (Anderson
and others, 1988b). Barometry for four (Santa
Catalina) or five (Whipple) pluton emplacement and
deformation events, plus ages of earliest unroofing,
track the upward transport of crustal sections since
the Late Cretaceous, with a notable acceleration of
decompression coincident with onset of mid-Tertiary
mylonitization and extension. The depth-time his-
tory of the Whipple complex is well constrained.
Older plutons were emplaced at 73-89 Ma when the
crustal section resided at 29-33 km. Tracking the
upward ascent of the complex, subsequent regional
mylonitization at 26 Ma occurred at 165 km,
followed by Miocene (17-19 Ma) plutonism with
emplacement depths of 5-7 km. Unroofing occurred
after 16 Ma as a consequence of Miocene decom-
pression rates that exceeded 2 mm/yr.
We are using three analytical approaches to
constrain evolution of granitic magmas within the
area (Figure 1): (1) major and trace element ana-
lyses of whole rock samples; (2) Rb-Sr, U-Pb, and
common Pb isotopic analyses of selected samples;
and (3) detailed examination of mineral chemistry
212 J. L. Anderson, A. P. Barth, E. D. Young, and others
through characterization of critical substitutions, 40
evaluation and judicious use of thermobarometry,
and theoretical treatment of mineral-liquid equili-
bria.
.J:l
a..
Isotopic Analyses ; 39
U-Pb ages of zircon separates and Pb-Pb and
Rb-Sr data presented here are being acquired
through a collaborative investigation with Joe
Wooden and Dick Tosdal (U.S. Geological Survey,
Menlo Park, California). The Pb and Sr isotopic
data are critical to model source characteristics and
crustal interaction during ascent. Determination of
isotopic abundances is not only necessary to con-
strain crystallization ages but also to provide in-
formation required to understand magma evolution.
Existing Sr isotopic data are limited. Miller
(1985) and Miller and others (1990) showed that
Cretaceous plutons in the Old Woman and Piute
Mountains have elevated initial 87Sr/8SSr ratios, or
Sri >0.708, and are crustally derived based on iso-
topic Sr, Pb, and 0 data. Anderson and Cullers
(1990) described less radiogenic (average Sri =
0.7077 0.0013) Cretaceous plutons of the Whipple
Mountains core complex and concluded derivation
from a crust-enriched, oceanic reservoir. Likewise,
several of the Jurassic plutons have an elevated Sri
(averaging 0.7065), thus requiring a crustal com-
ponent in their derivation (Young and Wooden,
1989). The only initial Sr data available for the San
Gabriel terrane comes from the Triassic Lowe intru-
sion (Joseph and others, 1982). At 0.70456
0.00006, the value is sufficiently low to preclude an
older, feldspathic crustal component in its evolution
(Barth and Ehlig, 1988).
Wooden and others (1986) and Wooden and
Stacy (1987) provided an extensive set of feldspar
common Pb data for Jurassic and Cretaceous plu-
tons of the Mojave Desert region showing that
Mesozoic plutons have a distinctive isotopic trend,
defining a 207Pb/
2os
Pb pseudo-isochron with an age
similar to that of the Proterozoic basement. Work-
ing with Wooden, we have added new data that are
in concert with the above findings (Figure 3).
Ratios of 208Pb/
2
osPb for Mesozoic plutons also fall
within the field defined by Proterozoic plutons.
Taken together, the data require an extensive crus-
tal component in the origin of these Mesozoic plu-
tons. In contrast, common lead for the Lowe intru-
sion (Davis and Barth, 1985) does not have a crustal
N
.J:l
a..
IX)
o
N
38
17
15.7
.J:l
a..
.,.
o
N
'15.6
.J:l
a..
"-
o
N
15.5

Tujunga K
Granile Min. J/K
Tujunga Tr-J
17.5 18 18.5 19 19.5
206 Pb I 204 Pb
M2 Feldspars

206 Pb I 204 Pb
Figure 3. Comparision of common Pb isotopic composition
of Mesozoic plutons of the Mojave Desert region (North
America) and the Tujunga terrane. The Mojave field is
drawn from data (not shown) of Wooden and Stacey (1987)
and the data points originate from this study in collabora-
tion with Wooden. Cretaceous plutons in both terranes are
similar in thorogenic Pb. Early Mesozoic plutons in the
Tujunga terrane are lower in thorogenic Pb relative to
Mojave. A single pluton in the Granite II Mountains also
shows low 2osPb/
2
0
4
Pb and is anomalous relative to other
Mesozoic plutons of the Mojave Desert.
signature (Figure 3), which is consistent with the
inferred subcrustal origin of this large batholith.
Plutonism across the Tujunga-North American terrane boundary 213
Apart from common Pb, there has been no dedi-
cated study to amass the isotopic composition of
Mesozoic plutons in this region of southern Califor-
nia. The Pb data are indicative of a major crustal
role, but it is unclear why the same Pb trend exists
for both Cretaceous and Jurassic igneous suites of
the Mojave, given their compositional differences
(see below).
Mineral Chemistry
Compositions of granitoid mineral phases are
sensitive to P, T, fluid composition, and interaction
with coexisting liquid and other solid phases.
Chemical variation in biotite, calcic amphibole,
white mica, and garnet are particularly useful in
elucidating changes in these parameters because
they exhibit multi-site exchanges. Dominant sub-
stitutions that are bulk composition-independent
can be attributed to changes in intensive para-
meters and equilibria with coexisting phases.
Several differences in the chemistry of the
above mineral phases occur, due in large part to
temperature and depth of crystallization. Because
magmas can crystallize over a range of pressure
during their crustal ascent, pressure-sensitive ex-
changes will be modified as a mineral grows from
core to rim. Other variations can be correlated with
changes in phase assemblage along the crystalliza-
tion path. With these factors in mind, the following
sections review the approach taken in this study to
extract crystallization history as recorded in the
composition of the magmatic mineral phases.
Thermobarometry
The extensive mineral compositional data accu-
mulated during the course of this research afford the
opportunity to not only apply but also to test and
refine numerous thermometers and barometers in
granitoids. Several thermobarometers we use have
been thoroughly tested only in metamorphic rocks
but, thermodynamically, are rigorously applicable
to liquid-bearing systems. To accurately estimate
emplacement depth, the phase compositions must
represent near solidus conditions, as the solid
phases in a magmatic system could have been ac-
quired at any stage during ascent, if not from the
restitic source. Complete characterization of com-
positional variations in the phase assemblage is a
fundamental prerequisite to any of the barometric
applications described below.
Many igneous rocks lack pressure-sensitive
phase assemblages, yet a prominent example in
peraluminous granites is the assemblage garnet,
muscovite, plagioclase, and biotite (Ghent and
Stout, 1981; Hodges and Royden, 1984). The equili-
bria are fluid independent and involve an increase
in the grossular component in garnet at the expense
of the anorthite component in plagioclase with
increasing pressure. Although originally intended
for pelitic schists, the barometer appears to work
well for granitic rocks (Anderson, 1985; Anderson,
1988; Anderson and others, 1988b) and can be used
in concert with the garnet-biotite thermometer of
Ferry and Spear (1978) or, for more manganiferous
garnets, with a suggested modified formulation by
Ganguly and Saxena (1984). For the same rocks, we
obtain similar, though less precise, pressure esti-
mates with the phengite barometer (based on the
silica content of muscovite in equilibrium with bio-
tite and K-feldspar) of Massonne and Schreyer
(1987). A previous calibration of this barometer by
Powell and Evans (1983), however, appears to yield
estimates 1-3 kb too high. We are uncertain of the
cause, but the Massonne and Schreyer formulation
also gives high results when the ferrimuscovite
component in muscovite exceeds a weight fraction of
O.l.
The presence of magmatic epidote is suggestive
of minimum crystallization pressures of 4-6 kb, the
lower limit variation being a function of oxygen
fugacity (Zen and Hammarstrom, 1984; Zen, 1985).
The significance of this mineral as an indicator of
pressure must be evaluated as both supersolidus
and subsolidus growth stages can occur. Chemo-
graphic relations (Figure 4) demonstrate pressure
dependence and paragenetic growth of magmatic
epidote. With decreasing temperature, epidote
should have a reactive relationship with horn-
blende, leading to its stable coexistence with biotite
(as shown by Naney, 1983). Many of the high
pressure plutons encountered in this study have
magmatic, or probable magmatic, epidote, including
the Lowe, Waterman, and Josephine intrusions of
the San Gabriel Mountains (Barth, 1990), the Old
Woman granodiorite of the Old Woman Mountains
(Young, this report), the Corn Springs granodiorite
of the Chuckwalla Mountains (Davis and Farber,
1989; Farber and Davis, 1989), and the Whipple
Wash and Axtel intrusions of the Whipple Moun-
tains (Anderson, 1988).
214
J. L. Anderson, A. P. Barth, E. D. Young, and others
[Iiq]
k

[hb]

k



A


[bio]
KCFFASH
temperature
Figure 4. Chemographic analysis of the KCFFASH system projected from quarz, feldspar, magnetite, and H
2
0, elucidating
possible equilibria amongst analyzed epidote (e), hornblende (h), biotite (b), plagioclase (p), and granitic liquid (1). darkened
field encloses the compositon of a metaluminous granodiorite from the Old Woman Mountains. A = A1
2
0
3
-K
2
0-CaO; F =
FeO-Fe203; and C = CaO.
For metaluminous granites, barometric esti-
mates are possible based on the solubility of total
aluminum in hornblende (Hammarstrom and Zen,
1986; Hollister, and others, 1987; Rutter and Wyllie,
1988). We observe that total Al and Fe/(Fe + Mg)
are positively correlated and that iron-rich horn-
blendes yield erroneously high pressures. We are
working on an empirical modification of the baro-
meter that will compensate for the apparent Fe/Mgl
influence, but meanwhile we are restricting its
usage to hornblendes of intermediate composition
(Fe/(Fe+Mg) = 0.42 to 0.58). The revision by Hol-
lister and others (1987) lessened the uncertainty to
1 kb over a pressure range of 2-8 kb. Subse-
quently, Rutter and Wyllie (1988) confirmed that
calibration at 10 kbar for hornblende coexisting
with quartz and K-feldspar. As a test case, we com-
pleted a comparison with pressures derived from
contemporaneous low-variance metamorphic min-
eral assemblages. In the southwestern San Gabriel
Mountains, amphibolite to granulite-facies gneisses
are intruded by synkinematic hornblende tonalites.
Metamorphic pressures were estimated at 4 1kb
based on GAR-CORD, GAR-BIO, and GAR-PLAG-SIL-
QZ equilibria (Barth and May, 1987). Hornblende
barometry yielded 4.20.3 kb. We have established
similar, within 1 kb, correspondence of hornblende
crystallization and contact metamorphic barometry
for two other plutons, one in the eastern San Gabriel
Mountains and the other in the Coxcomb Moun-
tains.
To estimate crystallization temperatures, we
have had some success with integrated two-feldspar
thermometry (Whitney and Stormer, 1977; Hazel-
ton and others, 1982; Fuhrman and Lindsley, 1988)
and with the Aliv hornblende thermometer of
Nabelek and Lindsley (1985). Where we obtained
good integrated compositions of K-feldspar, the two
thermometers have yielded similar results with the
Aliv thermometer averaging = 40C higher.
Plutonism across the Tujunga-North American terrane boundary 215
Mineral-Liquid Equilibria
Early crystallizing phases that exhibit growth
zoning (e.g., garnet, hornblende, and plagioclase)
potentially record changing intensive parameters
during large parts of the ascent and crystallization
history of magmas. Past approaches to recovering
this information utilized mineral-melt exchange
reactions (e.g., Helz, 1979) and are not generally
applicable to plutonic rocks where equilibrium melt
compositions (cf. Criscenti and Ghiorso, 1985) and
melt activity models are difficult to obtain.
The utility of simultaneous consideration of
heterogeneous equilibria (Gibbs method) has been
demonstrated in metamorphic rocks (Rumble, 1976;
Grew, 1981; Spear and Selverstone, 1983). The
strength of this technique lies in its relative insen-
sitivity to uncertainties in thermodynamic data and
compositional variability ofthe melt phase.
An example of application to a two-mica + gar-
net granodiorite from the Whipple Mountains is
summarized in Figure 5. The calculated (ap/aT)x for
isopleths of grossular component in garnet (Xgr)
and albite component in plagioclase (Xab) are
parallel with lines of constant Kn defined by the
empirically derived thermobarometer of Ghent and
Stout (1981). The method (not shown) demonstrated
that zoning in hornblende from foliated quartz
diorites of the Granite II Mountains is consistent
with formation prior to crystallization of quartz
(Young, in prep.). Such comparisons also aid in
identifying mineral compositions of magmatic ori-
gin, where deformation and/or thermal overprinting
occurred, and in the identification of probable dis-
equilibrium assemblages.
MAGMATIC CONTRASTS ACROSS THE
TERRANE BOUNDARY
Early Proterozoic Plutonism
On a global basis, the late Early Proterozoic
(1.7-1.9 Ga) was a time of major orogenic continen-
tal growth characterized by regional metamorphism
of variable grade, construction of mafic to felsic
metavolcanic belts, and synorogenic intrusion -
predominately mantle-derived calcalkaline plutons.
The Proterozoic portion of the North American cra-
ton records this same event; age provinces ranging
from 1.8-1.9 Ga occur in Canada and the Lake Su-
perior region and young southwestward to 1.7-1.8
'"

::IE

"
'"
'"
'"
a:
Gornet- Plog/oc/o s.

800
700
600
0.02 (eJp/dXGr

500 600 700 800 900
Temperature (OC)
Figure 5. Gibbs method calculations for garnet-
plagioclase composition isopeths in peraluminous granitic
magma. The sytem is composed of melt + GAR + BIO +
PLAG + MUSC + KSP + QTZ, and the isopleths are for
mole fraction grossular component (XGr). The heavy
isopleth (XGr = 0.26) is an average value for garnet in a
Cretaceous pluton from the Whipple Mountains core
complex. Similar isopleths can be drawn for the albite
component in plagioclase (XAb), the spacing of which is
designated. The dashed curve is the approximate high-
temperature limit of KSP + QZ (Naney, 1983), the high-
temperature curve is the terminal stability of MUSC +
QTZ, and the low-temperature curve is the water-satura-
ted granodiorite solidus.
and 1.6-1.7 Ga. The Mojave crustal province of
Wooden and Miller (1989) and Anderson and others
(1992) encompasses southeastern California and
adjacent portions of western Arizona and southern
Nevada and is strikingly atypical relative to the
remainder of the continent. There are no mafic
metavolcanic suites and orogenic plutonism (U-Pb
dated at 1.66-1.74 Ga) is void of any calc-alkaline
members. Variably deformed to foliated granitic
and augen gneisses, the plutons are all high or
ultra-high K. Bender (1989) recently completed a
comparitive study of Early Proterozoic plutons in
the southwestern United States and found that
those of the Joshua Tree and Tujunga/San Gabriel
terrane are indistinquishable from those of the
regionally unique Mojave province. Figure 6 com-
pares the Si0
2
-K
2
0 and Rb-(Y +Nb) compositional
216 J. L. Anderson, A. P. Barth, E. D. Young, and others
fields of Mojave, Joshua Tree, and San Gabriel
plutons to those of central Arizona, the latter being
akin to the remainder of the craton. The correspon-
dence between the Mojave, Joshua Tree, and San
Gabriel plutons is further tied by other data (not
shown), demonstrating that all are A-type grani-
toids (as defined by Whalen and others, 1987), inclu-
ding high Ti and other high field-strength elements,
Ba, rare earth, and other large ion lithophile
elements, and high FeIMg ratios at any silica level.
Triassic Plutonism
Plutons of Triassic age are not abundant in the
Cordillera. U-Pb dated plutons in the Sierra Neva-
da have ages of 200 to 215 Ma (Bateman, 1983;
Saleeby, 1981; Stern and others, 1981; Chen and
Moore,1982). Triassic plutons in the Mojave Desert
region include a K-Ar dated, 194 Ma pluton in the
San Bernardino Mountains, a U-Pb dated, 230 Ma
pluton in the adjacent Granite I Mountains (Miller,
1977, 1978), and an = 220 Ma dated pluton (U-Pb
model age based on an assumed upper intercept, J.
Wooden, pers. comm., 1987) in the Joshua Tree
National Monument (Brand and Anderson, 1982;
Brand, 1985). [Note: there are four Granite Moun-
tains in the Mojave Desert: the Granite I Mountains
occur just to the west of the area depicted in Figure
1. The areal extent of two others is shown in Figure
1; these have been termed Granite II and Granite III
Mountains.] The Mojave-occurrence rocks are
strikingly distinctive relative to the far more abun-
dant, typically calcalkaline, Cretaceous plutons by
which they are often intruded. Metaluminous, high
K, and low silica, the plutons are principally mon-
zonite, with lesser amounts of monzodiorite and
diorite. Megacrystic K-feldspar is common. The
principal mafic minerals are clinopyroxene, horn-
blende, and biotite; a quartz monzonite in the
Granite I Mountains contains calcic (andraditic)
garnet + hornblende. Compositionally, most plu-
tons are alkalic, with a shoshonitic affinity (Figure
7). High abundances of Ba, Sr (> 1000 ppm), and
LREE are characteristic. For plutons in the San
Bernardino and adjacent Granite I Mountains,
Miller (1978) convincingly argued that the magmas
were derived from a LILE-enriched, quartz eclogite
Transverse Ranges/MOjave/Arizona
ORG
55 60 65
Si02 (wt. %) Y + Nb (ppm)
San Gabriel IiIJoshua Tree [Z]Arizona
Figure 6. Comparision of KzO versus SiO
z
and Rb versus Y + Nb for Early Proterozoic plutons of Tujunga/San Gabriel and
Joshua Tree terranes (southeastern California), Mojave Desert region (California, western Arizona, and southern Nevada),
and central Arizona. Data from Bender (1989). Solid lines on KzO-SiOz diagram separate low-K, medium-K, high-K, and
ultra high-K fields of Gill (1981). Compositional field labels on the Rb + Y + Nb diagram from Pearce and others (1984): SYN-
COLG = syncollisional granites; VAG = volcanic arc granites; WPG = within-plate granites; ORG = ocean-ridge granites.
Plutonism across the Tujunga-North American terrane boundary 217
source. Hornblende barometry (calibration of Hol-
lister and others, 1987 ) for the Triassic(?) pluton in
the Joshua Tree Monument indicates emplacement
at 3.9 0.9 kb.
If the Tujunga terrane is far traveled with
respect to North America, then it is a remarkable
coincidence that the terrane contains Triassic plu-
tonism with affinity to those described above. The
Mount Lowe intrusion is an immense, batholith-
sized (>300 km2), zoned plutonic complex that
occurs in the San Gabriel Mountains (Ehlig, 1975).
Dated at 220 10 Ma (Silver, 1971), the magmatic
evolution and emplacement history of the complex
served as the basis of an ongoing dissertation study
(Barth, 1987; Barth and Ehlig, 1988; Barth, 1990).
In addition, Barth and others (1988) demonstrated
petrologic correlation to other exposures east of the
San Andreas (Chocolate, Little Chuckwalla, Mule,
and Trigo Mountains), as originally suggested by
Dillon (1976) and Crowell (1981). The zoned mar-
ginal facies of the complex is composed of horn-
blende quartz monzodiorite and quartz monzonite,
grading inward into garnet-biotite quartz mon-
zodiorite and granodiorite. A younger, central
Mojave-Tujunga Mesozoic Plutons
10
15
c
c
c
0

0
I
10
Alkal ic
0
0

0
-'
N

'" 5
Z
0

+
0
5

sub - Alkalic
0
0
40 50 60 70 So 40 50 60 70 80
SiO:a wt70 Si0
2
wt70
10
3000
+ Mojave K
0

,
0
c


Tujunga K
od' c *
*
oC
++
'l,

,
0 Mojave J
0

2000 c

Tujunga J
0
0
d'
0
Qd

S
:::a
"0
0 0
0
0 Mojave Tr
.........
0
5
8 s
0
' 0
+ cc
9", 0
*
Tujunga Tr
1-0
0

CIl til C tJ 0
'"
Cz.
."0
0d'
c
'b
o

'"
0
a 0
40 50 60 70 80 a 100 200 300 400
Si0
2
wt70 Rb ppm
Figure 7. Compositional comparison of Triassic, Jurassic, and Cretaceous plutons of the Tujunga terrane and the Mojave
Desert region. Sources of data include this study and others given in the text. Grid on Si0
2
-K
2
0 plot is from Gill (1981);
alkalic/subalkalic boundary on the Si0
2
-(K
2
0 + Na20) plot is from Irvine and Baragar (1971); and calcalkaline/tholeiitic
boundary on the Si0
2
-(FeOIMgO) plot is from Miyashiro (1974).
218 J. 1. Anderson, A. P. Barth, E. D. Young, and others

:;t 1.5
Brilltol lakfl Fi:e9ion _ _ ______
1,
Gronllo II
Old Wom"n Jurassic:
_ ... ... Cretaceous_ ... ... ................ _6""
C,anitell
C,etaceou$

----- --Providonco - - - - .l_
JurassIc
------- ------------- p-
o
o
Teutonic Baloolith
_________ el<J>

C,etaceOuS
A)- RoCk Spring
0Y'
______ j_ _ __________ .l _
Mid Hills
o
- - - - - - - K .... i., Spring - - - - _0 _
Joshua 1re8 ................................. ... Ill\b
_ _ ______
Palma
Gold P.rk
_ ____________________ -'-
______________________ 0_
0.5 '-__ -'-___ -'-__ --' ___ -'-__ -.J
o 0,2 0.4 0.6 0.8 o 0.2
0.' 0.'
0.8 I 0 0.2 0.' 0.6 0.8
FE2+ / MG + fE2+ FE2+/ MG + FE2 + ,E2+ / MG + ,E2+
2.5 r-----,----.,-----y----,----
[ogles .......................... _ ...

ChuCkwOllos ............... ... _ ............ _ .8lsb
_______________ _6_
COl'qo I.!ucl'locl\o ... _ ............ _ ...... _ ...... 8.kb
... Jc
dloril.
- - - - - - - - - - - - - _lI..
:;t 1.5

- - - - - - - - - - - - -------- _.-
________ __ ______ -'-

_____________________ -'-
_____________________ Sl_
--------------------- _Q
a.5
0
L ----,a:': .':-----,o=" . ,---o=' . :----::o.'= . - --.J
0.2 0. ' 0.6 0.8 I 0 0. 2
0.' 0.6 0 .
FE2+ / MG + FE2+ FE2+/ MG + ,E2+
FE2+ / MG + FE2+
2.5,--_-..,. ___ -,-___ .__---,----
),Aule - Trigo ......... _ ... _ ............... ...
__________________ ___ 61<1>.
:;t 1.5
____ _ ____ -'"
_ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ ..2!1b_
_____________ _______ Sl __
Son ",,"". , ?:# ______ aJ ____ __ u _lIJ<b
cr. T,_1c
CN
__________ _____ _ _____ _____ _ lI..
....-c x
_________ _____ _ _ ______ __ ______ _ 4_
___________ ____ __ ________ ______ .:t _
- - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - _Q
. ,----::0.'=6- --::a': .8,----..J1 .
FE2 +/ MG + 'E2+
Fe +2/ (Mg + Fe2 +}
Figure 8. Composite of hornblende rim chemisty and barometry for eight areas of the Mojave Desert and the Tujunga terrane.
Sources of data include Anderson and others (1988a), Young and Wooden (1988), Barth and others (1988), Farber and Davis
(1989), and Hayes (1989). Al = total aluminum atoms per 23 oxygens and isobars from the calibration of Hollister and others
(1987).
Plutonism across the Tujunga-North American terrane boundary 219
intrusion is composed of leucocratic biotite and
diorite and monzodiorite. Like the similar-aged plu-
tons of the craton, the Mount Lowe is metalumin-
ous, low in silica, and enriched in alkalis, Ba, and Sr
(>1000 ppm in the more primitive members). The
main differences from those described above include
lower K (Figure 7) and higher Na (to >7 wt %);
some of the more evolved rock types could be
trondhjemitic except for their low Si0
2
. Once again,
there is the unusual (with respect to this portion of
the Cordiliera) occurrence of calcic garnet ($ 45
mole % grossular + andradite) + hornblende.
Barth and Ehlig (1988) have independently, but for
the same general reasons given in Miller (1978),
suggested an eclogite-melting model for the origin of
the magmatic suite.
One of the most striking findings of Barth's on-
going study has been documentation that the Lowe
and all younger phases of Mesozoic plutonism in the
San Gabriel Mountains were emplaced in the
middle crust. Calculated pressures range from 5.5
to 6.6 kb (Figure 8), consistent with several attri-
butes of deep-seated crystallization, including
markedly aluminous hornblende (to > 11 % AI
2
0
3
),
calcic grarnet, magmatic epidote, and siliceous pri-
mary muscovite (in Cretaceous granites). In con-
trast, the offset exposures of the Lowe east of the
San Andreas fault record shallower levels of em-
placement at == 5-6 kb in the Little Chuckwalla
Mountains (Farber and Davis, 1989) and 4-5 kb in
the Mule Mountains (Barth and others, 1988). In
this region, the Lowe complex resides above the Red
Cloud and Mule Mountains thrusts, implying that
prior to San Andreas disruption, these faults de-
tached the upper portions of an intrusion that once
vertically occupied some 6-10 km of crust.
Jurassic Plutonism
By Middle Jurassic time, magmatic arc con-
struction in the southern Cordillera and in the
Tujunga terrane became a major feature of the oro-
gen. Figure 1 depicts the regional extent of the
event broadly dated at 180-155 Ma, plus a minor
pulse at 145 Ma (John, 1981; Powell, 1981; D. M.
Miller and others, 1982; Allen and others, 1983;
Tosdal and others, 1989; Young and Wooden, 1988).
Magmatism here is part of a well-defined arc trend-
ing southeast from the Klamath Mountains and
northern Sierras into the White-Inyo Mountains
and extending south across the Mojave Desert re-
gion into southern Arizona and Sonora. The Juras-
sic plutons throughout the area shown in Figure 1
comprise a compositionally expanded suite with
silica contents that vary continuously (48-76 wt %).
Rocks from all areas are predominantly metalumin-
ous and range from hornblende-clinopyroxene
gabbro and diorite to biotite-hornblende quartz
monzodiorite, quartz monzonite, and biotite sphene
monzogranite or syenogranite. The granitic rocks
are commonly coarse grained and seriate to por-
phyritic with large, lavender-colored alkali feldspar
phenocrysts. Like the Triassic suite, the Jurassic
plutons have high Ba and alkalis, modest levels of
Si0
2
, and intermediate FeIMg ratios (data largely
straddle the calcalkaline/tholeiitic boundary of
Miyashiro, 1974). A major difference is lower Sr,
usually < 700 ppm.
Although the data show considerable overlap,
some distinction can be made between plutons of the
Tujunga relative to those of the adjacent Mojave
Desert. A few Tujunga plutons, such as those of the
Cargo Muchacho Mountains (Hayes, 1989), have
lower K, total alkalis, and FeIMg ratios, but others
in the Chocolate, Chuckwalla, and Eagle Mountains
(Tosdal and others, 1987; Farber and Davis, 1989;
Johnson, in prep.) are as potassic and iron-rich as
those of the Mojave Desert region (Figure 7). In
terms of age, the Jurassic magmatic suite is com-
positionally distinct and, despite differences noted,
we see minimal contrast between plutons of the
Tujunga terrane and those of the Mojave Desert
region.
A key feature of the Jurassic magmatic event is
emplacement of the Independence dike swarm
(Moore and Hopson, 1961). Dated at 148 Ma in the
Sierras (Chen and Moore, 1979), this north-trending
swarm varies from diabase through dacite and rhyo-
lite and can be traced semi-continuously into the
western Mojave and the eastern Tranverse Ranges,
including the San Gabriel, Eagle, and Chuckwalla
Mountains portion of the Tujunga terrane (Powell,
1981; Karish and others, 1987; James, 1989; Barth,
1990). The swarm post-dates all of the Jurassic
plutons described above but is intruded by younger
Cretaceous plutons. As noted by James (1989), and
important to the goals of this study, is the obser-
vation that correlative Independence dikes can be
followed along strike across the inferred Tujunga-
North American terrane boundary.
Thermobarometric estimates of emplacement
conditions (all from hornblende barometry using the
calibration of Hollister and others, 1987) show a
220 J. L. Anderson, A. P. Barth, E. D. Young, and others
surprising variation (Figure 8) because we expected
most of the intrusive centers to be shallow. That
prejudgment came from the fact that several plutons
intrude Jurassic volcanic or unmetamorphosed
Paleozoic sections, including complexes in the Pro-
vidence, Marble, Bristol, and Eagle Mountains.
These specific plutons, in fact, yield shallow (2-3 kb)
emplacement barometric determinations consistent
with their geologic setting. Other shallow intru-
sions occur in the Chuckwalla (2-3 kb for plutons
with U-Pb ages < 169 Ma) and the Mule and Trigo
Mountains (3-4 kb). Yet, three deep-seated Jurassic
complexes are identified - specifically the Granite
II (south of the Teutonia batholith, Figure 1), the
Chuckwalla Mountains (the oldest Jurassic pluton
in that range, with a U-Pb age of .., 169 Ma), and the
Cargo Muchacho Mountains - all yielding crystal-
lization pressures of 6-7 kb. These results have
recently been described by Anderson and others
(1988a), Barth and others (1988), Young and Wood-
en (1988), Farber and Davis (1989), Davis and Far-
ber (1989), and Hayes (1989).
Cretaceous Plutonism
After an extended magmatic lull, extensive ig-
neous activity resumed during the Late Cretaceous
in both the Tujunga terrane and the Mojave region,
beginning at "" 95 Ma and reaching a peak at 72-80
Ma (Carter and Silver, 1972; Miller and others,
1982; Beckerman and others, 1982; Wright and
others, 1986; Wright and others, 1987). The com-
position and geologic setting of several Cretaceous
intrusive complexes in the Mojave Desert have been
well documented: (1) the Joshua Tree National
Monument (Brand and Anderson, 1982; Brand,
1985; Anderson and others, 1988a); (2) the Cadiz
Valley batholith (John, 1981; D. Miller and others,
1981, 1982; Howard and others, 1982; Calzia, 1982;
Howard and John, 1984; D. Miller and Howard,
1985; Anderson, 1988); (3) the Old Woman-Piute
Range (Miller and others, 1982; Young and Miller,
1983; Foster and others, 1989; Young and Miller,
1989); (4) the Teutonia batholith (Beckerman and
others, 1982; Anderson and others, 1988a); (5) the
Chemehuevi Mountains core complex (John, 1982,
1987; John and Wooden, 1989; Howard and others,
1987); and (6) the Whipple Mountains core complex
(Davis and others, 1980; Anderson and Rowley,
1981; Anderson, 1985; Anderson and others, 1988a,
b; Anderson, 1988; Anderson and Cullers, 1990).
Compositionally and by rock type, the Mojave
Cretaceous suites are unlike those of the early and
middle Mesozoic. None are alkalic, the K
2
0 abun-
dance is lower, and most plutons are relatively
silicic (>66 wt. % Si02) and calcalkaline (Figure 7).
Being within the inner peraluminous belt of Miller
and Bradfish (1984), two-mica granites are common,
as are metaluminous hornblende-biotite-sphene
granodiorites. Geochemically and isotopically, most
appear to have been derived from melting of older
crust (Miller, 1985; Wooden and Stacey, 1987;
Miller and others, 1990),. Exception are the plutons
of the Whipple core complex (Anderson and Cullers,
1990).
Cretaceous plutons in the Tujunga terrane are
also widespread. Powell (1981) defined a belt of
Cretaceous granitic plutons in the Hexie, Eagle,
Chuckwalla, and Little Chuckwalla Mountains.
Correlative plutons also occur in the Orocopia
Mountains (Anderson, unpublished data). In the
San Gabriel Mountains, Carter and Silver (1972)
reported an 8010 Ma age for the Mt. Josephine
intrusion, and petrographically similar plutons in-
clude the Mt. Wilson, Mt. Waterman, and Vetter
Mountain of the same range (Barth and others,
1989). These intrusive complexes show strong com-
positional similarity to the Cretaceous plutons of
the Mojave (Figure 7). If the Tujunga terrane is far
traveled, .., 1000 km, and was sutured to North
America in the early Tertiary (Vedder and others,
1983), then some difference in composition of the
Mesozoic suites would be expected. However, no
differences are evident amongst the Cretaceous
plutons.
Depths of emplacement of Cretaceous plutons
were derived from a variety of thermobarometers.
Unlike those of Jurassic age, many of these Creta-
ceous complexes are deep seated (Figure 8). Mid-
crustal intrusions (depths of emplacement> 20 km)
reside in the San Gabriel, Old Woman, Granite II,
Chemehuevi, and Whipple Mountains. Upper crus-
tal intrusions are identified in the Teutonia and
Cadiz Valley batholiths, the Chuckwalla Moun-
tains, and the Sacramento Mountains core complex
(Anderson and others, 1988a; Young and Wooden,
1988; John and Wooden, 1989).
A major compositional difference among Creta-
ceous plutons is seemingly related to depth of em-
placement. Shallow-emplaced 12 km) complexes
tend to be fundamentally granitic and often leuco-
cratic, whereas deeper Mesozoic plutons in both
areas (Tujunga and craton) tend to be more mafic,
Plutonism across the Tujunga-North American terrane boundary 221
including metaluminous diorite to quartz diorite
and marginally metaluminous to calcic peralumin-
ous tonalite and granodiorite. Exceptions occur
(e.g., Old Woman Mountains), but this generaliza-
tion may extend to older elements of the orogen as
well. For example, deeper portions of the Tujunga
terrane also contain the more mafic components of
the Triassic Lowe intrusion and Proterozoic an-
orthosite and leucogabbro. The implication is that a
fundamental contrast existed between the upper
15 km) and middle (15-27 km) portions of the
crust in this region of the Cordillera. Although most
of the upper crust intrusives appear to have been
derived from crustal sources, plutonism in the
deeper complexes is more varied in origin, including
crust-derived two-mica granitoids of the Old Woman
Mountains, mixed oceanic/continental crust-derived
plutons of the Whipple Mountains, and mantle-
derived Lowe intrusion ofthe Tujunga terrane.
CONCLUSIONS
The primary results of this study are (1) assess-
ment of fundamental differences in magmatic arc
construction of Early Proterozoic age through Meso-
zoic arc magmatism at upper and mid-crustal levels,
and (2) constraints on the timing and nature of the
accretionary history of the southern Cordillera.
Crustal Depth History
The above sections have delineated the relative
barometric evidence we have collected from a num-
ber of ranges. Figure 9 is a composite of all data;
depth was calculated from derived pressures with an
assumed geobarometric gradient (from average
crustal density) of 3.7 kmlkb. For some ranges
(Eagle, Sacramento, Joshua Tree), we used a "best
guess" crystallization age; all others are from U-Pb
(zircon) geochronology and those paths are fairly
well constrained. Evident is a broad range of ex-
posed crustal depth for both Tujunga and native
complexes. Although preliminary, two levels of
crust were sampled in this study. Shallow intrusive
centers 12 km) include the Teutonia and Cadiz
Valley batholiths, the Sacramento core complex,
plutons of the Eagle, Mule, and Trigo Mountains,
and younger members of the Chuckwalla plutonic
complex. The granites of the Joshua Tree Monu-
ment appear to have been emplaced at moderate
10
o
(km) 20
30
I "Ii
250 200
J K
T
150
Ma
Figure 9. Crustal depth history of southern California
based on barometry of plutonic emplacement and deforma-
tional events. Tujunga terrane complexes (dashed paths)
include those exposed in the San Gabriel (SG), Chuckwall
(Ch), Mule and Trigo (M-T), and Eagle (E) Mountins.
North American complexes (solid paths) include Whipple
(W), Old Woman and Granite II (OW/GM) , and Sacramento
(SAC) Mountains, Joshua Tree National Monument (JT),
and the Cadiz Valley (CVB) and Teutonia (TB) batholiths.
The sharp bend in the Chuckwalla path represents change
in determined crustal level relative to pre- and post-late
Red Cloud thrusting. From Anderson and others (l988a),
Young and Wooden (1988), Barth and others (1988), and
Farber and Davis (1989).
depths ("" 12-16 km). Deep-seated, middle crust
appears to have developed in two contrasting tec-
tonic regimes - the lower plate of core complexes
(the Whipple Mountains), and basement-cored
thrust complexes (Old Woman, Granite II, San
Gabriel, Cargo Muchacho, and Chuckwalla Moun-
tains). The abrupt change for the Chuckwalla path
is constrained by deep emplacement barometry of a
syn-Red Cloud thrust ("" 159 Ma) pluton and shal-
low barometry for post-thrust, Jurassic 159 Ma)
and Cretaceous plutons, implying that the thrusting
event involved upward ascent of 8-10 km of crust.
Some of these deeper crustal sections may have
originated in the middle crust, including the Tujun-
ga terrane. Proterozoic (1.2 Ga) anorthosite and
charnockite and Triassic to Late Cretaceous granitic
intrusives characterize the "allochthonous" complex
which, in part, resided at depths> 25 km during all
of the Mesozoic (if not since the Proterozoic). Uplift
commenced after early Tertiary underthrusting of
oceanic lithosphere and/or accretion to North
222 J. 1. Anderson, A. P. Barth, E. D. Young, and others
America. In contrast, the Old Woman path must be
part of a protracted depth-time loop. Originating at
surface conditions during the Paleozoic, the Old
Woman crustal and supracrustal sections were tec-
tonically buried to depths > 20 km by 74 Ma, as
recorded by conditions of intrusion of granitic plu-
tons coeval with high-grade regional metamorphism
(Miller and others, 1982; Young and Wooden, 1988).
Rapid uplift occurred during the latest Cretaceous
(Miller and others, 1990; Foster and others, 1989).
We have yet to understand the tectonic mechanism
leading to the uplift.
Allochthoneity of the Tujunga Terrane:
A Native Terrane Model
A central objective of this study has been to eva-
luate the degree of allochthoneity of the Tujunga/
San Gabriel terrane. If the terrane was far traveled,
then marked differences in crustal history with
respect to native portions of the craton are antici-
pated up to the time of accretion. Yet, instead of
differences, we observe a striking similarity. The
Cretaceous plutons are equivalent in being calc-
alkaline and containing both meta luminous and
peraluminous members. The Independence dike
swarm has been traced > 500 km into the Tujunga
terrane (James, 1989). The Jurassic suite exhibits
minor differences but has compositional affinity in
being largely metaluminous, high to ultra high K,
with modest silica and Sr abundances. We are also
impressed by the similarity of pre-Jurassic history.
The Triassic Lowe intrusion exposed in the San
Gabriel, Mule, and Trigo Mountains is markedly
sodic but otherwise maintains the low silica, high
alkali and Sr (> 1000 ppm), and mineralogic char-
acter of similar-aged intrusions in the Granite I and
San Bernardino Mountains of autochthonous North
America (Barth and Ehlig, 1988; Barth, 1990;
Miller, 1977). Both have been explained as the re-
sult of melting of an LILE-enriched eclogitic source.
Finally, Bender (1989) demonstrated (1) that the
Early Proterozoic, = 1. 7 Ga, plutons of the Mojave
region are unusal (with respect to the remainder of
the craton) in lacking calcalkaline members and
being solely comprised of A-type granitoids, and
(2) that the same type of plutons also occur in
abundance in the Tujunga terrane.
The principal difference between Tujunga and
the craton resides in their Middle Proterozoic
history. The data of Silver and others (1963) for a
1.4 Ga granulite event in the Tujunga terrane has
never been published. It remains unclear whether
the age represents a granulite-grade orogenic epi-
sode or magma-crust interaction event, such as one
related to derivation of the anorogenic 1.4 Ga
plutons common to the craton. The occurrence of the
1.2 Ga anorthosite-charnockite in the Tujunga
terrane is certainly unique. Yet, the single largest
exposure of the complex occurs in the San Gabriel
Mountains, which resided in the middle crust
throughout the Mesozoic if not since the Proterozoic.
In conclusion, we suggest that the "suspect"
nature of the Tujunga terrane is due to derivation
from deep crust and see no reason for the terrane to
be far traveled relative to native portions of the
southwestern United States. Because the Cordi-
llera, like most orogenic belts, has along-strike
continuity of major lithotectonic units, strike-slip
faulting could lead to intra-arc displacement and
juxtaposition of lithologically similar terranes.
Thus, we cannot rule out that Tujunga originated
elsewhere in the Cordilleran orogen and may have
travelled hundreds of kilometers. Yet, great trans-
port on the scale of thousands of kilometers (Vedder
and others, 1983) and an extra-Cordilleran origin
for Tujunga appear improbable.
Figure 10 proposes a native terrane model for
Tujunga at 60 Ma. Three subterranes are defined:
a southern subterrane (San Gabriel Mountains)
of mid-crustal origin at 20-25 km;
a middle subterrane (Chuckwalla to MuleITrigo
Mountains region) that also originated at mid-
crustal depths but, by the latest Cretaceous,
was at 5-15 km; and
a northern subterrane, representing the mid-
crustal Old Woman/Piute Mountains, contain-
ing Late Cretaceous, west-directed thrust faults
structurally overlain by supracrustal sections of
the Providence and Marble Mountains.
The variable crustal levels now exposed within
Tujunga are shown in the model to have been suc-
cessively juxtaposed by early Red Cloud (ERCT),
late Red Cloud (LRCT), Mule Mountains (MMT),
and Vincent-Chocolate Mountains (VCMT) thrust
faults.
k
b

S
G

C
h
w

M
T

O
W
/
P


?
S

"
-


"
-
"
-
.
.
.
.
.
,
,


"

"
-
.
.
.
.
.
,
,

,

'
:
:
'
:
:


'
'
'
-
'
-
"
,


-
.
.
.
.
.
"

"
-
"
'
-
'

"
-
"
'
"


-
.
.
.
.
.
"
'
-
-
"
"
-
'
-
"
'
"

-
.
.
.
.
.
"
-
.
.
.
.
.
,
,

,
-
.
.
.
.
.
"

"
-
-
"
"
"
'
,
-
.
.
.
.
.
,
,
'

.

k
m

F
i
g
u
r
e

1
0
.

M
o
d
e
l

c
r
o
s
s

s
e
c
t
i
o
n

a
t

6
0

M
a
,

s
h
o
w
i
n
g

t
h
e

r
e
l
a
t
i
v
e

c
r
u
s
t
a
l

l
e
v
e
l
s

e
x
p
o
s
e
d

i
n

t
h
e

S
a
n

G
a
b
r
i
e
l

(
S
G
)
,

C
h
u
c
k
w
a
l
l
a

(
C
h
w
)
,

M
u
l
e
-
T
r
i
g
o

(
M
T
)
,

a
n
d

O
l
d

W
o
m
a
n
i
P
i
u
t
e

(
O
W
f
P
)

M
o
u
n
t
a
i
n
s

d
u
e

t
o

d
i
s
r
u
p
t
i
o
n

o
f

c
r
u
s
t

b
y

t
h
e

V
i
n
c
e
n
t
/
C
h
o
c
o
l
a
t
e

M
o
u
n
t
a
i
n
s

(
V
C
M
T
)
,

M
u
l
e

M
o
u
n
t
a
i
n
s

(
M
M
T
)
,

l
a
t
e

R
e
d

C
l
o
u
d

(
L
R
C
T
)
,

a
n
d

e
a
r
l
y

R
e
d

C
l
o
u
d

(
E
R
C
T
)

t
h
r
u
s
t

f
a
u
l
t
s
.

T
h
e

c
o
m
p
o
s
i
t
e

s
e
c
t
i
o
n

i
s

s
h
o
w
n

t
o

b
e

i
n

p
a
r
t
i
a
l

s
t
r
u
c
t
u
r
a
l

c
o
n
t
a
c
t

w
i
t
h

t
h
e

u
n
d
e
r
l
y
i
n
g

P
e
l
o
n
a
/
O
r
o
c
o
p
i
a

s
c
h
i
s
t

(
J
K
p
s
)

a
c
r
o
s
s

t
h
e

V
i
n
c
e
n
t
/
C
h
o
c
o
l
a
t
e

M
o
u
n
t
a
i
n
s

t
h
r
u
s
t

f
a
u
l
t
.

U
n
i
t

s
y
m
b
o
l
s
:

P
C
g
n
,

P
r
o
t
e
r
o
z
o
i
c

g
n
e
i
s
s
;

P
C
a
n
,

P
r
o
t
e
r
o
z
o
i
c

a
n
o
r
t
h
o
s
i
t
e

a
n
d

s
y
e
n
i
t
e
;

P
C
j
t
,

P
r
o
t
e
r
o
z
o
i
c

J
o
s
h
u
a

T
r
e
e

t
e
r
r
a
n
e

g
n
e
i
s
s
e
s
;

P
C
s
e
d
s
,

P
r
o
t
e
r
o
z
o
i
c

s
t
r
a
t
a
;

P
z
,

P
a
l
e
o
z
o
i
c

s
t
r
a
t
a
;

T
R
g
,

T
r
i
a
s
s
i
c

L
o
w
e

i
n
t
u
s
i
o
n
;

J
R
g
,

J
u
r
a
s
s
i
c

g
r
a
n
i
t
e
s
;

M
z
,

o
l
d
e
r

M
e
s
o
z
o
i
c

s
t
r
a
t
a
;

J
v
,

J
u
r
a
s
s
i
c

v
o
l
c
a
n
i
c
s
;

K
m
,

M
c
C
o
y

F
o
r
m
a
t
i
o
n
;

K
g
,

C
r
e
t
a
c
e
o
u
s

g
r
a
n
i
t
e
s
.

'
"
d


M
-
o

8
.

0
0

S


'"
"
'
:
3

0
0

0
0

:
;

(
1
)

g


1
:1

(
J
Q


Z

o

:
:
1
-
:
:
r
"

>

S

(
1
)

:
J
.

'"
"
'


1
:1

1
t
i

>
-
j

;
5

g

0
-
'

o


5
-


t
-
:
>

t
-
:
>

c
.
:
>

224 J. L. Anderson, A. P. Barth, E. D. Young, and others
ACKNOWLEDGMENTS
This research has been supported by NSF
grants EAR 86-18285 and EAR 89-04060 to Ander-
son; GSA Penrose grants individually to Barth, Ben-
der, and Young; and the Foss Foundation. Calvin
Miller and Brady Rhodes provided very helpful re-
views of this manuscript. Without the many dis-
cussions and collaborative work with Keith Howard,
Bob Powell, Dick Tosdal, and Joe Wooden, the scope
of this research would have been severely limited.
REFERENCES
ALLEN C. M., D. M. MILLER, and K. A. HOWARD, 1983,
Field, petrologic and chemical characteristics of Jurassic
intrusive rocks, eastern Mohave Desert, southeastern
California [abstract): Geological Society of America,
Abstracts with Programs, v. 15, p. 410-411.
ANDERSON, J. L., 1983, Proterozoic anorogenic granite
plutonism of North America; pp. 133-152 in L. G. Medaris,
Jr., D. M. Mickelson, C. W. Byers, and W. C. Shanks (eds.),
Proterozoic Geology: Selected Papers from an Inter-
national Proterozoic Symposium: Geological Society of
America, Memoir 161,315 p.
__ , 1985, Contrasting depths of "core complex" mylo-
nitization: barometric evidence [abstract): Geological
Society of America Abstracts with Programs, v. 17, p. 337.
__ , 1988, Core complexes ofthe Mojave-Sonoran Desert:
conditions of plutonism, mylonitization, and decompres-
sion; pp. 503-525 in W. G. Ernst (ed.), Metamorphism and
Crustal Evolution of the Western United States (Ru-
bey Volume 7): Prentice Hall, Englewood Cliffs, New
Jersey, USA, 1153 p.
__ , 1989, Proterozoic anorogenic granites of the south-
western United States; pp. 211-238 in J. P. Jenney and S.
J. Reynolds (eds.), Geologic Evolution of Arizona: Ari-
zona Geological Society Digest, v. 17, 866 p.
__ , and E. E. BENDER, 1989, Nature and origin of
Proterozoic A-type granitic magmatism in the south-
western United States of America: Lithos, v. 23, no. 112, p.
19-52.
__ , and R. L. CULLERS, 1990, Magmatic evolution of a
metamorphic core complex, Whipple Mountains, south-
eastern California; pp. 47-69 in J. L. Anderson (ed.), The
Nature and Origin of CordiUeran Magmatism: Geo-
logical Society of America, Memoir 174,414 p.
__ , and M. C. ROWLEY, 1981, Synkinematic intrusion of
two-mica and associated metaluminous granitoids,
Whipple Mountains, California: Canadian Mineralogist, v.
19, p. 83-101.
__ , A. P. BARTH, E. E. BENDER, M. J. DAVIS, D. L.
FARBER, E. M. HAYES, K. A. JOHNSON, E. D. YOUNG, J.
L. WOODEN, and R. M. TOSDAL, 1990, San Gabriel (Tujun-
ga) terrane - Coming home to Mojave [abstract): Geo-
logical Society of America Abstracts with Programs, v. 22,
p.303.
__ , A. P. BARTH, and E. D. YOUNG, 1988b, Mid-crustal
roots of Cordilleran metamorphic core complexes: Geology,
v. 16, p. 366-369.
__ , A. P. BARTH, E. D. YOUNG, M. J. DAVIS, D.
FARBER, E. M. HAYES, and K. A. JOHNSON, 1988a, Con-
trasting depth of Messozoic arc emplacement across the
Tujunga-North American terrane boundary [abstract):
Geological Society of America Abstracts with Programs, v.
20, p.139.
__ , J. L. WOODEN, and E. E. BENDER, 1992, Mojave
province of southern Claifornia and vicinity; Chapter 4 in
W. R. Van Schmus and M. E. Bickford (eds.), The Geology
of North America, Volume C2: Transcontinental Pro-
terozoic Provinces, Precambrian: Conterminous U.S.:
Geological Society of America, Boulder, Colorado, USA (in
press).
__ , E. D. YOUNG, and A. P. BARTH, 1988c, Cordilleran
middle crust plutonism: Core complexes and suspect ter-
ranes [abstract): 8th International Conference on Basement
Tectonics, Program with Abstracts, p. 20.
ANDERSON, T. H., and L. T. SILVER, 1979, The role of the
Mojave-Sonora megashear in the tectonic evolution of
northern Sonora; pp. 59-68 in T. H. Anderson and J.
Roldan-Quintana (eds.), Geology of Northern Sonora
(Geological Society of America, Cordilleran Section
Guidebook, Field Trip 27): University of Pittsburgh,
Pittsburgh, Pennsylvania, USA, 93 p.
BARTH, A. P., 1987, Mid-crustal plutonism in the San Gab-
riel Mountains, southern California [abstract): Geological
Society of America Abstracts with Programs, v. 19, p. 357.
__ ,1990, Mid-crustal emplacement of Mesozoic plutons,
San Gabriel Mountains, California, and implications for
the geologic history of the San Gabriel Terrane; pp. 33-45
in J. L. Anderson (ed.), The Nature and Origin of Cor-
dilleran Magmatism: Geological Society of America,
Memoir 174,414 p.
__ , and P. L. EHLIG, 1988, Geochemistry and petro-
genesis of the marginal zone of the Mount Lowe Instrusion,
central San Gabriel Mountains, California: Contributions
to Mineralogy and Petrology, v. 100, p. 192-204.
Plutonism across the Tujunga-North American terrane boundary 225
__ , and D. J. MAY, 1987, Contrasting Late Cretaceous
granulite terranes, southeastern San Gabriel Mountains,
California [abstract): Geological Society of America Ab-
stracts with Programs, v. 19, p. 581.
__ , R. M. TOSDAL, and J. L. WOODEN, 1988, Charac-
teristics and implications of Triassic and Jurassic grani-
toids in the San Gabriel, Mule, and Trigo Mountains,
southern California and southwestern Arizona [abstract):
Geological Society of America Abstracts with Programs, v.
20, p. 141.
__ , J. L. WOODEN, and R. M. TOSDAL, 1989, Josephine
Mountain Intrusion, California: a model for calcalkalic
cratonal arc plutonism at mid-crustal depths [abstract):
Geological Society of America Abstracts with Programs, v.
21, p. 54.
BATEMAN, P. C., 1983, A summary of critical relations in
the central part of the Sierra Nevada Batholith, California:
Geological Society of America, Memoir 159, p. 241-254.
BECKERMAN, G. M., and J. L. ANDERSON, 1981, The Teu-
tonia batholith: a large intrusive complex of Jurassic and
Late Cretaceous age in the eastern Mojave Desert, Califor-
nia [abstract): Geological Society of America Abstracts
with Programs, v. 13, p. 44.
__ , J. P. ROBINSON, and J. L. ANDERSON, 1982, The
Teutonia batholith: A large intrusive complex of Jurassic
and Cretaceous age in the eastern Mojave Desert, Califor-
nia; pp. 205-221 in E. G. Frost and D. L. Martin (eds.),
Mesozoic-Cenozoic Tectonic Evolution of the Colorado
River Region, California, Arizona, and Nevada (An-
derson-Hamilton Volume): Cordilleran Publishers, San
Diego, California, USA, 608 p.
BENDER, E. E., 1989, Petrology of Early Proterozoic
granitic gneisses from the Eastern Transverse Ranges:
implications for tectonics of southeastern California
[abstract): Geological Society of America Abstracts with
Programs, v. 21, p. 56.
__ , J. L. ANDERSON, J. L. WOODEN, K. W. HOWARD,
and C. F. MILLER, 1988, Correlation of 1.7 Ga granitoid
plutonism in the lower Colorado River region [abstract):
Geological Society of America Abstracts with Programs, v.
20, p.142.
BLAKE, M. C., D. G. HOWELL, and D. L. JONES, 1982, Pre-
liminary Tectonostratigraphic Terrane Map of Cali-
fornia: U.S. Geological Survey, Open-File Report 82-593.
BRAND, J. H., 1985, Mesozoic Alkalic Quartz Monzonite
and Peraluminous Monzogranites of the Northern
Portion of Joshua Tree National Monument, Southern
California: MS thesis, University of Southern California,
Los Angeles, California, USA, 187 p.
__ , and J. L. ANDERSON, 1982, Mesozoic alkalic mon-
zonites and peraluminous adamellites of the Joshua Tree
National Monument, southern California [abstract): Geo-
logical Society of America Abstracts with Programs, v. 14,
p.151-152.
CALZIA, J. P., 1982, Geology of granodiorite in the
Coxcomb Mountains, southeastern California; pp. 173-180
in E. G. Frost and D. L. Martin (eds.), Mesozoic-Cenozoic
Tectonic Evolution of the Colorado River Region, Cali-
fornia, Arizona, and Nevada (Anderson-Hamilton Vol-
ume): Cordilleran Publishers, San Diego, California,
USA,608p.
CARTER, B., and L. T. SILVER, 1972, Structure and petro-
logy of the San Gabriel anorthosite-syenite body, Califor-
nia: 24th International Geological Congress (Montreal,
Canada), Section 2, p. 303-311.
CHAMPION, D. E., D. G. HOWELL, and D. S. GROMME,
1984, Paleomagnetic and geologic data indicting 2500 km
of northward displacement of the Salinian and related
terranes, California: Journal of Geophysical Research, v.
89, p. 7736-7752.
CHEN, J. H., and J. G. MOORE, 1979, Late Jurassic Inde-
pendence dike swarm in eastern California: Geology, v. 7,
p.129-133.
__ , and J. G. MOORE, 1982, U-Pb isotopic ages from the
Sierra Nevada batholith, California: Journal of Geophysi-
cal Research, v. 87, p. 4761-4785.
CONEY, P. J., D. L. JONES, and J. W. H. MONGER, 1980,
Cordilleran suspect terranes: Nature, v. 288, p. 329-333.
CRISCENTI, L. J., and M. S. GHIORSO, 1985, A method for
estimating the composition and silicate melts from chemi-
cal analyses of coexisting mineral phases [abstract): EOS,
Transactions of the American Geophysical Union, v. 66, no.
46, p. 1120.
CROWELL, J. C., 1962, Displacement along the San An-
dreas fault, California: Geological Society of America,
Special Paper 71, p. 1-61.
__ , 1981, An outline of tectonic history of southeastern
California; pp. 253-283 in W. G. Ernst (ed.), The Geo-
tectonic Development of California (Rubey Volume 1):
Prentice Hall, Inc., Englewood Cliffs, New Jersey, USA,
706p.
DAVIS, G. A., J. L. ANDERSON, E. G. FROST, and T. J.
SHACKELFORD, 1980, Mylonitization and detachment
faulting in the Whipple-Buckskin-Rawhide Mountains
terrane, southeastern California and western Arizona; pp.
79-129 in M. Crittenden, G. J. Davis, and P. J. Coney,
(eds.), Cordilleran Metamorphic Core Complexes: Geo-
logical Society of America, Memoir 153, 490 p.
DAVIS, M. J., and D. L. FARBER, 1989, Early Mezozoic
plutonism and ductile deormation in the Chuckwalla
Mountains of southeastern California [abstract): Geologi-
cal Society of America Abstracts with Programs, v. 21, p.
72.
226 J. L. Anderson, A. P. Barth, E. D. Young, and others
DAVIS, T. E., and A. P. BARTH, 1985, Pb isotopic con-
straints on the petrogenesis of the Mount Lowe Intrusion,
San Gabriel Mountains, California [abstract): Geological
Society of America Abstracts with Programs, v. 17, p. 350.
DmBLEE, T. W., 1982, Geology of the San Gabriel Moun-
tains, southern California; pp. 131-147 in D. L. Fife and J.
A. Minch (eds.), Geology and Mineral Wealth of the
California Transverse Ranges (Mason Hill Volume):
South Coast Geological Society, Santa Ana, California,
USA,699p.
DILLON, J. T., 1976, Geology of the Chocolate and Car-
go Muchacho Mountains, Southeasternmost Califor-
nia: PhD dissertation, University of California, Santa
Barbara, California, USA, 380 p.
EHLIG, P. L., 1975, Basement rocks of the San Gabriel
Mountains, south of the San Andreas fault, southern Cali-
fornia; pp. 177-186 in J. C. Crowell (ed.), The San An-
dreas Fault in Southern California: California Division
of Mines and Geology, Special Report 118, 272 p.
__ ,1981, Origin and tectonic history of the San Gabriel
Mountains, central Transverse Ranges; pp. 254-283 in W.
G. Ernst (ed.), The Geotectonic Development of Cali-
fornia (Rubey Volume 1): Prentice Hall, Inc., Englewood
Cliffs, New Jersey, USA, 706 p.
FARBER, D. L., and M. J. DAVIS, 1989, Plutonism in the
Chuckwalla Mountains, California: Evidence for anoma-
lous crustal evolution and tectonic history of the northern
San Gabriel terrane [abstract): Geological Society of Am-
ericaAbstracts with Programs, v. 21, p. 77.
FERRY, J. M., and F. S. SPEAR, 1978, Experimental cali-
bration of the partitioning of Fe and Mg between biotite
and garnet: Contributions to Mineralogy and Petrology, v.
66, p. 113-117.
FOSTER, D. A., T. M. HARRISON, and C. F. MILLER, 1989,
Age inheritance and uplift history ofthe Old Woman-Piute
batholith, California, and implications for K-feldspar age
spectra: Journal of Geology, v. 97, p. 232-243.
FUHRMAN, M. L., and D. H. LINDSLEY, 1988, Ternary
feldspar modelling and thermometry: American Minera-
logist, v. 73, p. 201-215.
GANGULY, J., and S. SAXENA, 1984, Mixing properties of
aluminosilicate garnets: Constraints from natural and ex-
perimental data, and applications to geothermobarometry:
American Mineralogist, v. 69, p. 88-97.
GHENT, E. D., and M. Z. STOUT, 1981, Geobarometry and
geothermometry of plagioclase-biotite-granet-muscovite
assemblages: Contributions to Mineralogy and Petrology,
v. 76, p. 92-97.
GILL, J., 1981, Orogenic Andesites and Plate Tec-
tonics: Springer-Verlag, New York, New York, USA, 390
p.
GRAHAM, C. M., and P. C. ENGLAND, 1976, Thermal
regimes and regional metamorphism in the vicinity of
overthrust fault - An example of shear heating and
inverted metamorphic zonation from southern California:
Earth and Planetary Science Letters, v. 31, p. 142-152.
__ , and R. POWELL, 1984, A garnet-hornblende geo-
thermometer: Calibration, testing, and application to the
Pelona schist, southern California: Jounal of Metmorphic
Geology, v. 2, p. 13-31.
GREW, E. S., 1981, Granulite-facies metamorphiism at
Molodezhnaya Station, East Antartica: Journal of Petro-
logy, v. 2, p. 13-31.
HAMILTON, W. B., 1988, Mexozoic tectonics of southern
California [abstract): Geological Society of America Ab-
stracts with Programs, v. 20, p. 165.
HAMMARSTROM, J. M., and E-AN ZEN, 1986, Aluminum
in hornblende, an empirical igneous geobarometer: Ameri-
can Mineralogist, v. 71, p.1297-1313.
HARDING, L. E., and P. J. CONEY, 1985, The geology of the
McCoy Mountains formation, southeastern California and
southwestern Arizona: Geological Society of America
Bulletin, v. 96, p. 755-769.
HAXEL, G. B., and J. DILLON, 1978, The Pelona-Orocopia
schist and Vincent-Chocolate Mountain thrust system,
southern California; pp. 453-469 in D. J. Howell and K. A.
McDougall (eds.), Mesozoic Paleogeography of the
Western United States: Society of Economic Paleontolo-
gists and Mineralogists, Pacific Coast Paleogeography
Symposium 2,573 p.
__ , and R. M. TOSDAL, 1986, Significanced of the
Orocopia schist and Chocolate Mountains thrust in the late
Mesozoic tectonic evolution of the southeastern Califronia-
southwestern Arizona region; pp. 52-61 in B. Beatty and P.
A. K. Wilkinson (eds.), Frontiers in Geology and Ore
Deposits of Arizona and the Southwest: Arizona Geo-
logical Society Digest, v. 16, 554 p.
__ , J. R. BUDAHN, T. L. FRIES, B. W. KING, L. D.
WHITE, and P. J. ARUSCABAGE, 1987, Geochemistry of the
Orocopia schist, southeastern California: Summary; pp.
49-64 in W. R. Dickinson and M. A. Klute (eds.), Mesozoic
Rocks of Southern Arizona and Adjacent Areas: Ari-
zona Geological Society Digest, v. 18, 394 p.
HAYES, E. M., 1989, Mid-crustal Mesozoic plutonism in
the Cargo Muchacho Mountains, southeastern California
[abstract): Geological Society of America Abstracts with
Programs,v.21,p.92.
HAZELTON, H. T., G. L. HOVIS, B. S. HEMINGWAY, and R.
A. ROBIE, 1982, Calorimetric investigation of the excess
entropy of mixing in analbite-sanidine solid solutions:
Lack of evidence for Na, K short-range order and implica-
Plutonism across the Tujunga-North American terrane boundary 227
tions for two-feldspar thennometry: American Mineralo-
gist, v. 68, p. 398-413.
HELZ, R. T., 1979, Alkali exchange between hornblende
and melt: A temperature sensitive reaction: American
Mineralogist, v. 64, p. 953-965.
HODGES, K. V., and L. ROYDEN, 1984, Geologic thermo-
barometry of retrograded metamorphic rocks: An indi-
cation of uplift trajectory of a portion of the northern
Scandinavian caledonides: Jounal of Geophysical Re-
search, v. 89, p. 7077-7090.
HOISCH, T. D., 1985a, Metamorphism in the Big Maria
Mountains, Southeastern California: PhD dissertation,
University of Southern California, Los Angeles, Califor-
nia, USA, 264 p.
1985b Conditions of metamorphism detennined
from iow grade meta granites, Big Maria Mountains, south-
eastern California [abstract]: EOS, Transactions of the
American Geophysical Union, v. 66, p. 953-965.
1987, Heat transport by fluids during Late Creta-
ceous regional metamorphism in the Big Maria Moun-
tains, southeastern California: Geological Society of Am-
erica Bulletin, v. 98, p. 549-553.
__ , C. F. MILLER, M. T. HEIZLER, T. M. HARRISON, and
E. F. STODDARD, 1988, Late Cretaceous regional meta-
morphism in southeastern California; pp. 538-571 in W. G.
Ernst (ed.), Metamorphism and Crustal Evolution of
the Western United States (Rubey Volume 7): Prentice
Hall, Inc., Englewood Cliffs, New Jersey, USA, 1153 p.
HOLDAWAY, J. J., 1972, Thermal stability of AI-Fe epidote
as a function of O
2
and Fe content: Contributions to Min-
eralogy and Petrology, v. 37, no. 4, p. 307-340.
HOLLISTER, L .S., G. C. GRISSOM, E. K. PETERS, H. H.
STOWELL, and V. B. SISSON, 1987, Confirmation of the
empirical correlation of Al in hornblende with pressure of
solidification of calcalkaline plutons: American Mineralo-
gist, v. 72, p. 231-239.
HOWARD, K. A., and B. E. JOHN, 1984, Geologic map of
the Sheephole-Cadiz Valley wilderness study area (DCDA-
305), San Bernardino County, California: U.S. Geological
Survey, Miscellaneous Field Studies Map MF-1615-A,
1:62,500 scale.
__ , B. E. JOHN, and C. F. MILLER, 1987, Metamorphic
core complexes, Mesozoic ductile thrusts, and Ceonzoic
detachments: Old Woman-Chemehuevi Mountains tran-
sect, California and Arizona; pp. 365-382 in G. H. Davis
and E. M. Van den Dolder (eds.), Geologic Diversity of
Arizona and Its Margins: Excursions to Choice Areas
(GSA Guidebook): Arizona Bureau of Geology and Min-
eral Technology, Geological Survey Branch, Special Paper
5,422 p.
__ , C. F. MILLER, and B. E. JOHN, 1982, Regional
character of mylonitic gneiss in the Cadiz Valley area,
southeastern California; pp. 441-447 in E. G. Frost and D.
L. Martin (eds.), Mesozoic-Cenozoic Tectonic Evolution
of the Colorado River Region, California, Arizona, and
Nevada (Anderson-Hamilton Volume): Cordilleran
Publishers, San Diego, California, USA, 608 p.
HOWELL, D. G., E. R. SCHERMER, D. L. JONES, Z. BEN-
AVRAHAM, and E. SCHEIBNER, 1983, Tectonostratigra-
phic Terrane Map of the Circum-Pacific Region: U.S.
Geological Survey, Open File Report 83-716.
IRVINE, T. N., and R. A. BARAGAR, 1971, A guide to
chemical classification of the common volcanic rocks.
Canadian Journal of Earth Sciences, v. 8, p. 523-548.
JAMES, E. W., 1989, Southern extension of the Indepen-
dence dike swarm of eastern California: Geology, v. 17, p.
587-590.
JOHN, B. E., 1981, Reconnaissance study of Mesozoic
plutonic rocks in the Mojave Desert region; pp. 49-51 in K.
A. Howard, M. S. Carr, and D. M. Miller (eds.), Tectonic
Framework of the Mojave and Sonoran Deserts, Cali-
fornia and Arizona: U.S. Geological Survey, Open-File
Report 81-503, 129p.
__ , 1982, Geologic framework of the Chemehuevi
Mountains, southeastern California; pp. 317-325 in E. G.
Frost and D. L. Martin (eds.), Mesozoic-Cenozoic Tec-
tonic Evolution of the Colorado River Region, Califor-
nia, Arizona, and Nevada (Anderson-Hamilton Vol-
ume): Cordilleran Publishers, San Diego, California,
USA,608p.
__ , 1987, Geometry and evolution of a mid-crustal
extensional fault system: Chemehuevi Mountains, south-
eastern California; pp. 313-335 in M. P. Coward, J. F.
Dewey, and P. L. Hancock (eds.), Continental Extension-
al Tectonics: Geological Society of London, Special Pub-
lication 28, 637 p.
__ , and J. L. WOODEN, 1990, Petrology and geo-
chemistry of the Chemehuevi Mountains plutonic suite,
southeastern California; pp. 71-98 in J. L. Anderson (ed.),
The Nature and Origin of Cordilleran Magmatism:
Geological Society of America, Memoir 174,414 p.
JONES, D. J., D. G. HOWELL, P. J. CONEY, and J. W. H.
MONGER, 1983, Recognition, character, and analysis of
tectonostratigraphic terranes in western North America;
pp. 21-35 in M. Hashimoto and S. Uyeda (eds.), Accre-
tionary Tectonics in the Circum-Pacific Regions:
Terra Scientific Publishing Company, Tokyo, Japan, 358 p.
JOSEPH, S. E., J. J. CRISCIONE, T. E. DAVIS, and P. L.
EHLIG, 1982, The Lowe igneous pluton; pp. 307-309 in D.
L. Fife and J. A. Minch (eds.), Geology and Mineral
Wealth of the California Transverse Ranges (Mason
228 J. L. Anderson, A. P. Barth, E. D. Young, and others
Hill Volume): South Coast Geological Society, Santa Ana,
California, 699 p.
KARISH, C. R, E. L. MILLER, and J. F. SU'ITER, 1987,
Mesozoic and tectonic history of the central Mojave Desert;
pp. 15-32 in W. R Dickinson and M. A. Klute (eds.), Meso-
zoic Rocks of Southern Arizona and Adjacent Areas:
Arizona Geological Society Digest, v. 18, 394 p.
MASSONNE, J. J., and W. SCHREYER, 1987, Phengite
geobarometry based on the limiting assemblage with K-
feldspar, phylogopite, and quartz: Contributions to Min-
eralogy and Petrology, v. 96, p. 212-224.
MILLER, C. F., 1977, Alkali-Rich Monzonites, Califor-
nia: Origin of Near Silica-Saturated Alkaline Rocks
and Their SignifICance in the Calcalkaline Batholithic
Belt of California: PhD dissertation, University of Cali-
fornia, Los Angeles, California, USA, 283 p.
__ , 1978, Early alkalic plutonism in the calc-alkaline
batholithic belt of California: Geology, v. 5, p. 685-688.
__ , 1985, Are strongly peraluminous magmas derived
from pelitic sedimentary sources? Journal of Geology, v.
93, p. 673-689.
__ , and L. BRADFISH, 1980, An inner Cordilleran belt of
muscovite-bearing plutons: Geology, v. 8, p. 412-416.
__ , K. A. HOWARD, and T. D. HOISCH, 1982, Mesozoic
thrusting, metamorphism, and plutonism in the Old
Woman-Piute Range, southeastern California; pp. 561-581
in E. G. Frost and D. L. Martin (eds.), Mesozoic-Cenozoic
Tectonic Evolution the Colorado River Region, Cali-
fornia, Arizona, and Nevada: Cordilleran Publishers,
San Diego, California, 608 p.
__ . J. L. WOODEN, V. C. BENNE'IT, J. E. WRIGHT, G. C.
SOLOMAN, and R. W. HURST, 1990, Petrogenesis of the
composite peraluminous-meta luminous Old Woman-Piute
Range batholith, southeastern California: Isotopic con-
stratints; pp. 99-109 in J. L. Anderson (ed.), The Nature
and Origin of Cordilleran Magmatism: Geological
Society of America, Memoir 174, 414 p.
MILLER, D. M., K. A. HOWARD, and B. E. JOHN, 1982,
Preliminary geology of the Bristol Lake region, Mojave
Desert, California; pp. 91-100 in J. D. Cooper (ed.), Geo-
logic Excursions in the California Desert (Cordilleran
Section Guidebook): Geological Society of America,
Boulder, Colorado, USA, 159 p.
__ , and K. A. HOWARD, 1985, Bedrock Geologic Map
of the Iron Mountains Quadrangle, San Bernardino
and Riverside Counties, California: U. S. Geological
Survey Map MF-1736, 1:62,500 scale.
__ , K. A. HOWARD, and J. L. ANDERSON, 1981, Myloni-
tic gneiss related to emplacement of a Cretaceous batho-
lith, Iron Mountains, southern California; pp. 73-75 in K.
A. Howard, M. S. Carr, and D. M. Miller (eds.), Tectonic
Framework of the Mojave and Sonoran Deserts, Cali-
fornia and Arizona: U.S. Geological Survey, Open File
Report 81-503,129 p.
MILLER, W. J., 1934, Geology of the western San Gabriel
Mountains of California: University of California, Pub-
lications in Mathematics and Physical Sciences, v. 1, p. 1-
114.
MIYASHIRO, A., 1974, Volcanic rock series in island arcs
and active volcanic margins: American Journal of Science,
v. 274, p. 321-355.
MOORE, J. G., and C. A. HOPSON, 1961, The Independence
dike swarm in eastern California: American Journal of
Science, v. 259, p. 241-259.
MORRIS, L. K., S. P. LUND, and D. J. BO'ITJER, 1986,
Paleolatitude drift history of displaced terranes in
southern and Baja California: Nature, v. 321, p. 844-847.
NABELEK, C. R, and D. H. LINDSLEY, 1985, Tetrahedral
Al in amphibole: A potential thermometer from some
mafic rocks [abstract]: Geological Society of America
Abstracts with Programs, v. 17, p. 673.
NANEY, M. T., 1983, Phase equilibria of rock-forming
ferromagnesian silicates in granitic systems: American
Journal of Science, v. 283, p. 993-1033.
ORRELL, S. E., J. L. ANDERSON, J. L. WOODEN, and J. E.
WRIGHT, 1987, Proterozoic crustal evolution of the lower
Colorado River region: Rear arc orogenesis to anorogenic
crustal remobilization [abstract]: Geological Society of
America Abstracts with Programs, v. 19, p. 795.
PEARCE, J. A., N. B. W. HARRIS, and A. G. TINDLE, 1984,
Trace element discrimination diagrams for the tectonic
intrepretation of granitic rocks: Journal of Petrology, v.
25, p. 956-983.
POWELL, R, and J. A. EVANS, 1983, A new geobarometer
for the assemblage biotite-muscovite-chlorite-quartz: Jou-
rnal of Metamorphic Geology, v. 1, p. 331-336.
POWELL, R E., 1981, Geology of the crystalline basement
complex, eastern Transverse Ranges, southern California:
Constraints on regional tectonic interpretation; pp. 87-89
in K. A. Howard, M. D. Carr, and D. M. Miller (eds.),
Tectonic Framework of the Mojave and Sonoran De-
serts, California and Arizona: U.S. Geological Survey,
Open-File Report 81-503, 129 p.
__ , 1982, Crystalline basement terranes in the southern
eastern Transverse Ranges, California; pp. 109-136 in J. D.
Cooper (compiler), Geologic Excursions in the Trans-
verse Ranges, Southern California (Cordilleran Sec-
tion Guidebook: Field Trip 11): Geological Society of
America" Boulder, Colorado, USA, 51 p.
Plutonism across the Tujunga-North American terrane boundary 229
RUMBLE, D., III, 1976, The use of mineral solid solutions to
measure chemical potential gradients in rocks: American
Mineralogist, v. 61, 1167-1174.
RU'ITER, M. J., and P. J. WYLLIE, 1988, Experimental
calibration of hornblende as a proposed empirical geo-
barometer [abstract): EOS, Transactions of the American
Geophysical Union, v. 68, no. 27, p. 628.
SALEEBY, J. B., 1981, Ocean floor accretion and volcano-
plutonic arc evolution of the Mesozoic Sierra Nevada; pp.
132-181 in W. G. Ernst (ed.), The Geotectonic Develop-
ment of California (Rubey Volume 1): Prentice-Hall,
Inc., Englewood Cliffs, New Jersey, USA, 706 p.
SIBERLING, N. J., D. L. JONES, M. C. BLAKE, and D. G.
HOWELL, 1987, Lithotectonic Terrane Maps of the
North American Cordillera: U.S. Geological Survey,
Miscellaneous Field Studies Map MF-1874-C.
SILVER, L. T., 1971, Problems of crystalline rocks of the
Transverse Ranges [abstract): Geological Society of Ameri-
ca Abstracts with Programs, v. 3, p. 193-194.
__ , 1982, Evolution of cystalline rock terrains of the
Transverse Ranges, southern California [abstract): Geo-
logical Society of America Abstracts with Programs, v. 14,
p.234.
__ , C. R. McKINNEY, S. DUETSCH, and J. BOLINGER,
1963, Precambrian age determiniation in the western San
Gabriel Mountains, California: Journal of Geology, v. 71,
p.196-214.
SPEAR, F. S., and J. SELVERSTONE, 1983, Quantitative P-
T paths from zoned minerals: Theory and tectonic appli-
cations: Contributions to Mineralogy and Petrology, v. 83,
p.348-357.
SPENCER, J. E., S. J. REYNOLDS, J. L. ANDERSON, G. A.
DAVIS, S. E. LAUBACH, S. M. RICHARD, and S. MARSHAK,
1987, Field trip guide to parts of the Harquahala, Granite
Wash, Whipple, and Buckskin Mountains, west-central
Arizona and southeastern California; pp. 351-364 in G. H.
Davis and E. M. Van den Dolder (eds.), Geologic Diversity
of Arizona and Its Margins: Excursions to Choice
Areas (GSA Guidebook): Arizona Bureau of Geology and
Mineral Technology, Geological Survey Branch, Special
Paper 5, 422 p.
STERN, T. W., P. C. BATEMEN, B. A. MORGAN, M. F.
NEWELL, and D. L. PECK, 1981, Isotopic Ages of Zircon
from the Granitoids of the Central Sierra Nevada,
California: U.S. Geological Survey, Professional Paper
1185,17p.
STONE, P., K. A. HOWARD, and W. HAMILTON, 1983,
Correlation of metamorphosed Paleozoic strata of the
souteastern Mojave Desert region, California and Arizona:
Geological Society of America Bulletin, v. 94, p. 1135-1147.
THOMAS, W. M., H. S. CLARKE, E. D. YOUNG, S. E.
ORRELL, and J. L. ANDERSON, 1988, Precambrian granu-
lite facies metamorphism in the Colorado River region,
Nevada, Arizona, and California; pp. 527-537 in W. G.
Ernst (ed.), Metamorphism and Crustal Evolution of
the Western United States (Rubey Volume 7): Prentice
Hall, Inc., Englewood Cliffs, New Jersey, USA, 1153 p.
TOSDAL, R. M., 1982, The Mule Mountains thrust in the
Mule Mountains, California, and its probable extension in
the southern Dome Rock Mountains, Arizona: A prelimin-
ary report; pp. 55-60 in E. G. Frost and D. L. Martin (eds.),
Mesozoic-Cenozoic Tectonic Evolution of the Colorado
River Region, Californa, Arizona, and Nevada (Ander-
son-Hamilton Volume): Cordilleran Publishers, San
Diego, California, USA, 608 p.
__ , 1984, Mesozoic crystalline rocks of the Mule
Mountains and Vincent-Chocolate Mountains thrusts,
southern California and southwest Arizona: Juxtaposed
pieces of the Jurassic magmatic arc of sothwest North
America; pp. 193-194 in D. G. Howell, D. L. Jones, A. Cox,
and A. M. Nur (eds.), Proceedings of the Circum-Pacific
Terrane Conference (Stanford, California, USA, 1983):
Stanford University, Publications in Geological Sciences
18,248 p.
__ , G. B. HAXEL, and J. E. WRIGHT, 1989, Jurassic
geology of the Sonoran Desert region, southern Arizona,
southeast California and northernmost Sonora: Construc-
tion of a continental-margin magmatic arc; pp. 397-434 in
J. Jenney and A. Reynolds (eds.), The Geologic Evolution
of Arizona: Arizona Geological Socierty Digest, v. 17,866
p.
VEDDER, J. G., D. G. HOWELL, and H. McLEAN, 1983,
Stratigraphy, sedimentation, and tectonic accretion of exo-
tic terranes, southern Coast Ranges, California; pp. 471-
496 in J. S. Watkins and C. L. Drake (eds.), Continental
Margin Geology: American Association of Petroleum
Geologists, Memoir 34, 805 p.
WHALEN, J. B., K. L. CURRIE, and B. W. CHAPPELL, 1987,
A-type granites: Geochemical characteristics, discrimina-
tion, and petrogenesis: Contributions to Mineralogy and
Petrology, v. 95, p. 407-419.
WHITNEY, J. A., and J. C. STROMER, Jr., 1977, The
distribution of NaAlSi
3
0
8
between coexisting microcline
and plagioclase and its effect on geothermometric calcula-
tions: American Mineralogist, v. 62, p. 687 -69l.
WOODEN, J. L., and D. M. MILLER, 1990, Chronologie and
isotopic framework for Early Proterozoic crustal evolution
in the eastern Mojave Desert region, SE California: Jour-
nal of Geophysical Research, v. 95, p. 20133-20146.
___ , andJ. S. STACEY, 1987, Lead isotopic constraints on
the origin of Cordilleran granitic magmatism in the west-
230
J. L. Anderson, A. P. Barth, E. D. Young, and others
ern U.S. [abstract): Geological Society of America Ab-
stracts with Programs, v. 19, p. 465.
__ , J. S. STACEY, K. HOWARD, and D. MILLER, 1986,
Crustal evolution in southeastern California: constraints
from Pb isotopic studies [abstract): Geological Society of
America Abstracts with Programs, v. 18, p. 200.
WRIGHT, J. E., J. L. ANDERSON, and G. A. DAVIS, 1986,
Timing of plutonism, mylonitization and decompression in
a metamorphic core complex, Whipple Mountains, Califor-
nia [abstract): Geological Society of America Abstracts
with Programs, v. 18, p. 201.
__ , K. A. HOWARD, and J. L. ANDERSON, 1987, Isotopic
systematics of zircons from Late Cretaceous intrusive
rocks, southeastern California: Implications for a verti-
cally stratified crustal column [abstract): Geological
Society of America Abstracts with Programs, v. 19, p. 898.
YOUNG, E. D., J. L. ANDERSON, H. S. CLARKE, and W. M.
THOMAS, 1989, Petrology of biotite-cordie rite-garnet
gneiss of the McCullough Range, Nevada, I: Evidence for
Proterozoic low pressure fluid-absent granulite grade
metamorphism in the southern Cordillera: Journal of
Petrology, v. 30, p. 30-60.
__ , and J. P. WOODEN, 1988, Mid-crustal emplacement
of Mesozoic plutons, eastern Mojave Desert: Evidence from
crystallization barometry [abstract): Geological Society of
America Abstracts with Programs, v. 20, p. 244.
__ , and C. F. MILLER, 1983, The geology of the Florence
pluton area, and implications for pluton emplacement, Old
Woman Mountains, southeast California [abstract]: Geo-
logical Society of America Abstracts with Programs, v. 15,
p.41.
ZEN, E-AN, 1985, Implications of magmatic epidote-
bearing plutons on crustal evolution in the accreted ter-
ranes of northwestern North America: Geology, v. 13, p.
266-269.
__ , and J. M. HAMMARSTROM, 1984, Magmatic epidote
and its petrologic significance: Geology, v. 12, p. 515-518.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Geophysical investigations of the cratonic margin in the
Pacific Northwest (USA)
RICHARD 1. THIESSEN
l
, KENT R. JOHNSON
l
,2, and GREGORY B. MOHL
3
lGeology Department, Washington State University, Pullman, Washington 99164-2812, USA
2Aquila Geosciences, P.O. Box 544, Potlatch, Idaho 83855, USA
3Conoco Inc., 800 Werner Court, Casper, Wyoming 82601, USA
(received November 6,1988; revision accepted May 24, 1989)
ABSTRACT
The cratonic margin of North America in west-central Idaho has abeen delineated by strontium isotope
and field studies. Outboard is the Blue Mountains Province, a complex amalgamation of arc-related terranes.
Gravity studies indicate a classic suture-style paired anomaly with a gradient into a low over North America
and isolated maxima within and just outboard of the Mesozoic suture. The latter appear to be due to mafic and
ultramafic pods. Magnetic anomalies change from relatively quiet over the continent to noisy over the Blue
Mountains Province. These geophysical features change character along the length of the suture and do not
trend exactly parallel to it. Near Orofino, Idaho, the suture changes from a N-S-trend to due west and extends
under the Miocene Columbia River Basalts into southeastern Washington. The geophysical signature
parallels this trend. Near Pomeroy, Washington, this signature again turns northward, but the anomalies
split. The magnetic boundary and the minor gravity maxima trend northward into the western edge of the
Kootenai Arc - the surface boundary between miogeoclinal North American lithologies and accreted ter-
ranes. The gravity gradient trends northeastward and extends far into the craton in northern Idaho, where it
corresponds to a change in lower crustal character as observed on COCORP reflection seismic data. In
northeastern Washington and southeastern British Columbia, the gravity and magnetic signatures appear to
follow the suture, but the gravity effects oflocal geologic features are overprinted on them.
INTRODUCTION
The N/S-trending Mesozoic suture zone in west-
central Idaho between cratonic North America and
accreted terranes has been well studied using field
mapping and geochemical studies. Near Orofino,
Idaho (Figure 1), the suture turns abruptly west-
ward into southeastern Washington, where it is
covered by the Miocene Columbia River Basalts.
Here, its location is poorly constrained by a few
isolated pre-basalt outcrops. Farther to the west,
the structure is totally buried. In northeastern
Washington, it re-emerges from beneath the basalts
(heavy line on Figure 1). The continuation of this
crucial suture zone under the basalts may be ana-
lyzed using geophysical techniques.
Globally, gravity surveys have repeatedly
shown that suture zones are typically characterized
by a paired anomaly of a line of maxima paralleled
by a gradient into a low (Gibb and Thomas, 1976;
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 231-240. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
231
232 R. L. Thiessen, K. R. Johnson, and G. B. Mohl
EXPLANATION
o
[] REPUBLIC CAAflEN AND RItLATED ROCKS
L] METAMORPHIC CORE COMPLEXES
rn MESOZOIC AND CENOZOIC INTRUSIVES
El

POMEROY
o
Em LOWeR PALEOZOIC MIOGEOCliNE
II METASEDIMENTS, POSSIBLY hl lllf\, " I.' .... , 9 ......i
LOWER PAlEOZOIC O A
Figure 1. Geologic map of the
study area, based on Huntting and
others (1961), Bond and Wood
(1978), Aadland and Bennett
(1979), Cheney (1980), Utterback
(1984), Carlson (1985), and Holder
and Holder (1988). Heavy solid
line marks the western edge of the
Kootenai arc. The thin, solid lines
labelled ".704" and ".708" show
contoured values of initial stron-
tium isotope ratios in post-accre-
tionary plutons, as determined by
Armstrong and others (1977) and
Fleck and Criss (1985).
13 __ ....... -:-____
Fountain and Salisbury, 1981; Price and Hatcher,
1983; Karner and Watts, 1983; Johnson and others,
1988; Thomas and others, 1988a, 1988b). The exact
relationship of the suture zone and accreted terrains
to the gravity maxima, minima, and gradient varies
from location to location. Aeromagnetic surveys
have also been utilized to examine crustal boun-
daries (Zietz and others, 1971; Mabey and others,
1978). Typically, the change in crustal type across a
major suture is paralleled by a change in intensity
and/or wavelength of magnetic anomalies. Cra-
tonic areas usually show relatively quiet magnetic
signatures, whereas the more diverse natures of
accreted terranes produce more chaotic anomaly
patterns.
In the present study, the geophysical charac-
teristics of the suture zone are established north and
south of the Columbia River Basalts (Figure 1),
where the location of the surface outcrop of the
western edge of the North American craton is well
exposed and studied. We propose an extension ofthe
cratonic boundary across the Columbia Plateau,
using regional geophysical databases.
CRUSTAL BOUNDARY ZONE OF WEST-
CENTRAL IDAHO AND SOUTHEASTERN
WASHINGTON
The suture zone between the accreted Blue
Mountains Province and the North American craton
has been delineated in west-central Idaho by field
mapping (Myers, 1982; Onasch, 1987; Aliberti,
1988; Aliberti and Manduca, 1988a,b; Lund, 1988;
Strayer, 1988; Davidson, 1988; Lund and Snee,
1989; Blake and others, 1989). West of this crustal
boundary zone, the Blue Mountains Province con-
sists of an accreted complex of late Paleozoic to early
Mesozoic island arcs and related oceanic terranes.
East of the suture zone, North America is charac-
Geophysics of the cratonic margin in the Pacific Northwest (USA) 233
terized by possible cratonic basement rocks (Arm-
strong, 1975) overlain by Late Proterozoic Belt
Supergroup metasediments and extensively intru-
ded by the Cretaceous-early Tertiary Idaho Batho-
lith. Hooper and Webster (1982), Lund (1988), and
Lund and Snee (1989) hypothesized the occurrence
of early Paleozoic miogeoclinal metasediments
deposited on the continental margin following latest
Precambrian rifting that formed the western edge of
North America.
Armstrong and others (1977) and Fleck and
Criss (1985) used initial 87Sr/86Sr ratios of post-
accretionary plutonic rocks to show that the suture
extends northward to Orofino, Idaho, where it turns
due west into the southeastern corner of our study
area. Their suture is defined as the zone between
the 0.704 and 0.706 lines (Figure 1). The location of
this geochemical boundary becomes poorly con-
strained because of ubiquitous Miocene Columbia
River Basalt cover. Strontium isotope analyses on
the basalts are not useful in delineating the location
of the suture (Hooper, 1984; Carlson, 1984). Pre-
basalt lithologies (at "C" on Figure 1) yield con-
tinental strontium values (Armstrong and others,
1977; Fleck and Criss, 1985), whereas those at "A"
are similar to lithologies observed in the accreted
terranes. The boundary zone is therefore limited to
a 40-km-wide zone between these outcrops.
Numerous mafic and ultramafic lithologies
have been mapped in the suture zone in west-central
Idaho (Hietanen, 1962; Hamilton, 1969; Myers,
1982; Bonnichsen, 1987; Onasch, 1987; Aliberti and
Manduca, 1988a,b). These tend to be within and
just outboard of the strontium-defined crustal
boundary zone. These dense lithologies should pro-
duce a series of gravity maxima and, in fact, cor-
respond to a discontinuous series of maxima on the
Idaho state gravity map (Bankey and others, 1984)
These highs follow the suture northward, and then
westward from near Orofino, Idaho, into southeast-
ern Washington. This line of maxima is paralleled
by a gradient into a low over the Idaho Batholith.
This gradient also appears to swing westward near
Orofino; however, its magnitude decreases from
=75 to 45 mgal. Magnetic anomalies (USGS, 1978)
east of this boundary are subdued and quiet,
whereas those to the west are much stronger and
more chaotic, showing that this zone corresponds to
a major magnetic crustal domain boundary. These
geophysical signatures change in nature and locally
cross the trace of the strontium-defined boundary
zone, which may be due to variations in geology or to
the poor resolution of the regional geophysical
surveys.
In order to determine if more detailed gravity
data could be used to better constrain the location of
the cratonic margin in the Columbia Plateau, a
detailed gravity survey (Mohl and Thiessen,
1985a,b,c, in press) was run in southeastern Wash-
ington (diamond patterned line on Figure 1). The
line of gravity maxima observed by Bankey and
others (1984) in west-central Idaho continues into
Washington and is well displayed in the detailed
data set as a discontinuous line of maxima. This
feature also appears on Lyons and O'Hara's (1982)
regional gravity database (anomaly B on Figure 2).
This anomaly trends between outcrops of accreted
terranes ("A" locations on Figure 1) and cratonic
lithologies (locations labelled C in Figure 1), just
outboard of the projected location of the strontium-
defined crustal boundary zone. This corresponds
with the general location of mafic and ultramafic
lithologies seen elsewhere in this suture zone.
Paleomagnetic studies on a 12 Ma basalt flow in-
dicate up to 10.3 of rotations associated with this
structure near Lewiston, Idaho (Thiessen and
others, 1986). The line of gravity maxima continues
westward through the detailed study area and ap-
pears to split into two anomalies (labelled G and H
on Figure 2) that extend to the north and north-
west. Paleomagnetic studies do not reveal appreci-
able rotation of a 12 Ma basalt flow at either ano-
maly G or H, although anomaly G corresponds to a
major domain boundary in lineament orientations
(Thiessen and others, 1986). The gradient that
extends from Idaho (Bankey and others, 1984) into
Washington (gradient D on Figure 2) is also defined
by the detailed gravity survey and continues to
parallel the maxima.
Immediately north of our detailed study area,
three detailed gravity profiles were collected by
Prieto and others (1985) and Johnson and others
(1989). Both studies defined the continuation of the
gradient and anomalies G and H, with G being the
stronger anomaly. Both anomalies are also associa-
ted with magnetic maxima as well. Prieto and
others (1985) collected a series of magnetotelluric
soundings which indicated a change from more
conductive basement (100 Ohm-m) in the west to
more resistant basement (1000 Ohm-m) to the east.
The transition occurred between anomalies G and
H; they ascribed it to a continuation of the Pasayten
Fault from northwestern Washington and did not
mention the cratonic margin.
234
R. L. Thiessen, K. R. Johnson, and G. B. Mohl
Figure 2. Bouguer gravity map of study area.
Data are directly from Lyons and O'Hara (1982).
Pertinent lines of gravity maxima are shown with a
dashed line and gradients are indicated with a very
heavy solid line. Both are based on Lyons and
O'Hara (1982), with modifications from Cady and
Meyer (1976), Finn and others (1984), and GSA
(1987a). The Olympic-Wallowa Lineament (OWL),
the Lewis and Clark Line (LCL), and the Klamath-
Blue Mountains Lineament (KB) are major crustal
structures in the plateau.
Farther to the north, the location of the geo-
physical anomalies paralleling the cratonic margin
are poorly constrained and must be extended into
the Columbia Plateau from known locations of the
margin in northern Washington.
CONTINUATION OF THE CRUSTAL BOUNDARY
ZONE FROM THE NORTH
Using initial 87Sr/86Sr ratios, Armstrong and
others (1977) and Fleck and others (1989) delineated
the western edge of the North American craton as
extending through north-central or northwestern
Washington. This conflicts with many terrane maps
of British Columbia (Coney and others, 1980;
Monger and others, 1982; Saleeby, 1983), which
show the surface trace of the cratonic margin along
the western edge of the Kootenai Arc, between the
accreted terrane of Quesnellia and miogeoclinal
o 25 50 KM
0,==== =2;. 5 _..-50 MI
CONTOUR INTERVAL' 5 MGALS
N

lithologies associated with North America (heavy
line on Figure 1). A slice of North American crust
may extend westward, below the accreted terranes
(Price, 1986; Brown and others, 1986; Orr and
Cheney, 1987), and COCORP deep crustal seismic
reflection profiles, interpreted by Potter and others
(1986) and Sanford and others (1988), support this
hypothesis. Alternately, the accreted terranes
themselves may have cores of continental litho-
logies that are totally exotic to North America, as
hypothesized for the Quesnellia terrane (Monger
and others, 1982; Howell and others, 1985).
Monger and Price (1979) summarized geophysi-
cal evidence for the western boundary of the Koo-
tenai Arc being the cratonic margin, including data
from electrical resistivity surveys, magnetic ano-
maly patterns, and seismic velocity structure of the
crust and mantle indicating a 10 to 15 km eastward
thickening of the crust along this structure. Price
and Hatcher (1983) observed an associated 70 to 100
Geophysics of the cratonic margin in the Pacific Northwest (USA) 235
4 6
120
o 25 50
I KM
0Ci = =2.5 _ ... 50 MI
CONTOUR INTERVAL 200 GAMMAS
- DNAG Ae(omagnetics
........... DNAG Aeromag" al te(nate pick
- l\'URE Aeromagnetics
- - - - - Zielz and oThers. 1971
mgal decrease in the gravity across a 100 to 140 km
wide zone. Stacey (1973) modeled this as being due
to a 20 km thickening of the crust not related to
accretionary events, and included increasing mantle
and crustal densities toward the craton.
Examination of a regional gravity map (GSA,
1987a) reveals a discontinuous line of gravity
maxima in British Columbia that marks the surface
trace of the suture zone between cratonic North
America and the accreted terranes. The gravity
gradient is to the east, over North American crust.
Analyses of compilation maps of magnetic anoma-
lies (GSA, 1987b; GSC, 1987) reveal relatively quiet
magnetic signatures over the craton and more ir-
regular, noisy patterns over the accreted terranes to
the west. The magnetic domain boundary between
these regions follows the gravity maxima and map-
ped suture very closely in central British Columbia.
Figure 3. Magnetic anomalies for the study area,
based on a compilation of six 1:250,000 scale maps
(NURE, 1983). Boundaries are based on regional
magnetic maps including the DNAG map (GSA,
1987b), Zietz and others (1971), and NURE (1983).
However, both become less well defined in south-
eastern British Columbia and northeastern Wash-
ington, where the geophysical signatures of Creta-
ceous and Paleocene intrusives and gneiss domes
are overprinted on them (Cady, 1980; Cady and Fox,
1984; Thomas and others, 1988b).
A line of gravity maxima just outboard of the
mapped location of the miogeocline-accreted terrane
boundary (heavy line on Figure 1) can be seen in
northeastern Washington (anomaly M in Figure 2).
This anomaly correlates with a gravity maxima
which Cady and Fox (1984) hypothesized is due to
the presence of denser lithologies in the Kettle
gneiss dome. They noted, however, that the south-
ern part of our anomaly M is actually to the east of
the gneiss dome and lies over denser lithologies that
correlate with Quesnellia.
236 R. L. Thiessen, K. R. Johnson, and G. B. Mohl
LOCATION OF THE CRATONIC MARGIN WITHIN
THE COLUMBIA PLATEAU
The line of gravity maxima just outboard of the
cratonic margin (anomaly M on Figure 2) splits into
four potential traces (anomalies I through L on Fig-
ure 2) in the Columbia Plateau, any of which
eventually joins with the margin-defining anoma-
lies to the south (anomaly B). Path L cannot be the
cratonic margin, as it crosses major outcrop belts of
North American miogeoclinal and Belt Supergroup
rocks (Figure 1). Cady (1980) and Cady and Fox
(1984) showed that the maxima which define path L
correspond to the mapped location of the Spokane
gneiss dome (Figure 1). Path K also crosses
through continental sedimentary rocks, but exact
correlation of these with the Belt, miogeocline, or
other lithologic group is unknown. Path I and Path
J both extend through the Columbia River Basalts
where there are no outcrops of pre-basalt lithologies.
Therefore, either path may correspond to the loca-
tion of the cratonic margin.
In order to further constrain the location of the
margin through the eastern Columbia Plateau,
three regional magnetic maps (Zietz and others,
1971; NURE, 1983; GSA, 1987b) were examined and
major domainal boundaries that may reflect crustal
changes were denoted (Figure 3) and used to illus-
trate the boundary locations. The domain boun-
daries observed on the various magnetic databases
do not differ significantly. In the northern portion of
this region (Figure 3), the boundary separates a
region of generally higher magnitude anomalies to
the west from relatively quiet continental signa-
tures to the east. In the southeastern part of the
region, the boundary is very well constrained
between the magnetically diverse Blue Mountains
Province and North America. In the central portion
of the study area, the domain boundary appears to
be more diffuse, with several possible paths (Figure
3). In general, possible locations of the magnetically
defined crustal boundary zone (Figure 3) appear
parallel to and just inboard of the lines of gravity
maxima.
NATURE OF PERTINENT GRAVITY GRADIENTS
In southeastern British Columbia, a major gra-
vity gradient (Stacey, 1973; Monger and Price, 1979;
Price and Hatcher, 1983) parallels the surface trace
ofthe cratonic margin, producing the paired gravity
anomaly typical of major suture zones. As noted by
Cady (1980) and Cady and Fox (1984), this gradient
extends southward into the United States (anomaly
N on Figure 2) but swings eastward into northern
Idaho (anomaly 0), away from the cratonic boun-
dary. The gradient is sharp, leading Cady (1980)
and Cady and Fox (1984) to conclude that it pro-
bably is due to effects in the upper 12 km of the
crust. However, their computer model (Cady and
Fox, 1984) of a profile across the gradient uses three
different causes to simulate the gradient: (1) an
eastward deepening of high-grade metamorphic
rocks from the surface to a depth of = 10 km; (2) a
west-to-east change in upper mantle densities from
3.21 glcc to 3.27 g/cc; and (3) a deepening of the
Moho from =32 km (W) to 37 km (E). We agree
with Cady and Fox (1984) that the gradient in
Washington and Idaho corresponds to the eastern
margin of the Eocene extensional systems of the
gneiss domes. A series of COCORP seismic reflec-
tion profiles across the gradient shows that it
corresponds to a deepening and major change in the
character of the lower crust (Johnson and others,
1989; Pruss en and Potter, 1989). Therefore, gra-
dient 0 does not directly relate to the cratonic mar-
gin but, instead, indicates later tectonic events
overprinted on it.
Gradient P (Figure 2) may be the southern
continuation of gradient O. If so, then it appears to
have been offset by the Lewis and Clark Line (LCL).
Gradient D was observed in our detailed gravity
survey. This anomaly parallels the gravity maxi-
mum at this location (anomaly B, Figure 2). These
two serve to define a paired gravity anomaly that
resembles those observed at suture zones globally.
Gravity modeling across this pair of anomalies
(Mohl and Thiessen, 1985a,b, in press) indicates
that the maximum is caused by north-dipping pods
of denser lithologies caught within the suture zone
and that the gradient to a low is created by shallow
level intrusions. Strayer (1988) showed that the
87Sr/86Sr initial ratio boundary at its pronounced
bend near Orofino, at the eastern end of gravity
gradient (D), is associated with a major NE-dipping
zone between dioritic accreted terrane on the south
and more felsic continental North America on the
north. This gravity gradient appears to wrap into
gradient P. Alternately, if P represents the eastern
limit of Eocene extension, older gravitational effects
of the cratonic margin may be truncated by more
recently formed gradient P.
Geophysics of the cratonic marg;in in the Pacific Northwest (USA) 237
Riddihough and others (1986) defined the
Klamath-Blue Mountains lineament (KB on Figure
2) as a major crustal boundary based on a 50-80
mgal gravity gradient and a 1000 m topographic
differential across it. This major structure cor-
responds to a magnetic domain boundary between
chaotic anomalies to the southeast and intermediate
ones to the northwest (Figure 3). This feature was
previously defined as the northwestern margin of
the Columbia Arc (Barrash and others, 1983) and
the southeastern edge of the Columbia Embayment
(Hamilton and Myers, 1966). The geophysically
defined cratonic margin trends northward, away
from the KB (Figures 2 and 3). Mohl and Thiessen
(1985c) showed that, at its northeastern end, the KB
corresponds with the Hite Fault, a major normal
fault that offsets the Columbia River Basalts.
Gravity anomalies observed in the detailed survey
area indicate that this fault propagated northward
across the trace of the cratonic margin.
Several workers have postulated that the
Olympic-Wallowa lineament (OWL on Figure 1)
may be the trace of a continental-oceanic crustal
boundary (Skehan, 1966; Hammond, 1979). Others
believe that the OWL is a major zone of dextral
shear associated with the boundary of crustal blocks
(Bentley and Anderson, 1979; Reidel and others,
1984). Regional gravity maps (Figure 2) and EM
soundings (Mitchell and Bergstrom, 1983) do not
show any major anomalies that correspond to the
OWL. It is aligned with regional aeromagnetic
anomalies as shown by Zietz and others (1971), GSA
(1987b), and Figure 3, but it is not aligned with a
major magnetic domain boundary. At the western
end of our detailed gravity study area (Figure 2),
the gravity defined margin extends north- or north-
westward, not southward toward the OWL. The
southern extension of the OWL, near the Wallowa
Mountains, trends through the accreted Blue Moun-
tains Province and therefore is definitely not the
cratonic margin at this location.
CONCLUSION
The boundary between the western edge of the
North American continent and accreted terranes is
well constrained by field mapping in west-central
Idaho, northeastern Washington, and southeastern
British Columbia. In these regions, the suture zone
exhibits a classic paired gravity anomaly of parallel
maxima (over the accreted terrains) and minima,
with a gradient between the two. The boundary also
aligns with a major aeromagnetic domain boundary,
with the accreted terranes being characterized by
chaotic anomalies and continental North America
being magnetically quiet. This suture zone extends
under the Miocene Columbia River Basalts, where
it is lost to surface mapping techniques. Using
regional gravity and magnetic databases, the
location of the western edge of the North American
craton through the eastern Columbia Plateau is
approximately delineated. However, the geophysi-
cal signature becomes somewhat confused in the
plateau, with the gravity maximum splitting into
several possible paths. The aeromagnetic domain
bOUItdary is not as well defined but appears to follow
one of the gravity maxima, suggesting that it is a
slightly more probable location of the cratonic
margin.
REFERENCES
AADLAND, R. K., and E. H. BENNETT, 1979, Geologic
Map of the Sandpoint Quadrangle, Idaho and Wash-
ington: Idaho Bureau of Mines and Geology, Geologic Map
Series, Sandpoint 2 Degree Quadrangle, scale 1:250,000.
ALIBERTI, E. A., 1988, A Structural, Petrographic, and
Isotopic Study of the Rapid River Area and Selected
Mafic Complexes in the Northwestern United States:
Implications for the Evolution of an Abrupt Island
Arc-Continent Boundary; PhD dissertation, Harvard
University, Cambridge, Massachusetts, USA, 194 p.
__ , and C. A. MANDUCA, 1988a, A transect across an
island arc-continent boundary in west-central Idaho; pp.
99-107 in P. K. Link and W. R Hackett (eds.l, Guidebook
to the Geology of Central and Southern Idaho: Idaho
Geological Survey, Bulletin 27,319 p.
__ , and C. A. MANDUCA, 1988b, Field guide to a tran-
sect across an island arc-continent boundary in west-
central Idaho; pp. 181-190 in S. E. Lewis and R. B. Berg
(eds.), Precambrian and Mesozoic Plate Margins:
Montana Bureau of Mines and Geology, Special Paper 96,
195p.
ARMSTRONG, R. L., 1975, Precambrian (1500 m.y. old)
rocks of central Idaho - The Salmon River Arch and its
role in Cordilleran sedimentation and tectonics: American
Journal of Science, v. 275-A, p. 437-467.
__ , W. H. TAUBENECK, and P. O. HALES, 1977, Rb-Sr
and K-Ar geochronometry of Mesozoic granitic rocks and
their Sr isotopic composition, Oregon, Washington, and
Idaho: Geological Society of America Bulletin, v. 88, p. 397-
411.
238
R. L. Thiessen, K. R. Johnson, and G. B. Mohl
BANKEY, V., M. WEBRING, D. R. MABEY, and D. KLEIN-
KOPF, 1984, Complete Bouguer Gravity Anomaly Map
of Idaho: Idaho Bureau of Mines and Geology, scale
1:500,000.
BARRASH, W., J. BOND, and R. VENKATAKRISHNAN,
1983, Structural evolution of the Columbia Plateau in
Washington and Oregon: American Journal of Science, v.
283, p. 897 -935.
BENTLEY, R. D., and J. L. ANDERSON, 1979, Right-lateral
strike-slip faults in the western Columbia Plateau [ab-
stract]: EOS, Transactions of the American Geophysical
Union, v. 60, p. 961.
BLAKE, D. E., R. L. THIESSEN, and A. J. WATKINSON,
1989, Structural analysis of the West Idaho suture zone in
the Salmon River gorge, Riggins region, west-central
Idaho [abstract]: Geological Society of America Abstracts
with Programs, v. 21, p. 58.
BOND, J. G., and C. H. WOOD, 1978, Geologic Map of
Idaho: Idaho Bureau of Mines and Geology, scale 1:
500,000.
BONNICHSEN, B. 1987, Pre-Cenozoic geology of the West
Mountain-Council Mountain New Meadows area, west-
central Idaho: U.S. Geological Survey, Professional Paper
1436, p. 151-170.
BROWN, R. L, J. M. JOURNEAY, L. S. LANE, D. C.
MURPHY, and C. J. REES, 1986, Obduction, backfolding,
and piggyback thrusting in the metamorphic hinterland of
the southeastern Canadian Cordillera: Journal of Struc-
tural Geology, v. 8, p. 255-268.
CADY, J. W., 1980, Gravity highs and crustal structure,
Omineca Crystalline Belt, northeastern Washington and
southeastern British Columbia: Geology, v. 8, p. 328-332.
__ , and K. F. FOX, Jr., 1984, Geophysical Inter-
pretation of the Gneiss Terrane of Northern Washing-
ton and Southern British Columbia, and Its Implica-
tions for Uranium Exploration: U.S. Geological Survey,
Professional Paper 1260,29 p.
__ , and R. F. MEYER, 1976, Bouguer Gravity Map of
the Okanogan, Sandpoint, Ritzville, and Spokane 1 x 2
Quadrangles, Northeastern Washington and Northern
Idaho: U.S. Geological Survey, Geophysical Investigations
Map GP914, scale 1:250,000.
CARLSON, D. H., 1985, Geology and Petrochemistry of
the Keller Butte Pluton and Associated Intrusive
Rocks in the South Half of the Nespelem and North-
ern Half of the Grand Coulee Dam Quadrangles, and
the Development of Cataclasites and Fault Lenses
Along the Manila Pass Fault, Northeastern Washing-
ton: PhD dissertation, Washington State University,
Pullman, Washington, USA, 181 p.
__ , 1984, Isotopic constraints on Columbia River flood
basalt genesis and the nature of the subcontinental
mantle: Geochemica et Cosmochimica Acta, v. 48, p. 2357-
2372.
CHENEY, E. S., 1980, Kettle Dome and related structures
of northeastern Washington; pp. 463-483 in M. D. Critten-
den, Jr., P. J. Coney, and G. H. Davis (eds.), Cordilleran
Metamorphic Core Complexes: Geological Society of
America, Memoir 153, 496 p.
CONEY, P. J., D. L. JONES, and J. W. H. MONGER, 1980,
Cordilleran suspect terranes: Nature, v. 288, p. 329-333.
DAVIDSON, G. F., 1988, Field guide of mylonitic rocks west
of Orofino, Idaho; pp. 171-174 in S. E. Lewis and R. B. Berg
(eds.) Precambrian and Mesozoic Plate Margins:
Montana Bureau of Mines and Geology, Special Publication
96, 195p.
FINN, C., W. M. PHILLIPS, and D. L. WILLIAMS, 1984,
Gravity Maps of the State of Washington and Adjacent
Areas: U.S. Geological Survey, Open File Report 84-416,
scale 1:250,000.
FLECK, R. J., and R. E. CRISS, 1985, Strontium and oxygen
isotopic variations in Mesozoic and Tertiary plutons of
central Idaho: Contributions to Mineralogy and Petrology,
v. 90, p. 291-308.
__ , R. W. KISTLER, and R. E. CRISS, 1989, Location and
isotopic characteristics of the Mesozoic continental margin
of North America in Idaho and Washington [abstract]:
Geological Society of America Abstracts with Programs, v.
21, p. 78-79.
FOUNTAIN, D. M., and M. H. SALISBURY, 1981, Exposed
cross-sections through the continental crust: Implications
for crustal structure, petrology and evolution: Earth and
Planetary Science Letters, v. 56, p. 263-277.
GSC, 1987, Magnetic Anomaly Map of Canada: Geo-
logical Survey of Canada, Map 1-255A (5th Edition), scale
1:5,000,000.
GIBB,R. A., and M. D. THOMAS, 1976, Gravity signature of
fossil plate boundaries in the Canadian shield: Nature, v.
262, p. 199-200.
GSA, 1987a, Gravity Anomaly Map of North America:
Geological Society of America, Continent-Scale Map-002,
scale 1:5,000,000.
__ , 1987b, Magnetic Anomaly Map of North Ameri-
ca: Geological Society of America, Continent-Scale Map-
003, scale 1:5,000,000.
HAMILTON, W. B., 1969, Reconnaissance Geologic Map
of the Riggins Quadrangle, West-Central Idaho: U.S.
Geological Survey, Map 1-579, scale 1:125,000.
__ , and W. B. MYERS, 1966, Cenozoic tectonics of the
western United States: Reviews of Geophysics, v. 4, p. 509-
549.
HAMMOND, P. E., 1979, A tectonic model for the evolution
of the Cascade Range; pp. 219-237 in J. M. Armentrout, M.
Geophysics of the cratonic margin in the Pacific Northwest (USA) 239
R. Cole, and H. Ferbest (eds.), Cenozoic Paleogeography
of the Western United States: Society of Economic
Paleontologists and Mineralogists, Pacific Paleogeography
Symposium 3, 335 p.
HIETANEN, A., 1962, Metasomatic Metamorphism in
Western Clearwater County, Idaho: U.S. Geological
Survey, Professional Paper 344-A, 116 p.
HOLDER, G. A. M., and HOLDER, R. W., 1988, The Cove
batholith: Tertiary plutonism in north-east Washington
associated with graben and core-complex (gneiss dome)
formation: Geological Society of America Bulletin, v. 100,
p.1971-1980.
HOOPER, P. R., 1984, Physical and chemical constraints on
the evolution of the Columbia River Basalt: Geology, v. 12,
p.495-499.
__ , and G. D. WEBSTER, 1982, Geology of the Pul-
lman, Moscow West, Colton and Uniontown 7112 Min-
ute Quadrangles, Washington and Idaho: Washington
Department of Natural Resources .. Geologic Map GM-26,
scale 1:62,500.
HOWELL, D. G., D. L. JONES, and E. R. SCHERMER, 1985,
Tectonostratigraphic terranes of the Circum-Pacific
region; pp. 3-30 in D. G. Howell (ed.), Tectonostrati-
graphic Terranes of the Circum-Pacific Region:
Circum-Pacific Council for Energy and Mineral Resources,
Earth Science Series 1, 581 p.
HUNTTING, M. T., W. A. G. BENNETT, Y. E. LIVINGSTON,
Jr., and W. S. MOEN, 1961, Geologic Map of Washington:
Washington Division of Mines and Geology, scale 1:
500,000.
JOHNSON, K. R., G. B. MOHL, and R. L. THIESSEN, 1988,
Gravity signatures of suture zones: A comparison between
the island arc-continent boundary in Idaho and the Mot-
agua suture of Guatemala [abstract]: Geological Society of
America Abstracts with Programs, v. 20, p. 423.
__ , R. L. THIESSEN, and M. R. PARODI, 1989, Geo-
physical constraints on the location of the cratonic margin
beneath the Columbia Plateau [abstract]: Geological
Society of America Abstracts with Programs, v. 21, p. 98.
KARNER, G. D., and A. B. WATTS, 1983, Gravity ano-
malies and flexure of lithosphere at mountain ranges:
Journal of Geophysical Research, v. 88, p. 10449-10477.
LUND, K., 1988, The Salmon River suture, western Idaho
- An island arc-continental boundary; pp. 103-110 in S. E.
Lewis and R. B. Berg (eds.), Precambrian and Mesozoic
Plate Margins: Montana Bureau of Mines and Geology,
Special Publication 96, 195 p.
__ , and L. W. SNEE, 1989, Metamorphism, structural
development and age of the continent-island arc juncture
in west-central Idaho; pp. 296-330 in W. G. Ernst (ed.),
Metamorphism and Crustal Evolution, Western Con-
terminous United States (Rubey Volume 7): Prentice-
Hall, Inc., Englewood Cliffs, New Jersey, USA, 1153 p.
LYONS, P. L., and N. W. O'HARA, 1982, Gravity Anomaly
Map of the United States: Society of Exploration
Geophysicists, scale 1:2,500,000.
MABEY, D. R., L. ZIETZ, G. P. EATON, and M. D. KLEIN-
KOPF, 1978, Regional magnetic patterns in part of the
Cordillera in the western United States; pp. 93-106 in R. B.
Smith and G. P. Eaton (eds.), Cenozoic Tectonics and
Regional Geophysics of the Western Cordillera: Geo-
logical Society of America, Memoir 152,395 p.
MITCHELL, T. H., and K. A. BERGSTROM, 1983, Pre-
Columbia River Basalt group stratigraphy and structure
in the central Pasco basin; pp. 56-74 in J. A. Caggianno
(ed.), Preliminary Interpretation of the Tectonic Stabi-
lity of the Reference Repository Location, Cold Creek
Syncline, Hanford Site: Rockwell Hanford Operations,
Publication SD-BWl-TI-ll1, 175 p.
MOHL, G. B., and R. L. THIESSEN, 1985a, Study of the cra-
tonic margin in eastern Washington as defined by gravity
measurements [abstract]: EOS, Transactions of the Ameri-
can Geophysical Union, v. 66, p. 26.
___ , and R. L. THIESSEN, 1985b, Gravity measurements
in southeastern Washington and implications as to the
location of the cratonic margin [abstract]: Geological
Society of America Abstracts with Programs, v. 17, p. 257.
__ , and R. L. THIESSEN, 1985c, Subsurface structure in
southeastern Washington as defined by Bouguer gravity
measurements [abstract]: Geological Society of America
Abstracts with Programs, v. 17, p. 370.
__ , and R. L. THIESSEN (in press), Gravity studies of an
island arc-continental suture in west-central Idaho and
adjacent Washington: U.S. Geological Survey, Professional
Paper 1438, in press.
MONGER, J. W. H., and R. A. Price, 1979, Geodynamic
evolution of the Canadian Cordillera - Progress and pro-
blems: Canadian Journal of Earth Science, v. 16, p. 770-
791.
__ , R. A. PRICE, and K. J. TEMPELMAN-KLUIT, 1982,
Tectonic accretion and origin of the two major metamor-
phic and plutonic welts in the Canadian Cordillera:
Geology, v. 10, p. 70-75.
MYERS, P. E., 1982, Geology of the Harpster Area, Ida-
ho County, Idaho: Idaho Bureau of Mines and Geology,
Bulletin 27, 46 p.
NURE (National Uranium Resource Evaluate Program),
1983, Residual Intensity Magnetic Anomaly Contour
Map: U.S. Department of Energy Open File Map (Okano-
gan Quadrangle, GJM-460; Pullman Quadrangle, GJM-
445; Ritzville Quadrangle, GJM-461; Sandpoint Quad-
rangle, GJM-443; Spokane Quadrangle, GJM-444; Walla
Walla Quadrangle, GJM-462), scale 1;250,000.
240 R. L. Thiessen, K. R. Johnson, and G. B. Mohl
ONASCH, C. M., 1987, Temporal and spatial relations be-
tween folding, intrusion, metamorphism, and thrust fault-
ing in the Riggins area, west-central Idaho: U.S. Geo-
logical Survey Professional Paper 1436, p. 139-149.
ORR, K. E., and E. S. CHENEY, 1987, Kettle and Okanogan
Domes, northeastern Washington and southern British
Columbia: Washington Division of Geology and Earth
Resources, Bulletin 77, p. 55-71.
POTTER, C. J., W. E. SANFORD, T. R. YOOS, E. L.
PRUSSEN, R. W. KEACH, J. E. OLIVER, S. KAUFMAN, and
L. D. BROWN, 1986, COCORP deep seismic reflection
traverse of the interior of the North American Cordillera,
Washington and Idaho: Implications for orogenic evolu-
tion: Tectonics, v. 5, p. 1007-1025.
PRICE, R. A., 1986, The southeastern Canadian Cordillera:
Thrust faulting, tectonic wedging, and delamination of the
lithosphere: Journal of Structural Geology, v. 8, p. 239-
254.
__ , and R. D. HATCHER, Jr., 1983, Tectonic significance
of similarities in the evolution of the Alabama-Penn-
sylvania Appalachians and the Alberta-British Columbia
Canadian Cordillera; pp. 149-160 in R. D. Hatcher, Jr., H.
Williams, and 1. Zietz (eds.), Contributions to the Tec-
tonics and Geophysics of Mountain Chains: Geological
Society of America, Memoir 158,228 p.
PRIETO, C., C. PERKINS, and E. BERKMAN, 1985, Colum-
bia River Basalt Plateau - An integrated approach to
interpretation of basalt-covered areas: Geophysics, v. 50, p.
2709-2719.
PRUSSEN, E., and C. POTTER, 1989, The seismic reflection
character of the Moho beneath the northwestern U.S. -
WA, ID, and MT [abstract]: Geological Society of America
Abstracts with Programs, v. 21, p. 131.
REIDEL, S. P., G. R. SCOTT, D. R. BAZARD, R. W. CROSS,
and B. DICK, 1984, Post- 12 million year clockwise rotation
in the central Columbia Plateau, Washington: Tectonics,
v. 3, p. 251-273.
RIDDIHOUGH, R., C. FINN, and R. COUCH, 1986, Klamath-
Blue Mountain lineament, Oregon: Geology, v. 14, p. 528-
531.
SALEEBY, J. B., 1983, Accretionary tectonics of the North
American Cordillera: Annual Reviews of Earth and Plane-
tary Sciences, v.15, p. 45-73.
SANFORD, W. E., C. J. POTTER, and J. E. OLIVER, 1988,
Detailed three-dimensional structure of the deep crust
based on COCORP data in the Cordilleran interior, north-
central Washington: Geological Society of America Bul-
letin, v. 100, p. 60-71.
SKEHAN, J. W., 1966, Olympic-Wallowa lineament: A
major deep-seated tectonic feature in the Pacific North-
west; pp. 158-159 in Abstracts for 1965: Geological Socie-
ty of America, Special Paper 87,366 p.
STACEY, R. A., 1973, Gravity anomalies, crustal structure,
and plate tectonics in the Canadian Cordillera: Canadian
Journal of Earth Science, v. 10, p. 615-628.
STRAYER, L. M., IV, 1988, Field guide to deformation and
sense of displacement in mylonitic rocks near Orofino,
Idaho; pp. 165-169 in S. E. Lewis and R. B. Berg (eds.),
Precambrian and Mesozoic Plate Margins: Montana
Bureau of Mines and Geology, Special Publication 96, 195
p.
THIESSEN, R. L., G. B. MOHL, and J. B. LIM, 1986, Geo-
physical investigations of southeastern Washington, p. 17-
18 in P. C. Beaver (ed.), Cenozoic Geology of Moscow,
Idaho and Surrounding Areas: Proceedings of the Ele-
venth Annual Meeting of the Tobacco Root Geological
Society, 74 p.
THOMAS, M. D., R. A. F. GRIEVE, and V. L. SHARPTON,
1988a, Gravity domains and assembly ofthe North Ameri-
can continent by collisional tectonics: Nature, v. 331, p.
333-334.
__ , D. W. HALLIDAY, and B. FELIX, 1988b, Preliminary
results of gravity surveys along the LITHOPROBE
southern Canadian Cordilleran transect; pp. 111-116 in
Current Research, Part E: Geological Survey of Canada,
Paper 88-IE, 254 p.
USGS, 1978, Aeromagnetic Map of Idaho: U.S. Geologi-
cal Survey, Geophysical Investigations Map GP-919, scale
1: 500,000.
UTTERBACK, W. C., 1984, Revised Geology and Mineral
Potential of the Colville Reservation: Generalized
Bedrock Map of the Colville Reservation, Okanogan
and Ferry Counties, Washington: Washington Map
Folio, scale 1:250,000.
ZIETZ, 1., B. C. HEARN, Jr., M. W. HIGGINS, G. D.
ROBINSON, and D. A. SWANSON, 1971, Interpretation of
an aeromagnetic strip across the northwestern United
States: Geological Society of America Bulletin, v. 82, p.
3347-3372.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Sedimentation and basin evolution along the Mesozoic margin
of western North America
WILLIAM L. BILODEAU
Department of Geology, California Lutheran University, 60 West Olsen Road, Thousand Oaks, California 91360-2787, USA
(received and accepted December 5, 1988)
INTRODUCTION
During the Mesozoic, the western, or Cordi-
lleran, margin of North America was dominated by
convergent plate tectonics. Sedimentary basins
within and peripheral to this margin developed in a
variety of tectonic settings (Dickinson, 1976a) and
contain basin-fill sequences that reflect the complex
interplay of basin margin (and interior) tectonism,
erosion, basin subsidence, sedimentation, eustasy,
and climate. Along this margin, large and small
lithospheric plates interacted in a zone of subduc-
tion, oblique convergence, and terrane accretion
that affected the interior of the North American
plate far to the east. Accretion and lateral strike-
slip segmentation and dispersal of primarily
composite intra-oceanic island-arc terranes during
the Mesozoic along different segments of the
continental margin - i.e., Sonomia in the Early
Triassic (Sonoma orogeny); Stikinia in the Late
Jurassic; Klamath-western Sierra Nevada terrane
in the Late Jurassic (Nevadan orogeny); and the
Alexander-Wrangellia-Peninsular composite ter-
rane in the mid- to Late Cretaceous - produced
regionally diverse evolutionary histories (Howell
and others, 1985, 1987).
The several types of Mesozoic basins can be
classified in a variety of ways based on their plate-
tectonic setting, proximity to a particular type of
plate margin, and the style of local and/or regional
tectonism that caused the basin to form, fill, and be
preserved for our eventual study (Dickinson, 1976b;
Reading, 1982; Kingston, Dishroon and Williams,
1983; Ingersoll, 1988). Typical basin types present
in the Mesozoic western Cordillera are: (1) ocean
basin and trench deposits in accretionary prism/sub-
duction zone complexes; (2) forearc basins; (3) intra-
arc or backarc rift basins floored by oceanic crust;
(4) successor basins or small peripheral foreland-
style basins developed on accreted terranes during
their amalgamation and final accretion; (5) backarc
continental rift or sag basins; (6) strike-slip or
wrench basins in forearc or intra-arc settings; (7) a
large retroarc foreland-basin; and (8) broken fore-
land, Laramide-style basins. The remainder of this
summary is devoted to brief descriptions of a
selection of these basins as they developed along and
inboard of the Mesozoic margin of western North
America (Figure 1). Cited references are not
intended to be comprehensive but to suggest recent
papers that contain important new data and much
more thorough reference lists.
TRIASSIC
The Triassic began, in western North America,
with accretion of the Sonomia "arc" terrane (Speed,
1979, 1982) and emplacement of the Golconda
allochthon (Miller and others, 1982, 1984; Brueck-
ner and Snyder, 1985). The western margin of
North America was a passive margin continental
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 241-247. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
241
242
W. L. Bilodeau
Figure 1. Map of western North America showing
locations of the Mesozoic basins discussed in the text.
shelf facing an open ocean or backarc basin. This
lower Triassic marine shelf was overthrust east-
ward by ocean basin/accretionary prism material
(Havallah-Schoonover sequence in the Golconda
allochthon) as the Sonomia arc terrane accreted to
the continent and the intervening ocean or backarc
basin was closed by subduction beneath the
Sonomia arc (Figure 2A).
Following accretion, the Sonomia arc terrane in
northwestern Nevada apparently subsided and was
overlapped by Middle to Upper Triassic sediments.
The shallow- to deep-marine deposits of the Upper
Triassic Auld Lang Syne Group (Heck and Speed,
1987; Elison and Speed, 1988), both west of and,
depositionally, on the Golconda allochthon, have
been correlated with the continental margin and
nonmarine alluvial plain red beds of the Chinle
Formation on the craton to the east (Lupe and
Silberling, 1985). This sedimentation and basinal
development took place to the east-northeast of an
evolving continental margin volcanic arc (Schweick-
ert, 1978; Dickinson, 1981). The fluvial and lacus-
trine Chinle Formation contains a large volume of
volcanic material likely derived from this develop-
ing arc terrane (Bilodeau, 1986; Stewart and others,
1986). This continental-crust- floored backarc basin
continued to develop into the Early Jurassic as the
volcanic arc evolved (Dickinson, 1981; Bilodeau,
1986). Areas of extensional tectonics and rift basin
development have been suggested for the backarc of
this east-dipping subduction zone (Oldow and Bar-
tel, 1987). In the Early Jurassic, following red bed
deposition, a large eolian sand sea or erg developed
behind the arc (Kocurek and Dott, 1983). This erg
was slowly flooded by shallow marine waters from
the north, producing a backarc basin with fluc-
tuating shorelines through the remainder of the
Jurassic.
JURASSIC
As the continental backarc basin in the western
Great Basin area continued to evolve into the Late
Jurassic, oceanic-crust-floored backarc or intra-arc
rift basins developed behind or within an offshore
island arc. This basin type is exemplified by the
Late Jurassic Galice Formation, which overlies the
Josephine ophiolite in the western Klamath Moun-
tains in northern California (Harper and others,
1985; Pinto-Auso and Harper, 1985; Wyld and
Wright, 1988). The Galice Formation is composed
primarily of deep-marine pelagic sediments capped
by a thick volcaniclastic flysch sequence, all resting
on pillow lavas of the Josephine ophiolite (Pinto-
Auso and Harper, 1985). The accretion of this off-
shore island arc and collapse of the Galice/Josephine
basin (with the subsequent deformation of the
basinal sediments) occurred as part of the Nevadan
orogeny (Harper and others, 1985).
The Nevadan orogeny in the western Sierra-
Klamath ophiolite belt closely resembles the Late
Jurassic terrane accretion event in British Colum-
bia where the much larger Stikinia composite-
terrane collided with North America in the Middle
Jurassic (Harper and others, 1985; Howell and
others, 1985). The Bowser basin, the type successor
Mesozoic basin evolution in western North America 243
A. EARLY TRIASSI C
LITHOSPHERE
Figure 2. Schematic west to east cross-
sections across the central part of the
western Cordillera at various times during
the Mesozoic.
...... ---
// ------
- - - ASTHENQSpHERE -
./
./
/
B. MICI-CRE'rA
C. LATEST CRETACEOUS - PALEOGENE
CORDI LLERAN
THRUST BElT
LARAMIDE
BROKEN-FORELAND BASINS
-----,
"
,
"-
,
- LITHOSPHERE
- ASTHENOSPHERe
---
--.. ............
basin, is situated on the Stikine terrane and con-
tains 4-5 km of basinal sediment (Eisbacher, 1981).
The lower half of the basinal section is Middle to
Late Jurassic in age and contains clasts eroded from
the adjacent Cache Creek terrane only, which thus
signifies a terrane amalgamation event prior to
collision of the super-terrane with North America.
Higher in the section, uppermost Jurassic and
Lower Cretaceous strata had a source from the Omi-
neca Crystalline Belt, part of the western margin of
North America at that time (Eisbacher, 1981), in-
dicating that accretion of the composite terrane had
occurred.
Following the Jurassic terrane accretions, east-
dipping subduction and a continental margin mag-
matic arc became established along the western
margin of North America (Dickinson, 1981). The
backarc region of the southern Cordillera was
complicated by the opening of the Gulf of Mexico
and NW-propagating rifting from the Gulf, right up
to the back of the volcanic arc in southern Arizona
(Dickinson, 1981; Bilodeau, 1982; Dickinson and
others, 1986). Late Jurassic rifting and sedimenta-
tion in the Chihuahua Trough, an inland extension
of the Gulf of Mexico depression, continued north-
westward into southern Arizona to produce the
Bisbee basin (Bilodeau, 1982; Dickinson and others,
1986) and possibly the McCoy basin, a rift basin in
SW Arizona/SE California (Harding and Coney,
1985; Stone and others, 1987). The Bisbee basin is
filled with over 3000 m of Bisbee Group strata, of
which the basal syntectonic Glance Conglomerate is
244 W. L. Bilodeau
Late Jurassic to Early Cretaceous (Bilodeau and
Lindberg, 1983; Bilodeau and others, 1987). The
McCoy Mountains Formation contains :57500 m of
nonmarine clastic sedimentary rocks and was depo-
sited in an elongated, WNW-trending rift basin
(Harding and Coney, 1985). How these strata act-
ually correlate with the Bisbee Group rocks to the
southeast is difficult to show, as the McCoy Moun-
tains Formation is hard to date and is still a
controversial rock unit (Stone and others, 1987).
In the forearc region of the developing Late
Jurassic-Cretaceous continental margin magmatic
arc, sedimentation was taking place in a large fore-
arc basin (Figure 2B), largely preserved today as
the Great Valley Sequence of northern and central
California (Ingersoll, 1983). This forearc basin
developed along the western edge of the Sierra
Nevada magmatic arc, initially as a residual forearc
basin formed on top of young oceanic crust, then as a
composite forearc on both oceanic and continental
crust (Ingersoll, 1982). Deposition and subsidence
within the forearc basin responded to changes in
plate interactions along the margin, especially
during the Late Cretaceous-Tertiary (Moxon and
Graham, 1987). Remnants of this forearc basin may
exist farther to the north as the Hornbrook and
Ochoco basins of central Oregon (Nilsen, 1984,
1986). The Great Valley forearc basin, though ini-
tially formed in the Late Jurassic, reached its
maximum extent during the Cretaceous and Early
Tertiary (Ingersoll, 1982).
CRETACEOUS
By the mid-Cretaceous, active westward ad-
vancement of the North American plate caused
formation of a backarc thrustbelt (Cordilleran
thrustbelt; Figure 2B) with westward under-
thrusting of the North American craton beneath the
supracrustal miogeoclinal sediments and accreted
terranes just east of the magmatic arc (Lowell, 1977;
Scholten, 1982). Crustal thickening within and
beneath the Cordilleran thrust belt structurally
loaded and down-flexed the western edge of the cra-
ton, forming a large, asymmetric retroarc foreland-
basin to the east (Jordan, 1981). This retroarc fore-
land-basin (Figure 2B) is commonly termed the
Western Interior Cretaceous basin (Alberta basin in
Canada), having at times extended from the Arctic
to the Gulf of Mexico, and contains a thick strati-
graphic section which illustrates the various con-
trols that tectonics, eustasy, and climate have on
transgressions and regressions in shallow marine
basins (Kauffman, 1982; Weimer, 1986).
Following a time of increased plate convergence
and the possible arrival of an aseismic ridge with
anomalously buoyant crust at the subduction zone
along the western margin of North America, the dip
of the subducting oceanic plate lessened during the
Late Cretaceous (Henderson and others, 1984;
Engebretson and others, 1985). This change in
plate-margin tectonics caused the locus of deforma-
tion to spread eastward from the Cordilleran thrust
belt into the area occupied by the retroarc foreland-
basin, an area of thick continental crust covered by
a relatively thin sedimentary veneer (Figure 2C).
The structural response was of thick-skinned, cry-
stalline basement-involved thrusting forming the
Late Cretaceous to Eocene Laramide Rocky Moun-
tains and associated broken-foreland basins (Cross,
1986; Bilodeau, 1987; Dickinson and others, 1988).
These broken-foreland basins are typically asymme-
trical in cross-section, contain thick nonmarine
clastic sections, and subside due to basement thrust-
loading along one margin (Hagen and others, 1985).
As the Cretaceous waned and the Laramide
orogeny was developing to the east, a new batch of
exotic terranes was arriving at the subduction zone
along the western margin of North America. The
Late Cretaceous arrival of the Wrangellia terrane
signaled the beginning of a complex phase of plate
interactions along the western margin of the con-
tinent (Howell and others, 1987). This phase was
dominated by strike-slip tectonics. The Methow and
Tyaughton basins of southern British Columbia and
Washington may have begun as successor basins
similar to the Bowser basin, but they became modi-
fied by wrench tectonics in the Late Cretaceous
(Trexler and Bourgeois, 1984). The Nanaimo basin,
in the Puget Sound area, was formed in the Late
Cretaceous by strike-slip tectonics in a forearc
setting, similar to the Methow and Tyaughton
basins (Johnson, 1984; Pacht, 1984). These basins
are filled with both marine and nonmarine clastic
rocks that have both east and west source terranes.
Some of these basins were structurally dismem-
bered by subsequent wrench tectonics, and even-
tually parts of them became widely dispersed along
the western margin of North America (Howell and
others, 1987).
Mesozoic basin evolution in western North America 245
REFERENCES
BILODEAU, W, L" 1982, Tectonic models for Early Cre-
taceous rifting in southeastern Arizona: Geology, v, 10, p,
466-470,
__ , 1986, The Mesozoic Mogollon Highlands, Arizona:
An Early Cretaceous rift shoulder: Journal of Geology, v,
94, p, 724-735,
__ , 1987, Laramide tectonics of the central Rocky
Mountain Foreland province: Colorado School of Mines
Quarterly, v, 82, no, 4, p, 48-64,
__ , and F, A, LINDBERG, 1983, Early Cretaceous tec-
tonics and sedimentation in southern Arizona, south-
western New Mexico, and northern Sonora, Mexico; pp,
173-188 in M, W, Reynolds and E. D, Dolly (eds,),
Mesozoic Paleogeography of the West-Central United
States: Society of Economic Paleontologists and Minera-
logists, Rocky Mountain Paleogeography Symposium 2, 391
p,
__ , C, F, KLUTH, and L, K. VEDDER, 1987, Regional
stratigraphic, sedimentologic and tectonic relationships of
the Glance Conglomerate in southeastern Arizona; pp,
229-256 in W, R Dickinson and M, A, Klute (eds,), Meso-
zoic Rocks of Southern Arizona and Adjacent Areas:
Arizona Geological Society Digest, v, 18,394 p,
BRUECKNER, H, K., and W, S, SNYDER, 1985, Structure of
the Havallah sequence, Golconda allochthon, Nevada:
Evidence for prolonged evolution in an accretionary prism:
Geological Society of America Bulletin, v, 96, p, 1113-1130,
CROSS, T, A" 1986, Tectonic controls of foreland basin
subsidence and Laramide style deformation, western
United States; pp, 15-39 in p, A, Allen and p, Homewood
(eds,), Foreland Basins: International Association of
Sedimentologists, Special Publication 8, 442 p,
DICKINSON, W, R, 1976a, Sedimentary basins developed
during evolution of Mesozoic-Cenozoic arc-trench system
in western North America: Canadian Journal of Earth
Sciences, v, 13, p, 1268-1287,
__ , 1976b, Plate tectonic evolution of sedimentary
basins; pp, 1-62 in W, R Dickinson and H, Yarborough
(eds,), Plate Tectonics and Hydrocarbon Accumula-
tions: American Association of Petroleum Geologists,
Continuing Education Course Note Series 1, 195 p,
__ 1981, Plate tectonic evolution of the southern
Cordillera; pp, 113-135 in W, R Dickinson and W, D,
Payne (eds,), Relations of Tectonics to Ore Deposits in
the Southern CordiUera: Arizona Geological Society Di-
gest, v, 14, 288 p,
__ , M, A, KLUTE, and p, N, SWIFT, 1986, The Bisbee
basin and its bearing on late Mesozoic paleogeographic and
paleotectonic relations between the Cordilleran and Carib-
bean regions; pp, 51-62 in p, L, Abbott (ed,), Cretaceous
Stratigraphy of Western North America: Society of
Economic Paleontologists and Mineralogists (Pacific
Section), v, 46, 233 p,
__ , M, A, KLUTE, M, J, HAYES, S, U, JANECKE, E, R
LUNDIN, M, A, McKITTRICK, and M, D, OLIVARES, 1988,
Paleogeographic and paleotectonic setting of Laramide
sedimentary basins in the central Rocky Mountain region:
Geological Society of America, Bulletin, v, 100, p, 1023-
1039,
EISBACHER, G, H., 1981, Late Mesozoic-Paleogene Bowser
basin molasse and Cordilleran tectonics, western Canada;
pp, 125-151 in A, D, Miall (ed,), Sedimentation and Tec-
tonics in AUuvial Basins: Geological Association of
Canada, Special Paper 23,272 p,
ELISON, M,W" and R C, SPEED, 1988, Triassic flysch of
the Fencemaker allochthon,East Range, Nevada: Fan fa-
cies and provenance: Geological Society of America Bul-
letin, v, 100, p, 185-199,
ENGEBRETSON, D, C" A, COX, and G, R GORDON (eds,),
1985, Relative Motions Between Oceanic and Contin-
ental Plates in the Pacific Basin: Geological Society of
America, Special Paper 206,59 p,
HAGEN, E. S" M, W, SHUSTER, and K. p, FURLONG,,1985,
Tectonic loading and subsidence of intermontane basins:
Wyoming foreland province: Geology, v, 13, p, 585-588,
HARDING, L, F" and p, J, CONEY, 1985, The geology of the
McCoy Mountains Formation, southeastern California and
southwestern Arizona: Geological Society of America Bul-
letin, v, 96, p, 755-769,
HARPER, G, D" J, B. SALEEBY, and E. A, S, NORMAN,
1985, Geometry and tectonic setting of sea-floor spreading
for the Josephine ophiolite, and implications for Jurassic
accretionary events along the California margin; pp, 239-
257 in D, G, Howell (ed.), Tectonostratigraphic Ter-
ranes of the Circum-Pacific Region: Circum-Pacific
Council for Energy and Mineral Resources, Earth Science
Series 1, 581 p,
HECK, F, R, and R C, SPEED, 1987, Triassic olistostrome
and shelf-basin transition in the western Great Basin:
Paleogeographic implications: Geological Society of Am-
erica Bulletin, v, 99, p, 539-551.
HENDERSON, L, J" R G, GORDON, and D, C, ENGEB-
RETSON, 1984, Mesozoic aseismic ridges on the Farallon
plate, and the Farallon plate and southward migration of
shallow subduction during the Laramide orogeny: Tec-
tonics, v, 3, p, 121-132,
HOWELL, D, G" D, L. JONES, and E. R SCHERMER, 1985,
Tectonostratigraphic terranes of the circum-Pacific re-
gion; pp, 3-30 in D, G, Howell (ed,), Tectonostratigraphic
Terranes of the Circum-Pacific Region: Circum-Pacific
Council for Energy and Mineral Resources, Earth Science
Series 1, 581 p,
246 W. L. Bilodeau
__ , G. W. MOORE, and T. J. WILEY, 1987, Tectonics and
basin evolution of western North America - An overview;
pp. 1-15 in D. W. Scholl, A. Grantz, and J. G. Vedder (eds.),
Geology and Resource Potential of the Continental
Margin of Western North America and Adjacent
Ocean Basins - Beaufort Sea to Baja California:
Circum-Pacific Council for Energy and Mineral Resources,
Earth Science Series 6, 799 p.
INGERSOLL, R V., 1982, Initiation and evolution of the
Great Valley forearc basin of northern and central Cali-
fornia, U.S.A.; pp. 459-467 in J. K. Leggett (ed.), Trench-
Forearc Geology: Geological Society of London, Special
Publication 10,576 p.
__ , 1983, Petrofacies and provenance of late Mesozoic
forearc basin, northern and central California: American
Association of Petroleum Geologists Bulletin, v. 67, p. 1125-
1142.
__ , 1988, Tectonics of sedimentary basins: Geological
Society of America Bulletin, v. 100, p. 1704-1719.
JOHNSON, S. Y., 1984, Evidence for a margin-truncating
transcurrent fault (pre-late Eocene) in western Washing-
ton: Geology, v. 12, p. 538-541.
JORDEN, T. E., 1981, Thrust loads and foreland basin
evolution, Cretaceous, western United States: American
Association of Petroleum Geologists Bulletin, v. 65, p. 2506-
2520.
KAUFFMAN, E. G., 1982, Paleobiogeography and evolu-
tionary response dynamic in the Cretaceous Western
Interior seaway of North America; pp. 273-306 in G. E. G.
Westermann (ed.), Jurassic-Cretaceous Biochronology
and Paleogeography of North America: Geologic Asso-
ciation of Canada, Special Paper 27,315 p.
KINGSTON, D. R, C. P. DISHROON, and P. A. WILLIAMS,
1983, Global basin classification system: American Asso-
ciation of Petroleum Geologists Bulletin, v. 67, no. 12, p.
2175-2193.
KOCUREK, G., and R H. DOTT, Jr., 1983, Jurassic paleo-
geography and paleoclimate of the central and southern
Rocky Mountain region; pp. 101-116 in M. W. Reynolds
and E. D. Dolly (eds.), Mesozoic Paleogeography of the
West-Central United States: Society of Economic Paleon-
tologists and Mineralogists, Rocky Mountain Paleogeo-
graphy Symposium 2, 391 p.
LOWELL, J. D., 1977, Underthrusting origin for thrust-fold
belts with application to the Idaho-Wyoming belt: Wyom-
ing Geological Association, 29th Annual Field Conference
Guidebook, p. 449-455.
LUPE, R, and SILBERLING, N. J., 1985, Genetic relation-
ship between lower Mesozoic continental strata of the
Colorado Plateau and marine strata of the western Great
Basin: Significance for accretionary history of Cordilleran
lithotectonic terranes; pp. 263-271 in D. G. Howell (ed.),
Tectonostratigraphic Terranes of the Circum-Pacific
Region: Circum-Pacific Council for Energy and Mineral
Resources, Earth Science Series 1, 581 p.
MILLER, E. L., B. K. HOLDSWORTH, W. B. WHITEFORD,
and D. RODGERS, 1984, Stratigraphy and structure of the
Schoonover sequence, northeastern Nevada: Implications
for Paleozoic plate-margin tectonics: Geological Society of
America Bulletin, v. 95, p. 1063-1076.
___ , L. R. KANTER, D. K. LARUE, R. J. TURNER, B.
MURCHEY, and D. L. JONES, 1982, Structural fabric of the
Paleozoic Golconda allochthon, Antler Peak Quadrangle,
Nevada: Progressive deformation of an oceanic sedimen-
tary assemblage: Journal of Geophysical Research, v. 87,
p.3795-3804.
MOXON, I. W., and S. A. GRAHAM, 1987, History and
controls of subsidence in the Late Cretaceous-Tertiary
Great Valley forearc basin, California: Geology, v. 15, p.
626-629.
NILSEN, T. H., 1984, Stratigraphy, sedimentology, and
tectonic framework of the Upper Cretaceous Hornbrook
Formation, Oregon and California; pp. 51-88 in T. H. Nil-
sen (ed.), Geology of the Upper Cretaceous Hornbrook
Formation, Oregon and California: Society of Economic
Paleontologists and Mineralogists (Pacific Section), v. 42,
257p.
__ , 1986, Cretaceous paleogeography of western North
America; pp. 1-39 in P. L. Abbott (ed.), Cretaceous Stra-
tigraphy of Western North America: Society of Econo-
mic Paleontologists and Mineralogists (Pacific Section), v.
46,233p.
OLDOW, J. S., and R L. BARTEL, 1987, Early to Middle(?)
Jurassic extensional tectonism in the western Great Basin:
Growth faulting and synorogenic deposition of the Dunlap
Formation: Geology, v.15, p. 740-743.
PACHT, J. A., 1984, Petrologic evolution and paleogeo-
graphy of the Late Cretaceous Nanaimo basin, Wash-
ington and British Columbia: Implications for Cretaceous
tectonics: Geological Society of America Bulletin, v. 95, p.
766-788.
PINTO-AUSO, M., and G. D. HARPER, 1985, Sedimenta-
tion, metalogenesis and tectonic origin of the basal Galice
Formation overlying the Josephine ophiolite, northwest-
ern California: Journal of Geology, v. 93, p. 713-725.
READING, H. G., 1982, Sedimentary basins and global
tectonics: Proceedings of the Geologists Association, v. 93,
no. 4, p. 321-350.
SCHOLTEN, R, 1982, Continental subduction in the north-
ern Rockies - A model for back-arc thrusting in the west-
ern Cordillera; pp. 123-136 in R B. Powers (ed.), Geologic
Studies of the Cordilleran Thrust Belt, Volume 1:
Rocky Mountain Association of Geologists Guidebook, 474
p.
Mesozoic basin evolution in western North America 247
SCHWEICKERT, R. A., 1978, Triassic and Jurassic paleo-
geography of the Sierra Nevada and adjacent regions,
California and western Nevada; pp. 361-384 in D. G. Ho-
well and K. A. McDougall (eds.), Mesozoic Paleogeo-
graphy of Western United States: Society of Economic
Paleontologists and Mineralogists, Pacific Coast Paleogeo-
graphy Symposium 2, 573 p.
SPEED, R. C., 1979, Collided Paleozoic microplate in the
western United States: Journal of Geology, v. 87, p. 279-
292.
__ , 1982, Evolution of the sialic margin in the central
western United States; pp. 457-468 in J. S. Watkins and C.
L. Drake (eds.), Continental Margin Geology: American
Association of Petroleum Geologists, Memoir 34,805 p.
STEWART, J. H., T. H. ANDERSON, G. B. HAXEL, L. T.
SILVER, and J. E. WRIGHT, 1986, Late Triassic paleo-
geography of the southern Cordillera: The problem of a
source for voluminous volcanic detritus in the Chinle
Formation of the Colorado Plateau region: Geology, v. 14,
p.567-570.
STONE, P., V. M. PAGE, W. HAMILTON, and K. A.
HOWARD, 1987, Cretaceous age of the upper part of the
McCoy Mountains Formation, southeastern California and
southwestern Arizona, and its tectonic significance: Re-
conciliation of paleobotanical and paleomagnetic evidence:
Geology, v. 15, p. 561-564.
TREXLER, J. H., Jr., and J. BOURGEOIS, 1984, Evidence for
mid-Cretaceous wrench-faulting in the Methow basin,
Washington: Tectonostratigraphic setting of the Virginian
Ridge Formation: Tectonics, v. 4, p. 379-394.
WEIMER, R. J., 1986, Relationship of unconformities,
tectonics, and sea level changes in the Cretaceous of the
western interior, United States; pp. 397-422 in J. A. Peter-
son (ed.), Paleotectonics and Sedimentation in the
Rocky Mountains Region, United States: American As-
sociation of Petroleum Geologists, Memoir 41,693 p.
WYLD, S. J., and J. E. WRIGHT, 1988, The Devils Elbow
ophiolite remnant and overlying Galice Formation: New
constraints on the Middle to Late Jurassic evolution of the
Klamath Mountains, California: Geological Society of
America Bulletin, v. 100, p. 29-44.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Thick-skinned overstep tectonics in the Jurassic Winnemucca
fold-and-thrust belt, north-central Nevada, USA:
Evidence from the Sonoma Range
STEPHEN D. STAHL
Department of Geology, Central Michigan University, Mt. Pleasant, Michigan 48859, USA
(received September 12, 1988; revision accepted May 11, 1989)
ABSTRACT
The Jurassic Winnemucca fold-and-thrust belt is characterized by westward transport of lower Paleozoic
to Upper Triassic rocks over upper Triassic shelf rocks. In the Sonoma Range, autochthonous Upper Triassic
rocks are exposed in two windows through the Jurassic allochthon. Fabric elements from the Triassic rocks
indicate that the Winnemucca thrust system developed in strict overstep fashion (structurally higher thrusts
are younger), whereas most fold-and-thrust belts are interpreted as piggy-back systems (structurally lower
thrusts are younger). Being a strict overstep system, the Winnemucca thrust system is apparently without
analogue. It is likely that the most significant factor leading to overstep tectonics in the thrust belt is
basement involvement.
Twenty km west of the Winnemucca fold-and-thrust belt is the apparently coeval Fencemaker belt. The
Fencemaker belt is characterized by eastward transport of Mesozoic basinal strata over Upper Triassic shelf
rocks. The Early Jurassic age of the Winnemucca-Fencemaker event is significant because south and west of
these belts deformation commenced in the Middle Jurassic. This temporal relationship requires com-
partmentalized shortening, perhaps accommodated in the Early Jurassic by a NW-striking, right-lateral,
strike-slip fault.
INTRODUCTION
The Winnemucca fold-and-thrust belt of Speed
and others (1982) trends roughly NNE for 150 km in
north-central Nevada (Figure 1). Its characteristics
are westward overturned folds of upper Triassic
shelf strata overridden by lower Paleozoic rocks of
the Mississippian-emplaced Roberts Mountains
allochthon (Roberts rocks), Permo-Pennsylvanian
Antler overlap sequence (Silberling and Roberts,
1962), and upper Paleozoic rocks of the Early
Triassic-emplaced Golconda allochthon (Golconda
rocks) (Figure 2). The exposed thrust of the belt has
been given a variety of names, most of them model-
dependent: the term "Willow Creek thrust" is used
herein because it is the most generally defined,
location-specific term and therefore carries the least
pedagogic baggage. The age of faulting is con-
strained to the Early to early Middle Jurassic by the
youngest rocks on the lower plate, the latest Triassic
Winnemucca Formation (Nichols and Silberling,
1977), and a 165 3 Ma post-faulting granite
(Silberling, 1975). The width of the Winnemucca
fold-and-thrust belt is unknown because of the pau-
city of exposures of Triassic and lower Jurassic rocks
east of the Sonoma Range (Figure 2).
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 249-261. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
249
250
S. D. Stahl
Figure 1. Generalized geologic map of Mesozoic
structural elements in Nevada. Geology: Wftb,
Winnemucca fold-and-thrust belt; Fftb, Fencemaker
fold-and-thrust belt; initial Sr 87/Sr86= 0.7060 and
= 0.7040 shown by solid lines; enclaves of lesser
Mesozoic deformation indicated by dashed lines;
Triassic shelf rocks shown by brickwork pattern,
Triassic to Early Jurassic basinal rocks shown by
dashed pattern; Jackson arc indicated; Sierra Neva-
da batholith indicated by random dash pattern;
major thrusts indicated by thick lines with saw teeth
on hanging wall.
West of the Winnemucca fold-and-thrust belt is
the Fencemaker fold-and-thrust belt. The Fence-
maker belt extends along the early Mesozoic shelf-
basin terrane boundary (Speed, 1978) for =200 km.
The Fencemaker is characterized by E-verging folds
of upper Triassic shelf strata overthrust by Late
Triassic to Early Jurassic basinal deposits (Elison
and Speed, 1988; Heck and Speed, 1987) (Figure 2).
Although the root thrusts of these belts are only
= 20 km apart at their closest approach in the East
Range (Figure 2), no unequivocal evidence exists
for their relative age. They may be nearly coeval,
however, because they share an autochthon. That
is, the youngest rocks on the footwalls to both the
Fencemaker and Willow Creek thrusts are the
latest Triassic Winnemucca Formation. In spite of
their opposite vergence, the Fencemaker and
Winnemucca fold-and-thrust belts are most likely
manifestations of the same tectonic regime because:
(l)oftheir probable synchronicity, and (2)they both
record E-W shortening.
0.7040
CA
I
{
I
I
I
I
I
I
I
I
I UT

<I
I
I
I
I
1
I
\. /1
I
" \
I
r
I
I
_ .."
'" \
" \ AZ
........ 1

Although the type area of the Willow Creek
thrust is in the East Range, field relations charac-
teristic of the Winnemucca fold-and-thrust belt are
best exposed in the Sonoma Range (Figure 3). Prior
to this study, the Sonoma Range was mapped by
Ferguson, Muller, and Roberts (1951), Gilluly
(1967), and Silberling (1975). The most significant
points upon which their interpretations differed
were (1) the depth to which rocks were allocht-
honous in the Mesozoic; and (2) the degree of
Jurassic shuffiing of tectonostratigraphic packages.
I have proposed that the Triassic shelf rocks are
autochthonous and that there is a single Jurassic
thrust exposed in the Sonoma Range (Stahl,
1987a,b, 1989); these conclusions are summarized in
Figures 3 and 4. Presented herein are the results
of detailed structural analysis of the Triassic rocks
exposed in the Sonoma Range and their bearing on
the mode of emplacement of the Willow Creek
allochthon.
Thick-skinned overstep tectonics in the Sonoma Range 251
J
I
1
,
,
I
%\
'0,
'1:.\
'- CJ.',
cS" ..... c..Q ...
<91'
Spring.
Figure 2. Simplified geologic map of
north-central Nevada showing Paleozoic
to middle Mesozoic tectonostratigraphic
packages and major faults. Box indicates
approximate area of Figure 3. Stratigra-
phy and symbols are indicated in the
legend; W, Winnemucca, Q, Quinn River
Crossing.
MAP UNITS
[lIT] Czi intrusive
tilll Jib basinal
is shelf
CONTACTS
______ Jurassic thruat
Golconda thrust
Roberts Mountains
thrust system
o PMg Golconda allochthon
o PlPo overlap sequence
____ other thrust
____ depositional contact
o DCh Harmony Formation
8ill DCro other Roberts Mountains allochthon
DCs shelf
STRUCTURAL DATA
Stratigraphic Units
With the exception of the Tallman Fanglo-
merate, which is the oldest formation, the strati-
graphic units examined in this work have been
extensively described in the literature (for example:
Nicholas and Silberling, 1977; Silberling and
Wallace, 1969; Speed, 1978) and only a cursory
overview of the basic lithologies is presented herein.
The Tallman Fanglomerate, unique to the Sonoma
Range, is a sedimentary conglomerate that was sub-
are ally deposited (Ferguson and others, 1951). It is
composed of clasts of Antler overlap sequence and
Golconda rocks; no definitive Roberts rock-clasts are
present (Stahl, 1987b). It lies unconformably above
Golconda rocks (Stahl, 1987b), and its upper contact
is gradational with the Koipato rhyolite (Gilluly,
1967). The Koipato rhyolite is tuffaceous and ano-
malously thin (75 m) in the Sonoma Range. Its con-
tact with the overlying Natchez Pass Formation is
structural.
The Natchez Pass Formation is very thickly
bedded carbonate. It is sharply and conformably
overlain by the Grass Valley Formation, a sequence
of sandstone and mudstone interbeds. The contact
between the Grass Valley and Dun Glen Formations
252
S. D. Stahl
o
.
Figure 3. Generalized geologic map of the Sonoma Range;
symbols as on Figure 2; A-F, referred to in text.
is not exposed in the Sonoma Range. The Dun Glen
Formation is composed of very thick to massive beds
of carbonate. The Dun Glen-Winnemucca contact is
sharply gradational. Interbeds of mudstone, silt-
stone, and sandstone, as well as minor amounts of
limestone, make up the Winnemucca Formation.
Structural Setting
The Triassic rocks are exposed in two windows
through the Willow Creek allochthon (Figure 3)
that are connected at depth (Stahl, 1987a,b). The
Willow Creek thrust cuts up-section from the east
side of the central window (within the Natchez Pass
,--
I
--l
PMg
OCh
--______ A
Figure 4. Schematic tectonostratigraphic column of the
Sonoma Range; symbols as before except thick line with
hatchures indicates low-angle normal fault (hatchures are
on hanging wall).
Formation, "A" on Figure 3) to the range-front win-
dow (top of the Winnemucca Formation, "C" on Fig-
ure 3). In addition, the Willow Creek thrust cuts
up-section from the north end of the range-front
window (within the Tallman Fanglomerate, "E" on
Figure 3) to the southern end of the range-front
window (top of the Winnemucca Formation, "C" on
Figure 3). Farther south, only Roberts rocks, Ant-
ler overlap sequence, and Golconda rocks within the
Willow Creek allochthon are exposed. The western
margin of the central window and eastern margin of
the range-front window are littered with thrust
slices of Triassic shelf rocks.
Major folds are present in both windows and are
truncated by the Willow Creek thrust. Outcrop pat-
terns in the central window define a major inclined,
W-overturned antiform. The predominant structur-
al feature of the Triassic rocks in the range-front
window is a major tight, recumbent, W-verging
synformal fold of the Dun Glen and Winnemucca
Formations. The Grass Valley and Natchez Pass
Formations exposed at the north end of the window
("F" on Figure 3) lie above the axial surface of the
synform but in the hinge region of a related re-
cumbent antiform.
Folding that post-dates emplacement of the
Willow Creek allochthon is evident from map rela-
tions. The axial trace of the major fold in the range-
front window and the trace of the Willow Creek
thrust trends north in the southern half of the
Thick-skinned overstep tectonics in the Sonoma Range 253
range-front window ("C" on Figure 3) but trend
WNW in the northern half of the window ("F" on
Figure 3). In the central window, the axial trace of
the antiform trends north until the very northern
reaches of the window where it trends northwest.
The N-S shortening of this post-emplacement fold-
ing is inconsistent with the E-W shortening of the
Willow Creek-Fencemaker event; it may be part of a
progressive deformation with orientation controlled
by irregularities in the continental margin; alter-
natively, it may represent a distinct younger de-
formation.
The major structural elements provide the rela-
tive time frame by which the minor fabric elements
(presented below) were analysed. The rocks were
mapped at the 1 :8000 scale. Relative ages of minor
folds were determined by field relations when
possible. However, the paucity, or localized develop-
ment, of planar fabric elements during pre- and syn-
Willow Creek allochthon-emplacement deforma-
tions hinders the development of a completely un-
ambiguous relative age scale for deformation events
in the Triassic rocks. Only detailed domainal ana-
lysis provides the control necessary to allow de-
ciphering of the deformation history.
Data from the range-front window are presen-
ted in two sections because of the reorientation of
the major synform and Willow Creek thrust by post-
emplacement folding. The entire Triassic section is
exposed in the northern section of the window and
the structural trace trends west of north. The
southern section of the window includes the Win-
nemucca and Dun Glen Formations with N-trending
structural trace. The window is contiguous.
Relations in the Northern Section of the Range-
Front Window
The Tallman Fanglomerate, at the base of the
Triassic shelf package, is in depositional contact
with the underlying Paleozoic rocks. The structural
fabric of the Tallman Fanglomerate is dominated by
minor bedding-parallel faults with well developed
slickensides and striae. At least one fault was
folded with similar geometry to the major synform.
The Koipato rhyolite is foliated parallel to bedding,
with penetrativeness of foliation increasing toward
its fault contact with the overlying Natchez Pass
Formation. The foliation is folded in the fault zone;
the asymmetry of these folds indicates south-
westward overriding.
The contact between the Koipato rhyolite and
the Natchez Pass Formation is a low-angle normal
fault because upright younger rocks are superposed
over upright older rocks. This geometry is incon-
sistent with thrust faulting and suggests instead
thinning of the stratigraphic section. Two lines of
evidence support the interpretation of this fault as a
low-angle normal fault. First, nowhere else in
north-central Nevada is the Natchez Pass Forma-
tion directly above the Koipato rhyolite; this
stratigraphic position is normally occupied by the
Prida Formation. Second, near the base of the
Natchez Pass Formation in the range-front window
is a marker layer of chert pebble conglomerate that
is indicative of the lower part of the upper Natchez
Pass Formation (Silberling and Wallace, 1969).
Therefore, the lower and middle units of the
Natchez Pass Formation are also absent. Removal
of the section must have been tectonic because,
although the rates of deposition of the Triassic shelf
strata were not uniform and there were periods of
nondeposition (Nichols and Silberling, 1977), the
thickness of the Prida Formation increases north-
ward toward the Sonoma Range (Silberling and
Wallace, 1969).
Two lines of evidence indicate that the low-
angle normal fault pre-dates major folding. First, a
synthetic reverse fault rooted in the low-angle
normal fault can be traced upward through the
section to within the Grass Valley Formation ("A"
on Figure 5A). The splay cannot be traced into the
Dun Glen Formation; while this mayor may not be
due to the cessation of exposure at the range-
bounding, high-angle, Cenozoic normal fault, the
fact that the Dun Glen Formation is not reoriented
by the fault indicates that faulting occurred prior to
Dun Glen time. Second, map relationships show
that the low-angle normal fault was folded by the
recumbent antiform (Figure 5B).
Minor folds (F 1) related to faulting are common
in the Natchez Pass and Grass Valley Formations.
In the Natchez Pass Formation, Fl folds are in-
clined, open, asymmetric, and SW-verging folds that
die away rapidly from the fault (Figure 5B). Grass
Valley Formation F 1 folds are recumbent and tight
to isoclinal with a well-developed foliation (Sl)
subparallel to bedding; Sl is folded by the major fold.
Domainal analysis indicates that the orientation of
Sl is not parallel on opposite sides of the reverse
fault synthetic to the Koipato-Natchez Pass fault
(Stahl, 1987b); this suggests that F 1 is related to
low-angle normal faulting.
254
8. D. 8tahl
Figure 5A. Geologic map of the northern part of the
range-front window. Stratigraphy indicated on legend;
contour interval is 400 feet; diagonal lines indicate edge of
the Willow Creek allochton; A, referred to in text.
Four phases of minor folding (F 2 to F 5) not F
related to low-angle normal faulting but having
8W-vergence are present in the shelf rocks in this
domain (Figure 6). F2 folds are tight to isoclinal,
W-verging, recumbent folds of Winnemucca Forma-
tion sandstones and mudstones. These folds are not
abundant and some have a poorly developed axial
planar foliation (8
2
) in the hinge region. They were
160
folded by the major recumbent synform (Figure
5B). The relative age of F2 and Fl is based on the
interpretations that (1) F 1 is related to low-angle
normal faulting, and (2) low-angle normal faulting
occurred prior to Dun Glen deposition. The Winne-
mucca Formation overlies the Dun Glen Formation,
and therefore any deformation that affects the
Winnemucca Formation must be younger than the
low-angle normal faulting that pre-dated Dun Glen
depositon.
F3 folds are recumbent to shallowly inclined
folds within the Natchez Pass Formation near the
Willow Creek thrust, as well as inclined, close-folds
of beds and 8
1
of the Grass Valley Formation. The
age of F3 relative to major folding cannot be fully
ascertained. However, F3 folds are correlated to the
major folding because of similar in style, orienta-
tion, and vergence.
F 4 folds are present only in the Winnemucca
Formation; they are close, W-verging, and inclined
with a very well developed foliation (8
5
) in and
remote to the hinge region. 8
5
is present in the
Natchez Pass and Grass Valley Formations. No
cross-cutting relationships between 8
4
and 8
5
were
observed; F 5 is interpreted as younger than F 4
because 8
4
, and not 8
5
, shows a cryptic distribution
about a vertical axis (the 8
4
-8
5
intersection) (Fig-
ure 6). An analog to F5 is present in the Jurassic-
allochthonous Roberts rocks and the same relation-
ship to F 4 is observed (8tahl, 1989).
Fs folds represent a distinct phase of deforma-
tion and are close, asymmetric, and steeply inclined
to vertical. They are present in the Natchez Pass,
Grass Valley, and Winnemucca Formations. N-8
closure is indicated by minor folds and supported by
the definition of an EIW-striking vertical plane by
K'
900
1800
1700
1500
Figure 5B. Geologic cross sections along KK' and LL' in Figure 5A. Elevation in meters, no vertical exaggeration.
Thick-skinned overstep tectonics in the 8onoma Range 255
fold axes. The spread of axes is not due to subse-
quent folding but reflects instead the intersection of
a vertical plane (8
6
) and diverse orientations of beds
as a result of the five earlier phases of folding (F 1-
F5)
Relations in the Southern Section of the Range-
Front Window
The southern section of the range-front window
comprises exposures of Dun Glen and Winnemucca
Formations (Figure 7). The traces of the Willow
Creek thrust and major synform trend north-south.
The hinge of the synform is faulted at the southern
end of the window, and displacement decreases
northward to the vanishing point over 2 km. A
thrust slice of probable Dun Glen Formation of
unknown facing is present in the hinge at the south-
ernmost end of the window. The Dun Glen-Winne-
mucca contact is faulted similarly to the hinge of the
major synform.
Five generations of minor folds (F
2
to F
6
) are
recognized in this domain (Figure 8). Rare F2 tight
to isoclinal folds in the Winnemucca Formation are
proximal to minor thrusts (Figure 7B). Minor
thrust-related F2 folds (open, W-verging, die out
away from the fault) are present in the Dun Glen
Formation as well. All minor thrusts and related
folds were folded by the major synform (Figure 7B).
Recumbent, tight, W-verging F3 folds are
present in both the Winnemucca and Dun Glen
Formations but are not abundant. F3 folds are
interpreted as parasites of the major synform. No
foliation was formed in the major synform (a non-
penetrative axial planar foliation, 8
3
, was recog-
nized in three minor F3 folds) in spite of the
tightness of the fold (25
0
interlimb angle), thickness
of section folded (975 m), and abundance of mud-
stone (Winnemucca Formation) in the hinge.
In marked contrast to the major and minor F3
folds, F 4 folds (1) constitute minor folds of the Win-
nemucca Formation; and (2) record less shortening
because they have greater apical angles (close) and
shorter short limbs than F3 folds; but (3) have well
developed, penetrative axial planar foliation (8
4
),
F 4 folds are W -verging and are common in the upper
limb of the major synform but rare in the lower
limb.
F5 folds are open and upright with well deve-
loped axial planar foliation (8
5
) in and away from
the hinge region. F5 abundance is not dependent
Figure 6. Structural fabric
elements from the northern
section of the range-front win-
dow. Nets are lower hemi-
sphere equal area projections.
A, Fl in Natchez Pass For-
mation; B, Fl in Grass Valley
Formation; C, F
z
in the Win-
nemucca Formation; D, F z in
the Dun Glen Formation;
E, F
3
; F, F
4
; G, F
5
; H, F
6
.
o told aXIs
o pole \0 ax,al plana
:) pole to tollallon
OpOlalob,seCllnllpl8rl8
256
S. D. Stahl
o 400
-==>
melalS
Figure 7A. Geologic map of the southern section of the
range-front window.
Figure 7B. Geologic cross sections of EE', FF' , and GG'
in Figure 7 A.
F
,600
''0

1400
llOO
upon structural horizon and has been recognized in
the Jurassic autochthonous Paleozoic rocks as well
(Stahl, 1989). F 6 folds are close, steeply inclined to
vertical folds of the Winnemucca Formation. They
are consistent with N -S shortening and are interpre-
ted as parasites of the post-emplacement major fold.
Relations in the Central Window
Exposures of the Natchez Pass and Grass
Valley Formations in the central window define a
W-inclined antiform, correlated to F
3
, that is trun-
cated by the Willow Creek thrust (Figure 9) . The
Willow Creek thrust is steeply E-dipping on the
eastern side of the window and shallowly W-dipping
on the western margin. Although the central win-
dow does not display direct evidence for normal
faulting, folds interpreted as Flare present in both
the Grass Valley and Natchez Pass formations
(Figure 10). Domainal analysis shows that Grass
Valley Formation Sl is folded by the major F3
antiform (Stahl, 1987b). F 6 folds are present in both
formations; they are most abundant at the north end
of the window where the trace of the antiform trends
west of north.
E'
700
1900
lBOO
1100
Thick-skinned overstep tectonics in the Sonoma Range 257
A N 6
O::J
0
C 0
C
DO
E N N
0 0
0
oiJ
Figure 8. Structural fabric elements from the southern
section of the range-front windows; symbols as before.
A, F 2 in the Winnemucca Formation; B, F 2 in the Dun
Glen Formation; C, F 3; D, F 4; E, F 5; F, F 6'
The most spectacular minor folds in the central
window are synemplacement folds in the eastern-
most exposures of the Natchez Pass Formation (see
Figure 9B). These W-verging folds have Willow
Creek thrust-parallel axial surfaces and define a
deformation gradient in which apical angle de-
creases and short limb length increases toward the
overriding Paleozoic rocks. Because the upper limb
of these folds is determined by the attitude of the
Willow Creek thrust whereas the lower limb paral-
lels the antiform, the major folds must pre-date
emplacement of the Willow Creek allochthon.

. ..
e:..
Figure 9A. Geologic map of the central window.
Deformation History
The six generations of folding are correlated to
deformation events in Table 1. Low-angle normal
faulting and related folding is the earliest event
(D1). F2-F5 are grouped together in D2 and are
interpreted as a progressive, coaxial deformation
which was the direct result of the Winnemucca-
Fencemaker event. F 6 is the post-emplacement D3
event. Although it is possible that D3 was a con-
tinuation of D
2
, it is interpreted as a distinct event
because (1) it folds F
3
, S4, S5, and the Willow Creek
258
S. D. Stahl
Figure 9B. eologic cross sections along 00' and PP' in
Figure9A.
Q meters
Figure 10. Structural fabric elements from the central
window; symbols as before. A, F 1 in the Grass Valley
Formation; B, F3 in the Natchez Pass Formation; C, F
5
;
D,F6
thrust, and (2) its N -S shortening direction is rough-
ly orthogonal to the E-W shortening direction of F 2-
F5
The style of deformation of the Triassic rocks
may provide evidence for a younger upper boundary
to the Winnemucca-Fencemaker event than that
0'
imposed by the 168 3 Ma post-deformation granite
(SHberling, 1975). In particular, it is significant
that the major F3 folding did not produce an axial
planar foliation whereas younger minor folds (F 4,
F
5
) that record less severe shortening have a well-
developed axial planar foliation. This may suggest
that the Winnemucca Formation was poorly consoli-
dated at the time of F3 folding. In other words, F3
folding and Willow Creek thrusting more likely
occurred early in the Early Jurassic rather than
early in the Middle Jurassic. Tectonic dewatering
perhaps hindered foliation development during F
3
;
such dewatering would have facilitated foliation
development in F 4 and F
5
. The similar transport
directions of Dl and D2 and the potentially narrow
time gap between Dl (late Grass Valley) and D2
(late to post-Winnemucca?) suggest that Dl and D2
may have had the same driving tectonic regime in a
regional progressive deformation.
The major Fa folding is interpreted as the result
of blind thrusting. The evidence for this interpre-
tation is (1) the presence of a "blind" thrust in the
southern part of the range-front window; (2) the
length of the short limb of the major synform de-
creases toward the north as the offset on the "blind"
thrust decreases; and (3) the major folds are trun-
cated by the Willow Creek thrust and therefore pre-
date Willow Creek allochthon emplacement. F 4 and
F 5 are recognized in the Paleozoic rocks of the
Willow Creek allochthon (Brueckner and Snyder,
1985; Stahl, 1989) and therefore post-date Willow
Creek thrusting.
Thick-skinned overstep tectonics in the Sonoma Range 259
Table 1. Deformation history of the Triassic rocks in the Sonoma Range.
Folding
Event Description
Deformation
Event
Recumbent isoclinal folds in Grass Valley Formation
Inclined to upright folds in Natchez Pass Formation
Low-angle normal faulting
F 2 Recumbent tight to isoclinal folds in Winnemucca Formation
Shallowly inclined tight folds in Natchez Pass Formation
Inclined close folds in Grass Valley Formation
Major folds of all Triassic rocks; emplacement of the Willow Creek allochthon
F 4 Steeply inclined close folds of Winnemucca formation
F 5 Upright open to close folds of Winnemucca formation
Steeply inclined to vertical minor and major folds of all Triassic rocks
DISCUSSION
Two general end-member models for the deve-
lopment of thrust systems are (1) piggy-back (Dahl-
strom, 1970), and (2) overstep (Elliot and Johnson,
1980). In the piggy-back model, thrusting propa-
gates in the direction of fault displacement. Older
thrusts are incorporated into the allochthon and
transported above a sole thrust. In the overstep
model, thrusting propagates opposite the direction
of fault displacement. Older thrusts are accreted to
their autochthon and overridden by younger
thrusts. Most foreland fold-and-thrust belts are
interpreted as piggy-back systems; however, evalua-
tion of structural data from Triassic rocks in the
Sonoma Range leads to the conclusion that the Win-
nemucca thrust system developed in strict overstep
fashion.
An assumption must be made to determine the
age of the Willow Creek thrust in the Sonoma Range
relative to structurally higher thrusts based on fab-
ric elements of the autochthonous Triassic strata. It
is assumed that the relative proximity of thrusting
is reflected by the attitude of axial planes of a fold
generation. Axial surfaces of a single fold genera-
tion formed initially by buckling but were subjected
to progressive deformation in thrust belts and re-
oriented by inhomogeneous simple shear. The
original attitudes of axial surfaces are (roughly)
perpendicular to maximum tectonic compression:
because the maximum shortening direction is
assumed to be approximately horizontal at the time
of fold formation, the axial planes are approxi-
mately vertical. Axial surface orientations become
shallowly dipping near the thrust because they are
rotated toward parallelism with the thrust (simple
shear plane) by inhomogeneous simple shear.
Therefore, folds distal to the fault are steeply
inclined to upright and folds proximal to the fault
are shallowly inclined to recumbent.
The key condition imposed by field relations in
the Sonoma Range is that the Willow Creek thrust
is interpreted as the root thrust of the Winnemucca
fold-and-thrust belt (Stahl, 1987b, 1989). If the
Winnemucca thrust system developed in overstep
fashion, the root thrust is the oldest. Early folds
would then be recumbent and later stage folds
would progress through inclined to upright. If the
Winnemucca fold-and-thrust belt developed as a
piggy-back system, then the root fault is the young-
est of the belt. In this case, early folds would be
upright and subsequent generations would become
more inclined with the youngest folds being re-
cumbent. The observed trends in the Triassic rocks
exposed in the Sonoma Range match those predicted
for an overstep thrust system.
One potential weakness in the above argument
is the interpretation that the root thrust of the
Winnemucca belt is exposed in the Sonoma Range.
Although recent work (Heck and Speed, 1987;
Elison and Speed, 1988) indicates that this may be
unlikely, the possibility exists of an unexposed,
structurally lower thrust beneath the Willow Creek
autochthon. If this were the case, then fold orien-
260 S. D. Stahl
tation trends within the Triassic strata still indicate
that the thrusting which caused F 4 is younger than
the Willow Creek thrust. However, it does not
indicate whether the younger thrusting was struc-
turally higher (overstep) or lower (piggy-back).
Evidence indicates that the F 4-related thrust-
ing was structurally higher than the Willow Creek
thrust. F 4 folds define a deformation gradient in
which abundance offolding increases upward. Post-
Willow Creek emplacement F 4 folds are absent in
the Jurassic autochthonous Paleozoic rocks (Stahl,
1987b, 1989), absent in the Grass Valley Formation,
rare in the Winnemucca Formation exposed in the
lower limb of the major synform, abundant in the
Winnemucca Formation exposed in the upper limb
of the range-front synform, and prevalent in the
Jurassic allochthonous Paleozoic rocks (Brueckner
and Snyder, 1985; Stahl, 1987b, 1989). Whether or
not the Willow Creek thrust is the root thrust of the
Winnemucca fold-and-thrust belt, this is the defini-
tion of an overstep thrust system.
A second weakness in determining the nature of
Winnemucca thrust system development is the ab-
sence of a recognized Jurassic thrust above the
Willow Creek thrust. East of the Sonoma Range are
"enclaves of lesser Mesozoic deformation" (Speed
and others, 1988). These "enclaves" coincide with
areas virtually devoid of exposure of early Mesozoic
rocks. Much of the work done in the "enclaves" has
been stratigraphic in nature; detailed structural
analysis of these Paleozoic rocks might produce
evidence of Mesozoic fabric overprints leading to the
discovery of heretofore unmapped Mesozoic thrusts.
CONCLUSIONS AND REGIONAL IMPLICATIONS
The overstep Winnemucca fold-and-thrust belt
is, to my knowledge, without analog. Therefore, its
style of development could shed much light on the
tectonic history of the central Nevada foreland
during the Jurassic. The specific dynamics of em-
placement have yet to be worked out, but the belt
does have some eccentricities when compared to
other fold-and-thrust belts. These may offer clues to
the cause of the overstep development of the Winne-
mucca thrust system.
First, the belt is probably thick skinned because
the stratigraphic throw of the fault is on the same
order of magnitude as the horizontal displacement.
The entire Cambrian to Upper Triassic strati-
graphic section (7.5-16 km thick) comprises the
Willow Creek allochthon (Stahl, 1987b), whereas
the distance between the east margin of the central
window (where the Willow Creek ramps to within
the Triassic rocks) to the closest exposure of the
Fencemaker thrust is, not accounting for Cenozoic
extension, 23 km. A reasonable estimate of hori-
zontal displacement on the Willow Creek thrust is
therefore 10-15 km. This constrains the average dip
of the fault between 30 and 60.
Second, the sense of overriding in the belt, east
to west, is opposite the west to east overriding
dominant in the Cordilleran orogen to the east (for
example, the Sevier belt) and south (for example,
the Luning belt) of the Sonoma Range. This oppo-
site vergence is most clearly shown by the adjacent
and (apparently) coeval Fencemaker fold-and-thrust
belt to the west.
Third, the Sonoma Range lies just east of the
Mesozoic shelf-basinal terrane boundary. Recent
work (Elison and Speed, 1988; Speed and others,
1988) indicates that this boundary remained static
through the Early Jurassic thrusting events. The
Sonoma Range also lies above transitional crust and
west of the Precambrian crystalline North Ameri-
can basement (Dilles and Wright, 1988). The rocks
in the Winnemucca fold-and-thrust belt are thus
sandwiched by basement heterogeneities. Because
points one and three above indicate basement con-
trol of deformation, it would appear that the thick-
skinned tectonics of the belt are factors in this
overstep development.
It is significant that deformation in the
Winnemucca-Fencemaker belt is Early Jurassic.
This deformation is older than the wholly E-directed
Middle Jurassic deformation of the early Mesozoic
shelf terrane south of the Winnemucca fold-and-
thrust belt (for example, Dilles and Wright, 1988)
and Middle Jurassic deformation of Mesozoic arc
terrane rocks west of the Sonoma Range in the
Jackson Mountains (Russell, 1984; Maher and
Saleeby, 1986). If the interpretation - based on the
absence of S3 foliation in the Winnemucca Forma-
tion - that the Winnemucca-Fencemaker event
was early Early Jurassic, then this increases the
time lag between the Winnemucca-Fencemaker
event and Middle Jurassic deformations and in-
creases the likelihood that the different deforma-
tions record distinct events.
The age difference in deformation between the
Early Jurassic Winnemucca-Fencemaker event in
north-central Nevada and Middle Jurassic deforma-
tion to the south necessitates at least one of the
Thick-skinned overstep tectonics in the Sonoma Range 261
following to accommodate shortening: (1) two
"Fencemaker" faults - a northern (near) syn-
Willow Creek thrust and a southern of Middle
Jurassic age; (2) a Jurassic W/NW-striking strike-
slip fault between the Fencemaker and Luning
thrusts (see Figure 1); or (3) a prolonged wavelike
progressive deformation involving both southern
and northern regions spanning the Early and
Middle Jurassic. Of these, (1) has been proposed by
Speed and others (1988) and (3) seems unlikely. The
strike-slip faults required by (2) would be difficult to
recognize in the field; however, a fault of similar
orientation and offset, albeit younger than that
required by this model, was recognized elsewhere in
Nevada (Oldow, 1983).
REFERENCES
BRUECKNER, H. K., and W. S. SNYDER, 1985, Structure of
the Havallah sequence, Golconda allochthon, Nevada:
Evidence for prolonged evolution in an accretionary prism:
Geological Society of America Bulletin, v. 96, p. 1113-1130.
DAHLSTROM, C. D. A., 1970, Structural geology in the
eastern margin of the Canadian Rocky Mountains: Bulle-
tin of Canadian Petroleum Geology, v. 18, p. 332-406.
DILLES, J. H., and J. E. WRIGHT, 1988, The chronology of
early Mesozoic arc magmatism in the Yerington district of
western Nevada and its regional implications: Geological
Society of America Bulletin, v. 100, p. 644-652.
ELLIOT, D., and M. R. W. JOHNSON, 1980, Structural
evolution in the northern part of the Moine thrust belt,
NW Scotland: Transactions of the Royal Society of Edin-
brugh: Earth Sciences, v. 71, p. 69-96.
ELISON, M. W., and R. C. SPEED, 1988, Triassic flysch of
the Fencemaker allochthon, East Range, Nevada: Fan
facies and provenance: Geological Society of America Bul-
letin, v. 100, p. 185-199.
FERGUSON, H., S. MULLER, and R. ROBERTS, 1951, Geo-
logic Map of the Winnemucca Quadrangle, Nevada:
U.S. Geological Survey, Geologic Quadrangle Map GQ-ll,
scale 1:125000.
GILLULY, J., 1967, Geologic Map of the Winnemucca
Quadrangle, Pershing and Humboldt Counties, Neva-
da: U.S. Geologic Survey, Geologic Map GQ-656, scale
1:62500.
HECK, F. R., and R. C. SPEED, 1987, Triassic olistostrome
and shelf-basin transition in the western Great Basin:
Paleogeographic implications: Geological Society of Am-
erica Bulletin, v. 99, p. 539-551.
MAHER, K. A., and J. SALEEBY, 1986, Geology of the
Jackson Mountains, NW Nevada [abstract]: Geological
Society of America Abstracts with Programs, v. 18, p. 679.
NICHOLS, K. M., and N. J. SILBERLING, 1977, Strati-
graphy and Depositional History of the Star Peak
Group (Triassic), Northwestern Nevada: Geological
Society of America, Special Paper 178, 73 p.
OLDOW, J. S., 1983, Tectonic implications of a late Meso-
zoic fold and thrust belt in northwestern Nevada: Geology,
v. 11, p. 542-546.
RUSSELL, B. J., 1984, Mesozoic geology of the Jackson
Mountains, northwestern Nevada: Geological Society of
America Bulletin, v. 95, p. 313-323.
SILBERLING, N. J., 1975, Age Relations on the Golconda
Thrust Fault, Sonoma Range, Northwestern Nevada:
Geological Society of America, Special Paper 163, 28 p.
__ , and R. J. ROBERTS, 1962, Pre-Tertiary Strati-
graphy and Structure of Northwestern Nevada: Geo-
logical Society of America, Special Paper 72, 58 p.
__ , and R. E. WALLACE, 1969, Stratigraphy of the
Star Peak Group (Triassic) and Overlying Mesozoic
Rocks, Humboldt Range, Nevada: U.S. Geological Sur-
vey, Professional Paper 592, 50 p.
SPEED, R. C., 1978, Paleogeographic and plate tectonic
evolution of the early Mesozoic marine province of the
western Great Basin; pp. 253-270 in D. G. Howell and K. A.
McDougall (eds.), Mesozoic Paleogeography of the
Western United States: Society of Economic Paleonto-
logists and Mineralogists, Pacific Coast Paleogeography
Symposium 2, 27 p.
SPEED, R. C., M. W. ELISON, and F. HECK, 1988, Phanero-
zoic tectonic history of North America in the Great Basin;
pp. 572-606 in W. G. Ernst (ed.), Metamorphism and
Crustal Evolution of the Western United States (Ru-
bey Volume 7): Prentice-Hall, Englewood Cliffs, New
Jersey, USA, 1153 p.
SPEED, R. C., M. ELISON, J. ENGELN, S. STAHL, A. E.
V AVERK, and D. WIENS, 1982, Mesozoic Winnemucca fold
and thrust belt [abstract]: Geological Society of America
Abstracts with Programs (Cordilleran Section), v. 14, p.
236.
STAHL, S. D., 1987a, Mesozoic structural features in
Sonoma Canyon, Sonoma Range, Humboldt and Pershing
Counties, Nevada; pp. 85-90 in M. L. Hill (ed.), Cordi-
lleran Centennial Field Guide, Volume I: Geological
Society of America, Boulder, Colorado, USA, 490 p.
__ , 1987b, Pre-Cenozoic Structural Geology and
Tectonic History of the Sonoma Range, North-Central
Nevada: PhD dissertation, Northwestern University,
Evanston, Illinois, USA, 286 p.
__ , 1989, Recognition of Jurassic transport of rocks of
the Roberts Mountains allochthon: Evidence from the
Sonoma Range, north-central Nevada: Geology, v. 17, p.
645-648.
CHAPTER 3
PRECAMBRIAN MARGINS
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Generation of granitoids at Archaean continental margins
in southern Africa
D.R. HUNTER
Department of Geology, University of Natal, Pietermaritzburg, 3200, Natal, South Africa
(received August 12, 1988; revision accepted February 27, 1989)
ABSTRACT
Archaean terranes, worldwide, typically comprise remnants of low metamorphic-grade, supracrustal
rocks, dominated by mafic and ultramafic volcanics with lesser but variable volumes of sediments, preserved
in a sea of granitoids. Age relations of the oldest granitoid gneisses and supracrustal remnants are commonly
equivocal. Recent geochronological data from Swaziland provide support for the conclusion that certain
layered, tonalitic/leucotonalitic gneisses pre-date extrusion of mafic and ultramafic lavas of the Barberton
(formerly Swaziland) Sequence. The purpose of this paper is to demonstrate that subdivision of the granitoids
into three so-called cycles is an oversimplification. Evidence is accumulating that sodic and potassic granitoid
magmas were emplaced repeatedly over a period of at least 1000 Ma. The supracrustal remnants preserved in
the granitoid sea reflect complex thrusting and reclined folding. Synsedimentary tectonism resulted in
accumulation of clastic sediments in fault-bounded depositional basins. On a regional scale, this deformation
resulted in progressive northward uplift of the crust to the south of the Barberton supracrustal remnant. The
Barberton Sequence was, in part, obducted onto sialic crust lying to the northwest as a result of horizontal
compression and lateral shortening (up to 50%). New tonalitic magmas were generated and older
tonalitic/leucotonalitic crust deformed and metamorphosed. Preservation of clasts composed of potassic
granite in the upper sedimentary member of the Barberton Sequence implies that potassic as well as sodic
granitic magmas were generated at = 3480 Ma. Synchronous with this later deformation was intrusion of
tabular leucotonalitic batholiths south of Barberton, generated by partial melting of tectonically intersliced
metavolcanic remnants and older sialic crust. Thick sialic crust resulted from a combination of tectonic
interslicing and the emplacement of the tabular potassic Lochiel granite at = 3000 Ma as a consequence of
partial melting of old sialic crust. The Lochiel granite batholith is confined to the southeastern flank of
Barberton supracrustal remnant. It is proposed that this granite was generated at a continental margin
following closure of the ocean in which the Barberton Sequence was generated. Deformation of the Pongola
Sequence and subsequent intrusion of post-Pongola granitoids were the final events of Archaean age in
Swaziland and adjacent areas to the south.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 265-28l. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
265
266 D. R. Hunter
INTRODUCTION
Archaean terranes worldwide comprise rem-
nants of supracrustal sequences, dominated by
mafic and ultramafic volcanic rocks with lesser but
variable volumes of sedimentary rocks, preserved in
seas of granitoids. The supracrustal rocks are
typically metamorphosed at greenschist facies, in
contrast to the upper amphibolite facies of adjacent
grey, tonalitic to trondhjemitic layered gneisses.
Age relations of the oldest granitoid gneisses and
the supracrustal remnants are equivocal, contacts
being commonly obscured by high strains or by later
intrusive granitoids.
The southeastern Kaapvaal structural province
that comprises Swaziland and the adjacent parts of
the Republic of South Africa is no exception to this
generalization. However, recent geochronological
data (Compston and Kroner, 1988) have demon-
strated that at least part of the Ancient Gneiss
Complex (A.G.C.) in Swaziland pre-dates the supra-
crustal Barberton Sequence.
The purpose of this paper is to summarize the
geological evolution during the Archaean of Swazi-
land and adjacent areas in the Transvaal province of
South Africa south of Barberton (Figure 1). Fol-
lowing the first classification of Archaean grani-
toids in Swaziland (Hunter, 1957), several modifica-
tions have been proposed, the most recent being a
simple subdivision into three granitic cycles (An-
haeusser and Robb, 1981). This paper shows that in
Swaziland and adjacent areas, felsic magmatism
was almost continuous throughout that period of
time from 3600 to 3000 Ma (Tables 1 and 2).
SUPRACRUSTAL SEQUENCES
The most extensive supracrustal remnant is
that preserved in the vicinity of Barberton, but a
number of smaller relics are preserved for 250 km
south of Barberton (Figures 1 and 2).
Mafic and ultramafic lavas constituting the
lowermost (Onverwacht) group of the Barberton
Sequence were extruded in marine environments
and are distinguished by their lack of interbeds
composed of terrigenous detritus. Originally, this
volcanic pile was inferred to represent a strati-
graphic thickness of 15 km (Viljoen and Viljoen,
1970). This was questioned following reinterpre-
tation of stratigraphic relations that are considered
to reflect duplication by imbricate thrusting (de Wit
and Stern, 1980; de Wit, 1982; de Wit and others,
1983). Other authors (e.g., Dokka and Lowe, 1984;
Lowe and others, 1985) have recognized some dup-
lication which they have regarded as being confined
to the uppermost Onverwacht. The upper forma-
tions of the Onverwacht Group in the southern area
of the Barberton remnant are characterized by a
variety of pyroclastic, volcaniclastic, and chemical
sediments. Sedimentary structures in these rocks
have been interpreted to reflect accumulation in
subaerial, intertidal, and subtidal environments
(Lowe and Knauth, 1977; Lanier and Lowe, 1982;
Lowe, 1982), but Stanistreet and others (1981) have
argued for sedimentation at water depths of at least
2km.
The clastic sediments overlying the Onver-
wacht Group were deposited, at least in the south,
contemporaneously with major deformational
events involving initial N-S-shortening and subse-
quent SEINW-shortening (Lamb and Paris, 1988).
The sedimentary sequences are now preserved
within thrust sheets up to a kilometre thick, the
thrusting being regarded as a late phase of the
synsedimentary deformation (Lamb, 1987). Lamb
(1987) has suggested that the actively deforming,
convergent plate boundary zone in the North Island
of New Zealand could be a possible analogue for the
tectonic environment during deposition of these
sediments.
The Dwalile remnant (Figure 2) comprises talc
and tremolite schists, amphibolites, serpentinites,
and hornblende schists. Subordinate intercalations
Figure 1. Simplified geological map of Swaziland and surrounding areas of Transvaal and northern Natal. Greenstone
remnants are unshaded but amphibolite grade enclaves in gneisses are shown in black. Geology of the granitic terrane west
and north of the Barberton Sequence is modified after Anhauesser and Robb (1981). Gneissic granitoids are shown by
broken lines that parallel the gneissic fabric. Tabular granitoid intrusions are shown with vertical or inclined crosses, post-
tectonic plutons of various ages are shown with horizontal ruling, vertical and horizontal ruling, or random dashed lines
with dots.
Granitoids at Archaean continental margins in southern Africa
B Borberton
Mo Mankoyone
Mb Mbobone
M Monzini
N Nelspruit
P Pief Retief
....,_ Sheor zone
- -Foult
t:.:.:;;.:j Pongola Sequence a
Usushwona Complex
GNEISSIC
LEUCOTONALITE
8 TRONDHJEMITE
o 10 20 ;30 40 !)O
I I I
km
UNMAPPED
GRANITOIDS
o
N
rr>
267
T
a
b
l
e

1
.

S
u
m
m
a
r
y

o
f
g
e
o
l
o
g
i
c
a
l

a
n
d

s
t
r
u
c
t
u
r
a
l

e
v
e
n
t
s
,

S
w
a
z
i
l
a
n
d

a
n
d

n
o
r
t
h
e
r
n

N
a
t
a
l
.

C
o
m
m
o
n
d
a
l
e

A
s
s
e
g
a
a
i

S
w
a
z
i
l
a
n
d

(
S
m
i
t
h

1
9
8
7
)

(
T
a
l
b
o
t

a
n
d

o
t
h
e
r
s
!

1
9
8
7
)

(
J
a
c
k
s
o
n
!

1
9
8
7
)

G
e
o
l
o
g
i
c
a
l

S
t
r
u
c
t
u
r
a
l

G
e
o
l
o
g
i
c
a
l

S
t
r
u
c
t
u
r
a
l

G
e
o
l
o
g
i
c
a
l

S
t
r
u
c
t
u
r
a
l

B
i
m
o
d
a
l

l
a
y
e
r
e
d

D
1
:

G
e
n
e
r
a
t
i
o
n

o
f

B
i
m
o
d
a
l

l
a
y
e
r
e
d

D
1

(
e
a
r
l
y
)
:

S
u
b
-
B
i
m
o
d
a
l

l
a
y
e
r
e
d

D
1
:

F
l
a
t
-
l
y
i
n
g

r
e
-
g
n
e
i
s
s
e
s

r
e
c
u
m
b
e
n
t

W
-
v
e
r
g
i
n
g

g
n
e
i
s
s
e
s

h
o
r
i
z
o
n
t
a
l

a
n
d

l
i
s
t
r
i
c

g
n
e
i
s
s
e
s

c
u
m
b
e
n
t

F

1
f
o
l
d
s

C
o
m
m
o
n
d
a
l
e

F

1
f
o
l
d
s

A
s
s
e
g
a
a
i

s
u
p
r
a
c
r
u
s
t
a
l

z
o
n
e
s

o
f
c
a
t
a
c
l
a
s
i
s

D
w
a
l
i
l
e

s
u
p
r
a
c
r
u
s
t
a
l

s
u
p
r
a
c
r
u
s
t
a
l

s
e
-
T
h
r
u
s
t
i
n
g

p
a
r
a
l
l
e
l

t
o

s
e
q
u
e
n
c
e

D
1

(
l
a
t
e
)
:

T
r
a
n
s
i
t
i
o
n

s
u
i
t
e

q
u
e
n
c
e
.

F

1
a
x
i
a
l

s
u
r
f
a
c
e
s

t
o

d
u
c
t
i
l
e

f
l
o
w
;

M
a
t
s
h
e
m
p
o
n
d
o

g
e
n
e
r
a
t
i
o
n

o
f

P
e
r
i
d
o
t
i
t
e

S
u
i
t
e

r
e
c
u
m
b
e
n
t

f
o
l
d
s

I
n
t
r
u
s
i
o
n

o
f

B
r
a
u
n
-
I
n
t
r
u
s
i
o
n

o
f

T
s
a
w
e
l
a

s
h
w
e
i
g

t
o
n
a
l
i
t
e

t
o
n
a
l
i
t
e

M
p
o
p
o
n
o

a
n
o
r
t
h
o
s
i
t
e

s
u
i
t
e

I
n
t
r
u
s
i
o
n

o
f
B
a
z
a
n
e

I
n
t
r
u
s
i
o
n

o
f

B
a
z
a
n
e

b
a
t
h
o
l
i
t
h
.

b
a
t
h
o
l
i
t
h

M
a
f
i
c

d
y
k
e
s

D
2
:

I
n
t
e
n
s
e

F

2

f
o
l
d
-
D
2
:

C
o
n
c
e
n
t
r
i
c

F

2

D
2
:

I
n
t
e
n
s
e

F

2
f
o
l
d
i
n
g

i
n
g

N
E
-
t
r
e
n
d
i
n
g

a
x
i
a
l

p
a
r
a
l
l
e
l

f
o
l
d
s

s
t
e
e
p
l
y

o
n

f
l
a
t
-
l
y
i
n
g

p
l
a
n
e
s

o
f

s
u
r
f
a
c
e
s
;

p
e
n
e
t
r
a
t
i
v
e

d
i
p
p
i
n
g

N
N
E
-
b
u
l
k

s
t
r
a
i
n

S
2

r
e
f
o
l
i
a
t
i
o
n
;

S
l

t
r
e
n
d
i
n
g

a
x
i
a
l

D
3
:

L
o
c
a
l

d
u
c
t
i
l
e

t
r
a
n
s
p
o
s
e
d
.

D
u
c
t
i
l
e

s
u
r
f
a
c
e
s

s
h
e
a
r

z
o
n
e
s

b
r
i
t
t
l
e

s
h
e
a
r
s

p
a
r
a
l
l
e
l

0
3
:

M
i
n
o
r

d
i
s
-
D
4
:

T
i
g
h
t

t
o

i
s
o
c
l
i
n
a
l

t
o

F
2

a
x
i
a
l

s
u
r
f
a
c
e
s

h
a
r
m
o
n
i
c

c
o
n
j
u
g
a
t
e

F
4

v
e
r
t
i
c
a
l

t
o

g
e
n
t
l
y

0
3
:

N
E
-
t
r
e
n
d
i
n
g

f
o
l
d
s
;

N
E
-
t
r
e
n
d
i
n
g

d
i
p
p
i
n
g

N
E
-
t
r
e
n
d
i
n
g

d
u
c
t
i
l
e

a
n
d

b
r
i
t
t
l
e

s
t
r
i
k
e
-
s
l
i
p

f
a
u
l
t
s

i
n

a
x
i
a
l

s
u
r
f
a
c
e
s

s
i
n
i
s
t
r
a
l

s
h
e
a
r
s

i
n

m
e
t
a
s
e
d
i
m
e
n
t
s

M
a
f
i
c

d
y
k
e
s

D
5
:

G
e
n
t
l
y

d
i
p
p
i
n
g

C
o
m
m
o
n
d
a
l
e

s
u
p
r
a
-
d
u
c
t
i
l
e

s
h
e
a
r

z
o
n
e
s
,

c
r
u
s
t
a
l
s

d
i
s
p
l
a
c
e
m
e
n
t

t
o
w
a
r
d
s

0
4
:

O
p
e
n

t
o

t
i
g
h
t

N
W

f
o
l
d
i
n
g

a
b
o
u
t

s
t
e
e
p

N
W
-
t
r
e
n
d
i
n
g

a
x
i
a
l

s
u
r
f
a
c
e
s

M
i
n
o
r

i
n
t
r
u
s
i
o
n

o
f

I
n
t
r
u
s
i
o
n

o
f
L
o
c
h
i
e
l

I
n
t
r
u
s
i
o
n

o
f
L
o
c
h
i
e
l

D
s
:

F
o
l
d
i
n
g

a
n
d

m
y
-
g
r
a
n
i
t
e

a
n
d

q
u
a
r
t
z

b
a
t
h
o
l
i
t
h

b
a
t
h
o
l
i
t
h

l
o
n
i
t
i
c

r
e
f
o
l
i
a
t
i
o
n

o
f

m
o
n
z
o
n
i
t
e

D
2

n
e
a
r

t
h
e

b
a
s
e

o
f

L
o
c
h
i
e
l

t
a
b
u
l
a
r

g
r
a
n
i
t
e

P
o
n
g
o
l
a

S
e
q
u
e
n
c
e

D
5
:

S
i
n
i
s
t
r
a
l

P
o
n
g
o
l
a

S
e
q
u
e
n
c
e

0
4
:

N
W
-
t
r
e
n
d
i
n
g

s
h
e
a
r
i
n
g

a
n
d

f
o
l
d
i
n
g

m
y
l
o
-
n
i
t
e

z
o
n
e
s

a
b
o
u
t

s
t
e
e
p

N
W

a
x
i
a
l

s
u
r
f
a
c
e
s

-
_

.. -
-
-
-
-
D
e
f
o
r
m
a
t
i
o
n
a
l

e
v
e
n
t
s

a
r
e

n
u
m
b
e
r
e
d

s
e
q
u
e
n
t
i
a
l
l
y

f
r
o
m

t
h
e

f
i
r
s
t

r
e
c
o
g
n
i
z
a
b
l
e

e
v
e
n
t

i
n

t
h
e

s
u
p
r
a
c
r
u
s
t
a
l

s
e
q
u
e
n
c
e
s
.

D
e
f
o
r
m
a
t
i
o
n

a
f
f
e
c
t
i
n
g

t
h
e

g
n
e
i
s
s

t
e
r
r
a
n
e

p
r
e
-
d
a
t
i
n
g

t
h
e

a
c
c
u
m
u
l
a
t
i
o
n

o
f

t
h
e

s
u
p
r
a
c
r
u
s
t
a
l

s
e
q
u
e
n
c
e

i
s

n
o
t

p
r
e
c
l
u
d
e
d

b
y

t
h
i
s

s
c
h
e
m
e
.

"
"

0
)

(
X
l

t
;
:
I

?
J

:
:
r
:

c


M
-
t
t
l

.
,

A
G
E

(
M
a
)

2
5
0
0
-
2
8
0
0

2
8
4
8


3
1

2
8
7
6


3
0

2
9
4
0


2
2

3
0
2
8


1
4

3
1
4
9


1
2
5

3
1
6
0


1
0
0

3
2
6
8


3
8

3
2
8
7


4
6

-
3
3
5
0

-
3
2
0
0

3
3
3
7


1
3

-
3
4
0
0
-
3
3
5
0

3
4
8
0


3
0

-
3
4
2
0
-
3
4
8
0

3
4
5
0


2
6
4

?

3
6
4
4


4

-
3
6
3
0


3
9
5
0
?

M
E
T
H
O
D

R
b
-
S
r

R
b
-
S
r

S
m
-
N
d

U
-
P
b

R
b
-
S
r

R
b
-
S
r

U
-
P
b

R
b
-
S
r

R
b
-
S
r

U
-
P
b
/

R
b
-
S
r

S
m
-
N
d

U
-
P
b

3
9
A
r
A
O
A
r

U
-
P
b

S
m
-
N
d

U
-
P
b

T
a
b
l
e

2
.

S
u
m
m
a
r
y

o
f
g
e
o
c
h
r
o
n
o
l
o
g
i
c
a
l

d
a
t
a
,

S
w
a
z
i
l
a
n
d

a
n
d

e
a
s
t
e
r
n

T
r
a
n
s
v
a
a
l
.

R
E
M
A
R
K
S

P
o
s
t
-
P
o
n
g
o
l
a

g
r
a
n
i
t
o
i
d
s

(
e
r
r
o
c
h
r
o
n
s
)
.

B
o
e
s
m
a
n
s
k
o
p

s
y
e
n
o
g
r
a
n
i
t
i
c

p
l
u
t
o
n

(
w
h
o
l
e

r
o
c
k
)

-
e
m
p
l
a
c
e
m
e
n
t

a
g
e
?

U
s
u
s
h
w
a
n
a

m
a
f
i
c

s
u
i
t
e

(
w
h
o
l
e

r
o
c
k
)
.

P
o
n
g
o
l
a

S
e
q
u
e
n
c
e

(
z
i
r
c
o
n
s
,

N
s
u
z
e

G
r
o
u
p
)
.

L
o
c
h
i
e
l

t
a
b
u
l
a
r

b
a
t
h
o
l
i
t
h

(
w
h
o
l
e

r
o
c
k
,

R
o

=

0
.
7
0
1
3
)

-
-
l
a
t
e
-
t
e
c
t
o
n
i
c

D
a
l
m
e
i
n
-
t
y
p
e

p
l
u
t
o
n
s

N
e
l
s
p
r
u
i
t

b
a
t
h
o
l
i
t
h

(
w
h
o
l
e

r
o
c
k
,

R
o

=

0
.
7
0
1
4
)
.

B
a
z
a
n
e

t
a
b
u
l
a
r

b
a
t
h
o
l
i
t
h

(
w
h
o
l
e

r
o
c
k
,

R
o

=

0
,
7
0
0
1
)

-
s
y
n
t
e
c
t
o
n
i
c
.

M
i
d
d
l
e

M
a
r
k
e
r
,

O
n
v
e
r
w
a
c
h
t

G
r
o
u
p

(
w
h
o
l
e

r
o
c
k
,

R
o

=

0
.
7
0
1
6
)

m
e
t
a
m
o
r
p
h
i
s
m
?

K
a
a
p

V
a
l
l
e
y

p
l
u
t
o
n

(
3
2
2
9


5
;

U
-
P
b
)
;

N
e
l
s
h
o
o
g
t
e

p
l
u
t
o
n

(
3
2
2
0


8
0
;

U
-
P
b
)
;

S
t
e
n
t
o
r

p
l
u
t
o
n

(
3
3
4
7


6
0
;

U
-
P
b
)
;

G
r
a
n
o
d
i
o
r
i
t
e

S
u
i
t
e

(
3
3
5
0


5
7
;

R
b
-
S
r
,

w
h
o
l
e

r
o
c
k
)
.

M
a
t
s
h
e
m
p
o
n
d
o

p
e
r
i
d
o
t
i
t
e

s
u
i
t
e

(
w
h
o
l
e

r
o
c
k
)

D
e
p
o
s
i
t
i
o
n

o
f

F
i
g

T
r
e
e

a
n
d

M
o
o
d
i
e
s

G
r
o
u
p
s
;

d
e
t
r
i
t
a
l

z
i
r
c
o
n
s

i
n

F
i
g

T
r
e
e

G
r
o
u
p

h
a
v
e

u
n
i
f
o
r
m

U
-
P
b

a
g
e
s
,

y
i
e
l
d
i
n
g

C
o
n
c
o
r
d
i
a

i
n
t
e
r
c
e
p
t

a
g
e

o
f

3
4
5
3


9

M
a
.

K
o
m
a
t
i

F
o
r
m
a
t
i
o
n
;

l
o
w
-
g
r
a
d
e

m
e
t
a
m
o
r
p
h
i
s
m
?
;

s
e
a
-
f
l
o
o
r

a
l
t
e
r
a
t
i
o
n
?

D
e
p
o
s
i
t
i
o
n

o
f

O
n
v
e
r
w
a
c
h
t

G
r
o
u
p
,

p
r
o
b
a
b
l
y

c
o
n
t
e
m
p
o
r
a
n
e
o
u
s

w
i
t
h

d
e
f
o
r
m
a
t
i
o
n

a
n
d

m
e
t
a
-
m
o
r
p
h
i
s
m

i
n

D
w
a
l
i
l
e

s
u
i
t
e

a
n

A
.
G
.
C
.

A
g
e

y
i
e
l
d
e
d

b
y

m
e
t
a
m
o
r
p
h
i
c

g
a
r
n
e
t

a
n
d

m
e
t
a
q
u
a
r
t
z
i
t
e

i
n

D
w
a
l
i
l
e

s
u
i
t
e
;

a
g
e

i
s

i
n
d
i
s
t
i
n
g
u
i
s
h
a
b
l
e

w
i
t
h
i
n

e
r
r
o
r

f
r
o
m

a
g
e

o
f
d
e
t
r
i
t
a
l

z
i
r
c
o
n
s

i
n

m
e
t
a
q
u
a
r
t
z
i
t
e
;

a
g
e
s

o
f
z
i
r
c
o
n

r
i
m
s

a
r
o
u
n
d

o
l
d
e
r

z
i
r
c
o
n

c
o
r
e
s
,

a
n
d

o
f

d
i
s
c
r
e
t
e

z
i
r
c
o
n
s

y
i
e
l
d

C
o
n
c
o
r
d
i
a

i
n
t
e
r
c
e
p
t

o
f
3
4
8
0


2
0

i
n
d
i
c
a
t
i
v
e

o
f

h
o
m
o
g
e
n
e
o
u
s

s
o
u
r
c
e

(
t
o
n
a
l
i
t
e

i
n
t
r
u
s
i
o
n
)
;

z
i
r
c
o
n
s

i
n

p
o
t
a
s
s
i
c

g
r
a
n
i
t
e

p
e
b
b
l
e
s

i
n

M
o
o
d
i
e
s

G
r
o
u
p

c
o
n
g
l
o
m
e
r
a
t
e
s
.

D
e
p
o
s
i
t
i
o
n

o
f

D
w
a
l
i
l
e

s
u
i
t
e
.

S
i
n
g
l
e

z
i
r
c
o
n

i
n

l
e
u
c
o
t
o
n
a
l
i
t
i
c

l
a
y
e
r

o
f
B
i
m
o
d
a
l

S
u
i
t
e

(
A
.
G
.
C
.
)
.

S
m
-
N
d

i
s
o
t
o
p
i
c

d
a
t
a

s
u
g
g
e
s
t

p
o
s
s
i
b
l
e

p
r
e
s
e
n
c
e

o
f
v
e
r
y

o
l
d

c
r
u
s
t

n
o
t

r
e
c
o
g
n
i
z
e
d

i
n

d
a
t
e
d

z
i
r
c
o
n
s
.

S
e
e

t
e
x
t

f
o
r

d
a
t
a

s
o
u
r
c
e
s
.

I
n
t
r
u
s
i
o
n

o
f
p
o
t
a
s
s
i
c

g
r
a
n
i
t
e
s

M
a
f
i
c

p
l
u
t
o
n
i
s
m

S
e
d
i
m
e
n
t
a
t
i
o
n

a
n
d

v
o
l
c
a
n
i
s
m

I
n
t
r
u
s
i
o
n

o
f
p
o
t
a
s
s
i
c

g
r
a
n
i
t
i
c

m
a
g
m
a
s

D
e
f
o
r
m
a
t
i
o
n

a
n
d

m
e
t
a
m
o
r
p
h
i
s
m

I
n
t
r
u
s
i
o
n

o
f
t
o
n
a
l
i
t
e
s

i
n
t
o

B
a
r
b
e
r
t
o
n

S
e
q
u
e
n
c
e

S
e
d
i
m
e
n
t
a
t
i
o
n

V
o
l
c
a
n
i
s
m

M
e
t
a
m
o
r
p
h
i
s
m

D
e
f
o
r
m
a
t
i
o
n
,

i
n
t
r
u
s
i
o
n

o
f
t
o
n
a
l
i
t
i
c

a
n
d

p
o
t
a
s
s
i
c

g
r
a
n
i
t
i
c

m
a
g
m
a
s

V
o
l
c
a
n
i
s
m
/
s
e
d
i
m
e
n
t
a
t
i
o
n

T
o
n
a
l
i
t
i
c

m
a
g
m
a
t
i
s
m

o


S
.

M
-
o

s
:

0
0

I
I
I

M
-
>

>
o
o
j
g
-
I
I
I

C
D


(
'
)

o


M
-
S
'

C
D


g

S

I
I
I

a
a

S
'

0
0

S
'


C
D

g

>

:
:
;
>


I
>
:
)

0
>

'
"

270
D. R. Hunter
I
8AR8ERTON

0$7 0/
DWALILE

eft r PDI Rel Ief
rCOMMONDALE
D GreensTone remnanfS
Vryheld
t>

(}[!,
RI> 80bonongo

2700"
2800
01 STR 1 BUTION OF
GREENSTONE REMNANTS
Figure 2. Distribution of main greenstone remnants south
of Barberton. The southernmost greenstone remnant in
Figure 1 is that at Commondale at the southern limit of
Figure 1. Immediately south of the Commondale remnant,
Mesozoic cover blankets the Archaean basement. The
Nondweni remnant appears as an inlier in this cover.
of calc-silicate gneisses, metapelites, and meta-
quartzites are present. Dwalile rocks have litho-
logical similarities with the Onverwacht Group at
Barberton, but their lower-amphibolite facies of
metamorphism contrasts with the typical green-
schist facies of the Onverwacht Group. Still farther
south, the Assegaai remnant comprises mafic
volcanic rocks similar to those at Dwalile but clastic
and chemical sediments are prominent (Hunter and
others, 1983). No radiometric dating of either of
these remnants has been attempted, but a tabular
leucotonalite (Bazane granite) intrusive into the
Assegaai rocks has yielded a Rb-Sr isochron age of
3268 38 Ma (R. E. Harmer, oral commun., 1987).
The Commondale remnant contrasts with the
small remnant southeast of Piet Retief (Figure 2) in
that only minor metasedimentary and calc-silicate
interlayers are present in a pile of mafic and ultra-
mafic rocks. The Commondale remnant is preserved
in two synformal keels separated by a major shear
zone. A rhythmically alternating sequence of spin i-
fex-textured and cumulate layers comprises the
distinctive Matshempondo Peridotite Suite (Smith,
1987) that is preserved in an area of low strain in
the core of the southern synform. Textures and
structures diagnostic of repeated lava flows are
absent. Rapid cooling of repetitively emplaced
ultramafic magmas in a high-level subvolcanic
intrusion could account for the observed features,
but this cannot be confirmed as contacts are not
exposed. Alternatively, the Matshempondo Suite
could represent repeated filling (and perhaps
draining?) of a lava lake. Whatever its origin, the
Matshempondo Suite was deformed by the major
event that imposed the regional NE-trending
structural grain on the Archaean rocks. This rela-
tion of deformation is significant, as samples from
this suite define a Sm-Nd isochron with an age of
3337 13 Ma (A. H. Wilson, oral commun., 1986).
The southernmost metavolcanic remnant is
preserved in the vicinity of Nondweni (Figure 2).
The dominant lithologies are pillowed basalts inter-
layered with komatiitic basalts. Intercalations of
acid lavas, pyroclastics, and tuffaceous sedimentary
rocks are minor but provide evidence of deposition
in shallow water, evaporitic environments. Sm-Nd,
Pb-Pb, and Rb-Sr data from the Nondweni remnant
are equivocal. It is not possible to discriminate
between extrusion at = 3500 Ma with resetting at
=3100 Ma, or formation at =3100 Ma with con-
tamination by 3500 Ma-old sialic crust (A. H.
Wilson, oral commun., 1986).
Evidence for low-angle thrusting and recum-
bent nappes is present in the Assegaai, Common-
dale, and Nondweni remnants (Smith, 1987; Talbot
and others, 1987; A. Versfeld, oral commun., 1987).
Granitoids at Archaean continental margins in southern Africa 271
The sequence of deformational events in Swaziland
and adjacent areas is summarized in Table 1.
GRANITOIDS
The oldest quartzofeldspathic gneisses con-
stitute the Bimodal Gneiss Suite of the Ancient
Gneiss Complex (A.G.C.), which yielded an Sm-Nd
isotopic age of 355948 Ma (Carlson and others,
1985) and a U-Pb age on a single zircon of 36444
Ma (Compston and Kroner, 1988). Whereas the Sm-
Nd date is indistinguishable from the Sm-Nd ages of
351060 Ma and 353050 Ma reported from the
Barberton Sequence (Hamilton and others, 1979,
1983), the U-Pb data from the Onverwacht Group
indicate a probable extrusion age of 3450 Ma (Arm-
strong and others, 1988), implying that at least
some of the Bimodal Gneiss Suite pre-dates the
Barberton Sequence. Despite the fact that the mor-
phology and texture of zircons in the Bimodal
Gneisses are consistent with precipitation from the
original magma (Compston and Kroner, 1988), the
stratigraphic relation of the Bimodal Gneiss Suite to
the Barberton Sequence remains equivocal. The
gneisses yielding the oldest zircons underlie a dis-
crete area some 70 km north of the type area of the
Bimodal Suite in southwestern Swaziland. There is
no conclusive proof that all the Bimodal Suite is of
the same age, but Jackson (1984) and Talbot and
others (1987) deduced from structural data that the
Dwalile and Assegaai supracrustal sequences post-
date the Bimodal Gneiss Suite.
Three groups of quartzofeldspathic gneiss have
been recognized (Hunter and others, 1984). The
most common has Rb and (Nb + Y) contents (Figure
3) similar to Phanerozoic subduction-related gra-
nites (Pearce and others, 1984) but is distinguished
from them by heavy rare-earth element (HREE) and
high field-strength element (HFSE) depletion (SWZ-
7,9; WR4, 7, 9; TGI4: Figures 4 and 5). A second
group comprises high Si0
2
gneisses with large nega-
tive Eu anomalies, high Th/Ba ratios, enriched
contents of HFSE, and flat HREE slopes (SWZ-5, 8,
28, 30, 31: Figures 4 and 5). Rare, potassic quart-
zofeldspathic gneisses are also present in the
Bimodal Gneiss Suite (Table 3) that have geo-
chemical characteristics (except for K
2
0) similar to
the high-Si02 gneisses (SWZ-24; Figures 4 and 5).
The third type is characterized by very strongly
fractionated REE patterns, small to large positive
.,.,,------------:r-----,.OO
100
.0
VAG ORG

Y + Nbppm
Figure 3. Rb vs (Y +Nb) plot for granitic rocks from
Swaziland and adjacent areas. The Lochiel batholith
occupies the field marked'" 3.0 Ga, the Bazane granite and
Dalmein type plutons occupy the field that lies in an
intermediate position (3.3 to 3.0 Ga), and layered gneisses
and intrusive leucotonalite plutons occupy the field
marked >3.3 Ga. SYN-COLG, syn-collision; VAG, vol-
canic arc; WPG, within plate; ORG, ocean ridge granite.
After Pearce and others (1984).

>-
'" 0
z
0
"
"-
"
'" 0
'"
100
50
10
La (@' Nd
>c,' SI/1-8. 30. 31 510,-77%
rrm 5111-5.18 5io, 73 -77%
of WR7
225111- 7.9 5iO,69 -71%
100
50
Sm Eu Gd Tb Oy Tm Yb
Figure 4. Chondrite-normalized rare-earth element
patterns for quartzofeldspathic gneisses in the Ancient
Gneiss Complex. The grouping into the respective fields is
identified in the text. Samples prefixed SWZ are from
Swaziland; samples prefixed WR and TG are from Com-
mondale.
272
D. R. Hunter
1.0
0.1
SWZ - 5
swz - 31
b ,WR4
SWZ -14
5 WZ - 9
10
0.1
0.01 0 0 1
K,O "b B. Th c. Hf Zr Sm Yb
Figure 5. Geochemical patterns of selected A.G.C. gneis-
ses normalized against ocean ridge granite (Pearce and
others, 1984).
Eu anomalies, and high contents of Ba and Sr
(Figure 6).
The Bimodal Gneiss Suite has a low initial
87Sr/86Sr ratio of 0.6999, REE patterns, moderate to
low ThlEa ratios, and low 5
18
0 values indicative of
derivation by partial melting of mantle-derived
mafic rocks metamorphosed to quartz eclogite, mafic
granulite, or amphibolite (Barker and Arth, 1976;
Barker and others, 1976; Barton and and others,
1980; Condie and Hunter, 1976; Hunter and others,
1978; Gower and others, 1982). The high Si0
2
,
REE-enriched gneisses may reflect partial melting
of a mafic source with residual pyroxene and plagio-
clase (Hunter and others, 1984). The strong HREE
depletion of the third group of gneisses suggests that
garnet was a significant residual phase.
The tectonic setting in which this early sialic
crust developed is uncertain, but its existence prior
to accumulation of at least some metavolcanic
supracrustal sequences is presumed because of the
presence of interbedded clastic sediments (e.g.,
Assegaai).
Further support for a complex pre-Barberton
Sequence history is based on the timing of meta-
morphism in the Dwalile suite and the presence of
potassic granite clasts in the basal Moodies conglo-
merate of the Barberton Sequence. Detrital zircons
w
e-
o-
o
Z
o
I
100
50
" 10
'" o
o
0:
Field of HI138 1 H 1147 and
y HI241
To 13

j I
i c] Ce Pr Nd Sm Eu Gd Dy Ho Er yb
Figure 6. Chondrite-normalized rare-earth element
patterns for layered gneisses from Manzini area, Swazi-
land (H series). TG13 is a quartzofcldspathic gneiss from
Commondale.
from metaquartzites and metamorphic garnet from
the Dwalile suite yield an array indicating a meta-
morphic event at "" 3450 Ma (Kroner and others,
1987). Granitic pebbles and gneiss clasts from the
basal quartzite of the Moodies Group have geo-
chemical characteristics of evolved continental crust
(Reimer and others, 1985). Zircons from these clasts
and pebbles yielded U-Pb ages of 3570 to 3518 Ma
(Kroner and Compston, 1988). It is reasonable that
a major deformational and metamorphic event
occurred at ""3500 Ma during which potassic gra-
nite melts were also generated. No trace of this old
potassic granite crust has been found to date, but its
generation would seem to be contemporaneous with
development of the oceanic environment in which
Barberton Sequence accumulated (see Table 2).
Furthermore, the ensialic Dwalile supracrustal
remnant is probably somewhat older than the
Barberton Sequence, (Kroner and Compston, 1988).
Leucotonalitic intrusions around the southern
and northwestern margins of the Barberton rem-
nant were equated with the Bimodal Gneiss Suite
(Viljoen and Viljoen, 1969; Anhaeusser and Robb,
1981). These rocks yielded a wide range of Rb-Sr
isotopic ages from 348092 to 277551 Ma with
initial 87Sr/86Sr ratios typically increasing from
0.7002 to 0.7029 with decreasing age. U-Pb isotopic
Granitoids at Archaean continental margins in southern Africa 273
data for zircons, sphene, and apatite indicate ages of
325080 and 322080 Ma for the Thesspruit and
Nelshoogte intrusions, respectively (Oosthuysen,
1970). These latter ages are close to the 328746
Ma age for the Middle Marker in the Onverwacht
Group (Hurley and others, 1972; Rb-Sr isochron)
which McLennan and others (1983) interpret as the
age of deformation and metamorphism of the Onver-
wacht Group (Table 3). Armstrong and others
(1988) reported a U-Pb age for zircon from the
Theespruit pluton of 3437 3 Ma which is indistin-
guishable from the age obtained for zircons from
felsic lavas of the Hooggenoeg Formation (i.e., the
unit immediately overlying the Middle Marker in
the conventional stratigraphy) and the Theespruit
Formation (i.e., in the lower part of the Onverwacht
Group). These data suggest a somewhat younger
age for the Onverwacht Group than given by whole-
rock Sm-Nd studies and contemporaneity of extru-
sion of felsic lavas with emplacement of leu co-
tonalitic magmas. Robb and others (1986) sugges-
ted that leucotonalitic plutons and the hornblende-
bearing Kaap Valley intrusion, which have im-
precise Rb-Sr whole rock ages, were emplaced at
"" 3450 Ma and subsequently remobilized by solid-
state diapirism at "" 3200 Ma. These authors con-
sider that their proposal is supported by (1) intense
lineation plunging radially down-dip in the leuco-
tonalites, in their mafic xenoliths, and in the
country rocks; (2) lack of large-scale cross-cutting
relations; and (3) presence of layered gneisses
within the plutons.
Recent isotopic data for the Kaap Valley quartz
dioritic pluton cast some doubt on this model.
Tegtmeyer and Kroner (1987) reported a U-Pb
zircon age of 3223 5 Ma for the Kaap Valley
pluton. They drew attention to the morphology of
the zircon population, which they consider to be
typical of magmatic growth. Tegtmeyer and Kroner
(1987) are reluctant to accept their zircon age as
reflecting Pb loss associated with solid-state dia-
pirism and interpret the age as the time of the
magmatic intrusive event. This has implications for
the age of mineralogically and chemically similar
Tsawela gneiss in Swaziland and Braunschweig
gneiss in northern Natal. The Tsawela gneiss,
which is intensely deformed, yielded an Rb-Sr age
3323 86 Ma. This date is assumed to reflect Sr
isotopic rehomogenization because the less de-
formed, post-Tsawela, Granodiorite Suite in central
Swaziland has an isotopically similar Rb-Sr age
(335057 Ma, Barton and others, 1983). The
Granodiorite Suite comprises two NE-elongated
intrusions differentiated from gabbro through
quartz diorite and leucotonalite to granodiorite.
Leucotonalite constitutes "" 60% of the suite (Hun-
ter, 1973).
The new geochronological data seem to suggest
that the leucotonalitic, domical intrusions adjacent
to the Barberton remnant were emplaced at "" 3450
Ma and were genetically related to contemporane-
ous felsic volcanism in the Onverwacht Group. A
subsequent period between "" 3300 Ma and "" 3200
Ma appears to reflect a major deformational event
when additional leucotonalitic and tonalitic mag-
mas were generated as a tabular intrusion; but in
the north the leucotonalitic and tonalitic magmas
solidified as crudely circular, diapiric bodies such as
the Kaap Valley and other plutons. :No tonaliticl
leucotonalitic intrusions with similar geometries
are found adjacent to the Assegaai, Commondale,
and Nondweni remnants. This absence of domical
granitoid intrusions here may relate to the large-
scale heterogeneous simple shear motion deduced
from structural studies in southwestern Swaziland
(Jackson, 1984).
Subsequent granitoid activity involved the
emplacement of multi-phase tabular batholiths and
high-level, discordant plutons (see Table 2). The
sodic Bazane batholith is now dismembered by ero-
sion and is represented by three discrete masses
that crop out south of Swaziland (see Figure 1). The
Nelspruit and Lochiel batholiths are adamellitic in
composition (Table 3). The former occurs only
north of the Barberton supracrustal remnant,
whereas the Lochiel batholith extends southwest-
ward along the southeastern margin of the Bar-
berton remnant.
Although the Bazane and Lochiel batholiths
have contents of Rb and (Nb+ Y) similar to those
reported from Phanerozoic subduction-related gra-
nites (see Figure 3), the former displays distinctive
HREE depletion (Figure 7). Six of the Bazane
granite samples have a prominent positive Eu ano-
maly (Figure 7). The Lochiel granite is distin-
guished by a negative Eu anomaly with a flat to
gentle HREE slope. Both the Bazane and Lochiel
granites have variable Th/Ba ratios which are atypi-
cal of Phanerozoic collision-related granites (Figure
8; Pearce and others,1984). Whereas the Bazane
sodic granitoids were generated by partial melting
of garnet eclogite or mafic granulite source rocks,
the presence of a positive Eu anomaly in one sample
suggests a non-homogeneous source. The Lochiel
274
D. R. Hunter
Table 3a. Representative analyses of granitoids.
1 2 3 4 5 6 7 8 9
SiO
z
73.37 67.81 76.19 76.81 68.42 65.44 68.37 72.96 72.63
TiO
z
00.26 00.61 00.26 00.31 00.32 00.68 00.36 00.25 00.32
Al
z
0
3
15.46 16.20 12.47 10.97 15.45 15.86 15.12 15.22 14.96
Fez03
00.16 00.47 00.36 00.61 00.79 00.55 3.15*1 00.18 00.23
FeO 1.33 3.79 1.70 2.44 2.35 4.46 1.45 1.83
MnO 00.01 00.07 00.04 00.04 00.06 00.08 00.01 00.03 00.04
MgO 00.49 1.59 00.42 00.69 2.03 2.10 1.81 00.55 00.49
CaO 2.64 4.09 00.89 1.00 3.23 4.79 3.04 12.9 1.88
NaOz 5.19 4.32 4.84 2.57 4.81 4.33 5.35 5.16 4.22
KzO 1.17 1.30 2.14 4.06 1.88 1.39 1.90 00.96 3.17
P
Z
0
5
00.08 00.25 00.03 00.04 00.09 00.23 00.13 00.09 00.10
HzO+ - - 00.28 00.53 00.80 - - - -
HzO- - - 00.04 00.04 00.07 - - - -
Total 100.16 100.50 99.72 100.08 100.30 99.91 100.37 99.04 99.87
Lor 00.73 00.77 - - - 00.39 1.13*2 00.79 00.87
Ba 263 147 136 288 303 343 148 416 718
Rb 48 52 66 58 67 50 51 47 131
Sr 673 434 50 85 426 403 467 646 346
Nb 3.1 17.4 - - - 4.6 - 1.3 3.5
Y 5.0 31.8 - - -
11.4
- 7.3 12.1
Zr 161 206 437 595 124 109 - 135 181
Sc 1.4 9.6 3.1 7.1 6.5 8.9 - 1.6 3.4
Th 4.2 4.2 18.3 5.9 5.2 0.8 -
-0.9 14.7
U 0.6 2.2 1.7 0.7 1.3 - - 1.5 0.6
Key: 1, Leucotonalitic gneiss (A.G.C.) (Smith, 1987); 2, Tonalitic gneiss (A.G.C.) (Smith, 1987); 3, High-Si leucotonalitic
gneiss (A.G.C.) (Hunter and others, 1984); 4, Potassic high-Si gneiss (A.G.C.) (Hunter and others, 1984); 5, Tsawela gneiss
(A.G.C.) (Hunter and others, 1984); 6, Braunschweig gneiss (Smith, 1987); 7, Kaap Valley granite (Robb, 1981); 8, Leuco-
tonalite, Bazane granite (Smith, 1987); 9, Granodiorite, Bazane granite (Smith, 1987). *1, Total Fe as Fez03; *2, Lor
included in total.
magmas owe their origin to shallow-level melting
where plagioclase was a residual phase (Condie and
Hunter, 1976). The low initial ratios indicate some
mixing with melts derived from a mantle or a
crustal source having a short residence time.
A number of small, megacrystic granodiorite
plutons (Table 3), collectively referred to as the
Dalmein type, intrude the margins of the Barberton
Sequence (Figure 1). These intrusions, together
with the Heerenveen granite, were probably em-
placed contemporaneously with the Lochiel granite.
The Dalmein-type granodiorites occupy the field
identified as that of the 3.0-3.3 Ga granites (Figure
3).
Emplacement of the Lochiel batholith at "='3000
Ma (Barton, 1983) marked the termination of a
period characterized by the dominance of events
resulting from mantle processes to one when crustal
processes assumed an increasingly more important
role. After 3000 Ma, tholeiitic volcanism was suc-
ceeded by sedimentation that was largely controlled
by repeated interaction of braided alluvial plains
and macrotidal basins (Pongola Sequence). This
period of volcanism and sedimentation was ter-
minated by intrusions including gneiss domes
(Nhlangano and Hluti), a tubular batholith (Hlati-
kulu), and a number of high-level plutons (Kwetta,
Mooihoek, Sicunusa, etc.) (Figure 1). The post-
Pongola granitoids have Rb and (Nb + Y) contents
appropriate to Phanerozoic within-plate granitoids.
Isotopic data yield only errorochrons, which are
considered to reflect incomplete homogenization
during partial melting of the sialic basement to the
Pongola depository (Hunter and Wilson, 1988).
Granitoids at Archaean continental margins in southern Africa 275
Table 3b. Representative analyses of granitoids.
10 11 12 13 14 15 16 17 18
Si0
2
70.60 68.73 69.72 68.89 74.14 73.52 69.14 73.48 78.19
Ti0
2
00.25 00.40 00.32 00.64 00.13 00.26 00.76 00.22 00.11
Al
2
0
3
15.33 15.37 15.02 14.57 13.29 12.19 14.05 13.39 11.09
Fe203
00.55 2.64*1 1.21 00.80 00.14 00.36 00.49 00.21 00.13
FeO 1.53 1.09 2.12 1.12 2.91 3.94 1.71 1.05
MnO 00.04 00.07 00.05 00.05 00.03 00.05 00.06 00.02 00.04
MgO 1.18 1.26 00.95 00.62 00.11 00.13 00.99 00.22 00.00
CaO 2.59 2.13 1.72 2.39 1.00 1.02 2.65 00.82 00.53
Na02 5.38 4.09 4.72 4.05 3.99 3.53 3.41 3.03 2.73
K
2
0 1.78 4.01 3.46 3.85 4.62 5.84 4.36 6.25 5.22
P205
00.08 00.05 00.18 00.23 00.02 00.03 00.26 00.06 00.01
H
2
O+ 00.34 - - - - - - - -
H
2
O- 00.04 - - - - - - - -
Total 99.69 99.33 99.51 99.37 98.59 99.84 100.11 99.41 99.10
LOI -
0.58*2 1.08*2 1.15*2 00.68 0.52 0.98 1.12 0.69
Ba 337 973 700 910 234 231 957 371 292
Rb 54 130 121 168 261 37 219 323 268
Sr 589 519 483 447 60 861 304 68 20
Nb - -
15.0 21.0 21.2 40.3 35.0 33.1 28.5
Y - - 18.0 16.0 45.3 115.1 58.0 27.8 101.2
Zr 147 - 77 130 116 634 339 232 158
Sc 3.1 - 5.0 5.0 1.8 2.7 8.7 4.1 3.0
Th 6.3 - 7.3 12.0 36.0 28.2 35.6 55.9 21.2
U - -
2.8 2.8 3.5 3.3 3.0 11.4 1.3
Key: 10, Theespruit pluton (Hunter and others, 1987); 11, Nelspruit porphyritic granite (Robb and others,1983);
12, Dalmein pluton (Glikson, 1976); 13, Lochiel granite (Glikson, 1976); 14, Hlatikulu granite, (unpubl. data); 15, Nhlan-
gano gneiss, (unpubl. data); 16, Kwetta pluton, (unpubl. data); 17, Sinceni pluton, (unpubl. data); 18, Mhlosheni pluton,
(unpubl. data). *1, Total Fe as Fe203; *2, LOI included in total.
STRUCTURE
Early Archaean supracrustal remnants and
spatially associated granitoids were complexly
folded. Although this has long been recognized,
detailed structural studies are mainly confined to
Swaziland and adjacent areas immediately south of
that country (Jackson, 1984; Smith, 1987; Talbot
and others, 1987).
Jackson (1984) demonstrated that the gneiss
terrane in southwestern Swaziland reflected evo-
lution from early homogeneous ductile strain when
the rocks were at upper amphibolite facies, and
hence in a state of maximum ductility, to late in-
homogeneous brittle strain, indicating deformation
at successively higher crustal levels.
Ductile distortion of the Swaziland terrane and
its small remnants of metavolcanic rocks, followed
by displacements along shear zones (D
5
, see Table
1), and possibly thrusts, lifted it northward and
northwestward toward the Barberton supracrustal
remnant (Jackson, 1984). A vertical displacement
of about 20 km to within a few kilometres of the
surface is considered probable (Jackson, 1984). Ele-
vation of this gneiss complex above sea level would
provide a source for the clastic sediments that com-
prise the stratigraphically higher parts of the
Barberton Sequence. In northwestern Swaziland,
these sedimentary sequences were tectonically
transported northwestward as a consequence of ex-
tensive low-angle thrusting and kilometre-scale
asymmetric synsedimentary folding (Lamb and
Paris, 1988).
276
D. R. Hunter
"-
"-
" "-
"-
"-
"-
"-
"-
fi eld of lodH, l granite(lO Ga)
F,.ld of e.,.." 9....,;1. (J. 2 G.,
" - 8.aunQo or.ande IPOH
\ r-- ......
\ I -_
\ ;'
--
--
Figure 7. Chondrite-normalized rare-earth element
patterns for the Bazane, Lochiel, and Hlatikulu granites.
C!>
Q:
o
"-
10
1.0
o
Q:
0.1
Kp Rb 80 Th Nb Ce Hf Zr Sm Y Yb
Figure 8. Geochemical patterns of the Lochiel (l08, 114)
and Bazane, (NLDl, NLD2) normalized against ocean
ridge granite (Pearce and others, 1984).
The late-tectonic intrusion of the multiphase
Lochiel batholith is demon-strated by the lack of
foliation that characterizes early phases. Later,
more potassic phases are devoid of a strong foliation.
The earliest deformation in the Assegaai
Sequence involved thrusting along brittle ramp and
crush zones. With increasing temperature, ductile
deformation led to development of an S1 foliation in
the supracrustal rocks. Early crush zones were
locally overprinted by mylonites, and reclined folds
gave rise to recumbent nappes with northerly
aligned axes. The Bazane tabular batholith was
emplaced in the interval between this D1 event and
the subsequent D2 event, which involved refolding
of older structures about NNE-trending, near-
vertical axial surfaces. Based on the provisional Rb-
Sr age of the Bazane batholith, D2 must have oc-
curred == 3200 Ma.
DISCUSSION
The Archaean rocks in Swaziland, southeastern
Transvaal, and northern Natal reflect periods of
dominantly mafic or ultramafic volcanism inter-
spersed with periods during which massive volumes
of granitoid magmas were intruded. During the
interval from earliest sialic crust formation at
== 3600 Ma to == 3000 Ma, Archaean rock suites re-
peatedly deformed and metamorphosed at amphi-
bolite facies. Subsequent to emplacement of the
Lochiel batholith, the Pongola Sequence accumula-
ted (Hegner and others, 1984) and was deformed
into typically open folds with gently dipping limbs.
Steeper dips are a feature where the Pongola
Sequence is involved in a major zone of north-
westerly aligned refoliation and faulting. Sedi-
mentation was terminated by emplacement of
potassic granitoids, largely within the central area
of the Pongola sedimentary basin.
The worldwide ubiquity of high-Mg volcanic
rocks (komatiites, basaltic komatiites) has been
interpreted as evidence for an Archaean mantle that
was hotter than the modem mantle (e.g., Sleep,
1979). In contrast, there is an apparent similarity of
Archaean continental geothermal gradients to
modem gradients (Bickle, 1978), and it has been
argued that Archaean continents formed above
subduction zones (Campbell and Jarvis, 1984). As
the evolution of early Archaean crust in the south-
eastern Kaapvaal province is intimately related to
the generation of large volumes of granitoid mag-
Granitoids at Archaean continental margins in southern Africa 277
mas, this proposal carries the implication that these
magmas were produced along subduction zones.
Acceptance of the circumstantial evidence that
the Bimodal Gneiss Suite constitutes the oldest, pre-
greenstone sialic crust, provides the basis for one
possible scenario. It has been proposed (Hunter and
others, 1984) that this crust evolved as a conse-
quence of deformation, metamorphism, and partial
melting of an early unstable, thick basaltic crust
yielding tonalitic/trondhjemitic liquids that were
intruded or extruded contemporaneously with con-
tinuing mafic volcanism. With the passage of time
this sialic crust became thicker and more extensive,
possibly as a result of coagulation and tectonic
interslicing of mini-plates. As these mini-plates en-
larged by accretion they were rifted. Ensialic
komatiitic volcanism with interbedded terrigeneous
sediments (e.g., Nondweni) probably developed
contemporaneously with komatiitic volcanism in
marine environments (e.g., Barberton). Such a scen-
ario would account for the subtle differences ob-
served in the Barberton remnant and its southern
counterparts.
Deformation of the sialic crust and its ensialic
supracrustal rocks involved an early stage of thrust-
ing along brittle ramp and flat crush zones. In-
creasingly ductile behaviour with rising tempera-
tures resulted in some early Dl crush zones being
represented by late Dl mylonitic zones. Intrusion of
thin but extensive sheets of sodic granitoid magmas
(Bazane granite) along subhorizontal contacts at the
interface of the supracrustal cover and older sialic
basement was a consequence of partial melting of
these ensialic metavolcanic supracrustal rocks. In-
trusion of these granitoid sheets, together with
subhorizontal thrusting, resulted in an initial thick-
ening of the continental crust.
The larger (wider?) marine environments repre-
sented by the Barberton remnant gradually closed
(Figure 9), resulting in partial melting of the mafic
meta volcanics and generation of syntectonic leuco-
tonalitic/trondhjemitic plutons at "'" 3450 Ma. Con-
tinued compression resulted in the development of
diapiric intrusions along the northwestern flank of
the closure (Figure 9). The presence of layered
tonalitic gneisses tectonically intersliced with meta-
volcanic rocks is a further consequence of this com-
pression. When the rising sialic crust to the south
became elevated above sea level, it was weathered
and eroded. The clastic detritus was transported
northward into depositional basins affected by
synsedimentary tectonism (Lamb and Paris, 1988).
I.
N
SL
Accumulation 01 Onv.rwacht 5
with formation 01 new oceanic f loor --------
2.
N 5
SL
"I .. If 10 .II 1'1 110 10: II ;It
"I :Ii :II II "I II .II '"
.II :0; "I It II II "I If 10:
3.
N
SL
Compre .. ion, initiation 01 obduction of Onverwachf :
reel I ned ibid 1"Il 01 Onverwacht
depolition
11.10 .... ..
. .. :Ii ... II
:00: }I; .. "I II ...
Collision 01 siolie plat.s:
Elevation above HQ- I ... I 01 sialic c,ullin south:
Oeposit lon 01 Fig" .. and Moodles Groups.
s
Figure 9. Schematic representation of closure of the
Barberton greenstone belt. In 3, new tonalitic magma can
also be generated in addition to reactivation of older sialic
crust.
On complete closure, partial melting of the
depressed tonalitic/trondhjemitic crust resulted in
generation of great volumes of more potassic
magmas (Lochiel batholith). It is proposed that the
Lochiel granite magmas were generated as a
consequence of collision of sialic crustal plates, but
it is further suggested that the chemistries of
Archaean collision-related products differ from
278
D. R. Hunter
those associated with younger collision boundaries.
Younger cover obscures much of the Archaean
geology of the central Kaapvaal province, but
borehole data indicate that potassic granites with
ages of ==: 3000 Ma occur south and southeast of the
line defined in Figure 10. Greenstone remnants are
preserved along the northern flank of this zone of
potassic granitoids, suggesting that they mark the
site of an original collision boundary. The 3000 Ma
granitoids underlie an area of ==: 50,000 km
2
, which
is comparable to the areal extent of the collision-
related potassic granites in the Higher Himalayan
batholith (Wilks, 1988).
Peak metamorphic conditions are considered in
this scenario to have been attained at 3300-3200 Ma
I!;J Granitoid roc:leI (nat _. but Ga)
rz;;J roclel Ga)
ago when the Kaap Valley tonalite in Barberton,
the late-tectonic Granodiorite Suite in Swaziland,
and the Matshempondo Peridotite Suite at Com-
mondale were emplaced (see Table 2).
Emplacement of the 3000 Ma Lochiel batholith
effectively terminated widespread granitoid intru-
sion in the southern and southeastern Kaapvaal
province. High-standing, relatively stable continen-
tal crust sustained a series of basins in which thick
sedimentary sequences with intercalated low MgO
volcanic rocks accumulated. The presence of 3300
Ma diamonds from the Kaapvaal province confirms
that, by this time, a 150-200-km-thick cold litho-
sphere existed beneath this cratonic area (Richard-
son and others, 1984).
Figure 10. Simplified geological map of the Kaapvaal province showing distribution of granitoids of different ages and of
greenstone remnants. Broken line extending across the province marks northern limit ofthe 3000 Ma granitoids, the presence
of which has been detected in boreholes beneath the Mesozoic cover (Burger and Coertze, 1973).
Granitoids at Archaean continental margins in southern Africa 279
REFERENCES
ANHAEUSSER, C. R., and L .J.ROBB, 1981, Magmatic
cycles and the evolutionofthe Archean granitic crust in the
eastern Transvaal and Swaziland; pp. 457-467 in J. E.
Glover and D. 1. Groves (eds.), Archaean Geology: Geo-
logical Society of Australia, Special Publication 7,515 p.
ARMSTRONG, R. A., W. COMPSTON, and M. J. DE WIT,
1988, Precise zircon ion microprobes ages in the Barberton
greenstone belt: An evaluation of stratigraphic relation-
ships [abstract): Geological Society of South Africa,
Extended Abstracts of the 22nd Earth Sciences Congress, p.
7-10.
BARKER, F., and J. G. ARTH, 1976, Generation of trond-
hjemitic-tonalitic-basalt suites: Geology, v. 4, p. 596-600.
1. FRIEDMAN, D. R. HUNTER, and J. D. GLEASON,
1976, Oxygen isotopes of some trondhjemites, siliceous
gneisses and associated mafic rocks: Precambrian Re-
search, v. 3, p. 547-557.
BARTON, Jr., J. M., 1983, Isotopic constraints on possible
tectonic models for crustal evolution in the Barberton
granite-greenstone terrane, southern Africa; pp. 73-79 in
C. R. Anhaeusser (ed.), Contributions to the Geology of
the Barberton Mountainland: Geological Society of
South Africa, Special Publication 9, 223 p ..
D. R. HUNTER, M. P. A. JACKSON, and A. E.
WILSON, 1980, Rb-Sr ageand source of the Bimodal Suite
of the Ancient Gneiss Complex, Swaziland: Nature, 283, p.
756-758.
D. R. HUNTER, M. P. A. JACKSON, and A. E.
WILSON, 1983, Geochronologic and Sr-isotopic studies of
certain units in the Barberton granite-greenstone terrane,
Swaziland: Geological Society of South Africa Transac-
tions, v. 86, p. 71-80.
BURGER, A. J., and F. J. COERTEZ, 1973, Radiometric
Age Measurements on Rocks from Southern Africa to
the end of 1971: Geological Survey of South Africa, De-
partment of Mines (Pretoria), Bulletin 58, 46 p.
BICKLE, M. J., 1978, Heat loss from the Earth: A con-
straint on Archean tectonics from the relation between
geothermal gradients and the rate of plate production:
Earth and Planetary Science Letters, v. 40, p. 301-315.
CAMPBELL, 1. H., and G. T. JARVIS, 1984, Mantle con-
vection and early crustal evolution: Precambrian Re-
search, v. 26, p. 15-56.
CARLSON, R. W., D. R. HUNTER, and F. BARKER, 1985,
Geochronologic investigation of the Archean Ancient
Gneiss Complex, Swaziland [abstact): EOS, Transactions
of the American Geophysical Union, v. 66, p. 419.
COMPSTON, W., and A. KRONER, 1988, Multiple zircon
growth within early Archean tonalitic gneiss form the
Ancient Gneiss Complex, Swaziland: Earth and Planetary
Science Letters, v. 87, p. 13-28.
CONDIE, K. C., and D. R. HUNTER, 1976, Trace-element
geochemistry of Archean granitic rocks form the Barberton
region, South Africa: Earth and Planetary Science Letters,
v. 29, p. 389-400.
DE WIT, M. J., 1982, Gliding and overthrust nappe tec-
tonics in the Barberton greenstone belt: Journal of Struc-
tural Geology, v. 4. p. 117-136.
and C. R. STERN, 1980, A 3500 Ma ophiolite
complex from the Barberton greenstone belt, South Africa:
Archan oceanic crust and its geotectonic implications
[abstract): Geological Society of Australia, Extended
Abstracts of the 2nd International Archaean Symposium
(Perth), p. 85-87.
___ , R. E. P. FRIPP, and 1. G. STAISTREET, 1983, Tectonic
and stratigraphic implications of new field observations
along the southern part of the Barberton greenstone belt;
pp. 21-29 in C. R. Anhaeusser (ed.), Contributions to the
Geology of the Barberton Mountainland: Geological
Society of South Africa, Special Publication 9, 223 p.
DOKKA, R. K., and D. R. LOWE, 1984, Large-scale
Phanerozoic-style thrust faulting in the Archaean: Evi-
dence from the southwestern Barberton greenstone belt,
South Africa [abstract): Geological Society of America Ab-
stracts with Programs, v. 16, p. 49.
GLIKSON, A. Y., 1976, Trace element geochemistry and
origin of early Precambrian acid igneous series, Barberton
Mountain Land, Transvaal: Geochimica et Cosmochimica
Acta, v. 40, p. 1261-1280.
GOWER, C. F., D. K. PAUL, and J. H. CROCKET, 1982,
Protoliths and petrogenesis of Archean gneisses from the
Kenora area, English River subprovince, northwest On-
tario: Precambrian Research, v. 7, p. 245-274.
HAMILTON, P. J., N. M. EVENSEN, R. K. O'NIONS, H. S.
SMITH, and A. J. ERLANK, 1979, Sm-Nd dating of Onver-
wacht Group volcanics, southern Africa: Nature, v. 279,
p.298-300.
R. K. O'NIONS, D. BRIDGWATER, and A. NUTMAN,
1983, Sm-Nd studies of Archaean metasediments from
west Greenland and their implications for the Earth's
early history: Earth and Planetary Science Letters, v. 62, p.
263-272.
HEGNER, E., A. KRONER, and A. W. HOFMANN, 1984, Age
and isotopic geochemistry of the Archaean Pongola and
Usushwana suites in Swaziland, South Africa: A case for
crustal contamination of mantle-derived magmas: Earth
and Planetary Science Letters, v. 70, p. 267-269.
HUNTER, D. R., 1957, The geology, petrology and classi-
fication of Swaziland granites and gneisses: Geological
Society of South Africa Transactions, v. 60, p. 85-120.
280
D. R. Hunter
__ , 1973, The granitic rocks of the Precambrian in
Swaziland; pp. 131-145 in L. A. Lister (ed.), Symposium
on Granites, Gneisses and Related Rock: Geological
Society of South Africa, Special Publication 3, 509 p.
__ , and A. H. WILSON, 1988, A continuous record of
Archaean evolution from 3.5 Ga to 2.6 Ga in Swaziland and
northern Natal: South African Journal of Geology, v. 91, p.
57-74.
__ , F. BARKER, and H. T. MILLARD, Jr., 1978, The
geochemical nature of the Archean Ancient Gneiss Com-
plex and Granodiorite Suite, Swaziland: A preliminary
study: Precambrian Research, v. 7, p. 105-127.
__ , A. R. ALLEN, and P. MILLIN, 1983, A preliminary
note on Archaean supracrustal and granitoid rocks west of
Piet Retief: Geological Society of South Africa Trans-
actions, v. 86, p. 301-306.
__ , F. BARKER, and H. T. MILLARD., Jr., 1984, Geo-
chemical investigation of Archaean bimodal and Dwalile
metamorphic suites, Ancient Gneiss Complex, Swaziland:
Precambrian Research, v. 24, p. 131-155.
HURLEY, P.M., W. H. PINSON, B. NAGY, and T. M.
TESKA, 1972, Ancient age of the Middle Marker horizon:
Onverwacht Group, Swaziland Sequence, South Africa:
Earth and Planetary Science Letters, v. 14, p. 360-366.
JACKSON, M. P. A., 1984, Archaean structural styles in
the Ancient Gneiss Complex of Swaziland, South Africa;
pp. 1-18 in A. Kroner and R. Greiling (eds.), Precambrian
Tectonics Dlustrated: E. Schweizerbart'sche Verlags-
buchhundlung, Stuttgart, Gemany, 419 p.
KRONER, A., and W. COMPSTON, 1988, Ion microprobe
ages of zircons from early Archaean granite pebbles and
greywacke, Barberton greenstone belt, southern Africa:
Precambrian Research, v. 38, p. 367-380.
__ , W. COMPSTON, A. TEGTMEYER, C. MILISENDA, and
T. C. LIEW, 1987, Growth of early Archaean crust in the
Ancient Gneiss Complex of Swaziland and adjacent Bar-
berton greenstone belt, southern Africa; pp. 85-87 in L. D.
Ashwal (ed.), Workshop on the Growth of Continental
Crust (Oxford, UK, July 1987): Lunar and Planetary
Institute (Houston, Texas, USA), Technical Report 88-02,
174p.
LAMB, S., 1987, Archean synsedimentary tectonic defor-
mation - A comparison with the Quaternary: Geology, v.
15, p. 565-568.
__ , and I. PARIS, 1988, Post-Onverwacht Group strati-
graphy in the SE part of the Archaean Barberton green-
stone belt: Journal of African Earth Sciences, v. 7, p. 285-
306.
LANIER, W. P., and D. R. LOWE, 1982, Sedimentology of
the Middle Marker (3.4 Ga), Onverwacht Group, Trans-
vaal, South Africa: Precambrian Research, v. 18, p. 237-
260.
LOWE, D. R., 1982, Comparative sedimentology of the
principal volcanic sequences of Archean greenstone belts
in South Africa, western Australia, and Canada: Im-
plications for crustal evolution: Precambrian Research, v.
17, p.1-29.
__ , and L. P. KNAUTH, 1977, Sedimentology of the
Onverwacht Group (3.4 billion years), Transvaal, South
Africa, and its bearing on the characteristics and evolution
of the early Earth: Journal of Geology, v. 85, p. 699-723.
__ , G. R. BYERLEY, B. L. RANSOM, and B. W. NOCITA,
1985, Stratigraphic and sedimentological evidence bearing
on structural repetition in early Archean rocks of the
Barberton greenstone belt, South Africa: Precambrian
Research, v. 27, p. 165-186.
McLENNAN, S. M., S. R. TAYLOR, and A. KRONER, 1983,
Geochemical evolution of Archean shales from South
Africa, I: The Swaziland and Pongola Supergroup: Pre-
cambrian Research, v. 22, p. 93-124.
OOSTHUYSEN, E. J., 1970, The Geochronology of a
Suite of Rocks from the Granitic Terrain Surrounding
the Barberton Mountain Land: PhD dissertation, Uni-
versity ofthe Witwatersrand, Johannesburg, South Africa,
94p.
PEARCE, J. A., N. B. W. HARRIS, and A. G. TINDLE, 1984,
Trace element discrimination diagrams for the tectonic
interpretation of granitic rocks: Journal of Petrology, v.
24, p. 956-983.
REIMER, T. 0., K. C. CONDIE, G. SCHNEIDER, and A.
GEORGI, 1985, Petrography and geochemistry of granitoid
and metamorphite pebbles from the early Archaean
Moodies Group, Barberton Mountain Land, South Africa:
Precambrian Research, v. 29, p. 383-404.
RICHARDSON, S. H., J. J. GURNEY, A. J. ERLANK, and J.
W. HARRIS, 1984, Origin of diamonds in old enriched
mantle: Nature, v. 310, p. 198-202.
ROBB, L. J., 1981, The Geological and Geochemical
Evolution of Tonalite-Trondhjemite Gneisses and
Migmatites in the Barberton Region, Eastern Trans-
vaal: PhD dissertation, University of the Witwatersrand,
Johannesburg, South Africa, 342 p.
_, J. M. BARTON, Jr., E. J. D. KABLE, and R. C.
WALLACE, 1986, Geology, geochemistry and isotopic char-
acteristics of the Archaean Kaap Valley pluton, Barberton
Mountain Land, South Africa: Precambrian Research, v.
31,p.1-36.
Granitoids at Archaean continental margins in southern Africa 281
_, C. R. ANHAEUSSER, and D. A. VAN NIEROP, 1983,
The recognition of the Nelspruit batholith north of the
Barberton greenstone belt and its significance in terms of
Archaean crustal evolution; pp. 117-130 in C. R An-
haeusser (ed.), ContributionB to the Geology of the Bar-
berton Mountainland: Geological Society of South Africa,
Special Publication 9, 233 p.
SLEEP, N. H., 1979, Thermal history and degassing of the
Earth: Some simple calculations: Journal of Geology, v.
87, p. 671-686.
SMITH, RG., 1987, Geochemistry and Structure of the
Archaean Granitoid-Supracrustal Terrane, South-
eastern Transvaal and Northern Natal: PhD disserta-
tion, University of Natal, Pietermaritzburg, Natal, South
Africa, 305 p.
STANISTREET, I. G., M. J. DE WIT, and R. E. P. FRIPP,
1981, Do graded units of accretionary spheroids in the
Barberton greenstone belt indicate Archaean deep water
environments?: Nature, v. 293, p. 280-284.
TALBOT, C. J., D. R HUNTER, and A. R ALLEN, 1987,
Deformation of the Assegaai supracrustals and adjoining
granitoids, Transvaal, South Africa: Journal of Structural
Geology, v. 9, p. 1-12.
TEGTMEYER, A. R., and A. KRONER, 1987, U-Pb zircon
ages bearing on the nature of early Archaean greenstone
belt evolution, Barberton mountain land, southern Africa:
Precambrian Research, v. 36, p. 1-20.
VILJOEN, M. J., and R P. VILJOEN, 1969, A proposed new
classification of the granitic rocks of the Barberton region;
pp. 153-180 in S. H. Haughton (ed.), Upper Mantle Pro-
ject: Geological Society of South Africa, Special Publica-
tion 2, 484 p.
__ , and R P. VILJOEN, 1970, Archaean volcanicity and
continental evolution in the Barberton region, Transvaal;
pp. 27-49 in T. N Clifford and I. G. Gass (eds.), African
Magmatism and TectonicB: Oliver & Boyd, Edinburgh,
Scotland, UK, 461 p.
WILKS, M. E., 1988, The Himalayas - A modern analogue
for Archaean crustal evolution: Earth and Planetary Sci-
ence Letters, v. 87, p. 127-136.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
The northern Wyoming Province: Contrasts in
Archean crustal evolution
DAVIDW.MOGK
1
, PAULA. MUELLER
2
, JOSEPHL.WOODEN
3
, and DONALDR.BOWES
4
lDepartment of Earth Sciences, Montana State University, Bozeman, Montana 59717, USA
2Department of Geology, University of Florida, Gainesville, Florida 32611, USA
3Branch ofIsotope Geology, U.S. Geological Survey, MS912, 345 Middlefield Road, Menlo Park, California 94025, USA
4Department of Geology and Applied Geology, Glasgow University, Glasgow G 12 8QQ, Scotland, UK
(received December 31,1988; revision accepted March 14, 1989)
ABSTRACT
The northern Wyoming Province can be subdivided into two fundamentally distinct regions: a dominantly
magmatic region (e.g., Bighorn and Beartooth Mountains), and a dominantly metasupracrustal region to the
west (e.g., Gallatin, Madison, and Tobacco Root Mountains). Although not generally exposed, the boundary
between these regions has been described as a mobile belt in the North Snowy Block, northwestern Beartooth
Mountains. This boundary zone separates regions that are clearly distinguished by differences in their
lithologic components, petrologic relations, structural style, whole-rock geochemistry, and isotopic ages.
Within the supracrustal region, however, a marked change also occurs from dominantly plagioclase-rich
gneisses with interlayered mafic amphibolites/granulites in the Gallatin and northeastern Madison Ranges to
K-feldspar-rich gneisses associated with marbles, pelitic schists, quartzites, iron formation, and metabasites
in the northwestern Madison Range and Tobacco Root Mountains. These metasupracrustal sequences are
separated by a 0.5 km wide ductile shear zone in the northern Madison Range. The major lithologic
associations of the northern Wyoming Province appear to have separate and distinct geologic histories.
Present evidence suggests that many of the magmatic and metamorphic lithologic associations were
amalgamated into their present configuration before the end of the Archean. This late Archean episode of
crustal growth in the northern Wyoming Province exhibits many geologic relations normally associated with
with modern accretionary boundaries oflarger scale.
INTRODUCTION
The Archean basement of the northern Wyom-
ing Province contains a variety of lithologic se-
quences with different petrologic relations, struc-
tural styles, and isotopic ages. Exposures of these
rocks are restricted to the cores of a series of fore-
land block uplifts (Figure 1) (e.g., Foose and others,
1961). It is difficult to reconstruct the evolution of
this ancient continental crust because of discontinu-
ous exposures of Archean rocks in the northern
Wyoming Province (as opposed to Archean shield
areas; e.g., Wooden and others, 1988a), and due to
the inherent lack of stratigraphic, paleontologic,
and paleomagnetic controls in Archean terranes in
general. However, recent attempts at establishing a
geochronologic framework for the northern Wyom-
ing Province, coupled with detailed petrologic and
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 283-297. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
283
284
D. W. Mogk, P. A. Mueller, J . 1. Wooden, and D. R. Bowes
N
North Snowy Block
ill Dominantly Plutonic Rocks
CJ Dominantly Metasupracrustal Rocks
o Other Archean Rocks
Montana
Wyoming
8;,ho," """o ..
+
o T"oo R .. "
o SO 100

Km
Wind River Range
Figure 1. Index map of exposures of Archean rocks in the northern Wyoming Province.
geochemical studies, demonstrated that adjacent
packages of Archean rocks experienced significantly
different geologic histories. The recognition of
abrupt discontinuities in lithology between these
adjacent sequences of Archean rocks allows the
interpretation that these lithologic packages are
largely allochthonous; geochronologic constraints
require that these lithologic packages were
amalgamated in the late Archean.
A fundamental boundary in the nature of the
Archean crust of southwestern Montana was recog-
nized in the North Snowy Block, Beartooth Moun-
tains (Mogk and others, 1988a; Mogk and Henry,
1988). The North Snowy Block is an Archean
mobile belt that contains tectonically interleaved
meta-igneous and metasupracrustal rocks (Table
1). To the east of the mobile belt the central and
eastern Beartooth Mountains and Bighorn Moun-
tains encompass a region that is comprised domin-
antly of late Archean igneous and meta-igneous
rocks. To the west of this mobile belt the Gallatin,
Madison, Tobacco Root, Ruby, and Blacktail Moun-
tains are comprised dominantly of meta supracrustal
rocks. Previous contributions toward the under-
standing of these Archean meta supracrustal rocks
include the work of Spencer and Kozak (1975) and
Salt (1987) in the northern Gallatin and Madison
Ranges, Erslev (1983) in the southern Madison
Range, Vitaliano and others, (1979) in the Tobacco
Root Mountains, Garihan (1979) in the Ruby Range,
and Clark (1987) in the Blacktail Mountains. The
geologic relations of the northern Wyoming Pro-
vince are presented in an E-W cross section through
the Beartooth, Gallatin, Madison, and Tobacco Root
Mountains (Figure 2). The rocks of each area are
characterized in terms of their lithologic associa-
tions, petrologic relations, structural style, whole-
rock chemistry, and geochronology.
MAGMA TIC ROCKS OF THE NORTHEASTERN
WYOMING PROVINCE
The Beartooth Mountains and Bighorn Moun-
tains are composed dominantly of late Archean
granitoids with inclusions of older metamorphic
rocks. The petrology, geochemistry, and geochrono-
logy of the Beartooth Mountains are described in
detail by Mueller and others (1985, and references
therein), Wooden and others (1988a,b; Wooden and
Mueller, 1988) and of the Bighorn Mountains by
Barker and others (1979) and Arth and others
(1980). The important crust-forming events in this
magmatic province are summarized below.
The eastern Beartooth Mountains contain a
voluminous suite of late Archean calcalkaline
granitoids and andesitic amphibolites that intrude a
sequence of middle to early Archean metasupra-
crustal rocks. The metasupracrustal rocks include
pelitic schists, quartzites, iron formation, mafic
The northern Wyoming Province: Contrasts in Archean crustal evolution 285
Table 1. Characteristics of the lithologic units of the North Snowy Block.
Unit Metamorphic Grade Structural Style Isotopic Ages
Supracrustal- Upper amphibolite Transposition foliation, intrafolial 3.4 Ga Rb-Sr whole rock
Migmatite 650-700C isoclinal folds on injected migmatites
Complex (garnet-biotite)
Pine Creek Mid-Upper Amphibiolite Isoclinal folding on all scales 3.2 Ga Sm-Nd chondritic
Nappe Complex 600-650C model age on amphibolite
(garnet-biotite) 2.5 Ga U-Pb shpene
Trandhjemite- Epidote-oligoclase zone Ductile shear zone, blastomylonitic, 3.55 and 3.26 Ga Sm-Nd
Amphibolite 500C passive flow folds chrondritic model age
Complex (coexisting albite- 3.4 Ga Rb-Srwhole rock
oligoclase)
Davis Creek Greenschist facies Phy llitic, local isoclinal fold, late- Not determined
Schist (chlor-musc-albite-qtz) stage kinks
Augen Gneiss Greenschist facies Granitic augen gneiss 2.55 Ga Rb-Sr whole-rock
Sill isochron
Paragneiss Unit Upper amphibolite Anastomosing shear zones, high 2.8 Ga Rb-Sr whole-rock
700C degree of internal tectonic mixing errorchron on quartzo-
(garnet-biotite) feldspathic gneiss in
Yankee Jim Canyon
Mount Cowen Greenschist facies Granitic augen gneiss 2.74 Ga Rb-Sr whole-rock
Augen isochron
granulites, and meta ultramafic rocks. They were
metamorphosed in the granulite facies at peak
metamorphic pressures of 5-6 kb and temperatures
of 750-800C (Henry and others, 1982; Mogk and
Henry, 1988). Minimium ages of 3.2-3.4 Ga were
determined based on U-Pb (zircon), Rb-Sr, and Nd-
Sm isotopic systematics (Wooden and others,
1988a). The metamorphism of this suite of rocks
Blacktail Tobacco Rootl Spanish Peaks
North Snowy Block Beartooth
w
Mountains ; ... - .... Ruby Ranges
,: \
E
Mountains
Figure 2. Interpretive cross section of the northern Wyoming Province showing principal lithologic associations and their
boundaries. Prominent features from east to west include: Beartooth Mountains, voluminous 2.78-2.74 Ga calcalkaline rocks
with inclusions of older, high-grade metasupracrustal rocks; North Snowy Block, linear belts of tectonically juxtaposed units
and superposed thrust sheets; Gallatin Range, ductile shear zones and mafic granulites (black pads) in tonalitic gneiss;
Spanish Peaks, magmatic rocks emplaced into granulite-grade metasupracrustal rocks, bounded on the west by a O.5-km-wide
ductile shear zone; Tobacco Root Mountains, K-feldspar-rich gneisses and granulite-grade, isoclinally folded, platform-type
metasedimentary rocks.
286
D. W. Mogk, P. A. Mueller, J. L. Wooden, and D. R. Bowes
requires burial to crustal depths of "" 20 km prior to
2.8 Ga and is interpreted as the consequence of an
episode of continental collision (Mogk and Henry,
1988). Thus, there is direct evidence for the exis-
tence of thick continental crust in the Beartooth
Mountains by the middle Archean.
Late Archean magmatism in the Beartooth
Mountains commenced with large volumes of ande-
sitic magmas at 2.79 Ga (Mueller and others, 1988).
These rocks have both tholeiitic and calcalkaline
affinities, with trace element signatures indicative
of an enriched mantle source (Mueller and others,
1983; Wooden and others, 1988a,b). These rocks
were subsequently metamorphosed in the amphibo-
lite facies. Syn-kinematic intrusion of the Long
Lake Granodiorite occurred at 2.78 Ga (Mueller and
others, 1988). These rocks have a typical calcalka-
line mineralogy and major element chemistry. The
younger Long Lake granite is the volumetrically
dominant rock type in the eastern Beartooth Moun-
tains. It is a post-kinematic unit emplaced ""2.74
Ga (Mueller and others, 1988). This granitic series
actually varies from a high silica tonalite to a typi-
cal granite (Mueller and others, 1985; Wooden and
others, 1988a) (Figure 3). Magmatic rocks in the
central Beartooth Mountains (Lake Plateau area)
also include calcalkaline andesitic amphibolites, but
the granitoids are most commonly granites to grano-
diorites; tonalites are subordinate (Richmond and
Mogk, 1986). Overall, the granites of the central
Beartooth Mountains appear to be contemporaneous
with the Long Lake granite (2.75 Ga; LaFrenz and
others, 1986). In aggregate, the magmatic rocks of
the eastern and central Beartooth Mountains are
interpreted as products of a magmatic episode that
developed as a result of subduction-like processes in
the late Archean (Mueller and others, 1985, Mueller
and Wooden,1988; Wooden and others, 1988).
The southern Bighorn Mountains also contain
extensive exposures of late Archean magmatic
rocks. Barker and others (1979) and Arth and
others (1980) distinguished two discrete magmatic
events. The first event (E
l
) produced early trond-
hjemitic melts, followed by deformation and meta-
morphism, syn-kinematic intrusions of tonalitic
magmas, and very late syn-kinematic intrusion of
andesitic magmas nearly identical to those of the
Beartooth area (Mueller and others, 1988). A com-
posite Rb-Sr whole-rock age for the El magmatism
is 3007 68 Ma. The second magmatic event (E
2
)
consisted of syn-kinematic intrusion of a trond-
hjemitic to leucogranodioritic pluton that sharply
55
62
% Si02
65
o
o
o 0
o
GRANODIORITE
LLGd
70
Si02, wI %
lOW Na llG
HIGH Na LlG
75
LOW No lLG
78
Figure 3. Harker diagrams showing compositional varia-
tion of andesitic amphibolite (AA), Long Lake Grano-
diorite (LLGd), and high- and low-Na granites ofthe Long
Lake Granite (LLG),
cuts structures of El rocks, followed by syn-
kinematic injection of tonalites, granodiorite, and
granite. A Rb-Sr whole-rock isochron for this suite
of rocks gives an age of 2801 62 Ma.
NORTH SNOWY BLOCK
The North Snowy Block occupies the north-
western margin of the Beartooth Mountains. It is
an Archean mobile belt characterized by tectonic
juxtaposition of both meta-igneous and metasedi-
mentary rocks, and it marks the transition between
magmatic rocks of the northern Wyoming Province
and metasupracrustal rocks to the west. The North
Snowy Block consists of four lithologically and
metamorphically distinct linear belts (from east to
west): (1) the Mount Cowen Augen Gneiss, (2) a
paragneiss unit, (3) the Davis Creek Schist, and
(4) a trondhjemitic gneiss-amphibolite unit. These
rocks are separated by faults, as indicated by mylo-
nitic contacts and abrupt discontinuities in meta-
The northern Wyoming Province: Contrasts in Archean crustal evolution 287
morphic grade. A granitic augen gneiss sill (dis-
tinctly younger than the Mount Cowen Augen
Gneiss) intruded the contact between the paragneiss
unit and the Davis Creek Schist. The trondhjemitic
gneiss-amphibolite unit is overlain by two E-verg-
ing thrust sheets (Mogk, 1984; Mogk and others,
1988a; Figure 3). These seven units are distin-
guished by abrupt discontinuities in metamorphic
grade, structural style, and isotopic age. A detailed
description of the geologic relations within the
North Snowy Block (summarized in Table 1) is
presented in Mogk (1984) and Mogk and others
(1988a).
Within the North Snowy Block, the wide
variety of rock types and abrupt discontinuities in
metamorphic grade, structural style, and isotopic
ages suggest significant tectonic displacements.
The paragneiss unit, Davis Creek Schist, and trond-
hjemitic gneiss-amphibolite complex were most
likely emplaced along transcurrent faults, as indi-
cated by subhorizontallineations. These units can-
not be readily restored to a prefaulting stratigraphy,
as would be expected if they currently represent a
series of stacked thrust sheets. In addition, the
paragneiss unit is a "broken formation," charac-
terized by a chaotic mixture of tectonically jux-
taposed, diverse rock types with a style of de-
formation of this unit analogous to tectonic mixing
associated with wrench faults. The minimum age
for juxtaposition of these units is 2.55 Ga - the age
of the protolith of the augen gneiss sill and a U-Pb
sphene age from amphibolite of the Pine Creek
Nappe that indicates the time of final cooling
through the 500-550C isotherm. Emplacement of
the two thrust sheets marks a significant change in
the tectonic style of the North Snowy Block and
appears to post-date transcurrent faulting. The
ductile shearing observed in the trondhjemitic
gneiss is absent in overlying thrust sheets. Meta-
morphic grade increases discontinuously upsection
from the Trondhjemitic Gneiss through the over-
lying Pine Creek Nappe and supracrustal-migma-
tite complex units. These numerous and varied
units of the North Snowy Block each experienced
unique geologic histories prior to amalgamation in
their present positions (= 2.55 Ga). Evolution of
this continental crust is dominated by tectonic
thickening as opposed to the magmatic growth that
occurred in the eastern and central Beartooth
Mountains. Thus, the North Snowy Block is ana-
logous, on a small scale, to accreted metamorphic
terranes observed in the western Cordillera of North
America and exposes a fundamental discontinuity
in the nature of the Archean crust in the northern
Wyoming Province (Mogk and others, 1988a).
METASUPRACRUSTAL ROCKS OF THE
NORTHWESTERN WYOMING PROVINCE
Archean rocks exposed west of the North Snowy
Block in the Gallatin, Madison, and Tobacco Root
Mountains include vast expanses of quartzofeld-
spathic gneiss with interlayered mafic granulite
and extensive platform-type metasedimentary
rocks. Gneisses are dominantly plagioclase-rich
with inter layered mafic amphibolites/granulites in
the Gallatin Range and eastern Madison Range. In
contrast, K-feldspar-rich gneiss is more common in
the western Madison Range and Tobacco Root
Mountains (Figure 4). A O.5-km-wide ductile shear
zone in the northern Madison Range separates these
two lithologic associations and demonstrates that
there are significant tectonic displacements within
the metasupracrustal region. Throughout the meta-
supracrustal region, metamorphism is dominantly
in the upper amphibolite to granulite facies. Iso-
clinal folding on all scales and nappe emplacement
charcterize the regional structural style. Plutonic
rocks are restricted to one narrow belt in the north-
ern Madison Range and are distinct in their com-
positions and ages compared with magmatic rocks of
the eastern and central Beartooth Mountains (Mogk
and others, 1988b). Greenstone belt assemblages
similar to those observed in the southern Wyoming
Province have not been observed in the meta-
supracrustal region. The characteristics of the
Archean exposures in the metasupracrustal region
are briefly discussed below.
Northern Gallatin Range
The northern Gallatin Range consists primarily
of a variety of migmatitic, quartzofeldspathic
gneisses with subordinate layers of orthoquartzite,
pelitic schists, and meter-scale boudins of mafic
granulite. The Archean rocks were mapped by
Spencer and Kozak (1975) and the petrology was
described in more detail by May (1985). The
gneisses exhibit compositional layering on a cm-
scale defined by wide variation of modal ratios of
plagioclase/K-feldspar, and by color index. This
fine-scale layering and the association with quartz-
288
D. W. Mogk, P. A. Mueller, J. L. Wooden, and D. R. Bowes
Plag
Figure 4. Ternary quartz-K-feldspar-plagioclase diagram
showing compositional range of quartzofeldspathic
gneisses in northern Wyoming Province. Rock units:
circles, gneisses of northern Gallatin Range; stars,
remobilized leucosome in gneisses of northern Gallatin
Range; triangles, gneisses of northern Madison Range east
of the shear zone; and squares, gneisses of northern
Madison Range west of the shear zone. Field lines based on
lUGS classification for igneous rocks for reference only.
(Sources: May, 1985; Salt, 1987; Mogk, unpubl. data.)
ites and pelitic schists suggest that these gneisses
are derived from a supracrustal protolith. The
modal mineralogy of the gneisses is dominantly
tonalitic, with individual layers showing extremely
quartz-rich compositions. Remobilized leucosomes
in migmatitic parts of the gneiss tend to be more
potassic than the host tonalitic gneisses (Figure 4).
Representative chemical analyses of this suite of
rocks are presented in Table 2. The dominance of
Na-rich gneisses in this area is illustrated in Figure
5. Metamorphism is in the upper amphibolite to
hornblende-granulite facies. The dominant meta-
morphic assemblage in mafic layers is clinopyro-
xene-garnet-plagioclase-hornblende-quartz scapo-
lite. Garnet-clinopyroxene temperatures in mafic
granulites (Dahl, 1979, 1980; Ellis and Green, 1979)
yield temperatures of 700-750C (Mogk and Henry,
1988). Garnet-biotite pairs in the gneissic and
schistose layers also yield temperatures of 700-
750C. Sillimanite is the stable aluminosilicate
polymorph in schists. Numerous layers of both
3.0
2.5
2.
..
0",
....
'"

1.5
..
'"
..

1.0



0.5

.. ......
.. - t

4
.. ..
50 55 60 65 70 75 80
WI% Si0
2
Figure 5. Geochemical variation diagram showing
compositional differences among quartzofeldpsathic
gneisses in the northern Wyoming Province. Rocks from:
triangles, Tobacco Root Mountains; circles, northern
Gallatin Range; squares, northern Madison Range. Rocks
with < 55 wt% Si0
2
are amphibolites and hornblende
gneisses.
concordant and discordant remobilized granitic
leucosome occur within the quartzofeldspathic
gneisses. These are interpreted as anatectic melts,
produced in response to vapor-present melting
during granulite facies metamorphism (Mogk and
Salt, 1986). Granulite-facies assemblages are over-
printed by amphibolite-facies assemblages. Intra-
folial isoclinal folds are present in the gneisses and
schists, and local detachment of nappes occurred.
High-temperature ductile shear zones cut the
gneisses and, in many cases, deformation is asso-
ciated with remobilization of the granitic leucosome.
Northern Madison Range
The northern Madison Range is lithologically
very similar to rocks of the Gallatin Range and is
included in the area mapped by Spencer and Kozak
(1975). The petrology was studied in more detail by
Salt (1987). The migmatitic gneiss-granulite asso-
ciation is present in both the Gallatin and Madison
Ranges. Garnet-clinopyroxene and garnet-biotite
temperatures are ""680-720C, and pressures de-
termined using the garnet-aluminosilicate-plagio-
Si0
2
Ti02
Al
2
0
3
Fe203
FeO
MnO
MgO
CaO
Na20
K
2
0
P
2
0
5
H
2
0+
O
2
Ba
Ce
Co
Cr
Cu
Ga
La
Ni
Pb
Rb
Sr
Th
U
Y
Zn
Zr
The northern Wyoming Province: Contrasts in Archean crustal evolution
Table 2. Representative chemical analyses of Archean rocks, northern Wyoming Province.
G1
70.50
0.26
15.42
0.72
1.33
0.02
0.91
2.93
4.96
1.61
0.09
0.82
0.15
451
31
25
15
10
19
14
10
22
54
505
5
1
5
45
112
G
G2
71.38
0.23
15.60
0.53
1.01
0.02
0.63
2.53
5.17
2.14
0.07
0.70
0.18
680
32
25
4
9
19
18
5
23
56
465
4
o
4
39
113
G
Northern Gallatin Range: Gallatin River Canyon
G3
50.02
1.27
14.31
2.13
10.75
0.22
7.09
10.75
2.31
0.25
0.09
1.47
0.38
95
14
69
179
247
20
3
119
7
7
115
1
2
27
130
81
A
G5
71.08
0.25
15.64
0.69
1.21
0.04
0.77
2.59
4.68
1.97
0.10
0.34
0.22
820
42
23
6
9
18
22
3
18
53
491
3
1
5
41
129
G
G6
49.17
0.90
14.15
2.74
9.26
0.17
7.49
10.88
2.51
0.85
0.06
1.17
0.31
114
8
62
262
123
19
4
112
10
9
103
o
2
18
103
53
A
G7
49.38
0.16
16.32
0.48
7.84
0.14
12.89
9.34
0.83
0.15
1.01
1.69
0.16
28
o
58
408
20
13
o
169
12
5
56
2
2
6
78
22
A
G8
66.43
0.36
12.51
1.10
7.60
0.18
5.10
0.68
0.94
2.20
0.05
1.74
0.66
741
43
38
302
35
17
24
141
10
85
60
13
3
22
68
101
G
G9
77.95
0.22
4.17
0.74
8.88
0.41
3.15
2.59
0.00
0.07
0.04
0.68
0.53
198
27
45
177
47
7
12
72
1
3
12
2
2
14
50
45
G
All analyses done using standard XRF techniques: glasses used for major elements, pressed powders used for trace
elements. FeO, H
2
0, and CO
2
from wet chemical techniques. A, amphibolite; G, gneiss; Q, quartzite; S, schist.
(continued on following pages)
289
290 D. W. Mogk, P. A. Mueller, J. L. Wooden, and D. R. Bowes
Table 2 (continued)
Northwestern Madison Range: Beartrap Canyon of Madison River
BC1 BC2 BC3 BC4 BC5 BC6 BC7
Si0
2
50.62 50.76 64.47 47.96 74.39 74.34 71.76
Ti0
2
1.19 0.89 0.84 1.29 0.21 0.10 0.23
Al
2
0
3
12.78 13.76 15.99 14.07 13.79 13.87 14.56
Fe203
2.86 2.35 1.24 1.39 0.42 0.27 0.75
FeO 11.73 9.18 3.89 10.96 0.86 0.75 1.18
MnO 0.23 0.18 0.08 0.19 0.00 0.02 0.02
MgO 5.25 6.88 2.17 7.10 0.59 0.69 0.50
CaO 9.41 9.83 3.22 11.49 1.71 1.47 1.73
Na20 2.02 2.41 3.85 1.94 3.67 3.48 3.60
K
2
0 0.67 0.45 1.95 0.33 3.22 3.70 3.52
P
2
0
5
0.11 0.07 0.06 0.10 0.04 0.03 0.06
H
2
O+
0.52 1.17 0.53 0.66 0.25 0.53 0.71
O2
0.39 0.32 0.19 0.40 0.30 0.19 0.68
Ba 303 188 665 168 1379 1553 1153
Ce 32 14 47 18 69 21 65
Co 71 60 43 74 33 33 35
Cr 53 182 194 184 6 16 12
Cu 177 129 27 180 11 15 16
Ga 20 18 24 21 17 17 19
La 15 4 26 11 36 10 33
Ni 62 85 97 117 17 14 13
Pb 7 7 23 4 22 26 27
Rb 26 12 86 12 93 77 97
Sr 126 127 146 217 380 246 230
Th 1 2 9 0 14 5 19
U 1 0 3 2 2 1 2
Y 34 24 21 21 4 4 8
Zn 129 100 126 110 28 21 45
Zr 121 69 193 88 129 56 120
A A G A G G G
All analyses done using standard XRF techniques: glasses used for major elements, pressed powders used for
trace elements. FeO, H
2
0, and CO
2
from wet chemical techniques.
A, amphibolite; G, gneiss; Q, quartzite; S, schist.
TR1
Si0
2
96.60
Ti0
2
0.05
Al
2
0
s
1.32
Fe20S 0.09
FeO 0.00
MnO 0.00
MgO 0.00
CaO 0.09
Na20 0.01
K
2
0 0.54
P
2
0
5
0.03
H
2
0+ 0.50
O
2
0.30
Ba 341
Ce 45
Co 80
Cr 34
Cu 344
Ga 1
La 24
Ni 31
Pb 6
Rb 9
Sr 10
Th 12
U 2
Y 6
Zn 8
Zr 322
Q
The northern Wyoming Province: Contrasts in Archean crustal evolution
TR2
98.75
0.20
0.36
0.08
0.00
0.00
0.00
0.08
0.00
0.16
0.00
0.19
0.19
174
11
162
7
211
1
5
4
5
3
7
2
2
1
4
25
Q
Table 2 (continued)
Tobacco Root Mountains: Ramshorn Creek Area
TR4
63.72
0.50
16.74
1.16
2.79
0.07
2.05
4.93
4.70
1.27
0.15
1.10
0.24
375
37
36
35
26
21
14
31
11
42
464
2
1
13
63
138
G
TR5
68.87
0.25
16.72
0.82
1.08
0.01
0.85
3.72
4.93
1.09
0.06
0.68
0.22
345
21
27
5
11
20
11
7
16
49
478
4
2
5
31
89
G
TR6
67.02
0.27
17.10
0.62
1.40
0.02
1.16
4.23
5.01
1.13
0.06
0.63
0.27
329
14
22
28
16
21
8
18
16
49
489
2
2
7
50
66
G
TR7 TR8a TR8b
52.64 74.21 74.44
0.57 0.03 0.05
15.31 13.58 13.65
2.13 0.18 0.15
6.76 0.18 0.24
0.17 0.00 0.01
7.53 0.08 0.16
9.02 0.99 1.60
3.30 2.29 2.78
0.43 6.94 5.06
0.04 0.02 0.05
1.37 0.70 0.76
0.34 0.26 0.64
118 1828 1349
24 13 11
61 35 42
259 0 0
14 10 5
20 16 16
12 4 7
192 6 0
6 42 34
4 143 105
259 259 251
000
122
17 3 2
92 7 11
98 88 64
A S S
TR9
67.59
0.43
15.69
1.40
2.21
0.06
1.46
4.06
3.92
1.37
0.12
1.14
0.28
233
45
39
18
4
20
18
21
14
54
216
5
1
30
52
155
G
291
TRlO
68.78
0.26
16.54
1.07
1.08
0.02
1.12
3.44
4.56
1.49
0.08
1.08
0.22
457
25
31
8
15
20
11
15
13
48
421
1
1
4
43
90
G
All analyses done using standard XRF techniques: glasses used for major elements, pressed powders used for trace elements.
FeO, H
2
0, and CO
2
from wet chemical techniques. A, amphibolite; G, gneiss; Q, quartzite; S, schist.
(continued)
292
TRll
Si0
2
48.49
Ti0
2
0.63
Al
2
0
3
13.46
Fe203 3.18
FeO 7.17
MnO 0.31
MgO 9.26
CaO 11.63
Na20 1.68
K
2
0 0.40
P
2
0
5
0.06
H
2
0+ 1.78
O
2
0.37
Ba 139
Ce 4
Co 67
Cr 798
Cu 49
Ga 15
La 4
Ni 126
Pb 4
Rb 7
Sr 127
Th 0
U 4
Y 17
Zn 128
Zr 51
A
D. W. Mogk, P. A. Mueller, J. L. Wooden, and D. R. Bowes
Table 2 (continued)
Tobacco Root Mountains: Ramshom Creek Area
TR12 TR13 TR14 TR14a TR15 TR16 TR17
44.39 52.88 76.03 75.22 67.22 73.16 51.10
0.32 0.34 0.32 0.30 0.88 0.39 1.93
18.58 6.39 11.54 11.53 13.53 13.26 23.49
1.92 1.57 1.01 1.49 2.09 0.92 1.48
7.17 8.15 0.78 0.88 2.98 1.04 4.17
0.22 0.27 0.03 0.02 0.07 0.04 0.10
11.57 19.78 0.21 0.20 1.08 0.32 2.51
11.63 7.17 0.96 1.10 3.04 1.64 5.12
1.33 0.71 3.04 2.74 3.11 2.80 4.61
0.29 0.10 4.42 4.88 3.91 5.22 3.04
0.02 0.04 0.03 0.04 0.29 0.08 0.05
2.15 2.36 0.68 0.67 0.86 0.64 1.53
0.18 0.18 0.02 0.13 0.02 0.08 0.05
86 47 666 1291 1466 1369 1444
4 3 142 130 153 126 110
70 70 46 57 35 40 66
217 1499 3 1 17 9 273
14 10 5 4 19 5 52
15 9 16 18 19 17 29
3 13 71 70 77 68 63
324 291 5 5 11 5 150
5 3 29 38 33 32 21
5 3 163 211 147 186 93
91 13 55 102 127 109 718
o 1 23 18 17 14 21
1 1 2 342 1
7 9 N M
82 91 18 37 82 50 115
31 45 281 228 493 284 345
A A S S G S A
TR18
51.18
0.73
13.63
2.91
6.65
0.13
8.31
8.23
3.39
1.19
0.08
1.51
0.21
342
16
56
793
46
17
10
230
10
45
185
1
2
20
86
88
A
All analyses done using standard XRF techniques: glasses used for major elements, pressed powders used for
trace elements. FeO, H
2
0, and CO
2
from wet chemical techniques.
A, amphibolite; G, gneiss; Q, quartzite; S, schist.
The northern Wyoming Province: Contrasts in Archean crustal evolution 293
clase-quartz barometer (Ghent, 1976) are ""7-8 kb
(Mogk and Henry, 1988).
In addition, a series of mesozonal intrusive
rocks is present in the central part of the northern
Madison Range (Salt and Mogk, 1985). The intru-
sive rocks include a 3.2 Ga suite of monzodiorites
and granodiorites, a 2.9 Ga suite of biotite tonalites,
and a 2.6 Ga suite of granites (Mogk and others,
1987, 1988b). The monzodiorite-granodiorite se-
quence occurs as syn-kinematic, sill-like intrusions
in high-grade tonalitic gneisses. The intrusive tona-
lite suite is also syn-kinematic but clearly exhibits
cross-cutting relations with the monzodioritic suite.
The granitic rocks are syn- to post-kinematic and
are spatially restricted to the ductile shear zone that
separates the plagioclase and K-feldspar-rich
gneisses. Intrusive rocks of the northern Madison
Range are volumetrically subordinate to the host
quartzofeldspathic gneisses. The protracted intru-
sive history of these rocks records episodic crust-
forming events not recognized elsewhere in the
northern Wyoming Province.
The 500-m-wide, NE-striking, steeply SE-dip-
ping, ductile shear zone is continuous across Arch-
ean exposures in the northern Madison Range. It is
characterized by anastomosing bands of mylonite
interleaved with meter-scale macrolithons of
relatively undeformed gneisses and amphibolite.
Deformational textures occur on the meso- and
microscopic scale and include incipient myloni-
tization, augen textures, blastomylonitic texture,
pronounced reduction of grain size due to dynamic
recrystallization, quartz ribbons, S-C mylonitic
surfaces, and local development of ultramylonites.
This zone defines the break in lithologies and meta-
morphic grade between the eastern and western
Madison Range (Salt, 1987).
Archean rocks to the west of the ductile shear
zone include K- rich gneisses (e.g., Figure 4),
sillimanite-bearing pelitic rocks, interlayered mafic
amphibolites/granulites, and post-kinematic gra-
nites adjacent to the shear zone. Representative
chemical analyses of this suite of rocks are pre-
sented in Table 2. Calculated temperatures of
metamorphism of this suite of rocks are somewhat
higher than those east of the shear zone, "" 750-
800C. Pressures, however, are somewhat lower,
5.5-7 kb (Mogk and Henry, 1988). The apparent
break in lithology and metamorphic grade across
this shear zone indicates significant tectonic dis-
placement occurred between these discrete blocks of
Archean rocks. Down-dip lineations suggest that
juxtaposition occurred along thrust faults, with
higher pressure rocks to the east emplaced over
lower pressure rocks to the west.
Tobacco Root Mountains
Archean exposures of the Tobacco Root Moun-
tains consist of dominantly quartzofeldspathic
gneisses and interlayered mafic amphibolites/
granulites in the eastern part of the range, and a
supracrustal assemblage of marble, pelitic schist,
iron formation, quartzite, amphibolite, and quartzo-
feldspathic gneiss in the western part of the range.
K-rich gneisses, subordinate plagioclase-rich
gneisses, and interlayered mafic amphibolites/
granulites of the western Madison Range occur in
almost continuous exposures through the eastern
Tobacco Root Mountains. Vitaliano and others
(1979) mapped and described Archean rocks in this
area. Compositional data for examples of the
gneisses, schists, quartzites, and metabasites are
presented in Table 2; the dominance of K-rich
gneisses in this area is illustrated in Figure 5.
The structure of these Archean rocks is charac-
terized by isoclinal folding observed on all scales.
Regional-scale isoclinal folds, overturned to the
east, are easily traced along the metasupracrustal
marker units such as marbles and iron formations.
Attenuation of lower limbs of isoclinal folds and
detachment of nappe structures is locally evident.
Metamorphism is in the upper amphibolite to
granulite facies. Diagnostic metamorphic assem-
blages include the following:
olivine-diopside-scapolite-phlogopite-calcite-
dolomite (marble);
garnet-biotite-kyanite-sillimanite-plagioclase-
quartz (pelitic schist);
garnet-orthopyroxene-quartz-magnetite-retro-
grade grunerite (banded iron formation);
olivine-orthopyroxene-spinel (ultramafites);
and
garnet-clinopyroxene-plagioclase-hornblende
orthopyroxene (metabasites).
Preliminary garnet-biotite and garnet-clinopyro-
xene geothermometry (Mogk, unpubl. data) are
"" 700-780C. Pressure estimates based on the
garnet-aluminosilicate-quartz-plagioclase barome-
ter are"" 7 kb.
A second metamorphic episode (Vitaliano and
others, 1979) resulted in re-equilibration in the
upper amphibolite or almandine-amphibolite facies.
294
D. W. Mogk, P. A. Mueller, J. L. Wooden, and D. R. Bowes
This overprinting metamorphism may simply be a
later phase of a single orogenic cycle, based on
patchy overgrowths of lower grade amphiboles and
micas on high-grade assemblages and the lack of a
younger penetrative fabric. Alternatively, a second
metamorphic event may have occurred, based on
preserved chill margins observed on some of the
metamorphosed dikes (H. L. James, pers. commun.,
1988). It is unlikely that this metamorphism is the
result of thermal effects of a Middle Proterozoic
event ("" 1600 Ma; Giletti, 1966) that reset K-Ar
isotopic clocks in this area. Effects of possible
formation of a 1.8 Ga gneiss dome in the adjacent
Highland Range (O'Neill and others, 1988) have not
yet been documented.
DISCUSSION
The Archean crust of the northern Wyoming
Province evolved by means of both magmatic and
tectonic thickening. Recent progress toward estab-
lishing a geochronologic framework, characterizing
whole-rock geochemistry, and quantifying the phy-
sical conditions of metamorphism have allowed new
interpretations of the timing and mechanisms of
crustal evolution in this area. The numerous crus-
tal components described above developed in
distinctly different geologic environments and fol-
lowed separate geologic histories prior to their final
emplacement. We interpret the present configura-
tion of these crustal segments as the result of
tectonic juxtaposition of metamorphic and mag-
matic associations in the late Archean. The critical
lines of evidence in support of this hypothesis are as
follows.
The two major regions are separated by a zone
containing the North Snowy Block mobile belt. The
dominantly magmatic region lies east of this boun-
dary zone and the dominantly metasupracrustal
region lies to the northwest. Within these regions
further subdivisions are established based on geo-
chronologic constraints and petrologic and geo-
chemical characteristics of different lithologic
packages.
Thick continental crust was established in the
Beartooth Mountains by middle Archean times.
Granulite-facies metamorphism of platform-type
sediments> 2.8 Ga, and probably during the Middle
Archean (Henry and others, 1982; Wooden and
Mueller, 1987), required metamorphism at crustal
depths of "" 20 km. This suggests that the crust may
have been at least twice as thick and is interpreted
as the result of tectonic thickening. This thick
continental crust was subsequently the site where a
magmatic province developed in the late Archean.
Emplacement of rocks of andesitic composition
across this magmatic province, as well as the calc-
alkaline plutonic rocks of the Beartooth Mountains,
is analogous in many ways to development of
magmatic arcs in Phanerozoic settings. Arguments
based on geochemical and isotopic systematics
presented by Wooden and Mueller (1988), Wooden
and others (1988), and Mueller and Wooden (1988)
provide strong evidence of crustal recycling through
subduction-like processes in the late Archean. Simi-
lar plutonic rocks exposed in the Bighorn Mountains
are somewhat older than the Beartooth rocks.
Overall, these magmatic rocks comprise the bulk of
this portion of the Wyoming Province.
The North Snowy Block mobile belt marks a
major discontinuity in the nature of the continental
crust in the northern Wyoming Province. The late
Archean, calcalkaline plutonic rocks east of the
mobile belt have no temporal or compositional equi-
valents west of the mobile belt; metasupracrustal
rocks are present in the mobile belt but are demon-
strably allochthonous. Individual units within the
North Snowy Block each have separate and distinct
geologic histories and were emplaced into their
present structural position> 2.55 Ga, after they had
attained their peak metamorphic mineral assem-
blages and fabrics.
The metasupracrustal region west of the North
Snowy Block evolved primarily by tectonic thick-
ening, with only minor magmatic additions to the
crust in Archean times. The dominant rocks in this
region are a variety of quartzofeldspathic gneisses
interpreted to have supracrustal protoliths because
of fine-scale compositional layering and intercala-
tion with quartzites, pelitic schists, banded iron
formation, marbles, and conformable metabasites.
These rocks must have been buried to mid-crustal
levels of "" 20-25 km, and tectonic thickening is
interpreted as the most reasonable mechanism to
accomplish this (Mogk and Henry, 1988). The
occurrence of E-verging, overturned isoclinal folds
and local detachment of nappes is consistent with an
interpretation of overall crustal shortening in this
region.
There are at least two distinct sequences of
rocks within the metasupracrustal region. Tonalitic
gneisses associated with basaltic amphibolitesl
granulites dominate the eastern area; granitic
The northern Wyoming Province: Contrasts in Archean crustal evolution 295
gneisses associated with platform-type metasedi-
mentary rocks characterize the western area. These
two sequences are separated by a 0.5-km-wide
ductile shear zone in the northern Madison Range.
A break in metamorphic grade across this shear
zone is established by mineral assemblages and
geothermobarometers. Other tectonic boundaries,
based on chronologic or petrologic discontinuities,
may also be present. For example, numerous ultra-
mafic pods, scattered throughout the Tobacco Root
Mountains and adjacent Ruby Range, are inter-
preted as diapiric serpentinites probably emplaced
along faults and metamorphased to olivine-ortho-
pyroxene assemblages rather than magmatic rocks
(either ultramafic flows or intrusions) (Desmarais,
1980); these may mark locations of "cryptic"
sutures. An important consequence of the possibi-
lity oftectonic stacking and interleaving of Archean
basement is that correlation of similar lithologic
packages, such as the "Cherry Creek," "Pony
Gneiss," or "Dillon Gneiss," should be reserved for
rocks that have been carefully characterized in their
geochronology, geochemistry, and petrologic rela-
tions (e.g., Vitaliano and others, 1979). Much of
these data are not available at this time.
The distance and direction of tectonic transport
of the separate units in the metasurpracrustal re-
gion remains problematic in the absence of strati-
graphic, paleomagnetic, and paleontologic data. It
is possible that individual sequences merely repre-
sent telescoped segments of a once continuous basin.
Alternatively, each may be truly exotic with respect
to its nearest neighbors. In either case, a mechan-
ism of crustal shortening resulted in tectonic
thickening in the metasupracrustal region as op-
posed to magmatic thickening to the east.
The rocks of the metasupracrustal region can-
not simply be a continuous sequence deposited in an
ensialic basin, with coarse clastic material deposited
in the east, adjacent to the Beartooth Mountains,
and more distal sedimentary facies deposited in the
Tobacco Root Mountains to the west as suggested by
Vitaliano and others (1979). Intrusion of 3.2 Ga
monzodiorite and granodiorite into tonalitic
gneisses, as well as compositions of these gneisses,
preclude derivation of them from the 2.78-2.74 Ga
plutonic rocks of the Beartooth Mountains.
Magmatic rocks of the northern Madison Range
provide further insight into the timing and style of
crustal evolution. The three distinct magmatic
events - monzodiorite-granodiorite at 3.2 Ga, bio-
tite tonalites at 2.9 Ga, and granites at 2.55 Ga -
clearly show that this crust was mobilized several
times during the middle to late Archean. None of
these intrusive rocks have coeval compositional
equivalents in the Beartooth Mountains to the east,
indicating separate magma-generating processes.
Numerous Archean components of the northern
Wyoming Province were assembled as either mag-
matic or metamorphic entities at the end of the
Archean (== 2.55-2.6 Ga). In an E-W cross section
across this province, adjacent rock packages have
different origin, as well as different metamorphic
grade and structural styles. Because these rock
sequences were amalgamated after either intrusion
or metamorphism, they are referred to as magmatic
or metamorphic terranes. Consequently, the dif-
ferent sequences of rocks may be analogous on a
small scale to accreted terranes of the Cordillera of
western North America (e.g., Howell and Jones,
1983). The 2.55-2.6 Ga age is significant because it
is the age of (1) intrusion of the augen gneiss sill in
the North Snowy Block, (2) the U-Pb sphene-age in
amphibolite of the Pine Creek Nappe, and (3) gra-
nites in the ductile shear zone in the northern
Madison Range. These events mark the final tec-
tonic stabilization of this crust in the late Archean.
Recognition of discrete terranes in Archean
basement of southwestern Montana has important
implications for Laramide (and younger) structural
evolution. The crystalline basement is not homo-
geneous in either composition or mechanical pro-
perties and, therefore, was unlikely to behave as a
mechanically isotropic body during Laramide-style
uplift as suggested by Stearns (1978). A further
implication is that Laramide structures do not
simply reactivate "basement structures," but rather
they may selectively reactivate the boundaries
between these Archean lithologic associations. The
best example of this is in the Yellowstone Valley
where Laramide and Basin and Range faulting
reactivated the western margin of the North Snowy
Block. Other examples of reactivated terrane boun-
daries may include the range-bounding faults along
the northwestern margin of the Gallatin and Madi-
son Ranges and the western side of the Tobacco Root
Mountains. In addition, the Cretaceous Tobacco
Root Batholith occupies the transition zone between
the quartzofeldspathic gneisses of the eastern
Tobacco Root Mountains and the platform-type
metasedimentary sequence of the western part of
the range, where this zone intersects the Proterozoic
Spanish Peaks Fault that was reactivated during
the Laramide Orogeny. Although the concept of
296 D. W. Mogk, P. A. Mueller, J. L. Wooden, and D. R. Bowes
reactivated terrane boundaries is purely hypo-
thetical at this point, integrated studies of the
geology of the basement rocks and analysis of
Laramide structures should be able to test this
hypothesis (e.g., McBride and others, this volume).
CONCLUSIONS
The northern part of the Wyoming Province is
composed of numerous lithologic packages that have
separate origins and geologic histories. The eastern
part of the province consists of late Archean
magmatic rocks that may have formed through sub-
duction-related processes. The western part of the
province consists of a variety of metasupracrustal
rocks that exhibit distinct lithologic, petrologic, and
geochemical characteristics. Deep burial of the pro-
toliths of these supracrustal rocks probably occurred
by means of crustal shortening and tectonic thick-
ening. Tectonic juxtaposition of these diverse rock
suites occurred after establishment of peak meta-
morphic conditions or after crystallization of the
magmatic rocks. Thus, the northern Wyoming
Province is a mosaic of magmatic and metamorphic
terranes amalgamated during the late Archean.
REFERENCES
ARTH. J. G., F. BARKER, and T. W. STERN, 1980, Geo-
chronology of Archean gneisses in the Lake Helen area,
southwestern Bighorn Mountains, Wyoming: Precam-
brian Geology, v. 11, p. 11-22.
BARKER, F., J. G. ARTH, and H. T. MILLARD, Jr., 1979,
Archean trondhjemites of the southwestern Bighorn
Mountains, Wyoming--a preliminary report; pp. 401-414 in
F. Barker (ed.), Trondhjemites, Dacites, and Related
Rocks: Elsevier, Amsterdam, The Netherlands, 659 p.
CLARK, M., and D. MOGK, 1985, Development and sig-
nificance of the Blacktail Mountains Archean meta-
morphic complex, Beaverhead County, Montana [abstract]:
Geological Society of America Abstracts with Programs, v.
17,p.212.
DAHL, P. S., 1979, Comparative geothermometry based on
major element and oxygen isotope distributions in Pre-
cambrian metamorphic rocks from southwestern Montana:
American Mineralogist, v. 64, p. 1280-1293.
__ , 1980, The thermal-compositional dependence of
Fe +2_Mg distributions between coexisting garnet and py-
roxene - Applications to geothermometry: American
Mineralogist, v. 65, p. 852-866.
DESMARAIS, N. R, 1980, Metamorphosed Precambrian
ultramafic rocks in the Ruby Range, Montana: Precam-
brian Research, v. 16, p. 67-101.
ELLIS, D. J., and D. H. GREEN, 1979, An experimental
study of the effect of Ca upon garnet-clinopyroxene ex-
change equilibria: Contributions to Mineralogy and Petro-
logy, v. 71, p. 13-22.
ERSLEV, E. A., 1983, Pre-Beltian Geology of the South-
ern Madison Range, Southwestern Montana: Montana
Bureau of Mines and Geology, Memoir 55, 26 p.
FERRY, J. M., and F. S. SPEAR, 1978, Experimental cali-
bration of the partitioning of Fe and Mg between biotite
and garnet: Contributions to Mineralogy and Petrology, v.
66, p. 113-117.
FOOSE, R M., D. U. WISE, and G. S. GARBARINI, 1961,
Structural geology of the Beartooth Mountains, Montana
and Wyoming: Geological Society of America Bulletin, v.
72, p. 1143-1172.
GARIHAN, J. M., 1979, Geology and structure of the cen-
tral Ruby Range, Madison County, Montana: Geological
Society of America Bulletin, Part II, v. 90, p. 695-788.
GHENT, E. D., 1976, Plagioclase-garnet-AI
2
Si0
5
-quartz -
A potential geobarometer-geothermometer: American
Mineralogist, v. 61, p. 309-340.
GILE'ITI, B., 1966, Isotopic ages from southwestern Mon-
tana: Journal of Geophysical Research, v. 71, p. 4029-4036.
GUY, R E., and A. K. SINHA, 1988, Geochronologic and
isotopic studies in the Yankee Jim and Lamar Canyon
areas, Montana and Wyoming; pp. 53-68 in S. E. Lewis and
R B. Berg (eds.), Precambrian and Mesozoic Plate
Margins: Montana Bureau of Mines and Geology, Special
Publication 96, 195 p.
HENRY, D. J., P. A. MUELLER, J. L. WOODEN, J. L.
WARNER, and R LEE-BERMAN, 1982, Granulite grade
supracrustal assemblages of the Quad Creek area, eastern
Beartooth Mountains, Montana; pp. 147-159 in P. A.
Mueller and J. L. Wooden (eds.), Precambrian Geology of
the Beartooth Mountains, Montana and Wyoming:
Montana Bureau of Mines and Geology, Special Publication
84,167p.
HOWELL, D. G., and D. L. JONES, 1983, Tectonostrati-
graphic terrane analysis and some terrane vernacular:
Proceedings of the Circum-Pacific Terrane Conference:
Stanford University Publications, v. 18, p. 6-8.
LAFRENZ, W. B., R D. SHUSTER, and P. A. MUELLER,
1986, Archean and Proterozoic granitoids of the Hawley
Mountain area, Beartooth Mountains, Montana; pp. 79-89
in YBRA Field Conference Guidebook: Joint Field
Conference and Sympsoium (50th Anniversary Edi-
tion): Montana Geological Society, Billings, Montana,
USA.
The northern Wyoming Province: Contrasts in Archean crustal evolution 297
MAY, K, 1985,Archean Geology ofa Part of the North-
ern Gallatin Range, Southwest Montana: MS thesis,
Montana State University, Bozeman, Montana, USA, 91 p.
MOGK, D. W., 1984, Petrology, Geochemistry and
Structure of an Archean Terrane in the North Snowy
Block, Beartooth Mountains, Montana: PhD disserta-
tion, University of Washington, Seattle, Washington,
USA,440p.
__ , and D. J. HENRY, 1988, Metamorphic petrology of
the northern Archean Wyoming Province, southwestern
Montana: Evidence for Archean collisional tectonics; pp.
363-382 in W. G. Ernst (ed.), Metamorphism and Crustal
Evolution in the Western United States (Rubey Vol-
ume 7): Prentice-Hall, Englewood Cliffs, New Jersey,
USA, 1153p.
__ , and P.A. MUELLER, 1987, Tectonic evolution of
Archean basement in the Spanish Peaks area, northern
Madison Range, Montana [abstract]: Geological Society of
America Abstracts with Programs, v. 19, p. 775.
__ , and K J. SALT, 1986, Granulite-migmatite
relations in the Archean basement of the northern Galla-
tin Range, SW Montana [abstract]: Geological Society of
America Abstracts with Programs, v. 18, p. 698.
__ , P. A. MUELLER, E. WEYAND, J. L. WOODEN, and D.
R. BOWES, 1988b, Protracted magmatic history of an
Archean collision zone, northern Madison Range, SW Mon-
tana [abstract]: Geological Society of America Abstracts
with Programs, v. 20, p. A50.
__ , P. A. MUELLER, and J. L. WOODEN, 1988, Archean
tectonics of the North Snowy Block, Beartooth Mountains,
Montana: Journal of Geology, v. 96, p. 125-141.
MUELLER, P. A., and J. L. WOODEN, 1988, Evidence for
Archean subduction and crustal recyling, Beartooth Moun-
tains, USA: Geology, v. 16, p. 871-874.
__ , R. D. SHUSTER, M. A. GRAVES, J. L. WOODEN, and
D. R. BOWES, 1988, Age and composition ofa late Archean
magmatic complex, Beartooth Mountains, Montana-
Wyoming; pp. 7-22 in S. E. Lewis and R. B. Berg (eds.),
Precambrian and Mesozoic Plate Margins: Montana
Bureau of Mines and Geology, Special Publication 96, 195
p.
__ , J. L. WOODEN, D. J. HENRY, and D. R. BOWES,
1985, Archean crustal evolution of the eastern Beartooth
Mountains, Montana and Wyoming; pp. 9-20 in G. K
Czamanske and M. L. Zientek (eds.), The Stillwater
Complex, Montana - Geology and Guide: Montana
Bureau of Mines and Geology, Special Publication 92, 396
p.
__ , J. L. WOODEN, K SCHULZ, and D. BOWES, 1983,
Incompatible element rich andesitic amphibolites from the
Archean of Montana and Wyoming: Evidence for mantle
metasomatism: Geology, v. 11, p. 203-206.
O'NEILL, J. M., M. S. DUNCAN, and R. E. ZARTMAN, 1988,
An Early Proterozoic gneiss dome in the Highland Moun-
tains, southwestern Montana; pp. 81-88 in S. E. Lewis and
R. B. Berg (eds.) , Precambrian and Mesozoic Plate Mar-
gins: Montana Bureau of Mines and Geology, Special Pub-
lication 96, 195 p.
RICHMOND, D. P., and D. W. MOGK, 1985, Archean
geology of the Lake Plateau area, Beartooth Mountains,
Montana [abstract]: Geological Society of America Ab-
stracts with Programs, v. 17, p. 262.
SALT, K, 1987, Archean Geology of the Spanish Peaks
Area, Southwestern Montana: MS thesis, Montana
State University, Bozeman, Montana, USA, 81 p.
__ , and D. W. MOGK, 1985, Archean goelogy of the
Spanish Peaks area, southwestern Montana [abstract]:
Geological Society of America Abstracts with Programs, v.
17, p. 263.
SPENCER, E., and S. V. KOZAK, 1975, Precambrian evo-
lution of the Spanish Peaks, Montana: Geological Society
of America Bulletin, v. 86, p. 785-792.
STEARNS, D. W., 1978, Faulting and forced folding in the
Rocky Mountains foreland; pp. 1-38 in V. Matthews, III
(ed.), Laramide Folding Associated with Basement
Block Faulting in the Western United States: Geologi-
cal Society of America, Memoir 151,370 p.
VITALIANO, C. J., W. S. CORDUA, D. F. HESS, H. R.
BURGER, T. B. HANLEY, and F. K ROOT, 1979, Ex-
planatory text to Accompany the Geologic Map of the
Southern Tobacco Root Mountains, Madison County,
Montana: Geological Society of America, Map and Chart
Series MC-31.
WOODEN, J. L., and P. A. MUELLER, 1987, Pb isotopic
evidence for Middle and Early Archean crust in the north-
ern Wyoming Province [abstract]: Geological Society of
America Abstracts with Programs, v. 19, p.896.
__ , and P. A. MUELLER, 1988, Pb, Sr, and Nd isotopic
compositions of a suite of late Archean igneous rocks,
eastern Beartooth Mountains: Implications for crust-
mantle evolutions: Earth and Planetary Science Letters, v.
87, p. 59-72.
__ , P. A. MUELLER, and D. W. MOGK, 1988a, A review
of the geochemistry and geochronology of the Archean
rocks of the northern part of the Wyoming Province; pp.
383-410 in W. G. Ernst (ed.), Metamorphism and Crustal
Evolution in the Western United States (Rubey Vol-
ume 7): Prentice-Hall, Englewood Cliffs, New Jersey,
USA,1153p.
__ , P. A. MUELLER, and D. W. MOGK, 1988b, A review
of the Archean rocks of the Beartooth Mountains, Montana
and Wyoming; pp. 23-42 in S. E. Lewis and R. B. Berg
(eds.), Precambrian and Mesozoic Plate Margins: Mon-
tana Bureau of Mines and Geology, Special Publication 96,
195p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Evidence for the amalgamation of Archean oceanic and
continental blocks to form the Beartooth Plateau
JAMES K. MEEN
Department of Chemistry, University of Houston, Houston, Texas 77204, USA
(received August 12, 1988; revision accepted March 14, 1989)
ABSTRACT
The structure and age of the lithospere in the western Beartooth Plateau was investigated by geochemical
study of Cretaeous igneous rocks derived from the deep crust and the lithospheric mantle and by study of
exposed basement rocks. This study shows that the upper crust is dominated by "'" 2740 Ma granitoids,
similar to those widespread in the Beartooth Mountains. These have inclusions of supracrustal amphibolites
and metasediments. Sm-Nd and Pb-Pb isotopic systematics of these rocks indicate that their protoliths were
separated from the mantle only a few hundred million years prior to 2700 Ma, and no evidence of early
Archean crust was found.
Cretaeous granitoids were derived from "",3300 Ma deep crust with relatively low 87Sr/86Sr, 206PbPo4Pb,
and eNd' This crust is believed to have been similar to exposed amphibolites. Trace-element and isotopic
considerations show, however, that the source region contained little felsic material (granitoids or
metapelites) and was dominated by amphibolites. The amphibolites and source regions of the granitoids had
negative Ce-anomalies, indicating that these rocks has passed through a weathering cycle. Protoliths may
have formed as mafic-intermediate lavas in an oceanic environment with little emergent land. The present-
day upper crust may have formed when this terrane was closer to the continent or was more evolved.
The mantle in the area records an early Archean event and a Proterozioc event. These produced LILE-
rich mantle with low Rb/Sr, SmlNd, and UIPb that developed unusual isotopic compositions. Some Cretaceous
magmas were derived from these sources.
The lithosphere structure dictates a history separate from the early Archean continent of the eastern
Beartooth Mountains. I propose that the western Beartooth Plateau developed in an oceanic environment
prior to "'" 2800 Ma and was accreted to the eastern plateau during subduction at 2740-2790 Ma. The two
terranes were amalgamated by "'" 2740 Ma, as a simultaneous magmatic event affected both at that time.
INTRODUCTION
The Beartooth Block is a Laramide uplift, "'" 130
by 60 km, exposing a wide variety of Precambrian
crystalline rocks. There are four recognized struc-
tural areas in the Beartooth Mountains - the Bear-
tooth Plateau, the North and the South Snowy
Blocks, and the Stillwater Complex (Figure 1).
This study discusses evidence provided by basement
and Cretaceous volcanic rocks in the western Bear-
tooth Plateau for the structure, age, and composi-
tion ofthe lithosphere.
Much of the chemical data used in this study
are for Phanerozioc igneous rocks that are used as
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 299-311. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
299
300
J. K. Meen
McLeod -
Livingslon
110' W
IDAHO
WYOMING
10 20 30 40
Scale in km
N
I
50
[:':':'1 Cretaceous and Tertiary
extrusive Igneous rocks
Cretaceous and Tertiary
,',',' i ntrusive igneous rocks
Phanerozoic sedimentary
rocks
Precambrian
complex
D
Precambrian
rocks
Stillwater
basement
Red
Figure 1. General geological map of Beartooth Mountains, Montana-Wyoming, showing positions of Beartooth Plateau,
North and South Snowy Blocks, and Stillwater Complex. BR, Broadwater River area; LL, Long Lake area; LP, Lake Plateau.
Box in Lake Plateau shows study area. Detailed maps in Meen and Eggler (1987, 1989), and Meen (1987a).
probes of their source regions in the deeper parts of
the lithosphere. This approach is necessitated by
the lack of lower crustal or mantle xenoliths in
volcanic rocks in the study area. Such a technique is
inherently limited for several reasons. Clearly,
refractory lithospheric assemblages that have not
yielded melts cannot be identified by this technique.
Depths of derivation of liquids are imperfectly
known, and the shape and volume of source regions
are unknown. Some volume of liquid approached
chemical homogeneity due to diffusive and convec-
tive processes during anatexis, collection, rise, and
eruption or intrusion, thereby potentially averaging
derived chemical effects of different lithologies. The
scale of homogenization is unknown and depends on
liquid properties and the duration of melting and
solidification processes. Furthermore, the presence
of any major tectonic and stratigraphic features is
obscured by this approach insofar as derived melts
cannot record evidence of such features. Major con-
trasts in lithologies over wide areas can, however,
be identified by changes in the chemistry of igneous
rocks.
Mineral assemblages in source regions of mag-
mas are postulated from mineralogical and chemical
features of the igneous rocks and cannot be defined
uniquely. Consequently, a wide variety of tech-
niques are required to develop a model of the litho-
sphere. Mantle lithologies are modeled from trace-
element and isotopic compositions of mafic igneous
rocks coupled with experimental study of plausible
parageneses in mantle source regions. Crustal
lithologies are illustrated by consideration of
phenocryst assemblages and chemical composition
of granitiods, and compared with exposed crustal
rocks.
Distinct differences inferred to exist between
the lithosphere for the western Beartooth Plateau
and those in neighboring areas require that the
histories of these terranes were separate during
Archean oceanic and continental blocks form the Beartooth Plateau 301
lithosphere formation. In particular, the crust in
the western Beartooth Plateau may have stabilized
in an oceanic environment distant from evolved
continental crust such as that present in the eastern
Beartooth Plateau. These areas were not proximal
at that time (3300-3400 Ma). A major episode of
magmatism affected the entire plateau at 2550-
2750 Ma, so the terranes were amalgamated prior to
that event, plausibly = 2800 Ma.
Mogk and others (1988) inferred that the North
Snowy Block was assembled and amalgamated with
the Beartooth Plateau 2560-2740 Ma, contempora-
neous with magmatism. This episode post-dates
amalgamation of the Beartooth Plateau and in-
dicates a lengthy period of complex processes of
terrane fusion. Amalgamation of the Beartooth
Plateau may have been due to subduction as
postulated by Wooden and Mueller (1988) ,
REGIONAL GEOLOGY
Over 5000 km
2
of Archean crust are exposed in
the Beartooth Mountains (Mueller and others
1982a) and can be separated into four
regions (Figure 1), as described by Mogk and Henry
(1988). An abbreviated account is presented here.
The Stillwater Complex is a layered mafic-
ultramafic intrusion on the northeastern margin of
the Beartooth Block. Sm-Nd dating (DePaolo and
Wasserburg, 1979) and Rb-Sr dating of hornfels
(Mueller and Wooden, 1976) indicate it was in-
truded at =2700 Ma. U-Pb zircon (Nunes and
Tilton, 1971) and Sm-Nd model ages (DePaolo and
Wasserburg, 1979) suggest that protoliths of the
aureole formed at 3100-3300 Ma.
The North Snowy Block, west of the Stillwater
Complex, was studied by Reid and others (1975) and
Mogk and others (1988). The western part of the
North Snowy Block consists of six major supra-
crustal units that were amalgamated tectonically
following metamorphism, at 2740-2560 Ma (Mogk
and others, 1988). The Mount Cowen augen gneiss
occurs southeast of the units and is similar to grani-
toids farther east. Mogk and others (1988) inferred
that a suture zone separates supracrustal rocks to
the west from the principally plutonic rocks to the
east.
The Beartooth Plateau, the largest part of the
uplift, is dominated by felsic intrusions containing
blocks of older supracrustal units. The eastern part
of the plateau is composed mainly of = 2700 Ma
'tl
Z
W
10
0
-10
-20
-30
-40
-50
0
CHUR
Model ages =
2500-3100 Ma
Granites
.",,"""",,",,"'"'' Amphibolites
1000 2000 3000
T (Ma)
Figure 2. Nd isotope growth curves for basement rocks of
Lake Plateau (Meen, 1987a). Model ages referenced to
CHUR are 2500-3199 Ma.
granitoids and migmatites that intruded meta-
morphosed sediments and ultramafic-to-interme-
diate igneous rocks (Wooden and others, 1988). The
older rocks were metamorphosed to granulite facies
at = 3400 Ma (Henry and others, 1982), and some
were retrogressed to amphibolite facies (Mogk and
Henry, 1988). Apparently, part of the Beartooth
continental crust was in place at = 3500 Ma (Muel-
ler and others, 1985). The Long Lake and Broad-
water River areas in the south-central Beartooth
Mountains are also dominated by granitic rocks
with included supracrustal units. The inclusions
are dominated by = 2800 Ma andesitic amphibolites
that may have formed near a convergent plate
margin (Wooden and other, 1988).
The Lake Plateau of the central Beartooth
Mountains is dominated by granitoids with supra-
crustal enclaves. The latter range from andesitic
amphibolites to matasediments and underwent
metamorphism at =600C, 7-8 kb (Mogk and Hen-
ry,1988). Granitoids formed at 2740 Ma (274820
Ma, U-Pb zircon, Lafrenz and others, 1986; 2737
72* Ma, Pb-Pb whole-rock, Meen, 1987a); Lafrenz
and others (1986) and Meen (1987a) found that
these granitoids had their Rb-Sr systems disturbed,
an effect attributed to Laramide activity. Inclusions
are undated but granitoids and amphibolites
studied by Meen (1987a) have Sm-Nd chondritic
model ages of 2500-3100 Ma (Figure 2), and model
ages determined relative to a depleted mantle are
All errors. quoted from my work are 20 and include
MSWD weighting if necessary
302 J. K. Meen
only 200-300 million years older. This suggests that
their precursors separated from the mantle no ear-
lier than middle Archean time and argues against
the presence of significant early Archean crust in
this region (Meen, 1987a).
The South Snowy Block contains a metasedi-
mentary pile intruded by late Archean granitoids
that are especially abundant in the eastern part of
the block (Casella and others, 1982). Montgomery
and Lytwyn (1984) dated the major metamorphism
at "'" 2700 Ma and interpreted the relatively low
87Sr/86Sr of metasediments at that time to mean that
their source rocks were not separated from the
mantle much before 3000 Ma.
LITHOSPHERIC STRUCTURE IN THE WESTERN
BEARTOOTH PLATEAU
Trace-element and isotopic compositions of
basement rocks and of Cretaceous igneous rocks
have been used to construct a lithospheric cross-
section of the Lake Plateau. Meen (1987a) has des-
cribed basement rocks from the upper Boulder River
valley (Figure 1). Meen and Eggler (1987) modeled
trace-element and isotopic chemistry of subcon-
tinental upper mantle (SCUM) beneath that area
using compositions of mafic mantle-derived rocks
and inferred (Meen and Eggler, 1989) minera-
logical, major- and trace-element, and isotopic com-
positions of the deep crust.

Nd
20
30
,'0
Amphlbolites
so +-___
07 08 09
87 86
SrI Sr
Figure 3. Measured Sr-Nd isotopic compositions of base-
ment samples from the Lake Plateau (Meen, 1987a). All
samples have eNd< -25 but show large 87Sr;86Sr range.
Nature of the Upper Crust
Major- and trace-element and isotopic composi-
tions of rocks in the Lake Plateau are provided in a
previous paper (Meen, 1987a). Basaltic to andesitic
amphibolitic rocks are enriched in large-ion litho-
phile elements (LILE) , especially alkaline earths
and light rare-earth elements (LREE). Zr contents
in amphibolites in the Lake Plateau area are much
higher than those in amphibolites in the eastern
Beartooth Mountains (Mueller and others, 1983,
1985). LREE abundances in the latter range widely
(50-500 X chondrite) and correlate strongly with Sr
contents. In contrast, REE patterns of amphibolites
in the Lake Plateau are alike (LREE = 75-100 X
chondrite), and LREE and Sr contents correlate
poorly. Such differences deny neither contempora-
neity of the two suites nor their formation in a simi-
lar tectonic setting, but do deny a direct genetic
relation (Meen, 1987a).
Sm-Nd systematics of the granitoids betray no
sign of crust older than 3200 Ma (Figure 2). In
addition, a Pb-Pb isochron defined by these rocks
indicates an equivalent single-stage growth curve
prior to 2740 Ma with pe
38
U/204Pb) = 8.04, indi-
cating that granitoid source regions contained little
16.0
:ii"
15.5
"
CI.
0
N
:0
15.0
CI.
...
0
N
14.5
14.0
1 2 1 4 16 18 20
206 204
( Pbl Pb)
Figure 4. 207PbPo4Pb_206Pb1204Pb isotopic compositions of
Cretaceous granitoids from Boulder River valley corrected
for in situ growth of radiogenic Pb (Meen and Eggler,
1989). Isochron obtained by the method of York (1969)
using error of reproducibility of standard measurements.
Error estimate is product of calculated 20 and square root
of MSWD to account for scatter. Growth curve indicated by
this isochron (11=8.25) emanates from primordial lead of
Tatsumoto and others (1973), as indicated by position at
particular points in time (in thousands ofMa).
Archean oceanic and continental blocks fonn the Beartooth Plateau 303
U-rich material. Such low values of 11 are inter-
preted to indicate that protoliths separated from the
mantle 200-500 million years before crustal stabi-
lization (e.g., Moorbath, 1977). Although there was
little U-rich material in the granitoid source, Pb
isotopic compositions do not exclude the possibility
that the granitoids had an old mafic provenance.
Figure 3 shows Sr-Nd isotopic compositions of
three groups of rock studied previously (Meen,
1987a). Although they have similar eNd, the fields
are largely discrete; amphibolites have relatively
low 87Sr/86Sr, whereas granitoids and biotite schists
have much higher ratio values. Compositions of the
granitoids reflect those ofthe crustal source regions.
Nature of the Lower Crust
Meen and Eggler (1989) studied Cretaceous
granitoids in the headwaters of the Boulder River
valley. They used major-element compositions of
rocks and their phenocrysts, in addition to trace-
element and isotopic compositions, to argue that the
granitoids were restite-melt mixtures and involved
very little material derived directly from the
mantle. Consequently, the chemical compositions of
the granitoids reflect those of crustal source regions.
Pb isotopic compositions of the granitoids cor-
rected for in situ growth since intrusion (Figure 4)
are non-radiogenic and define a linear array that
Meen and Eggler interpreted as an inherited iso-
chron, indicating crustal homogenization at 3320
220 Ma and llsource = 8.25. The low value of 11 in-
dicates that protoliths of source regions spent only a
brief time in U-rich environments and may have
separated from the mantle just before 3300 Ma.
This result is consistent with Sm-Nd and Pb-Pb
compositions of basement samples and indicates an
absence of voluminous early Archean crust. Fields
of Sr-Nd compositions of basement samples (Figure
3) are age-corrected for radiogenic growth since the
Cretaceous (Figure 5). The Cretaceous rocks clear-
ly have much lower (87Sr/86Sr)i than Archean grani-
toids or boitite schists and were not derived from
source regions containing these rocks. Similarly,
Archean granitoids have 206PbPo4Pb:547 and schist
values are ""40 (Meen, 1987a), whereas all Lara-
mide granitoids have e06Pb/204Pb)i < 19 (Figure 4).
The Sr-Nd field of Cretaceous granitoids over-
laps that of basement amphibolites (Figure 5), and
amphibolites have 206PbPo4Pb "" 16, similar to initial
ratios of granitoids. This is one line of evidence used
by Meen and Eggler (1989) to infer that granitoid
source regions were dominated by amphibolites
similar to those at the surface.
Neither Rb-Sr nor Sm-Nd compositions define
even crude isochrons. Initial Sr and Nd composi-
tions of the granitoids (Figure 5) broadly correlate,
but both (87Sr/86Sr)i and e43Nd/144Nd); are very poor-
ly correlated with (206PbPo4Pb)i (Meen and Eggler,
1989). Isotopic compositions do not, therefore, owe
their variance to two-component mixing. Sm-Nd
chondritic model ages of most granitoids are 1300-
2300 Ma, consistent with derivation of granitoids
from source regions that separated from undepleted
mantle at"" 3300 Ma, if fractionation of Sm from N d
between source and melt were 0.5-0.8. Such values
are expected of residues of melting contained much
amphibole or pyroxene, as indicated by restite as-
semblages (Meen and Eggler, 1989).
The restite assemblage in the largest granitic
intrusion is plagioclase, two pyroxenes, quartz, and
magnetite; pyroxenes have a reaction relationship
with the liquid, forming amphibole. The most plau-
sible anatectic reaction, therefore, is amphibole +
plagioclase + quartz = pyroxene + Ca-rich plagio-
clase, requiring an amphibolitic source with only
minor mica and leaving a granulitic residue. Mica-
ceous lithologies (e.g., schists or granitoids) would
-10
0
-20
0
0
m

::!i:
""
-30
"
z
w
-40
Basement
Amphlbolltes
-50
0.700 0.705 0.710 0.715 0.720
Figure 5. Sr and Nd isotopic compositions of Cretaceous
granitoids shown in Figure 4, corrected for in situ growth
of radiogenic Sr and N d and superimposed on fields defined
by basement rocks in Figure 3. Data from Meen and
Eggler (1989).
304 J. K. Meen
melt at lower temperatures and would not have
pyroxenitic restite (Meen and Eggler, 1989).
The REE pattern of the granitic melt is best
modeled by 11% melting, leaving a pyroxene-rich
granulitic residue, of a source with a REE pattern
similar to that of basement amphibolites (Figure 6;
Meen and Eggler, 1989). The pattern is LREE-
enriched, is less fractionated in MREE and HREE,
and has a small negative Eu-anomaly and a
negative Ce-anomaly. The best fit requires a source
with somewhat higher l: REE content and slightly
greater LREE-enrichment than possessed by
sampled amphibolites.
Contents of LILE in Cretaceous granitoids are
also consistent with derivation from amphibolites.
Elements, such as Th, that are uniformly distribu-
ted in amphibolites strongly correlate with Si0
2
in
the granitoids. Th contents imply 11-15% melting of
amphibolites (Meen and Eggler, 1989). The Cre-
taceous granitoids also demonstrate large ranges in
contents of Ba, Sr, and Rb, however. Meen (1988)
showed that these variations are not the result of
subsolidus alteration of the granitoids and regarded
them as reflections of the contents of these elements
in the felsic magmas. Similarly, these elements are
inhomogeneously distributed in the amphibolites,
200
La Co p, Nd Sm Eu Gd Dy Y Vb
Rare Earth Elements
Figure 6. Chondrite-normalized REE plot for calculated
melt components of granitoids in the Boulder River valley
(solid line joining open squares) determined by Meen and
Eggler (1989). Stippled line is REE pattern calculated for
11 % melting of an Archean amphibolite (1-82-019 of Meen,
1987a) to leave a granulite with clinopyroxene:plagioclase
= 4:1 using the partition coefficients of Arth (1976).
apparently because of strongly variable modal
abundances of alkali feldspar and mica. The cause
of the variability is unclear. It may reflect original
differences in the contents of these elements, in dif-
ferent lava flows, or intrusions. Alternatively, the
variability may result from mobility of alkalis and
alkaline earths during metamorphism.
Meen (1990) has shown that the granitoids have
negative Ce-anomalies - i.e., the chondrite-
normalized content ofCe (CeN) is depressed relative
to the geometric mean of LaN and PrN (Ce*). Values
of CeN/Ce* of granitoids in a stock are 0.88-0.97 (one
value of 1.07); another intrusion has CeN/Ce* =
0.84-0.88. Both intrusions have strong peaks at
values outside the error range around CeN/Ce* =
1.00 (0.05), as shown in Figure 7 (Meen, 1990).
Amphibolites have CeN/Ce* = 0.87-0.98, a range
similar to that of the granitoids. Mantle-derived
rocks described by Meen and Eggler (1987) fall in
the 0.95-1.05 band and thus have no recognizable
Ce-anomaly.
Fractionation of Ce from other LREE apparent-
ly occurs only in oxidizing, near surface environ-
\5
10

GranitoidS 01 Mill Craek

01 Ind81)endel'lCe

Archean

Postulatet1 mantle-derived
O.S< 0.86- 0.88 0.90' 0.92 0.94 0.96 0.98 1.00 1.02 1.04 1.06
Ce " ICe'
Figure 7. Histogram of Ce-anomaly values in rocks from
Boulder River valley. Ce-anomaly defined as ratio of nor-
malized content of Ce to geometric mean of normalized
contents of La and Pro Analytical uncertainty on anomaly
is 5%, so anomaly is only analytically reliable if outside
0.95-1.05.
Archean oceanic and continental blocks form the Beartooth Plateau 305
ments, especially during sub-seafloor metamor-
phism. The presence of Ce-anomalies in both
amphibolites and granitoids implies that protoliths
of the former and source regions of the latter had
been through a weathering cycle. They may have
formed as basaltic or andesitic lavas on the ocean
floor and experienced sub-seafloor metamorphism
and hydrothermal alteration (see Meen, 1990).
Masuda and Nagasawa (1975) described basalts
from the Shatsky Rise that were pervasively altered
by sub-seafloor metamorphism and that have CeNI
Ce* 0.23. CeN/Ce* values in the granitoids are
higher than this, implying that only a small per-
centage of the crustal pile underwent extensive
hydrothermal alteration, although altered rocks are
distributed throughout the granitoid source regions.
As altered rocks constituted only a small amount of
this source, major-element features expected in
altered rocks were obscured.
Meen (1990) and Meen and Eggler (1989) noted
that the time-integrated Th/u of granitoid source
regions (as exhibited by 208Pb/204Pb_2osPbPo4Pb
systematics) indicate fractionation of U and Th dur-
ing Archean times. Such fractionation is another
sensitive indicator of element mobilization (pre-
sumably in an oxidizing environment in which U
S
+
was stabilized) that major-element compositions
may fail to detect.
10
5
--
0
CII

!
5
"tI 10
Z
W
15 HATS
20
HAB
25
0.703 0.704 0.705 0.706 0.707
87 86
SrI Sr)
Figure 8. Sr-Nd isotopic compositions of mantle-derived
Cretaceous igneous rocks in the Boulder River valley
(Meen and Eggler, 1987).
Nature of the Lithospheric Mantle
Meen and Eggler (1987) showed that two
SCUM-derived Cretaceous magma types occur in
the upper Boulder River valley. One has parental
high-magnesium andesite that differentiated to
banakite (HAB); the other ranges from high-
alumina tholeiitic basalt to shoshonite and dacite
(HATS). Their initial Sr-Nd isotopic compositions
are far from the "mantle array" that defines abun-
dant asthenospheric sources (Figure 8). Values of
(87Sr/8SSr)i are close to that of bulk silicate earth
(0.7045-0.7055 versus =:;0.70465), whereas eNd is
less than zero, so the mantle sources and bulk sili-
cate earth had similar time-integrated Rb/Sr but
this mantle had higher LREEIHREE than CHUR.
Sm-Nd depleted mantle model ages are Ma,
a minimum age because anatexis enriches melts in
Nd with respect to Sm.
Meen and Eggler (1987) showed that Pb isotopic
compositions of RAB and HATS suites define in-
herited isochrons with ages of 3870 160 Ma and
2040 360 Ma, respectively (Figure 9). The paren-
tal HAB rocks are extremely primitive high-magne-
sium andesites that are little evolved from mantle
melts. Such compositions are produced only from
shallow hydrous mantle (e.g., Tatsumi, 1981) and
are interpreted to have formed by partial melting of
hydrated peridotie just below the Moho.
15.5
IS.'
HATS
D
2040 360 Ma
..,11.
15.3
I--Geochron
0
N
:c
II.
15.2
..
0
84 Ma
N
15.1
ago
15.0
15.0 15. 5 1S.0 1S.5 17. 0 17.5 18 .0
206 204
Pbl Pb)
Figure 9. 207Pb/204Pb_206Pb/204Pb isotopic compositions of
mantle-derived Cretaceous igneous rocks shown in Figure
8, corrected for in situ growth in radiogenic Pb (Meen and
Eggler,1987). Isochrons calculated by the method of York
(1969) and error estimate weighted by the square root of
MSWD. Growth curve indicated by HATS isochron
(11 = 8.10) emanates from the primordial lead of Tatsumoto
and others (1973) as in Figure 4.
306 J. K. Meen
The younger mantle age is defined by the HATS
suite that has basaltic parents. The primary mafic
liquids were formed by melting of mantle at un-
known depth. They were apparently 01- and hy-
normative, and Meen (1987b) has shown that evolu-
tion of these magmas requires a low water content.
The mantle source regions of these rocks were
anhydrous enriched pyroxenites or Iherzolites that
formed penecontemporaneously with crustal accre-
tion around the Wyoming Province (Nelson and
DePaolo, 1985; Patchett and Arndt, 1986). This
source region was probably at depths of 50-70 km
(almost certainly:::; 100 km), because temperatures
required to melt anhydrous peridotite at 30 kb
would require whole scale conversion of lithosphere
to asthenosphere. Either new lithospheric mantle
was added to the base of the SCUM or old depleted
material was modified by addition to asthenosphere-
derived melts. This activity is also reflected in local
granite production in the Lake Plateau (Lafrenz and
others, 1986).
The thickness of the Cretaceous lithosphere is
unknown. Lithosphere beneath the ColoradolWyo-
ming state line was ;:: 200 km thick in Devonian
time (Eggler and others, 1987). If SCUM beneath
the Wyoming Province was of similar thickness, it
was penetrated deeply by melts at "" 2000 Ma, im-
plying major resetting of its composition. Dudas
2.74.<0,07 Go
Granitoids
and others (1987) also inferred the presence of 1500-
1800 Ma mantle beneath the Crazy Mountains im-
mediately north of the Beartooth Mountains. This
lithosphere was again invaded by advection of heat
and, possibly material, in Cretaceous time. Appar-
ently, however, significant amounts of SCUM exist
beneath parts of the present-day Wyoming Province
(Eggler and others, 1988).
A Section Through the SubBeartooth
Lithosphere
Figure 10 illustrates the proposed model of the
Beartooth lithosphere prior to Cretaceous anatexis.
Exposed crustal rocks are 2740 Ma granites that
intruded supracrustal amphibolites and schists.
Presumably, the 2740 Ma granites have comple-
mentary depleted-residue in the deep crust, un-
sampled in Cretaceous time because of the required
high temperatures for anatexis. Not all the crust
was depleted, however, and much melted during
later events.
The deep crust under the Boulder River valley
was apparently dominated by amphibolites formed
from basalts or andesites, some of which experi-
enced sub-seafloor alteration. The supracrustal pile
that was isotopically homogenized at 3300 Ma
AmphlbolltiC crusl
Figure 10 Diagram of model developed
for lithosphere beneath the western
Beartooth Plateau prior to melting in
Cretaceous time. 3,32 .<0,22 Go
3.87 0 t6 Go
204 .. 036 Go
Asthenosphere
with inaeasing
amount oj felsIC
metasedimentS
upwardS
and. presumably, 0
granulites in
areas lhat metted
2700 Ma ago. at dOplh,

upper mantle
(SCUM) composed
01 depleted pendCllt8
an by metasomatIC
veins of enriched
pendobto and
pyrO):8n1t9.
o
o
tm
Archean oceanic and continental blocks form the Beartooth Plateau 307
contained little or no intercalated felsic material,
because the Cretaeous granitoids have low
(87Sr/86Srh and e0
6
PbFo4Pb);, but low (143Nd/144Nd)i.
There is, however, metasedimentary material near
the present-day surface. Thick piles of mafic rocks
with little intercalated sedimentary or felsic rock
are typical of oceanic rather than continental
environments. Both amphibolites exposed at the
surface and in postulated source regions of the
Cretaceous granitoids have compositions similar to
those of LILE-enriched basalts or andesites with
significant LREE-enrichment. These characteris-
tics are not similar to those of present-day "normal"
ocean-ridge basalts, but such traits may indicate
formation in an island-arc environment. Such an
environment is consistent both with formation of a
buoyant pile of igneous rocks and with development
from a totally submarine into a partially emergent
terrane.
Isotopic compositions of crustal rocks are con-
sistent with separation from the mantle only shortly
before 3300 Ma. There appears to be little early
Archean crust in the western Beartooth Plateau;
continental crust of that age was not sampled in the
Boulder River valley.
Comparison with Surrounding Areas
Estimated ages of crystallization or mantle
separation of rocks from the Beartooth Mountains
and, more generally, the Wyoming Province are
shown in Figure 11. Ages younger than "'" 3000 Ma
are omitted. A large part of the Wyoming Province
yields ages of 3000-3300 Ma, implying that much of
the crust was initially stabilized in middle Archean
time. There are two notable exceptions to this.
Some rocks in the eastern Beartooth Plateau are
considerably older (>3500 Ma). In addition, a
gneiss from the North Snowy Block has a Sm-Nd
model age of 3550 Ma, although another has a
model age of 3260 Ma and an amphibolite has one of
3200 Ma (Mogk and others, 1988).
The lithospheric section modeled in Figure 10
is different from that described from neighboring
areas, although a significant part of the Beartooth
crust was in place prior to ""' 3000 Ma. The wide-
spread granitoid magmatism that occurred at 2740-
2560 Ma was imposed on a largely amalgamated
crust. Mogk and others (1988) suggested that part
of the North Snowy Block and areas farther west
were accreted to the plutonic terrane during this
time.
Mueller and others (1982b, 1985) and Henry
and others (1982) showed that evolved continental
crust was present in the eastern Beartooth
Mountain at 3200-2500 Ma. This area is only 60-80
km east of the area considered here (see Figure 1).
It is highly unlikely that the terranes were this
close at 3200-3500 Ma. At that time, the crust in
the western Beartooth Plateau was apparently
being formed in an oceanic environment. The lack
Rb.Sc mpdetBqg
Sm-Nd model age
Pb-Pb lsochron
u-pb 111000 1'1' R
Figure 11. Compilation of times at which
various portions of Wyoming Province
crust separated from mantle or cry-
stallized at 3000 Ma. Data from De
Paolo and Wasserburg (1979), Fischer and
Stacey (1987), Leeman and others (1985),
Meen and Eggler (1987, 1989), Meen
(1987a), Mogk and others (1988), Mont-
gomery and Lytwyn (1984), Mueller and
others (1982, 1985), Nunes and Tilton
(1971), Reid and others (1975).
NO'I inCh.J(Mg y!lUllgaf Ihan 3000 Ma
17002000 Ma
2000-2300 M.l
308
J. K. Meen
30002800 Ma
OCEANIC CONTINENTAl.
TE.ARANE MARGIN
SUbdoctlon-ral3led

Wc:ST EAST
<2740 Ma
EARLY ARCI-EAN
OCEANIC MA TEAIAI..
REOISTRt!ruTE:ll 2300
Mil AGO. WITH EARLY
ARCHEAN SCUM
PRESERVED
E.ARI.Y ARCHEAN
CONTINENT Pl.US
PAATERW.AOOeO
30()0.2800 Ma AGO.
WITH 3000 2800 M3
SCUM
late AfcI1eiln grilmtoids
., r9'Sidut ograni!OiCI
6 Stillwater C<H'n!:Ilal.
l&anllo 1C$i(luO Of 1'nC'1lmg
Figure 12. Model for evolution of the Beartooth Plateau
during late Archean time. The upper diagram showns the
western and eastern parts of Beartooth Plateau separate at
""3000 Ma; the western part was oceanic terrane of un-
certain tectonic affinity, whereas the eastern part was
continental crust. The western Beartooth Plateau pre-
serves evidence of ancient lithospheric mantle not des-
troyed during late Archean activity. This area is inter-
preted as "rafted" on subducting ocean plate and accreted
to continent. The lower diagram shows that post-accretion
activity resulted in wholescale intra-lithospheric redis-
tribution of material.
of much felsic sedimentary material presumably
reflects a general lack of emergent crust. On the
other hand, the eastern Beartooth Plateau appar-
ently underwent crustal thickening at ==: 3400 Ma to
create continental crust =:: 40 km thick (Mueller and
other, 1985; MogkandHenry, 1988).
Wooden and Mueller (1988) inferred that the
late Archean mantle beneath the eastern Beartooth
Plateau was enriched by addition of crustal rocks
with elevated Sr and Pb isotopic ratios and lower
143Nd/144Nd. They believed this to be required by
compositions of late Archean intermediate to felsic
igneous rocks of the eastern plateau, and they
suggested that crustal recycling into the mantle
occurred during late Archean subduction. This sub-
duction, they proposed, was the driving mechanism
behind the formation of 2740-2790 Ma rocks.
Subduction of oceanic crust beneath the eastern
Beartooth Moutains could have caused accretion of
two rather different terranes at ==:2800 Ma, al-
though it is impossible to ascertain the direction of
subduction. A simple diagram (Figure 12) illus-
trates this model. The western Beartooth Moun-
tains are interpreted to have formed as thickened
oceanic crust (possibly oceanic arc) distant from the
continental eastern Beartooth Plateau. The two
may have been juxtaposed and amalgamated by
subduction ",,2750 Ma (Figure 12, top). This
subduction led to production of late Archean inter-
mediate to acid igneous rocks in the eastern Bear-
tooth Plateau (Wooden and Mueller, 1988). Fur-
thermore, the mantle was heated and fluxed.
Transport of the western Beartooth Plateau on
the oceanic plate carried ancient SCUM with it
(Figure 12, top). The lithospheric thickness at that
time is unknown, but it was probably < 200 km,
inferred for the Colorado-Wyoming lithosphere by
Eggler and others (1987). Following amalgamation
of the eastern and western Beartooth Plateau
blocks, "" 2750 Ma, subduction ceased and the area
was subject to voluminous granitic magmatism.
Wooden and others (1988) suggested that some late
Archean granitoids in the Beartooth Plateau may
have resulted from post-collisional adiabatic melt-
ing. Major crustal thickening is indicated for the
Lake Plateau at this time, possibly involving over-
thrusting within the oceanic terrane. On the basis
of metamorphic assemblages in supracrustal rocks,
Mogk and Henry (1988) proposed that metamor-
phism was at ==: 6000C and 7-8 kb.
In addition, the Stillwater Complex formed,
==: 2700 Ma (Figure 12, bottom), from magma that
almost certainly came from SCUM. The source has
been suggested to be comparable to sources of mag-
mas in both the western Beartooth Plateau (Me en
and Eggler, 1987) and the eastern Beartooth
Plateau (Wooden and Mueller, 1988).
Mogk and others (1988) proposed, on the basis
of structural and geochronological study of the
North Snowy Block, that the western part of the
block was accreted to the high-grade terrane at 2560
Ma. Rocks in the accreted terrane have Sm-Nd
model ages of 3200, 3260, and 3550 Ma. They may
have been derived from both the ==: 3300 Ma oceanic
and the >3500 Ma continental terrane.
Archean oceanic and continental blocks form the Beartooth Plateau 309
CONCLUSIONS
The western and eastern part of the Beartooth
Plateau have different natures and ages, and the
former has a vertically heterogeneous composition.
Deep crust in the west that was the source of Cre-
taceous granitoids was apparently dominated by
amphibolites of mafic-intermediate compostion.
These may be metamorphosed equivalents of lavas
that formed in an oceanic environment, possibly an
oceanic arc. Negative Ce-anomalies in the grani-
toids indicate that protoliths had been through an
episode of alteration.
The source regions of the Cretaceous granites
have low Rb/Sr (87Sr/86Sr = 0.703-0.710), u/Pb
e0
6
PbPo4Pb = 15-19), and Sm/Nd (eNd = -15 to - 30)
- consistent with the suggestion that they were
amphibolitic. These regions did not contain vol-
uminous granite or metapelite that would have
resulted in higher 87Sr/86Sr and 206Pb/
204
Pb. Fur-
thermore, REE and LILE systematics of the grani-
toids are consistent with formation from amphibo-
litic sources. The sequence in the exposed crust
contains amphibolites similar to those postUlated to
have dominated the deep crust. These are inter-
calated with metasedimentary material, including
metapelites, and intruded by voluminous 2740 Ma
granitoids. Pelites and granitoids are not signifi-
cantly represented at depth.
The Cretaceous granitoids yield a Pb-Pb iso-
chron with an age of "" 3300 Ma, interpreted to
reflect metamorphism of their amphibolitic source.
The equivalent single-stage growth curve of Pb in
the source prior to 3300 Ma and their Sm-Nd
systematics, as well as single-stage curves and Sm-
Nd model ages for upper crustal rocks, are con-
sistent with the crust having separated from the
asthenosphere only shortly before 3300 Ma.
The western Beartooth Plateau crust may have
developed in an oceanic environment, possibly an
arc, distant from continental crust. The crust now
exposed at the surface includes significant amounts
of metasedimentary rock originally deposited in
shallow water. This may reflect upward growth of
the crust. On the other hand, the western and
eastern parts of the Beartooth Plateau may have
been closer when these rocks formed. Two biotite
schists analyzed by Meen (1987a) possess negative
Eu-anomalies that imply an input from evolved
material.
The mantle in the western Beartooth Plateau
indicated a pre-history before development of the
crust, so the mantle was not disrupted by subduction
in late Archean time. Subduction may have occured
under the eastern Beartooth Plateau (Wooden and
Mueller, 1988) to produce the 2740-2790 Ma igneous
rocks, enrich the mantle, and destroy any older
mantle, resulting in accretion of the oceanic western
Beartooth Plateau on the ancient continent. This
may explain the juxtaposition of a wide range of
lithospheric units (Figure 12, bottom). Volumin-
ous magmas formed by melting during intra-
lithospheric redistribution at 2560-2740 Ma.
Mogk and others (1988) argued that a supra-
crustal sequence, now exposed in the western North
Snowy Block, was amalgamated at 2560-2740 Ma
and accreted to plutonic terrane at 2560 Ma. Many
Precambrian uplifts west of the Beartooth Moun-
tains are similar to the western North Snowy Block
and may be part of the same terrane (Mogk and
others, 1988).
Wyoming Province crust southwest of the Bear-
tooth Mountains may have formed "" 3300 Ma, and
this area may comprise an oceanic terrane accreted
to the continent "" 2800 Ma (compare with conclu-
sions of Leeman and others, 1985). A wide variety of
rock types were juxtaposed in the Beartooth Moun-
tains and since modified. Intra-lithospheric re-
distribution at 2560-2740 Ma resulted in the rise of
low-temperature material at depth (Figure 12,
bottom). The major event at 1700-2000 Ma resul-
ted in modification of SCUM beneath the Beartooth
Mountains (Meen and Eggler, 1987) and neigh-
boring areas (Dudas and others, 1987) and of crust
in the Beartooth Mountains (Reid and others, 1975;
Giletti, 1966; Lafrenz and others, 1986). In addi-
tion, the Cretaceous and Eocene thermal events
greatly modified the lithosphere.
ACKNOWLEDGMENTS
This paper was greatly improved by D. Mogk
and two anonymous reviewers. Field work was
funded by Sigma Xi, the Geological Society of
America Penrose Fund, and the Krynine Fund of
The Pennsylvania State University. The author
was funded in these studies by NSF Grant EAR-
8206769 to D. H. Eggler at The Pennsylvania State
University, by the Department of Terrestrial Mag-
netism, Carnegie Institution of Washington, and by
the Department of Geology, University of North
Carolina, Chapel Hill.
310
J. K. Meen
REFERENCES
ARTH, J. G., 1976, Behavior of trace elements during
magmatic processes: Summary of theoretical models and
their appplications: U.S. Geological Survey Journal of
Research, v. 4, p. 41-47.
CASELLA, C. J., J. LEVAY, E. EBILE, B. HmST, K.
HUFFMAN, V. LAHTI, and R. METZGER, 1982, Precam-
brian geology of the southwestern Beartooth Mountains,
Yellowstone National Park, Montana and Wyoming; pp. 1-
24 in P. A. Mueller and J. L. Wooden (eds.), Precambrian
Geology of the Beartooth Mountains, Montana and
Wyoming: Montana Bureau of Mines and Geology, Special
Paper 84,167 p.
DEPAOLO, D. J., and G. J. WASSERBURG, 1979, Sm-Nd
age of the Stillwater Complex and the mantle evolution
curve for neodynium: Geochimica et Cosmochimica Acta,
v. 43, p. 999-1008.
DUDAs, F. D., R. W. CARLSO, and D. H. EGGLER, 1987,
Regional mid-Proterozioc enrichment of the subcontinen-
tal mantle source of igneous rocks from central Montana:
Geology, v. 15, p. 22-25.
EGGLER, D. H., F. b. DUDAs, B. C. HEARN, JR., M. E.
McCALLUM, E. S. McGEE, H. O. A. MEYER, and D. J.
SCHULZE, 1987, Lithosphere of the continental United
States: Xenoliths in kimberlites and other alkaline mag-
mas; pp. 41-57 in P. H. Nixon (ed.), Mantle Xenoliths:
John Wiley and Sons, Chichester, England, UK, 844 p.
__ , J. K. MEEN, F. G. P. WELT, F. b. DUDAs, K. P.
FURLONG, M. E. McCALLUM, and R. W. CARLSON, 1988,
Tectonomagmatism of the Wyoming Province; pp. 25-40 in
J. Drexler and E. E. Larson (eds.), Colorado Volcanism:
Colorado School of Mines Quarterly, v. 2, 320 p.
GILETTI, B. J., 1966, Isotopic ages from southwestern
Montana: Journal of Geophysical Research, v. 71, p. 4092-
4136.
HENRY,D.J.,P. A. MUELLER, J. L. WARNER,andR. LEE-
BERMAN, 1982, Granulite grade supracrustal assemblages
of the Quad Creek area, eastern Beartooth Mountains,
Montana; pp. 147--156 in P. A. Mueller and J. L. Wooden
(eds.), Precambrian Geology of the Beartooth Moun-
tains, Montana and Wyoming: Montana Bureau of Mines
and Geology, Special Paper 84, 167 p.
LAFRENZ, W. B., R. D SHUSTER, P. A. MUELLER, and D.
R. BOWES, 1986, Archean and Proterozoic granitoids of the
Hawley Mountain area, Beartooth Mountains, Montana;
pp. 79-89 in YBRA Field Conference Guidebook: Joint
Field Conference and Sympsoium (50th Anniversary
Edition): Montana Geological Society, Billings, Montana,
USA,185p.
LEEMAN, W. P., M. A. MENZIES, D. J. MATY, and G. F.
EMBREE, 1985, Strontium, neodymium and lead isotopic
compositions of deep crustal xenoliths from the Snake
River Plain: Evidence for Archean basement: Earth and
Planetary Science Letters, v. 75, p. 354-368.
MASUDA, A., and S. NAGASAWA, 1975, Rocks with nega-
tive cerium anomalies, dredged from Shatsky Rise: Geo-
chemical Journal, v. 9, p. 227-223.
MEEN, J. K., 1987a, Sr, Nd, and Pb isotopic compositions of
Archean basement rocks, Boulder River region, Beartooth
Mountains, Montana: IsochronlWest, v. 50, p. 13-24.
__ , 1987b, Formation of shoshonites from ca1calkaline
basalt magmas: Geochemical and experimental con-
straints from the type locality: Contributions to Minera-
logy and Petrology, v. 97, p. 333-351.
__ , 1988, Petrological and chemical variations in the
Independence volcanic complex, Absaroka Mountains,
Montana: Northwest Geology, v. 17, p. 21-41.
__ , 1990, Negative Ce anomalies in Archean amphibo-
lites and Laramide granitoids, southwestern Montana:
Chemical Geology, v. 81, p. 191-207.
__ , and D. H. EGGLER, 1987, Petrology and geo-
chemistry of the Cretaceous Independence volcanic suite,
Absaroka Mountains, Montana: Clues to the composition
of the Archean sub-Montanan mantle: Geological Society
of America Bulletin, v. 98, p. 238-247.
__ , and D. H. EGGLER, 1989, Chemical and isotopic
compositions of Absaroka granitoids, southwestern Mon-
tana: Evidence for deep-seated Archean amphibolitic base-
ment in the Beartooth region: Contributions to Mineralogy
and Petrology, v.l02, p. 462-477.
MOGK, D. W., and D. J. HENRY, 1988, Metamorphic
petrology of the northern Archean Wyoming Province,
southwestern Montana: Evidence for Archean collisional
tectonics; pp. 363-382 in W. G. Ernst (ed.), Metamorphism
and Crustal Evolution of the Western United States
(Rubey Volume 7): Prentice Hall, Inc., Englewood Cliffs,
New Jersey, USA, 1153 p.
__ , P. A. MUELLER, and J. L. WOODEN, 1988, Archean
tectonics ofthe North Snowy Block, Beartooth Mountains,
Montana: Journal of Geology, v. 96, p. 125-141
MONTGOMERY, C. W., and J. N. LYTWYN, 1984, Rb-Sr
systematics and ages of principal Precambrian lithologies
in the South Snowy Block, Beartooth Mountains: Journal
of Geology, v. 92, p. 103-112.
MOORBATH, S., 1977, Ages, isotopes and evolution of
Precambrian continental crust: Chemical Geology, v. 20, p.
151-187.
MUELLER, P. A., and J. L. WOODEN, 1976, Rb-Sr whole
rock age of the contact aureole of the Stillwater igneous
complex, Montana: Earth and Planetary Science Letters, v.
29, p. 384-388.
Archean oceanic and continental blocks form the Beartooth Plateau 311
__ , J. L. WOODEN, and D. R. BOWES, 1982a,
Precambrian evolution of the Beartooth Mountains, Mon-
tana and Wyoming, U.S.A.: Revista Brasileira de
Geociencias, v. 12, p. 215-222.
__ , J. L. WOODEN, D. J. HENRY, and D. R. BOWES,
1985, Archean crustal evolution of the eastern Beartooth
Mountains, Montana and Wyoming; pp. 9-20 in G. K
Czamanske and M. L. Zienzek (eds.), The Stillwater
Complex, Montana - Geology and Guide: Montana
Bureau of Mines and Geology, Special Paper 92, 396 p.
__ , J. L. WOODEN, A. L. ODOM, and D. R. BOWES,
1982b, Geochemistry of the Archean rocks of the Quad
Creek and Hellroaring Plateau area of the eastern
Beartooth Mountains; pp. 69-82 in P. A. Mueller and J. L.
Wooden (eds.) Precambrian Geology of the Beartooth
Mountains, Montana and Wyoming: Montana Bureau of
Mines and Geology, Special Paper 84,167 p.
__ , J. L. WOODEN, K SCHULZ, and D. R. BOWES, 1983,
Incompatible-element-rich andesitic amphibolites from the
Archean of Montana and Wyoming: Geology, v. 11, p. 203-
206.
NELSON B. K, and D. J. DEPAOLO, 1985, Rapid produc-
tion of continental crust 1.7-1.9 m.y. ago: Nd isotopic
evidence from the basement of the North American mid-
continent: Geological Society of America Bulletin, v. 96, p.
746-754.
NUNES, P. D., and G. R. TILTON, 1971, Uranium-lead ages
of minerals from the Stillwater Igneous Complex and
associated rocks, Montana: Geological Society of America
Bulletin, v. 82, p. 2231-2250.
PATCHETT, P. J., and N. R. ARNDT, 1986, Nd isotopes and
tectonics of 1.9-1.7 Ga crustal genesis: Earth and Plane-
tary Science Letters, v. 78, p. 329-338.
REID, R. R., W. J. McMANNIS, andJ. C. PALMQUIST, 1975,
Precambrian Geology of the North Snowy Block,
Beartooth Mountains, Montana: Geological Society of
America, Special Paper 157, 135 p.
TATSUMI, Y., 1981, Melting experiments on a high-
magnesian andesite: Earth and Planetary Science Letters,
v. 54, p. 357-365.
TATSUMOTO, M., M. J. KNIGHT, and C. J. ALLEGRE, 1973,
Time differences in the formation of meteorites as deter-
mined from the ratio of lead-207 to lead-206: Science, v.
180, p. 1279-1283.
WARNER, J. L., R. LEE-BERMAN, and C. H. SIMONDS,
1982, Field and petrologic relations of some Archean rocks
near Long Lake, eastern Beartooth Mountains, Montana
and Wyoming; pp. 57-68 in P. A. Mueller and J. L. Wooden
(eds.), Precambrian Geology of the Beartooth Moun-
tains, Montana and Wyoming: Montana Bureau of Mines
and Geology, Special Paper 84,167 p.
WOODEN, J. L., and P. A. MUELLER, 1988, Pb, Sr, and Nd
isotopic compositions of a suite of late Archean igneous
rocks, eastern Beartooth Mountains: Implications for
crust-mantle evolution: Earth and Planetary Science
Letters, v. 87, p. 59-72.
__ , P. A. MUELLER, and D. W. MOGK, 1988, A review of
the geochemistry and geochronology of the Archean rocks
of the northern part of the Wyoming Province; pp. 384-410
in W. G. Ernst, (ed.), Metamorphism and Crustal Evolu-
tion of the Western United States (Rubey Volume 7):
Prentice Hall, Inc., Englewood Cliffs, New Jersey, USA,
1153 p.
YORK, D., 1969, Least-squares fitting of a straight line
with correlated errors: Earth and Planetary Science
Letters, v. 5, p. 320-324.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Precambrian geology and ductile normal faulting in the
southwest corner of the Beartooth Uplift, Montana
ERIC A. ERSLEV
Department of Earth Resources, Colorado State University, Fort Collins, Colorado 80523, USA
(received November 21,1988; revision accepted April 3, 1989)
ABSTRACT
The northwestern flank of the Beartooth uplift exposes the Snowy shear zone, a thick belt of NW-dipping,
ductilely-deformed Archean rocks. In the South Snowy Block north of Gardiner, Montana, the shear zone
separates high-grade sillimanite+cordierite-bearing biotite gneisses and augen gneisses (to the southeast)
from andalusite-bearing biotite-staurolite schists (to the northwest). The shear zone contains fine-grained,
mylonitized, and retrograded equivalents of these rocks interfingering with lenses of actinolite amphibolite,
metagabbro, and quartzite.
Foliation and fold axis orientations within the biotite gneisses and the biotite-staurolite schists are nearly
identical, suggesting a common geologic history and limited displacement in the Snowy shear zone since
prograde metamorphism. Asymmetrical fabrics are well developed in the shear zone, showing one stage of
oblique slip with normal and right-lateral components. The normal slip component is consistent with the
differences in metamorphic equilibration conditions recorded in the biotite gneisses and the biotite-staurolite
schists. These results contradict earlier hypotheses that suggested lateral shortening in these mylonites. The
Snowy shear zone probably formed during intra-cratonic extension during a late-stage, possibly Proterozoic,
orogen that uplifted an ancestral Beartooth block by ductile normal faulting.
INTRODUCTION
The great diversity of lithotectonic assemblages
in the Archean rocks of the northwestern Wyoming
Province promises to increase our knowledge of
early earth processes. These assemblages include
middle Archean gneiss terranes (Erslev, 1983; Mogk
and others, 1988), thick accumulations of pelitic and
semi-pelitic schists (Casella and others, 1982),
granitic batholiths (Mueller and others, 1988), and
unusual - at least for the Archean - carbonate-
pelite-quartzite assemblages (James and Hedge,
1980; Sumner and Erslev, 1988). These rocks are
well-exposed locally in Laramide foreland uplifts
(Figure 1), which are characterized by block motion
and limited penetrative deformation (Erslev, 1986).
However, Laramide and Basin and Range faulting
has resulted in discontinuous exposures, leading to
extensive debates concerning the validity of litho-
stratigraphic correlations between ranges (e.g.,
Vitaliano and others, 1979; James and Hedge, 1980;
Erslev, 1988).
In particular, correlations between the ranges
in southwestern Montana and the Beartooth block
in south-central Montana are unclear. Mogk (1988)
emphasized the differences between areas in the
northern and western Beartooth Mountains, stating
"the western margin of the Beartooth Mountains
lies along a fundamental discontinuity between two
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 313-322. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
313
314
E. A. Erslev
o Butte
Highland
Mountoins
Blacktail
Mountains
MONTANA
. '- .,'''-
Bridger
Range
o Bozeman

o
N
50 KM
-.' .. "
IDAHO
. " 11 10
Figure 1 Geologic map of the pre-Beltian
rocks of southwest Montana and the Bear-
tooth uplift.
distinct terranes in the Archean basement of the
northern Wyoming Province." In contrast, Erslev
(1983, 1988) has pointed out similarities in proto-
liths, metamorphic history, and geochemistry of
pelites overlying the Cherry Creek carbonate shelf
sequence in southwestern Montana with pelites in
the southwestern Beartooth Mountains. Lithostra-
tigraphic differences between these two areas can be
explained as facies changes, necessitating no major
crustal boundary between the areas. In the deposi-
tional model proposed by Erslev (1988), the car-
bonate-rich metasedimentary rocks represent a
continental shelf setting adjacent to the turbidite
trough presently represented by the Archean sedi-
mentary rocks in the Beartooth block.
A crucial locality in this debate is the belt of
deformed rocks extending along the western edge of
the Beartooth block (Figure 1). The northern sec-
tion of this belt is exposed in the North Snowy Block
southwest of Livingston, Montana. In this area,
Reid and others (1975) documented a complex his-
tory of deformation, which was expanded by Mogk
(1983) who proposed that the zone be called the
"North Snowy Block mobile belt." Extensively
sheared rocks from the southwestern part of this
belt in the South Snowy Block were described by
Erslev (1982) and Burnham (1982). As this zone of
ductile shearing and associated retrograde meta-
morphism traverses both the North and South
Snowy Blocks, it is called the "Snowy shear zone" in
this paper.
In the North Snowy Block, the NW-dipping
Snowy shear zone forms the northwest margin of the
Precambrian exposures. As a result, the hanging
wall of the shear zone is not exposed, making it
difficult to determine its significance and the type of
regional offset. The southern end of the shear zone
is not as well exposed as the northern end, but rocks
unaffected by greenschist-facies shearing flank both
sides of the zone, allowing the comparison of litho-
logies and their geologic histories across the shear
zone.
This study was initiated to determine the tec-
tonic significance of the Snowy shear zone. An area
of "" 90 km
2
along the southwest margin of the
Beartooth Range was mapped at a scale of 1:24,000
to delineate the units and geologic history of the
Precambrian rocks. Thin sections of oriented
Precambrian geology and ductile normal faulting in the Beartooth Uplift, Montana 315
samples of sheared rock and unoriented, unsheared
rocks were used to determine their mineralogy,
metamorphic assemblages, and approximate com-"
positions. Three polished sections were analyzed for
geothermometry using microprobes at Harvard
University and the U.S. Geological Survey in Den-
ver, Colorado. Thin sections oriented parallel and
perpendicular to major lineations were examined to
determine the axis of maximum stretch (X). Thin
sections parallel to the maximum stretch and per-
pendicular to the foliation were examined for asym-
metric fabrics indicative of shear sense using the
methods summarized by Simpson and Schmid
(1983).
Previous work in this area has consisted of re-
connaissance studies of limited areas of exposure.
Erslev (1982), acting on the suggestion of David
Fountain (University of Wyoming), noted the
existence of fine-grained, highly sheared rocks with
a moderate NW-dip in the Tom Miner Basin, Six
Mile Creek, and Emigrant Creek areas of the South
Snowy Block and the Mill Creek and Pine Creek
areas of the North Snowy Block. I suggested that
these rocks may be correlative with the Madison
mylonite zone, a major shear zone on strike to the
southwest in the southern Madison Range (Figure
1). Subsequent kinematic and isotopic analyses
have shown that the Madison mylonite zone is an
outlying basement thrust zone to a major 1800-1900
Ma orogen centered in the northwestern exposures
of Archean basement in southwest Montana (Erslev,
1988; O'Neill and others, 1988; Erslev and Sutter,
1990).
Burnham (1982) mapped exposures of the fine-
grained rocks and adjacent gneisses in Six Mile and
Yankee Jim Canyons. He noted the existence of
highly sheared protomylonite gneiss at both loca-
lities but attributed the fine-grained amphibolite,
quartzite, and phyllite in Six Mile Canyon (inter-
preted by Erslev (1982) as mylonites) to a separate,
unsheared sequence. Because the rocks between
these two localities were not mapped, Burnham
(1982) noted that correlations between the two
localities were tenuous.
Guy and Sinha (1988) sampled the gneisses in
Yankee Jim Canyon to determine if these rocks are
possible sources for the Yellowstone rhyolites. They
documented a high grade (775-800C, 4.5-6 kb) equi-
libration in pelitic rocks with thermobarometry
assemblages containing garnet, cordierite, biotite,
and sillimanite. Isotopic analyses show consider-
able scatter, with whole-rock Pb isotope data clus-
tering at == 2625-2650 Ma and the discordant zircon
U-Pb data giving an intercept age of ==2850 Ma.
Guy and Sinha (1988) attributed this scatter and
that of the Rb-Sr systematics to disturbance during
a later metamorphic event. Early Proterozoic Rb-Sr
and K-Ar mineral ages from Archean rocks in the
center of the South Snowy Block have also been
interpreted as reflecting a late period of basement
reactivation (Brookins, 1968; Montgomery and
Lytwyn, 1984).
LITHOTECTONIC UNITS
The Precambrian exposures in the south-
western corner of the Beartooth uplift are flanked
by numerous Tertiary and Quaternary units which
are not differentiated on the geologic map in Figure
2. In general, Quaternary alluvium and landslides
dominate valley bottoms and Tertiary volcanics of
the Absaroka igneous suite overlie the Precambrian
rocks. Intermediate and silicic intrusive rocks, pre-
sumably of Tertiary age, cut Precambrian exposures
in Yankee Jim Canyon and bound the exposures in
Emigrant Gulch. Paleozoic and Mesozoic rocks were
not seen within the mapping area. Recent faulting
along the eastern margin of Paradise Valley has
downdropped the valley to the west. However,
actual displacements may be rather minimal in the
study area since Precambrian exposures are nearly
continuous across Tom Miner Basin from Yankee
Jim Canyon to Crystal Cross Mountain.
The Precambrian rocks can be subdivided into
three main lithotectonic units on the basis of
mineral assemblages and shear textures. In the
southeastern side of the area, biotite gneiss and
intrusive augen gneiss in Yankee Jim Canyon and
the Six Mile Creek drainage area are characterized
by high-grade mineral assemblages lacking shear
fabrics and extensive retrograde metamorphic
effects (Burnham, 1982; Guy and Sinha, 1988). A
small domain along the northwestern side of the
study area around Crystal Cross Mountain is char-
acterized by medium-grade biotite schists con-
taining biotite, muscovite, garnet, staurolite, and
andalusite. The intervening Snowy shear zone con-
tains sheared and retrogressed pelitic Isemi-pelitic
rocks from the adjacent blocks as well as a distinc-
tive, partially retrogressed amphibolite-quartzite
assemblage. These rocks contain equilibrium bio-
tite+chlorite-bearing assemblages, with only one
sample containing garnet. The association of the
a

T
e
r
t
i
o
r
y

D
i
k
e
s

E
x
p
l
a
n
a
t
i
o
n


S
n
o
w
y

S
h
e
o
r

Z
o
n
e

f
i
!
i
r
l

C
h
l
o
r
i
t
e
-
B
i
o
t
i
t
e

S
c
h
i
s
t

_

Q
u
o
r
t
z
i
t
e

I
i
i
i
:
I

A
m
p
h
i
b
o
l
i
t
e

&

M
e
t
a
g
a
b
b
r
o

"
:
<
'
,


/

.
,

'
"
"

/
"
'
"

'
<

/

_
_

r
.

'
0

A
r
c
h
e
o
n

R
o
c
k
s

'
l


A
u
g
e
n

G
n
e
i
s
s


B
i
o
t
i
t
e
-
S
t
o
u
r
o
l
i
t
e

S
c
h
i
s
t

[
]
[
)

B
i
o
t
i
t
e

G
n

e
i
s
s

.
J
.
.

'I
>

/

'
"

-
S
:
"
.
,,:,,
.
,
/
'

,


.
"
.

,
.
,

/
"

,

.

.
.
.


,
.
,
;
.
;
!

.
/

.
.
.
.
.

-
-
-
v
-
F
o
l
i
a
t
i
o
n

O
n
e

K
i
l
o
m
e
t
e
r

F
i
g
u
r
e

2
.

G
e
o
l
o
g
i
c

m
a
p

o
f

b
a
s
e
m
e
n
t

l
i
t
h
o
l
o
g
i
e
s

i
n

t
h
e

s
o
u
t
h
w
e
s
t

c
o
r
n
e
r

o
f

t
h
e

B
e
a
r
t
o
o
t
h

u
p
l
i
f
t
.

<
:.:>

.
.
.
.
.

C
j
)

t
z
:
:
l

>

t
z
:
:
l

.
.
.

0
0

C
D

<
:

Precambrian geology and ductile normal faulting in the Beartooth Uplift, Montana 317
upper greenschist-facies assemblages and the high-
ly mylonitized fabrics in the Snowy shear zone sug-
gests mylonitization synchronous with retrograde
metamorphism.
High-Grade Gneisses
The high-grade gneisses are best exposed in
Yankee Jim Canyon, where part of the exposure was
studied by Burnham (1982) and Guy and Sinha
(1988). The biotite gneiss is also exposed in the Six
Mile Creek drainage area (Burnham, 1982), and
mylonitized equivalents make up the easternmost
exposures in Emigrant Gulch.
The dominant lithology in these rocks is a high-
grade biotite gneiss and migmatite with minor
amphibolite and rare meta-ultramafic pods. These
rocks are complexly folded, showing considerable
local compositional variability due to partial melt-
ing. Thin sections show AFM assemblages of garnet
+ biotite + cordierite (Figure 3A) and garnet + bio-
tite+cordierite+cummingtonite. Guy and Sinha
(1988) also reported pelitic assemblages containing
sillimanite as well as mafic rocks with hypersthene
and augite. They estimated prograde metamorphic
temperatures of 775-800C using biotite-garnet Fe-
Mg partitioning and pressures of 4.5-6 kb using
mineral assemblage data. However, a higher pres-
sure may be indicated by kyanite reported by D. W.
Mogk (pers. commun. in Guy and Sinha, 1988).
Microprobe analyses of coexisting garnet and biotite
in one specimen give temperatures of 730C using
the Ferry and Spear (1978) calibration of the Fe-Mg
exchange geothermometer, consistent with the
results of Guy and Sinha (1988). These high-grade
rocks are clearly highly aluminous and probably ori-
ginated as semi-pelitic to pelitic sediments. Quartz
grains within these gneisses show a well-developed
polygonal texture, indicating complete annealing
subsequent to the deformation associated with the
high-grade metamorphism.
A white augen gneiss sill intrudes biotite gneiss
at Yankee Jim Canyon. Microcline, quartz, biotite,
and sodic plagioclase dominate this unit, which is
clearly intrusive - containing xenoliths of the sur-
rounding biotite gneisses. Feldspar grain size is
variable, with finer grained samples exposed near
the plutonic contacts. The augen gneiss appears to
be truncated against the Snowy shear zone to the
northeast. The shear zone in this area is particu-
larly rich in feldspar clasts, probably due to the
abundance of feldspar in the augen gneiss.
Biotite-Staurolite Schist
The outcrops on Crystal Cross Mountain are
important since they expose the only rocks on the
hanging-wall of the Snowy shear zone showing no
evidence of the greenschist facies shearing which
characterizes the Snowy shear zone. These rocks
are well-bedded and rhythmically layered, grading
between pelitic and semi-pelitic compositions in a
fashion typical of turbidite deposits. Tops directions
were not clearly discerned on the outcrops, although
several exposures suggest younging to the north-
west. Important AFM assemblages include garnet
+ biotite + staurolite + andalusite. Garnet-biotite
geothermometry of two samples using rim garnet
compositions (Table 1) give temperatures of 553C
and 529C - which, in the presence of andalusite,
give a maximum pressure of "" 3 kb. Garnet zoning
profiles summarized in Table 1 show normal, pro-
grade zoning with Mn-rich and Mg-poor core com-
positions. Like the biotite gneisses in Yankee Jim
Canyon, mineral assemblages and preliminary geo-
chemistry point to a pelitic to semi-pelitic protolith
(Erslev, 1988). The ratio of pelitic to semi-pelitic
rocks appears to increase to the northwest, away
from the Snowy shear zone.
The contact with the chloritic amphibolite at
the base of Crystal Cross Mountain is partially
sheared and altered to greenschist-facies minerals.
However, alteration and shear fabrics are absent in
rocks on top of Crystal Cross Mountain, although
folds exists on all scales (e.g., Figure 3B). These
folds are defined by a micacaeous foliation that is
enclosed by post-tectonic staurolite crystals as sinu-
ous inclusion trails. These textures show that the
folds developed synchronous with prograde, middle
amphibolite-facies metamorphism.
Snowy Shear Zone
Between the high-grade gneisses and the
biotite-staurolite schists, the Snowy shear zone ex-
poses interfingering, variably sheared and altered
chlorite-biotite schist, amphibolite, metagabbro,
and quartzite. In the chlorite-biotite schist, a pelitic
to semi-pelitic lithology - the stable AFM assem-
F
i
g
u
r
e

3
.

A
)

P
h
o
t
o
m
i
c
r
o
g
r
a
p
h

o
f

c
o
r
d
i
e
r
i
t
e
-
b
i
o
t
i
t
e
-
g
a
r
n
e
t

g
n
e
i
s
s

f
r
o
m

Y
a
n
k
e
e

J
i
m

C
a
n
y
o
n

(
4

m
m
)
.
.

B
)

P
h
o
t
o
m
i
c
r
o
g
r
a
p
h

o
f

s
t
a
u
r
o
l
i
t
e

+

a
n
d
a
l
u
s
i
t
e

+

b
i
o
t
i
t
e

g
a
r
n
e
t

s
c
h
i
s
t

f
r
o
m

C
r
y
s
t
a
l

C
r
o
s
s

M
o
u
n
t
a
i
n

(
1
0

m
m
)
.

C

;
a
n
d

D
)

P
h
o
t
o
m
i
c
r
o
g
r
a
p
h
s

o
f
m
y
l
o
n
i
t
i
c

p
e
l
i
t
i
c

g
n
e
i
s
s

o
f
t
h
e

S
n
o
w
y

s
h
e
a
r

z
o
n
e

w
i
t
h

a
s
y
m
m
e
t
r
i
c

q
u
a
r
t
z

a
n
d

f
e
l
d
s
p
a
r

p
o
r
p
h
y
r
o
c
l
a
s
t
s

i
n
d
i
c
a
t
i
n
g

s
h
e
a
r

s
e
n
s
e

(
1
0

m
m
)
.

c
.
o

.
.
.
.
.
.

e
x
>

t
'
=
j


t
'
=
j

>
-
j

'
"

C
D

<
:

T
a
b
l
e

1
.

B
i
o
t
i
t
e
-
g
a
r
n
e
t

p
a
i
r

c
o
m
p
o
s
i
t
i
o
n
s

a
n
d

g
e
o
t
h
e
r
m
o
m
e
t
r
y
.

S
a
m
p
l
e

4
-
5
8

S
a
m
p
l
e

4
-
4
8

S
a
m
p
l
e

9
-
1
5
7

S
t
a
u
r
o
l
i
t
e

S
c
h
i
s
t

S
t
a
u
r
o
l
i
t
e

S
c
h
i
s
t

M
y
l
o
n
i
t
e

G
a
r
n
e
t

G
a
r
n
e
t

G
a
r
n
e
t

G
a
r
n
e
t

G
a
r
n
e
t

G
a
r
n
e
t

E
d
g
e

R
i
m

C
o
r
e

B
i
o
t
i
t
e

E
d
g
e

R
i
m

C
o
r
e

B
i
o
t
i
t
e

G
a
r
n
e
t

B
i
o
t
i
t
e

n
=
2

n
=
6

n
=
5

n
=
7

n
=
2

n
=
4

n
=
5

n
=
8

n
=
3

n
=
2

S
i
0
2

3
6
.
7
7

3
7
.
2
2

3
6
.
6
6

3
4
.
6
0

3
8
.
0
8

3
7
.
8
9

3
7
.
9
4

3
4
.
8
8

3
6
.
9
0

3
4
.
5
2

A
1
2
0
3

2
1
.
1
7

2
1
.
3
7

2
1
.
3
3

1
9
.
9
6

2
1
.
4
6

2
1
.
5
0

2
1
.
5
3

2
4
.
3
2

2
0
.
3
6

1
7
.
7
0

T
i
0
2

0
.
0
5

0
.
0
5

0
.
0
8

1
.
6
9

0
.
0
7

0
.
0
6

0
.
1
0

1
.
4
7

0
.
1
2

1
.
6
2

F
e
O

3
2
.
5
0

3
2
.
2
5

3
0
.
9
8

1
7
.
0
5

3
1
.
7
7

3
0
.
8
5

2
9
.
4
4

1
6
.
0
8

2
8
.
7
5

2
2
.
3
2

M
g
O

2
.
8
9

3
.
6
7

3
.
5
1

1
1
.
5
4

3
.
1
2

3
.
3
1

2
.
9
2

1
0
.
1
1

1
.
3
7

8
.
2
4

M
n
O

4
.
4
7

3
.
1
2

3
.
5
8

0
.
0
1

3
.
3
9

3
.
1
6

4
.
0
6

0
.
0
3

5
.
5
2

0
.
1
7

K
2
0

0
.
0
5

0
.
0
3

0
.
0
3

9
.
7
2

0
.
0
4

0
.
0
3

0
.
0
3

8
.
2
2

9
.
3
1

N
a
2
0

0
.
0
1

0
.
0
2

0
.
0
2

0
.
2
1

0
.
0
1

0
.
0
0

0
.
0
2

0
.
1
8

0
.
0
5

C
a
O

2
.
4
8

2
.
9
3

3
.
7
9

0
.
0
3

3
.
3
1

3
.
8
3

4
.
9
6

0
.
0
1

7
.
4
1

0
.
2
1

T
o
t
a
l

1
0
0
.
3
7

1
0
0
.
6
5

9
9
.
9
7

9
4
.
7
9

1
0
1
.
2
4

1
0
0
.
6
2

1
0
0
.
9
9

9
5
.
2
9

1
0
0
.
4
3

9
4
.
1
2

M
g
l
F
e

0
.
0
8
9

0
.
1
1
4

0
.
1
1
3

0
.
6
7
7

0
.
0
9
8

0
.
1
0
7

0
.
0
9
9

0
.
6
1
5

0
.
0
4
8

0
.
3
6
9

T
O
C

4
8
0

5
5
3

5
5
1

5
0
3

5
2
9

5
0
5

4
7
6

T
O
C
:

a
v
e
r
a
g
e

t
e
m
p
e
r
a
t
u
r
e
s
,

c
a
l
i
b
r
a
t
i
o
n

f
r
o
m

F
e
r
r
y

a
n
d

S
p
e
a
r

(
1
9
7
8
)

u
s
i
n
g

a

p
r
e
s
s
u
r
e

o
f

3

k
b
.

S
a
m
p
l
e
s

4
-
4
8

a
n
d

4
-
5
8

a
r
e

f
r
o
m

t
h
e

t
o
p

o
f
C
r
y
s
t
a
l

C
r
o
s
s

M
o
u
n
t
a
i
n
.

S
a
m
p
l
e

9
-
1
5
7

i
s

f
r
o
m

E
m
i
g
r
a
n
t

G
u
l
c
h
.

'
l
j

;
t
i

(
"
.
)


g
.

>
-o
j
>
"
.


o

0
"


0
.

0
.

[

>
"
.

c
o


o

8

e
.

S
'

2
.

c
+


(
!
)

t
o

(
!
)


:
+

o

o

c
+

:
:
r

c
:
:
:
:


>
"
.


c
:
.
o

.
.
.
.
.
.

<
0

320
E. A. Erslev
blage - is biotite + chlorite, with Mn-rich garnet
occurring in one sample. Electron microprobe
analyses of this sample (9-157, Table 1) yielded an
average temperature of 476C. The clear reaction
textures showing chlorite growth at the expense of
garnet, the high Mn composition of the garnet, and
the disequilibrium texture of the rock all suggest
that this value should be interpreted as a maximum
temperature during shearing. Pelitic/semi-pelitic
rocks commonly are either very fine grained or show
excellent asymmetric porphyroclasts, C-S foliations,
and/or feldspar augen (Figure 3C,D) . The fine-
grained rocks occur in the center of the shear zone
and contain no primary structures, suggesting that
they are the mylonitic equivalents of the higher
grade rocks on the shear zone margins. This inter-
pretation is contrary to that of Burnham (1982), who
considered them to be a separate group of low-grade
metasedimentary rocks.
The quartzites and amphibolites occur together
with minor interlayers of pelitic material. The
amphibolites contain actinolite or blue-green horn-
blende, plagioclase, chlorite, and quartz. Localized
pods of sheared metagabbro suggest a plutonic pro-
tolith for some of the amphibolites. However, am-
phibolite interlayered with quartzite appears to be
either extrusive or metasedimentary. One distinc-
tive amphibolite lithology contains stretched quartz
segregations up to a few centimeters in length. The
quartzites are commonly extremely monomimeralic,
with some biotite, muscovite, and plagioclase inter-
layers. Texturally, these quartzites range from
proto-mylonites with excellent asymmetric porphy-
roclasts to totally recrystallized, medium-grained
metamorphic quartzites. They are quite discon-
tinuous on a local scale, suggesting complex folding
and/or faulting.
STRUCTURAL GEOLOGY
Geologic mapping clearly shows the intrusive
nature of the augen gneiss and the correlation
between the greenschist facies assemblages and the
shear fabrics of the Snowy shear zone. These fabrics
were studied in detail to elucidate the kinematics of
the shear zone.
Poles to foliations and fold axes from the three
lithotectonic domains are plotted in Figure 4.
Foliations from the Snowy shear zone (Figure 4B)
are very well clustered, showing an average N53E-
46NW strike and dip orientation based on eigen-
vector analysis. This uniformity of foliation is seen
in other major shear zones like the Madison mylo-
Fabric Data From the SW Corner of the Beartooth Uplift
Foliations
Figure 4. Equal-area, lower-hemisphere
stereonet projections.
A, B, C: poles to foliation;
D: fold axes;
E: shear sense determinations from
oriented thin sections.
Contour intervals are 1%,3%, and 6% of
the data per 1% area.
A B c
NW of Sh'OI' Zou (82) Sh .... Zon. (300) Yonk ." W.tomo<phlc Compl (I H)
o Fol d Axes E Shear Sense Indicators
.... , Domoln. (151)

- X u-.eot.loft
_ WoO
51>_ Zone (,.) T' ........ 0"',,_ (X)
Precambrian geology and ductile normal faulting in the Beartooth Uplift, Montana 321
nite zone to the southwest (Erslev, 1983), because
simple shear deformation tends to rotate foliations
toward parallelism with the plane of shear.
Foliations from the domains adjacent to the
Snowy shear zone give great-circle distributions
about similar poles (20, N32E for Figure 4A and
19, N25E for Figure 4C based on eigenvector
analysis). This corresponds very closely to the aver-
age eigenvector orientation of 20, N37E for the fold
axes in Figure 4D, showing a well-defined homo-
tactic fabric common to both sides of the Snowy
shear zone. While it is tempting to attribute this
rotation and folding of foliation to the shear zone,
many of the folds, like those shown in Figure 3B,
originated during prograde metamorphism - not
greenschist-facies retrograde metamorphism. At
Crystal Cross Mountain, these folds are commonly
asymmetrical, with 19 out of 22 folds verging to the
southeast. Since these folds pre-date the growth of
spectacular staurolite crystals, they suggest earlier,
probably Archean, crustal shortening characterized
by northwest over southeast shearing.
Fabric asymmetries indicative of shear sense in
the Snowy shear zone were studied to determine the
tectonic motion within the shear zone. Thin sections
from 14 oriented samples were prepared parallel
and perpendicular to the major lineation. The maxi-
mum stretch (X) directions were determined by
observing grain ellipticity in the two sections. The
sections containing the X direction were carefully
observed for evidence of diagnostic asymmetrical
fabrics using the criteria outlined by Simpson and
Schmid (1982).
The results of this analysis are shown in Figure
4E. In all but one sample, the dip-slip component of
motion is normal, with the hanging wall moving
down-dip. A component of right-lateral strike-slip
motion is seen in all but one sample. '!'he rocks with
the most pronounced shear fabrics show primarily
normal slip, with less sheared rocks showing more
oblique slip. This may be the result of the com-
bination of earlier fold-axis parallel elongation seen
in the schists from Crystal Cross Mountain with
later normal fault fabrics or two stages of motion in
the shear zone.
CONCLUSIONS
The similarities in the pelitic/semi-pelitic com-
positions and fabrics from both the schists at Crystal
Cross Mountain and the gneisses at Yankee Jim
Canyon indicate a common geologic history. The
difference in metamorphic equilibration conditions
between these areas of == 200C and 1.5-3 kb is con-
sistent with ductile normal faulting in the Snowy
shear zone. This kinematic history is supported by
the oblique, part normal and part right-lateral
motion indicated by fabrics within the shear zone.
Thus, earlier hypotheses indicating thrusting and
crustal shortening in the shear zone (Erslev, 1982)
are not substantiated by this study.
The time of shearing is uncertain. It seems
clear that a substantial interval must have sepa-
rated the prograde crystallization of the pelitic rocks
and the subsequent mylonitization. According to
Guy and Sinha (1988), a later isotopic disturbance
has effected the Rb-Sr and Pb-Pb systems in Yankee
Jim Canyon.
Recent documentation of a 1800-1900 Ma age
for the Madison mylonite zone (Erslev and Sutter,
1990) and similar ages of metamorphism and de-
formation to the northwest (O'Neill and others,
1988) suggest the possibility that the Snowy shear
zone formed during the Early Proterozoic. The
textures in the Madison mylonite zone are distinctly
more thoroughly recrystallized than those in the
Snowy shear zone, which are more consistent with
the rapid cooling associated with extensional mylo-
nites. If these areas went through comparable ther-
mal histories during the Early Proterozoic, then the
mylonites from the Snowy shear zone must have
formed after the deformation and recrystallization
of the Madison mylonite zone. This hypothesis is
supported by K-Ar ages reported by Reid and others
(1975) for the North Snowy Block, which average
1780 170 Ma ( 1 standard deviation). These ages
could be the result of the uplift of an ancestral
Beartooth block by normal faulting along the Snowy
shear zone during the Early Proterozoic. The paral-
lelism of both the compressional (Madison mylonite
zone) and extensional (Snowy shear zone) zones in
the region suggest a genetic relationship. Gravita-
tional spreading of the orogenic welt created by
earlier crustal compression may have been accom-
plished by extension in the Snowy shear zone.
This study has shown a complex, polyphase
history in which late-stage oblique normal faulting
appears to be the last event. Earlier events and the
overall timing of the deformation continue to be
obscured from our view. But the identification of
pristine and reworked domains should aid further
studies in the Precambrian development of the
Beartooth basement complex.
322
E. A. Erslev
ACKNOWLEDGMENTS
This study has benefited by discussions with
Scott Miller, David Fountain, and Ana Vargo, as
well as helpful reviews by J. Michael O'Neill,
Christopher Schmidt, and David Mogk.. Karen
W ogsland and Kristin Gunckel provided excellent
field assistance. John C. Reed, Jr., and Ralph Chris-
tenson provided access and assistance with the
ARL-SEMQ electron microprobe at the U.S. Geo-
logical Survey in Denver, Colorado. Acknow-
ledgment is made to the Donors of The Petroleum
Research Fund, administered by the American
Chemical Society, for the support of this research.
REFERENCES
BROOKINS, G., 1968, Rb-Sr and K-Ar age determinations
from the Precambrian rocks of the Jardine-Crevase Moun-
tain area, southwestern Montana: Earth Science Bulletin,
v.1, p. 5-9.
BURNHAM, R., 1982, Mylonite Zones in the Crystalline
Basement Rocks of Six Mile Creek and Yankee Jim
Canyon, Park County, Montana: MS thesis, University
of Montana, Missoula, Montana, USA, 93 p.
ERSLEV, E. A., 1982, The Madison mylonite zone: A major
shear zone in the Archean basement of southwestern
Montana; pp. 213-221 in S. G. Ried (ed.), Geology of the
Yellowstone Park Area: 33rd Annual Field Confer-
ence Guidebook: Wyoming Geological Association, Cas-
par, Wyoming, USA, 221 p.
__ , 1983, Pre-Beltian Geology of the Southern Madi-
son Range, Southwestern Montana: Montana Bureau of
Mines and Geology, Memoir 55, 26 p.
__ , 1986, Basement balancing of Rocky Mountain fore-
land uplifts: Geology, v. 14, p. 259-262.
__ , 1988, Field guide to pre-Beltian geology of the
southern Madison and Gravelly ranges, southwest
Montana; pp. 141-150 in S. E Lewis and R. B. Berg (eds.),
Precambrian and Mesozoic Plate Margins: Montana
Bureau of Mines, Special Publication 96, 195 p.
__ , and J. SUTTER, 1990, Evidence for Proterozoic
mylonitization in the northwestern Wyoming Province:
Geological Society of America Bulletin, v. 102, p. 1681-
1694.
FERRY, J. M., and F. S. SPEAR, 1978, Experimental cali-
bration of the partitioning of Fe and Mg between biotite
and garnet: Contributions to Mineralogy and Petrology, v.
66, p. 113-117.
GUY, R. E., and A. K. SINHA, 1988, Petrology and geo-
chemistry of Archean basement lithologies in the Yankee
Jim and Lamar Canyon areas, Montana and Wyoming; pp.
53-68 in S. E Lewis and R. B. Berg (eds.), Precambrian
and Mesozoic Plate Margins: Montana Bureau of Mines,
Special Publication 96, 195 p.
JAMES, H. L., and C. E. HEDGE, 1980, Age of basement
rocks of southwest Montana: Geological Society of America
Bulletin, v. 91, p. 11-15.
MOGK, D. W., 1983, The Petrology, Structure, and Geo-
chemistry of an Archean Terrane in the North Snowy
Block, Beartooth Mountains, Montana: PhD disserta-
tion, University of Washington, Seattle, Washington,
USA,236p.
__ , 1988, Archean allochthonous units in the northern
and western Beartooth Mountains, Montana; pp. 53-52 in
S. E Lewis and R. B. Berg (eds.), Precambrian and
Mesozoic Plate Margins: Montana Bureau of Mines,
Special Publication 96, 195 p.
__ , P. A. MUELLER, and J. L. WOODEN, 1988, Archean
tectonics of the North Snowy Block, Beartooth Mountains,
Montana: Journal of Geology, v. 96, p. 125-141.
MUELLER, P. A., R. D. SHUSTER, M. A. GRAVES, J. L.
WOODEN, and D. R. BOWES, 1988, Age and composition of
a late Archean magmatic complex, Beartooth Mountains,
Montana-Wyoming; pp. 7-22 in S. E Lewis and R. B. Berg
(eds.), Precambrian and Mesozoic Plate Margins: Mon-
tana Bureau of Mines, Special Publication 96, 195 p.
MONTGOMERY, C. W., and J. W. LYTWYN, 1984, Rb-Sr
systematics and ages of principal Precambrian lithologies
in the South Snowy Block, Beartooth Mountains: Journal
of Geology, v. 92, p. 103-112.
O'NEILL, J. M., M. S. DUNCAN, and R. E. ZARTMAN, 1988,
An Early Proterozoic gneiss dome in the Highland Moun-
tains, southwestern Montana; pp. 81-89 in S. E Lewis and
R. B. Berg (eds.), Precambrian and Mesozoic Plate Mar-
gins: Montana Bureau of Mines, Special Publication 96,
195p.
REID, R. R., W. J. McMANNIS, and J. L. PALMQUIST,
1975, Precambrian Geology of the North Snowy Block,
Beartooth Mountains, Montana: Geological Society of
America, Special Paper 157,135 p.
SIMPSON, C., and S. M. SCHMID, 1983, An evaluation of
criteria to deduce the sense of motion in sheared rocks:
Geological Society of America Bulletin, v. 94, p. 1281-1288.
SUMNER, W., and E. A. ERSLEV, 1988, Late Archean thin-
skinned thrusting of the Cherry Creek metamorphic suite
in the Henry's Lake Mountains, southern Madison Range,
Montana and Idaho; pp. 7-22 in S. E Lewis and R. B. Berg
(eds.), Precambrian and Mesozoic Plate Margins: Mon-
tana Bureau of Mines, Special Publication 96, 195 p.
VITALlANO, C. J., W. S. CORDUA, D. F. HESS, H. R.
BURGER, T. B. HANLEY, and F. K. ROOT, 1979, Explana-
tory Text to Accompany Geologic Map of the Southern
Tobacco Root Mountains, Madison County, Montana:
Geological Society of America, Map and Chart Series MC-
31.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Geochemistry and origin of amphibolite and ultramafic rocks,
Branham Lakes area, Tobacco Root Mountains,
southwestern Montana
MICHAEL L. CUMMINGS and WILLIAM R. McCULLOCH
Department of Geology, Portland State University, Portland, Oregon 97207, USA
(received September 12, 1988; revision accepted April 12, 1989)
ABSTRACT
An Archean amphibolite unit, composed of amphibolite, garnet amphibolite, and meta-ultramafic rocks,
is exposed in the Branham Lakes area, Tobacco Root Mountains, southwestern Montana. These lithologies
are separated from other Archean lithologic associations by shear zones and an inferred thrust fault system
and they are intruded by the Cretaceous Tobacco Root Batholith.
The meta-ultramafic rocks in the amphibolite unit form either layers concordant to subconcordant to
compositional layering in amphibolite, or pods where layering is disrupted by shear zones. Hornblende-rich
blackwalls occur at the contacts between the amphibolite and meta-ultramafic layers and along shear zones
that disrupt layers. Amphibolite from the amphibolite unit is geochemically and texturally distinct from thin
amphibolite layers in another Archean unit, the intermediate gneiss and schist unit. Meta-ultramafic rocks
contain 45-55% Si0
2
, 3-7% A1
2
0
3
, 0.19-0.31% Ti0
2
, 22-32% MgO, 0.01-0.54% P
2
0
5
, and 1300-4000 ppm Cr;
CaO/MgO == 0.11-0.33, LaiCe == 0.14-0.56; La/Sm = 2.88-6.02.
The data distribution on Ti-Zr, Ti-Zr-Sr, and Ti-Zr-Y diagrams for amphibolite from the amphibolite unit
are consistent with low-K or ocean-floor tholeiites, or calcalkaline basalts. On a Nb-Y-Zr plot, the data for
amphibolites are consistent with N-type MORB or volcanic-arc basalts. The amphibolite unit may be a sliver
of oceanic crust tectonically juxtaposed against shelf sediments and the protolith of the intermediate gneiss
and schist unit.
INTRODUCTION
Archean age rocks of the Wyoming Province
(Condie, 1969) of the North American shield are
exposed in southwestern Montana within basement
uplifts in the foreland fold and thrust belt. Geo-
logic, geochemical, and geochronologic studies with-
in these uplifts indicate two major terranes within
the province. Mogk and others (1988a) and Mogk
(1988) propose that the North Snowy Block of the
northwestern Beartooth Mountains is a fundamen-
tal boundary between these two distinct terranes.
The eastern terrane in the Beartooth and Bighorn
Mountains (Figure 1) is underlain by late Archean
granitoids that contain blocks of older supracrustal
rocks. These older rocks yield a Rb-Sr isochron of
3390 55 Ma, an age interpreted to be that of
granulite facies metamorphism (Henry and others,
1982). The granitoid data fit isochrons between
2740 and 2780 Ma (Wooden and others, 1988). West
of the North Snowy Block, the terrane contains
high-grade metamorphic supracrustal rocks exposed
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 323-340. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
323
324 M. 1. Cummings and W. R. McCulloch
N
t
Montana
" . . '
Wyoming
Bi!ji'lorn Mountains 1_:'
so 100

o T .ton Ronge
Wind Range
Figure 1. Distribution of Archean basement blocks within
Montana and northern Wyoming (modified from Mogk and
others, 1988a).
in the Gallatin, Madison, Ruby, and Tobacco Root
Mountains. The final tectonism in the Madison and
Beartooth Ranges is marked by crustal shortening
and emplacement of granite intrusions between
2600 and 2550 Ma (Mogk and others, 1988b).
The Tobacco Root Mountains (TRM) are under-
lain by high-grade supracrustal rocks and, along
with the Highland Mountains to the west (O'Neill
and others, 1988) and the Ruby Range to the south-
west, contain the westernmost Archean-age crust in
the Wyoming Province. The Branham Lakes area
in the southern Tobacco Root Mountains is under-
lain by four Precambrian lithologic associations.
Our studies focus on metasomatism occurring be-
tween amphibolite and meta-ultramafic lithologies
within an amphibolite unit (McCulloch, 1989;
McCulloch and Cummings, 1987, 1988); however,
we have examined, on a reconnaissance basis, the
other lithologic associations that occur in the Bran-
ham Lakes cirque. In this paper, we describe the
lithologies and field relations among lithologic
associations in the Branham Lakes cirque and the
geochemistry of amphibolites from the various litho-
logic associations, and we propose a metamorphic
and tectonic history for the area.
GENERAL SETIING
The Archean rocks of the Tobacco Root Moun-
tains have long been separated into the Cherry
Creek Group (Runner and Thomas, 1928), the Pony
Group (Tansley and others, 1933), and the Spuhler
Peak Formation (Gillmeister, 1972). However,
Vitaliano and others (1979) argued that these strati-
graphic names "should be held in abeyance." The
lack of stratigraphic indicators for the group names
was cited as the reason for their advice.
These Archean rocks have a polydeformational
and polymetamorphic history. The apparent highe-
st grade of metamorphism occurred under condi-
tions of the upper amphibolite facies at 2667 66
Ma (Mueller and Cordua, 1976, Rb/Sr isochron) .
The conditions of metamorphism were determined
from iron formations by IInmega and Klein (1976)
for the southern Tobacco Root Mountains and by
Dahl (1979) for the northern Ruby Range. The
metamorphic conditions were 650-750C and 4-6 kb
in the Tobacco Root Mountains and 745 50C in
the Ruby Range. Friberg (1976) reported metamor-
phic conditions of 600-800C and 4-6 kb in the Sphu-
ler Peak area where conditions exceeded those of the
sillimanite-orthoclase isograd. Metamorphism un-
der almandine-amphibolite subfacies (Giletti, 1966;
Vitaliano and others, 1979) and under greenschist
facies (Vitaliano and others, 1979) conditions occur-
red at '"" 1,600 Ma and during post-Precambrian
time.
In addition to the geologic compilation and
mapping reported by Vitaliano and others (1979),
the geology of the Branham Lakes area has been
studied by Burger (1967), Hess (1967), and McCul-
loch (1989). Burger (1967) reported that inter-
layered amphibolite, hornblende gneiss, anthophyl-
lite gneiss, and quartzite were separated from an
intermediate gneiss assemblage by the Noble Fault.
However, no fault plane or shear zone was recog-
nized.
GEOLOGY OF THE BRANHAM LAKES CIRQUE
Four Precambrian lithologic associations sepa-
rated by structural boundaries or intrusive contacts
occur in the Branham Lakes cirque (Figure 2).
These lithologic associations are intruded by the
Tobacco Root Batholith and its satellite intrusions.
The intrusion of the Tobacco Root Batholith
occurred at 77-72 Ma (Giletti, 1966; Vitaliano and
others,1980). The Precambrian lithologic associa-
tions (Figure 2) include the following: an amphibo-
lite unit (Aa), an interlayered schist and quartzite
unit (Ams), an interlayered intermediate gneiss and
amphibolite unit (Aig), and a layered mafic intru-
sion(Ami).
Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana 325

0.0
: : ::
.'. '.:'; Ki - Tobacco Root Batholith
Ami - Layered Mafic Intrusion
Aig - Intermediate Gneiss and
Schist
j
,-,
18.5
V
1.0 mi
:5; II?
11 ,:-:, Aa - Amphibol ite
Ams - Interlayered
Quartzite and Schist
Shear Zones
Inferred Faults
Figure 2. Geologic map of the Branham Lakes cirque.
The Aa unit underlies ""4.5-7.0 km
2
of the
Branham Lakes cirque (Figure 2). Three litho-
logies occur in the Aa unit: amphibolite, garnet-
amphibolite, and meta-ultramafic layers and pods.
Within the amphibolite, compositional layering is
produced by modal variations of hornblende and
plagioclase and, in some layers, garnet (Table 1).
The modal percentage of hornblende is 10-40% in
light-colored layers and 70-90% in dark-colored
layers. Garnet-amphibolite is sparse; where pre-
sent, garnet makes up :5 35% of the rocks. Clino-
pyroxene makes up :510% in some samples. The
326 M. L. Cummings and W. R. McCulloch
Table 1. Modal data for amphibolite and blackwall from the amphibolite (Aa) unit.
Sample Total
Numbers Hbl Cpx Pc Ap Qtz Gar Tita Clz Op Unk Points
Amphibolite
85-34 65.8 33.6 0.4 0.2 500
85-35 69.6 24.2 0.4 1.0 0.2 1.3 0.4 2.9 1000
85-92-1 65.8 31.4 0.4 2.4 tr. 500
86-19 64.6 35.0 0.2 0.2 tr. 500
86-23 66.2 3.8 29.6 0.2 0.2 500
86-25d 56.0 7.8 35.6 0.2 0.4 500
Sample Total
Numbers Hbl Pc Ap Qtz Trem Points
BwckweU
85-31A 99.6 0.4 500
86-15 99.0 0.8 0.4 tr. 500
85-67 99.0 0.4 0.6 500
Sample Mg- Mg- Total
Numbers Opx 01 Hbl Trem Tc ChI Op Unk Points
Meta-
mtrama/ic
85-87E 3.5 94.0 2.0 0.5 200
85-39 75.6 23.6 0.2 0.6 500
85-41 0.8 18.8 69.7 10.7 1000
85-113 66.0 0.6 27.2 5.8 0.4 500
86-74 74.0 16.8 9.2 250
Hbl, hornblende; Cpx, clinopyroxene; Pc, plagioclase; Ap, apatite; Qtz, quartz; Gar, garnet; Tita, titanite; Clz, clino-
zoisite; Op, opaques; Trem, tremolite; Opx, orothopyroxene; 01, olivine; Mg-Hbl, Mg-hornblende;
Tc, talc; Mg-Chl, Mg-chlorite; Unk, unknown; tr, trace.
texture of the amphibolite is granoblastic; clino-
pyroxene, hornblende, plagioclase, and quartz aver-
age 2 mm in diameter.
Foliation is weakly developed to absent, but
boudins and isoclinal folds of compositional layering
are present. Axial surfaces offolds and the apparent
long dimension of boudins generally are subparallel
to compositional layering.
Equidimensional to elongate bodies of meta-
ultramafic rocks occur within the amphibolite (Fig-
ure 3). Thicknesses of these bodies range from 10 to
20 meters, and lengths range from 10 to a few
hundred meters. The longer meta-ultramafic bodies
are conformable to compositional layering in the
amphibolite. Three types of meta-ultramafic bodies
are distinguished by modal abundances of ortho-
pyroxene and olivine, those that contain (1) pre-
dominantly orthopyroxene, (2) predominantly oli-
vine, and (3) approximately equal proportions of
olivine and orthopyroxene. Chromite is a common
accessory.
In meta-ultramafic rocks containing similar
modal percentages of olivine and orthopyroxene
megacrysts, olivine was optically determined to be
Fo85, and orthopyroxene to be En85-87' The mega-
crysts are fractured and display highly irregular
Figure 3. Geologic map of parts of amphibolite unit (Aa) showing shear zones and meta-ultramafic bodies. Qd, Glacial
deposits; Ki, Tobacco Root Batholith and associated intrusions; Ami, layered mafic intrusion; Aa, amphibolite unit;
Aig, intermediate gneiss and schist unit; Ams, interlayered quartzite and schist unit.
Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana
Aig
+
v
0.1
Qd Glacial depolltl T I. I u I
KI Tobacco Root Batholith and al.oclated Intrusion.
Ami Layered mafic Intrullon
Aa Amphibolite Unit
amphibolite
meta-ultramafic
Aig Intermediate gnel.1 and .chllt Unit
Ami Interlayered quartzite and Ichllt Unit
zone.

2P-Strlke and dip of foliation
Sample locality (Table 2)
327
328 M. L. Cummings and W. R. McCulloch
grain boundaries. Mg-hornblende replaces orthopy-
roxene in rocks composed of Mg-hornblende.
Contacts between amphibolite and meta-ultra-
mafic layers are invariably marked by hornblende-
rich mineral rinds that contain modal traces of apa-
tite and plagioclase. Field relations, mineralogy,
and mineral assemblages of the rinds have been
described by McCulloch (1989). Although the rinds
obscure detailed contact relations, where the meta-
ultramafic rock is concordant to compositional
layering in amphibolite the contacts are of two
types. In the first type, rinds grade outward from a
sharp contact with meta-ultramafic to a gradational
contact with amphibolite. The percentage and grain
size of hornblende decrease into the amphibolite.
Large megacrysts of orthopyroxene, $15 cm "long,
occur within the meta-ultramafic at the contact. In
the second type of concordant contact, hornblende-
rich apophyses form sharp contacts where they ex-
tend into amphibolite. Similar apophyses contain-
ing Mg-hornblende extend into meta-ultramafic
rock.
Highly discordant contacts between meta-
ultramafic and amphibolite are produced by shear
zones that occur at high angles to compositional
layering in the amphibolite. Hornblende-rich bands
are developed along the contact between amphi-
bolite and meta-ultramafic and along extensions of
shear zones into amphibolite.
Amphibolite occurs within the other lithologic
associations of Archean age that crop out in the
Branham Lakes cirque. Amphibolite layers occur
within the Ams unit on the north and east cirque
walls. Amphibolite on Branham Peaks occurs as
slices within schist and quartzite (Figure 2). Large-
scale folds occur within the interlayered sequence
immediately above the contact with the Aa unit
(Figure 4). In the folds, amphibolite layers pinch
out along apparent faults that juxtapose the
amphibolite against lenses and layers of quartzite,
sulfide-rich and anthophyllite-bearing rocks, and
pelitic schists. The fine-scale layering in the amphi-
bolite is discordant to inferred faults. Amphibolite
occurs in elongate, concordant lenses higher on the
flanks of Branham Peaks where the large-scale folds
are absent. In this area, one apparent fault surface
separates tightly folded metasedimentary rocks
from underlying medium-grained amphibolite
wherein compositional layering is not folded.
Overall, except in areas offolding, the orientation of
compositional layering and contacts between com-
positional types is similar to that in the Aa unit.
Amphibolite is a minor lithology in the Aig unit
exposed on the south and west walls of the cirque.
Amphibolite layers are usually only a few meters
wide. Amphibolites display fine-scale mineral
streaking rather than well-developed compositional
layering.
SHEAR ZONES
Shear zones are an important lithologic sub-
group within the Aa unit and also separate the Aa
unit from the Aig unit. McCulloch (1989) examined
hornblende-rich lithologies developed along shear
zones within the Aa unit that cut meta-ultramafic
and amphibolite lithologies. At least two sets of
shear zones are present. A N/S-oriented set is cut by
a N500E-trending set. Hornblende-rich lithologies
are developed along shear zones in both orienta-
tions. The distribution of these shear zones pro-
duces the discontinuous and pod-like character of
meta-ultramafic bodies.
Other large-scale shear zones also occur within
the Aa unit. In one area, a prominent N60o/75W-
trending shear zone contains tectonic blocks of
meta-ultramafic and iron formation. Granitic peg-
matites and basalt dikes within the zone are tightly
folded and boudins are common. Silicification is
prominent, and coarse-grained clots of sillimanite
are present. This zone is at least 100 m wide. A se-
cond large-scale shear zone contains sheared horn-
blende-rich lithologies. Quartz veins and pegmatite
dikes generally are absent from this zone.
Between the Aig unit and the Aa unit along the
western wall of the cirque is a broad N50o/75W-
striking, 50
o
/60
o
N-dipping shear zone (Figure 2).
Tectonite fabrics are well developed and include
rootless intrafolial folds, feldspar augen, boudins,
and folds with sheared axial surfaces. The width of
the zone is ;:::: 300 m on the south flank of Leggat
Peak. Blocks of the Aig unit occur as rotated tec-
tonic blocks within the shear zone.
GEOCHEMISTRY
Amphibolite samples from the Aa, Aig, and
Ams units were analyzed by x-ray fluorescence at
Washington State University (Tables 2 and 4) and
were also analyzed by neutron activation (INAA) at
Portland State University (Table 3). Samples were
analyzed at 5, 15, and 150 days after irradiation.
Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana 329
Figure 4. South-facing flank of Branham Peak viewed across Upper Branham Lake. Light-colored rocks are quartzite. Top of
the lower tree-covered area is approximate location of possible thrust that separates amphibolite (Aa) and metasedimentary
(Ams) units.
Part of the data (columns 1-6, Table 2) are from
amphibolite samples (columns 1-9, Table 2) from
the Aa unit; these plot within a relatively restricted
range on the variation diagrams (Figure 5). For
these six samples, Si0
2
is 48-51% and, in five of the
six samples, Ti0
2
is 0.40-0.70%. The dendrogram
(Figure 6) determined by cluster analysis (Davis,
1986) of major-element concentrations, indicates
little difference for composition of samples in
columns 1 to 5. The sample in column 6 falls outside
this group. On the basis of the REE plots (Figure
7), two groups of samples are indicated. In one
group, REE plots are flat and a weakly developed
positive Eu anomaly may be present. The second
group shows enrichment in LREE.
Samples were collected from across the strike of
the compositional layering and throughout the Aa
unit (Figure 3). Compositions of three samples
(columns 7-9) plot outside the general field defined
by other amphibolite samples. For these three
samples, data points for at least two major elements
are distinct from other samples (Figure 5), as
confirmed by the dendrogram (Figure 6). Field
relations for these samples are as follows. Sample
86-14 was collected near a hornblende-rich black-
wall that separates a meta-ultramafic pod from
amphibolite. Sample 86-72 is near the intrusive
contact between the Aa unit and the Tobacco Root
Batholith. Sample 86-99 is a garnet-rich amphibo-
lite associated with garnet gneiss within a shear
zone along the southern edge of the cirque.
Two samples of amphibolite (Table 2, columns
10 and 11) were collected from outcrops on Branham
Peaks at an elevation above folds located at the base
of the peaks. At this elevation, amphibolite, schist,
and quartzite have locally concordant contact
relations but are believed to be fault-bounded slices.
Amphibolite in column 10 is geochemically similar
to those in columns 1-5, whereas amphibolite in
column 11 plots with those in columns 7-9 (Figure
6). Sample 86-28 contains higher concentrations of
CaO and FeO, and the Fe/(Fe+Mg) ratio is among
the highest in the sample suite.
On most variation diagrams compositions of
four samples of amphibolite from the Aig unit
(Table 4) form a group that is separate from data
points for amphibolite from the Aa unit (Figure 5).
This is also indicated in the dendrogram (Figure 6),
C
o
l
u
m
n
:

1

S
a
m
p
l
e
:

8
6
-
1
9

S
i
0
2

5
1
.
8
7

A
l
2
0
3

1
5
.
3
5

T
i
0
2

0
.
4
0

F
e
O
*

7
.
7
9

M
n
O

0
.
3
0

C
a
O

1
1
.
5
2

M
g
O

1
0
.
3
6

K
2
0

0
.
3
8

N
a
2
0

1
.
1
9

P
2
0
5

0
.
0
2
2

F
e
f
(
F
e
+
M
g
)

0
.
4
9

N
i

1
4
7

C
r

3
5
7

S
c

3
6

V

1
6
6

B
a

9
1

R
b

4

S
r

1
9
5

Z
r

5
3

Y

1
1

N
b

1

G
a

1
5

C
u

6
7

Z
n

5
4

T
a
b
l
e

2
.

G
e
o
c
h
e
m
i
s
t
r
y

o
f
a
m
p
h
i
b
o
l
i
t
e
s

f
r
o
m

t
h
e

a
m
p
h
i
b
o
l
i
t
e

u
n
i
t
.

2

3

4

5

6

7

8

9

8
6
-
2
1

8
6
-
2
5
D

8
6
-
2
3

8
5
-
3
4

8
5
-
3
5

8
6
-
7
2

8
6
-
9
9

8
6
-
1
4

4
9
.
8
8

5
0
.
1
1

4
9
.
1
9

4
8
.
6
8

5
0
.
4
9

4
7
.
7
2

4
8
.
0
7

4
5
.
8
3

1
6
.
1
5

1
4
.
4
5

1
6
.
0
7

1
6
.
4
8

1
3
.
8
7

2
1
.
1
1

1
7
.
0
2

1
7
.
5
5

0
.
6
9
3

0
.
6
3
1

0
.
4
3

0
.
4
9

1
.
8
5

0
.
3
9
6

1
.
7
0
0

0
.
4
2
1

1
0
.
6
9

1
0
.
2
2

8
.
4
5

9
.
2
7

1
3
.
2
6

6
.
9
5

1
1
.
9
8

8
.
2
6

0
.
2
8
1

0
.
2
6
2

0
.
2
7
0

0
.
2
6

0
.
2
6

0
.
1
7
1

0
.
3
8
8

0
.
2
4
4

1
2
.
3
4

1
2
.
7
9

1
3
.
6
5

1
2
.
4
9

1
0
.
6
3

1
3
.
1
8

1
3
.
1
5

1
5
.
9
0

7
.
2
4

9
.
2
4

1
0
.
2
9

9
.
8
1

6
.
6
0

8
.
4
3

4
.
9
2

1
0
.
2
0

0
.
3
3

0
.
3
5

0
.
1
7

0
.
2
2

0
.
4
3

0
.
2
3

0
.
8
6

0
.
3
8

2
.
3
5

1
.
9
1

1
.
4
4

2
.
1
9

2
.
4
0

1
.
7
8

1
.
7
4

1
.
1
9

0
.
0
5
7

0
.
0
4
0

0
.
0
0
8

0
.
0
8
7

0
.
1
9
1

0
.
0
3
9

0
.
1
7
5

0
.
0
1
6

0
.
6
6

0
.
5
9

0
.
5
1

0
.
5
5

0
.
7
2

0
.
5
2

0
.
7
6

0
.
5
1

5
5

9
6

1
1
4

1
8
2

7
9

1
1
4

6
7

2
0
3

9
3
4

4
5
6

2
1
7

4
2
5

3
9

4
3

4
0

3
0

4
4

4
7

2
2
4

2
4
8

1
8
3

1
2
7

3
4
7

1
9
9

6
2

6
5

3
6

3
9

1
4
1

5
3

3

3

4

4

2
2

4

1
4
5

1
2
8

2
0
8

1
6
2

9
7

2
2
6

6
1

3
7

3
5

3
3

1
1
1

3
4

1
5

1
4

1
0

1
0

3
1

1
0

3
.
9

2
.
3

N
D

1
.
7

9
.
3

3
.
0

1
8

1
4

1
3

1
4

2
2

1
6

1
0
9

1
5
3

N
D

1
5
1

3
8

1
6

1
0
4

8
0

9
6

6
0

2
2
4

6
8

D
e
t
e
r
m
i
n
e
d

b
y

X
R
F

a
t

W
a
s
h
i
n
g
t
o
n

S
t
a
t
e

U
n
i
v
e
r
s
i
t
y
.

O
x
i
d
e

w
t
%
;

t
r
a
c
e

e
l
e
m
e
n
t
s

i
n

p
p
m
.

C
o
n
c
e
n
t
r
a
t
i
o
n
s

i
n

w
t
%

r
e
c
a
l
c
u
l
a
t
e
d

t
o

1
0
0
%
.

*
T
o
t
a
l

i
r
o
n

a
s

F
e
O
.

1
0

8
6
-
3
3

4
8
.
6
2

1
4
.
4
0

0
.
4
8

1
1
.
6
6

0
.
2
4

1
3
.
8
6

9
.
4
7

0
.
1
0

1
.
1
6

0
.
0
0
5

0
.
6
1

8
4

1
8
6

4
4

2
4
8

1
1

5

7
1

3
0

1
2

1

1
6

8
2

8
7

1
1

8
6
-
2
8

4
5
.
2
2

1
7
.
9
3

0
.
9
8
7

1
4
.
0
8

0
.
3
5
0

1
2
.
1
6

7
.
2
8

0
.
5
2

1
.
4
2

0
.
0
7
2

0
.
7
1

6
5

1
7
5

4
3

2
9
1

2
3

1
2

1
3
7

7
4

2
7

6
.
6

2
0

8
0

9
7

c
.
o

c
.
o

o


t
-
'

(
"
)


S

S

5
'

(
J
q

U
l


0
-


p
:
:
I


n

(
"
)

a

0
'

n

:
:
r

Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana
Table 3. Selected trace elements for amphibolite and meta-ultramafic samples from the amphibolite unit (aa).
Sample
Numbers Cr Se Co La Ce Sm Eu Lu Hf Th
Amphi-
bolite
86-19 343 35.03 37.3 9.9 18.7 1.96 0.66 ND 1.0 1.3
86-25D 230 45.81 48.8 2.9 ND 1.77 0.73 0.35 ND ND
86-23 975 37.00 39.2 2.2 7.0 1.33 0.45 0.26 ND ND
85-34 324 41.27 40.4 8.1 18.8 1.82 0.69 0.25 1.4 1.8
85-35 122 40.78 47.0 9.9 24.8 4.01 1.49 0.56 4.1 0.9
81-33 202 56.80 56.9 1.4 ND 1.14 0.50 ND ND ND
Meta-
ultra-
mafic
85-32A 3224 19.38 87.2 1.9 8.3 0.66 0.14 0.14 ND 0.66
85-32D 1306 14.62 82.0 1.2 4.4 0.37 0.07 0.10 0.6 0.42
85-113 3205 15.77 96.1 6.1 17.2 1.09 0.29 0.14 ND 0.95
85-82 2636 14.72 95.2 7.1 20.8 1.18 0.23 0.14 0.8 1.12
85-85 2431 14.26 73.0 5.4 11.0 1.36 0.19 0.13 1.1 1.20
86-10 2685 21.36 85.0 4.1 ND 1.27 0.42 0.34 ND ND
Data were determined by INAA at Portland State University. Concentrations in ppm. ND, not detected.
Table 4. Geochemistry of amphibolites from interlayered intermediate gneisses and schists.
Sample: TRM-319 86-83 86-90D 86-94
Si0
2
51.74 51.07 53.01 52.99
Al
2
0
3
13.81 17.28 13.98 16.13
Ti0
2
1.250 0.923 1.149 0.697
FeO* 13.83 10.46 12.44 9.55
MnO 0.229 0.158 0.270 0.211
CaO 9.69 9.62 9.12 10.29
MgO 5.81 6.78 6.54 6.73
K
2
0 0.75 0.53 0.75 0.32
Na20 2.77 3.06 2.64 3.01
P
2
0
5
0.119 0.117 0.110 0.066
Fe/(Fe+Mg) 0.75 0.66 0.71 0.65
Ni
Cr
Sc
V
Ba
Rb
Sr
Zr
Y
Nb
Ga
Cu
Zn
31 124 44 44
33 210 45 54
39 30 38 33
327 219 312 196
183 113 184 130
19 7 21 3
147 136 167 185
101 83 93 70
26 25 26 15
6.6 5.7 7.6 3.9
15 20 17 16
77 45 132 16
113 103 109 71
Oxide in wt%; trace elements in ppm. Concentrations in wt% recalculated to 100%.
*Total iron as FeO.
331
332
M. L. Cummings and W. R. McCulloch
a.o
2 I.
......
,... 2 a. 0


2.0
-
0
-
><
><

0



0
0+
CII +
1 1.0
0
+-
0
+
0
0
1.0
x
+ III
0

t-
O
0 +
0
1 7.0
+
0
0
0'"
0
+
0.,0
0
1 1.0
40.0
10.0 10.0 '0.0 40.0 50.0 0.0
70.0
1.0
1 7. ....
......

4.0
1 I.
++ t
-

Xo

a.o

1 a. 0
lit 0
0
0
0
0
o.
X
CII
ell
0
2.0

X
.
1 1
0
+
0
0 ++
x
.,
+
CII
1.0
1.0
III
IL.
Z
7.0
4 0
, 0 10.0 '0.0
40.0
10.0 10.0 70.0
9102 ( Wt. 8102 ( Wt.
Figure 5. Selected chemical variation diagrams for amphibolite from the Aa and Aig units.
where these four samples form a group that includes
one sample from the Aa unit (Table 2, column 6).
These amphibolites contain higher Si0
2
, Na20 +
K
2
0
2
, and Ti0
2
concentrations and lower FeO +
MgO concentrations than samples from the Aa unit.
The Fe/(Fe + Mg) ratios are among the highest for
all amphibolite samples (Tables 2-4).
Compositions of the least altered meta-
ultramafic rocks (Table 5) that occur in the Aa unit
are of samples selected on the basis of low modal
percentages of Mg-hornblende, the mineral associa-
ted with alteration in meta-ultramafic rocks
(McCulloch, 1989). Si0
2
is 45-55%; lower concen-
trations of Si0
2
occur in samples with higher con-
centrations of MgO (32-22%). The CaO/Al
a
0
2
ratio
is 0.63-1.98, K
2
0 ranges from less than detectable
by XRF to 0.04% (one analysis at 0.32), and Ti0
2
isO.19-0.32%.
DISCUSSION
Masses of amphibolite and hornblende gneiss
occur throughout the southern Tobacco Root
Mountains (Vitaliano and others, 1979). The Aa
unit in the Branham Lakes cirque is one of the
Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana 333
2.1100 1,3901 . 6 7 0 2
8e 1 8
88 2 1
,-
88 250
88 33
I
88 2 3
85 34
(
85 35
TRM 318
88 800
r=
8 83
8 84
8. 8.
8. 2.
I

1 2
.8 1 4
2.1100 1.'301 1.2101 . 102
Val"., .'."\1 It .xl, .r.
Ilmll.rltl
Figure 6. Dendrogram for amphibolite samples from Aa
and Aig units. Values along x axis are similarities.
larger masses, if not the largest single mass, of
amphibolite in the range. Samples of amphibolite
within this unit are (1) compositionally and
texturally similar, (2) closely associated with meta-
ultramafic rocks, and (3) different in composition
and texture from thin amphibolite layers within the
Aig unit. On the basis of field relations and des-
criptions of amphibolite and hornblende gneiss from
other parts of the southern Tobacco Root Mountains
presented by Vitaliano and others (1979) and Bur-
1 00
50

Q. 5
E

.,
L C
S m E u L u
REE (Incro.alng n)
Figure 7. Chondrite-normalized REE diagrams for am-
phibolite samples from Aa unit. Chondrite concentrations
of C1 chondrite (Ekambaram and others, 1984) used for
normalization.
ger (1967,1969), the Aa unit is a distinctly different
unit.
The metamorphic history (Figure 8), structural
relations to other lithologic sequences, and geo-
chemistry of the Aa unit are discussed here, along
with the implications of these relations.
Orthopyroxene and olivine megacrysts in meta-
ultramafic bodies are the oldest textural features
preserved in the unit. Prismatic Mg-hornblende
within and along the margins of megacrysts and
along what are interpreted to be shear zones within
meta-ultramafic bodies developed later than the
megacrysts. The hornblende-rich blackwall be-
tween the two lithologies and hornblende-rich
material in apophyses and along shear zones that
truncate meta-ultramafic bodies are texturally and
mineralogically similar. Mineral textures are con-
sistent with contemporary crystallization of Mg-
hornblende and hornblende in blackwalls.
334
Sample:
Si0
2
Al
2
0
3
Ti0
2
FeO*
MnO
CaO
MgO
K
2
0
Na20
P20 5
Fe I(Fe+Mg)
Ni
Cr
Sc
V
Ba
Rb
Sr
Zr
Y
Nb
Ga
Cu
Zn
M A F I C
M.1. Cummings and W. R. McCulloch
Table 5. Geochemistry of meta-ultramafic rocks from the amphibolite unit.
85-32A 85-32D 85-113 85-111 85-112 85-78 85-82 85-85
52.38 56.29 51.96 49.03 46.54 51.26 49.29 49.70
6.89 3.61 5.12 4.72 5.76 5.90 4.79 4.94
0.30 0.19 0.28 0.31 0.30 0.32 0.25 0.26
12.83 11.98 10.61 7.11 9.84 11.21 11.54 11.71
0.28 0.28 0.28 0.29 0.37 0.27 0.27 0.29
4.34 2.62 4.84 9.37 4.23 7.03 5.62 4.80
22.23 24.54 25.71 28.46 32.69 23.07 27.39 27.60
0.03 0.01 0.02 0.04 ND 0.03 0.05 0.03
0.66 0.44 0.63 0.29 0.24 0.86 0.73 0.64
0.044 0.026 0.547 0.36 0.006 0.04 0.064 0.024
0.43 0.39 0.35 0.24 0.28 0.38 0.35 0.35
866 1354
224 3972
20 13
135 107
34 14
2 2
47 23
26 30
13 10
1 2
1 5
3 21
6 123
Oxide wt%; trace elements in ppm. Concentrations in wt% recalculated to 100%.
*Total iron as FeO. ND, not detected.
A
ULTRAMAFIC
ol-opx-chrom
AMPHIBOLITE BLACKWALL META-ULTRAMAFIC
-A tot PHI B 0 LIT E
I
I
RET R DIG R A 0 E.
relict ol-opx-chrom
METAMORPHISM
I
I I
contact
e f fee t s
I
c .0
--BLACKWALl
... -tot ETA -
ULTRAMAFIC
85'34
86-23 86-19
86-33- .85-35
0
86-10
47.54
7.24
0.36
12.15
0.48
6.65
24.81
0.32
0.41
0.034
0.39
829
2576
21
96
75
17
17
33
12
ND
7
ND
74
80
Figure 8. Chronology of development of mineral assem-
blages within amphibolite, blackwall, and meta-ultramafic
rocks within the Aa unit.
Figure 9. ACF diagram showing composition of amphibo-
lite, meta-ultramafic rocks, and intervening blackwall.
Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana 335
Compositions of amphibole-rich lithologies plot
on an ACF diagram between those of meta-
ultramafic bodies and amphibolite (Figure 9).
McCulloch and Cummings (1988) concluded that
metasomatic reactions occurred between the amphi-
bolite and meta-ultramafic lithology during meta-
morphism in which granoblastic textures in the
amphibolite crystallized.
Mineral textures in shear zones, which disar-
ticulate the meta-ultramafic layers, are distinct
from those of large-scale shear zones that occur
within the Aa unit and those that separate the Aa
unit from the Aig unit. Where these shear zones cut
the amphibolite, blackwall, and meta-ultramafic
lithologies, they clearly post-date development of
the blackwall lithology. These relations are best
documented in the area of Leggat Peak near the
shear zone that separates the Aa and Aig units.
One large-scale shear zone, within the Aa unit
contains deformed amphibole-rich assemblages;
however, a detailed investigation of the kinematic
and metamorphic histories of various shear zones in
the Branham Lakes cirque is lacking. We suggest
that at least three phases of deformation are de-
monstrable from textures in different shear zones.
The structural relations between the Aa and
the Ams units on Branham Peaks are interpreted to
be those of an imbricate thrust zone based on the
following factors.
The contact between the Aa unit and the over-
lying zone, in which large-scale folds are prominent-
ly developed, is covered but can be located within
"='10 meters.
Discontinuities in compositional layering and
lithology on opposite sides of this contact suggest a
fault contact.
Within the zone of folding, compositional layer-
ing in neighboring fault slices is discordant across
sharp contacts.
The different lithologies form large-scale sig-
moidally shaped lenses (Figure 4).
At the upper elevations of Branham Peaks,
quartzite, schist, and amphibolite display locally
concordant contacts, but the bodies are lens-shaped.
The imbricate thrust zone tectonically juxta-
posed rocks of the Aa unit with supracrustal rocks of
the Ams unit.
Similar compositions of amphibolite samples
from the thrust zone and from the Aa unit are con-
sistent with this interpretation.
The protolith of amphibolite metamorphosed
under conditions of the upper amphibolite facies is
difficult to determine because primary textures and
fabrics have been destroyed by metamorphic re-
crystallization. Based on geochemical data, the pro-
tolith of the amphibolite in the Aa unit is inter-
preted to be basalt.
The lithologies (quartzite, biotite and biotite-
garnet schist, anthophyllite-bearing and pyrrhotite-
rich rocks) of the Ams unit are interpreted as meta-
morphosed and metasomatised shelf sediments.
Marbles, garnet-biotite-sillimanite schist and
gneiss, and iron formation that occur elsewhere in
the Tobacco Root Mountains were interpreted by
Heinrich and Rabbitt (1960), Reid (1963), Burger
(1969), and Vitaliano and others (1979) as shelf
sediments.
The Aig unit contains bulk compositions that
are more mafic than those in the Ams unit, and
quartzites have not been indentified within the Aig
unit. The bulk compositions are more felsic than
those of the Aa unit. The overall composition of the
Aig unit is more similar to a protolith dominated by
graywacke or igneous rocks of intermediate com-
position.
The applicability of tectonic models derived
from modern geologic environments to Precambrian
systems, particularly Archean systems, is depen-
dent upon distinguishing those chemical processes
related to partial melting in different tectonic set-
tings from those produced by magmatic fractiona-
tion during crystallization, hydrothermal altera-
tion, and metamorphism. Redistribution of major
and trace elements during alteration prior to
metamorphism and under conditions of low-grade
metamorphism have been examined in many
studies, such as Gelinas and others (1982), Ludden
and others (1982), and Knoper and Condie (1988).
The high-field-strength elements (HFSE) and rare-
earth elements (REE) generally are relatively im-
mobile during alteration and metamorphic pro-
cesses.
From the data (Table 2 and 3) and sample
locations throughout the Aa unit, we conclude that
fractional crystallization is not represented by the
compositional variations in the data set. We did not
identify compositional changes due to metasomat-
ism or alteration except in three samples (Table 2,
columns 7, 8, and 9). Fully recognizing the dangers
and uncertainties inherent in tectonomagmatic
reconstructions, we examined the geochemistry of
the amphibolites using various tectonic discrimin-
ant diagrams. On each diagram, the points for
samples in columns 7-9 (Table 2) are indicated
M. L. Cummings and W. R. McCulloch
where they significantly differ from the fields in
which samples from columns 1-6 plot.
On the basis of the tectonic discriminant dia-
grams (Figure 10), the geochemistry of amphibo-
lites (Table 2, columns 1-6) is consistent with that
of oceanic crust. On Ti-Zr-Y diagrams (Pearce and
Cann, 1973), the data plot within the fields for low-
potassium tholeiites, ocean-floor basalts, or calc-
alkaline basalts. On the Ti-Zr-Sr and Zr-Ti dia-
grams, the samples plot within the calcalkaline
basalt and ocean-floor basalt fields. On Nb-Zr-Y
diagrams (Meschede, 1986), the data are consistent
with those of N -type MORB or volcanic-arc basalts.
The concentration of Ta is below detection limits of
INAA, thus Hf-Th-Ta diagrams were not construc-
ted. Comparison of the composition of amphibolite
TI/100
A LK T
to that of N-type MORB (Figure 11) indicates
enrichment in Rb and Ba and depletion in Nb, REE,
Ti, Y, Zr, and P. K is enriched in half of the samples
(Figure 11). Chondrite-normalized REE plots
(Figure 7) for most samples are flat and consistent
with low concentrations of REE. Similar patterns
are found in oceanic basalts. Those samples that are
enriched in LREE were collected near blackwall and
may have been enriched in LREE by metasomatic
reactions.
Amphibolites from the Aig unit plot more com-
monly within the volcanic arc-calcalkaline basalt
fields (Figure 10) than within other fields. On a
diagram of Si0
2
versus ZrITi0
2
(Figure 12), these
samples plot near the boundary between the sub-
alkaline basalt field and the andesite field (Win-
2Nb
B OFB. LKT. CAB
C CAB
AI WPB
All WPT
C PMORB. VAB
D NMORB. VAB
Zr
15000
E
10000
Co
Co

5000
0
0
100
Zr (ppm)
3 Y Z r
-4-
A L K T
B OFB. LKT. CAB
C CAB
D OFB
200
Y
Figure 10. Tectonic discriminant diagrams for amphibolite samples from Aa and Aig units. LKT, low-K tholeiite;
OFB, ocean-floor basalt; CAB, calcalkaline basalt; WPB, within-plate basalt; WPT, within-plate tholeiite; PMORB, plume-
type MORB; NMORB, normal-type MORB; V AB, volcanic arc basalt.
Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana 337
'"
II:
o
::I
z


t1>
1 0
Rb Ba K Nb La P Zr 8m TI Y
Figure 11. Geochemistry of amphibolite samples (Tables 2
and 3) normalized to N-type MORB. Data are normalized
to those of Sun and Nesbit (1977).
chester and Floyd, 1977). Amphibolites inter-
layered with lithologies of the Aig unit are more
calcalkaline-like than those of the Aa unit.
The genesis of meta-ultramafic bodies within
the Aa unit could have occurred by (1) intrusion of
dikes and sills into the amphibolite protolith prior to
metamorphism; (2) emplacement of cold ultramafic
bodies during tectonism and metamorphism; (3) er-
uption of ultramafic lava flows within a basaltic
volcanic sequence; or (4) separation of a restite pro-
duced by partial melting processes.
Vitaliano and others (1979) showed that meta-
ultramafic bodies occur throughout the Tobacco
Root Mountains as lenses, pods, and stock-like
masses in close association with metabasites. How-
ever, meta-ultramafic bodies in the Branham Lakes
cirque are distinct from those described by Vitaliano
and others (1979).
'"
o
60
ANDESITE
SUB-ALKALINE /"
0'
BAS A L T o'
50
40+---____ __ ______
o 00 1 o 0 1
Z r / T i 0
2
o 1
Figure 12. Si0
2
versus Zr/Ti0
2
diagram showing composi-
tion of amphibolite samples from Aa and Aig units.
Desmarais (1981) described ultramafic bodies
in the Ruby Range, = 50 km south of the Tobacco
Root Mountains, that contain mono- or bimineralic
metasomatic rocks at contacts between the bodies
and country rocks that consist of felsic and mafic
gneisses. Metamorphic conditions were determined
to be 710C and 5 kb. It is unclear whether the
meta-ultramafic rocks were emplaced as igneous
intrusions prior to metamorphism and tectonism, or
if tectonism and metamorphism accompanied their
emplacement as cold ultramafic bodies.
Several characteristics relate meta-ultramafic
bodies in the Branham Lakes cirque to their gene-
sis.
Olivine and orthopyroxene megacrysts form a
relic texture that formed prior to crystallization of
the Mg-hornblende-bearing metasomatic assem-
blage. There are no indications of relic primary tex-
tures of volcanic origin.
Olivine and orthopyroxene megacrysts lack de-
formation lamellae that may have formed during
cold emplacement. Although megacrysts are frac-
tured, various fragments of the crystals are in
optical continuity, indicating no relative movement
among fragments. Optical continuity among frag-
ments indicates that the texture is not due to
polygonization of strained megacrysts.
Although shear zones are spatially related to
meta-ultramafic bodies, the most prominent shear
zones were active during metasomatism/meta-
morphism and are commonly at a high angle both to
the strike of compositional layering in the am-
phibolite and to the long dimension of meta-
ultramafic bodies. Indications of shearing along
contacts between meta-ultramafic bodies and am-
phibolite (that may have developed during tectonic
emplacement of cold ultramafic bodies into amphi-
bolite) are obscured by development of black wall.
338 M. L. Cummings and W. R. McCulloch
Rodingites and serpentinite or metarodingites
are not associated with meta-ultramafic rocks as in
meta-ultramafic pods in the Ruby Range (Des-
marais, 1981).
The geochemistry of the least altered meta-
ultramafic rocks (Table 5) is similar to that for
komatiites for some elements (Arndt and Nesbet,
1982: Si0
2
<53%, MgO>9%, K
2
0 and Ti0
2
<O.9%,
high Cr) but the CaO:AI
2
0
3
ratio is < 1 for nearly
all samples of meta-ultramafic. The geochemistry,
the textures of megacrysts of olivine and ortho-
pyroxene, and the lack of relic spinifex textures are
not consistent with a komatiitic parentage for these
meta-ultramafic rocks.
Lithologic associations in the Branham Lakes
area are believed to contain a record of terrane
juxtaposition accompanied by upper amphibolite-
facies metamorphism. Recently, models interpre-
ting the Archean history of southwestern Montana
in terms of processes active at modern plate boun-
daries have been proposed. In relation to the model
proposed by Mogk and others (1988a), the Tobacco
Root Mountains occur west of the North Snowy
Block of the Beartooth Mountains. This block
separates largely supracrustal rocks in the Gallatin,
Ruby, Madison, and Tobacco Root Mountains ranges
from the complex terrane exposed to the east. The
eastern terrane contains calcalkaline granitoids and
associated rocks interpreted to be products of sub-
duction along a continental margin (Mueller and
others, 1985). Montgomery (1988) examined late
Archean quartz monzonite intrusions from the
South Snowy Block and found the geochemistry of
these rocks consistent with that of syn-collisional
and post-collisional granite groups.
The ranges to the west of the North Snowy
Block contain high-grade metasedimentary rocks
that were metamorphosed during the late Archean.
Heinrich (1960), Reid (1963), and Vitaliano and
others (1979) interpreted quartzite, marble, alumin-
ous schist, banded iron formation, and quartzo-
feldspathic gneiss as sedimentary deposits that had
formed in a stable platform environment. Mogk and
Henry (1988) proposed a working model for the
Archean system in the western ranges. In the
model, they proposed early formation of a rift-
bounded ensialic basin followed by collapse of the
basin and deep burial through stacking of nappes. A
small amount of oceanic crust may have formed
during development of the ensialic basin. They
argue that the metamorphic and structural styles
observed in these ranges require large-scale hori-
zontal shortening such as occurs during continental
collisions.
We concur with the model proposed by Mogk
and Henry (1988) on the basis of the geology of the
Branham Lakes cirque. The amphibolite and meta-
ultramafic rocks of the Aa unit are interpreted to be
a sliver of oceanic crust generated during ensialic
basin development, as proposed in this model. The
Aa unit was juxtaposed against the Aig unit along
the large-scale shear zone that separates the two
units. The shelf sediments of the Ams unit were
placed against the Aa unit, and probably the Aig
unit, in an imbricate thrust zone that tectonically
juxtaposed slices of amphibolite of the Aa unit with
slices of shelf sediments. Metamorphism occurred
during deep burial produced by stacking of nappes.
ACKNOWLEDGMENTS
Critical reviews of this paper by Jonathan
Husch and David Mogk greatly facilitated a re-
appraisal of its focus. We greatly appreciate the
quality and rigor of the reviews. Part of this study
was supported by the William and Edith Rockie
Fund, maintained by the Department of Geology
within the Portland State University Foundation.
XRF analyses at Washington State University were
done under the supervision ofP. R. Hooper.
REFERENCES
ARNDT, N. T., and E. G. NISBET, 1982, What is a koma-
tiite?; pp. 19-27 in N. T. Arndt and E. G. Nisbet (eds.),
Komatiites: George Allen & Unwin, London, England,
UK,539p.
BURGER, H. R., 1967, Bedrock Geology of the Sheridan
District, Madison County, Montana: Montana Bureau of
Mines and Geology, Memoir 41,22 p.
__ , 1969, Structural evolution of the southwestern
Tobacco Root Mountains, Montana: Geological Society of
America Bulletin, v. 80, p. 1329-1342.
CONDIE, K. C., 1969, Geologic evolution of the Precam-
brian rocks in northern Utah and adjacent areas: Utah
Geological Survey Bulletin, no. 82, p. 71-95.
DAHL, P. S., 1979, Comparative geothermometry based on
major element and oxygen isotope distributions in Precam-
brian metamorphic rocks from southwestern Montana:
American Mineralogist, v. 64, p. 1280-1293.
Amphibolite and ultramafic rocks, Tobacco Root Mountains, southwestern Montana 339
DAVIS, J. C., 1986, Statistics and Data Analysis in
Geology: John Wiley & Sons, New York, New York, USA,
646p.
DESMARAIS, N., 1981, Precambrian metamorphic rocks in
the Ruby Range, Montana: Precambrian Research, v. 16,
p.67-10l.
EKAMBARAM, V., 1. KAWABE, T. TANAKA, A. M. DAVIS,
and L. GROSSMAN, 1984, Chemical compositions ofrefrac-
tory inclusions in the Murchison C2 chondrite: Geochimica
et Cosmochimica Acta, v. 48, p. 2089-2105.
FRIBERG, L. V. M., 1976, Petrology of a metamorphic
sequence of upper-amphibolite facies in the central To-
bacco Root Mountains, southwest Montana [absract): Ab-
stracts International, v. 37, no. 4, p. 1592B.
GELINAS, L., M. MELLINGER, and A. TRUDEL, 1982,
Archean mafic metavolcanics from the Rouyn-Noranda
district, Abitibi Greenstone Belt, Quebec - 1, Mobility of
the major elements: Canadian Journal of Earth Science, v.
19, p. 2258 -2275.
GILETTI, B. J., 1966, Isotopic ages from southwestern
Montana: Journal of Geophysical Research, v. 71, p. 4029-
4036.
GILLMEISTER, N. M., 1972, Cherry Creek Group-Pony
Group relationship in the central Tobacoo Root Mountains,
Madison County, Montana: Northwest Geology, v. 1, p. 21-
24.
HEINRICH, E. W., and J. C. RABBITT, 1960, Pre-Beltian
Geology of the Cherry Creek and Ruby Mountains
Areas, Southwestern Montana: Montana Bureau of
Mines and Geology, Memoir 38, 40 p.
HENRY, D. J., J. L. WOODEN, P. A. MUELLER, J. L.
WARNER, and R. LEE-BERMAN, 1982, Granulite grade
supracrustal assemblages of the Quad Creek area, eastern
Beartooth Mountains, Montana; pp. 147-156 in P. A.
Mueller and J. L. Wooden, (eds.), Precambrian Geology
of the Beartooth Mountains, Montana and Wyoming:
Montana Bureau of Mines and Geology, Special Publication
84,167p.
HESS, D. F., 1967, Geology of Pre-Beltian Rocks in the
Central and Southern Tobacco Root Mountains, with
Reference to Superposed Effects of the Laramide-Age
Tobacco Root Batholith: PhD dissertation, Indiana
University, Bloomington, Indiana, USA, 333 p.
IMMEGA, I. P., and C. KLEIN, 1976, Mineralogy and petro-
logy of some metamorphic Precambrian iron formations in
southwestern Montana: American Mineralogist, v. 6], p.
1117-1144.
JAMES, H. L., and C. E. HEDGE, 1980, Age of the basement
rocks of southwest Montana: Geological Society of America
Bulletin, Part I, v. 91, p. 11-15.
KNOPER, M. W., and K. C. CONDIE, 1988, Geochemistry
and petrogenesis of Early Proterozoic amphibolites, west-
central Colorado, U.S.A.: Chemical Geology, v. 67, p. 209-
225.
LUDDEN, J., L. GELINAS, and P. TRUDEL, 1982, Archean
meta volcanics from the Rouyn-Noranda district, Abitibi
Greenstone Belt, Quebec - 2, Mobility of trace elements
and petrogenetic constraints: Canadian Journal of Earth
Science, v. 19, p. 2276-2287.
McCULLOCH, W. R., 1989, Metasomatism Between Am-
phibolite and Metaultramafic Rocks During Upper
Amphibolite Facies Metamorphism, Tobacco Root
Mountains, Southwest Montana: MS thesis, Portland
State University, Portland, Oregon, USA, 137 p.
and M. L. CUMMINGS, 1987, Metasomatism be-
Archean age metamorphosed mafic and ultramafic
rock, Tobacco Root Mountains, Montana [abstract]: Geo-
logical Society of America Abstracts with Programs, v. 19,
no. 5, p. 320.
and M. 1,. CUMMINGS, 1988, Metamorphic assem-
blages in mafic and ultramafic bulk compositions, Tobacco
Root Mountains, southwest Montana [abstract]: Geologi-
cal Society of America Abstracts with Programs, v. 20, no.
6, p. 431.
MESCHEDE, M., 1986, A method of discriminating between
different types of mid-ocean ridge basalts and continental
tholeiites with the Nb-Zr-Y diagram: Chemical Geology, v.
56, p. 207-218.
MIYASHIRO, A., 1974, Volcanic rock series in island arcs
and active continental margins: American Journal of
Science, v. 274, p. 321-355.
MOGK, D. W., 1988, Archean allochthonous units in the
northern and western Beartooth Mountains, Montana; pp.
43-51 in S. E. Lewis and R. B. Berg, (eds.), Precambrian
and Mesozoic Plate Margins: Montana Bureau of Mines
and Geology, Special Publication 96, 196 p.
__ , and D. J. HENRY, 1988, Metamorphic petrology of
the northern Archean Wyoming province, southwest Mon-
tana: Evidence for Archean collisional tectonics; pp. 362-
382 in W. G. Ernst, (ed.), Metamorphism and Crustal
Evolution in the Western United States (Rubey Vol-
ume VII): Prentice Hall, Inc., Englewood Cliffs, New
Jersey, USA, 1153 p.
P. A. MUELLER, and J. L. WOODEN, 1988a, Archean
tectonics of the North Snowy Block, Beartooth Mountains,
Montana: ,Tournai of Geology, v. 96, p. 125-14l.
___ , P. A. MUELLER, E. WEYEND, J. WOODEN, and D.
BOWES, 1988b, Protracted magamtic history of an Arch-
ean collision zone, northern Madison Range, SW Montana
[abstract): Geological Society of America Abstracts with
Programs, v. 20, p. A50.
340 M. L. Cummings and W. R. McCulloch
MONTGOMERY, C. W., 1988, Geochemical characterization
and possible petrogenesis of Archean quartz monzonite
from the southwestern Beartooth Mountains [abstract):
Geological Society of America, Abstracts with Programs, v.
20, p. 433.
MUELLER, P. A., and W. S. CORDUA, 1976, Rb-Sr whole
rock age of gneisses from the Horse Creek area, Tobacco
Root Mountains, Montana: IsochronlWest, no. 16, p. 33-36.
O'NEILL, J. M., M. S. DUNCAN, and R E. ZARTMAN, 1988,
An Early Proterozoic gneiss dome in the Highland Moun-
tains, southwestern Montana; pp. 81-88 in S. E. Lewis and
R B. Berg (eds.), Precambrian and Mesozoic Plate
Margins: Montana Bureau of Mines and Geology, Special
Publication 96, 196 p.
PEARCE, J. A., 1982, Trace element characteristics oflavas
from destructive plate boundaries; pp. 525-548 in R S.
Thorpe (ed.), Andesites; Orogenic Andesites and Rela-
ted Rocks: John Wiley & Sons, Chichester, England, UK,
724p.
__ , and J. R CANN, 1973, Tectonic setting of basic
volcanic rocks determined using trace element analyses:
Earth and Planetary Science Letters, v. 19, p. 290-300.
REID, R R, 1963, Metamorphic rocks of the northern
Tobacco Root Mountains, Madison County, Montana: Geo-
logical Society of America Bulletin, v. 74, p. 293-306.
RUNNER, J. J., and L. C. THOMAS, 1928, Stratigraphic
relationships of the Cherry Creek Group in the Madison
Valley, Montana [abstract): Geological Society of America
Bulletin, v. 39, p. 202-203.
SUN, S.-S., and R W. NESBIT, 1977, Chemical hetero-
geneity of the Archaean mantle, composition of the Earth
and mantle evolution: Earth and Planetary Science
Letters, v. 35, p. 429-448.
TANSLEY, W., F. A. SCHAFER, and L. H. HART, 1933, A
Geological Reconnaissance of the Tobacco Root
Mountains, Madison County, Montana: Montana
Bureau of Mines and Geology, Memoir 9,57 p.
VITALIANO, C. J., H. R BURGER III, W. S. CORDUA, T. B.
HANLEY, D. F. HESS, and F. K. ROOT, 1979, Explanatory
Text to Accompany Geologic Map of Southern Tobac-
co Root Mountains, Madison County, Montana: Geo-
logical Society of America, Map and Chart Series MC-31, 8
p.
__ , S. KISH, and D. G. TOWELL, 1980, K-Ar dates and
Sr isotope values for rocks of the Tobacco Root Batholith,
southwest Montana: IsochronlWest, no. 28, p. 13-15.
WINCHESTER, J. A., and P. A. FLOYD, 1977, Geochemical
distributions of different magma series and their differen-
tiation products using immobile elements: Chemical
Geology, v. 20, p. 325-343.
WOODEN, J. L., P. A. MUELLER, and D. W. MOGK, 1988, A
review of the geochemistry and geochronology of the
Archean rocks of the northern part of the Wyoming Pro-
vince; pp. 383-410 in W. G. Ernst (ed.), Metamorphism
and Crustal Evolution in the Western United States
(Rubey Volume 7): Prentice Hall, Inc., Englewood Cliffs,
New Jersey, USA, 1153 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Multiple reactivation of a collisional boundary:
An example from southwestern Montana
BARRY C. McBRIDE1, CHRISTOPHERJ. SCHMIDT2, GARY E. GUTHRIE
3
, and MARK K. SHEEDL0
4
lMOBIL-GIS, P.O. Box 650232, Dallas, Texas 75265-0232, USA; 2Department of Geology, Western Michigan University,
Kalamazoo, Michigan 49008, USA; 3Marathon Oil Company, P.O. Box 3128, Houston, Texas 77253, USA;
4Conoco Inc., 3500 General DeGaulle, New Orleans, Louisiana 70114, USA
(received August 12, 1988; revision accepted February 17, 1989)
ABSTRACT
Structural elements and stratigraphy of the central Snowcrest Range, southwestern Montana, reflect
multiple reactivation of what was probably a late Archean and/or Early Proterozoic collisional margin. The
oldest structural element is a N54E ,41 NW foliation in metamorphic basement rocks, on the western margin
of the range, which is inferred to have coincided with establishment of the NE-trending margin of the range as
an important zone of crustal weakness (Snowcrest-Greenhorn lineament). The second major structural event
is reflected in upper Paleozoic strata. A substantially thicker section of Mississippian through Triassic rocks
was deposited in a crustal depression (Snowcrest trough) on the northwest side of the lineament. The
lineament was probably a NW-dipping normal fault zone during late Paleozoic time, reflecting crustal
extension across the earlier boundary. During the Late Cretaceous, E-W shortening across the lineament
produced a system of right oblique-slip, NW-dipping thrusts (Snowcrest-Greenhorn thrust system). The
earlier Snowcrest trough facies were juxtaposed against thinner shelf facies during this basement-involved
thrusting. Foliation-parallel brittle fractures probably formed in basement rocks at this time. Beginning in
the Miocene, extension across the lineament produced a NW-dipping listric normal fault system (Sage Creek
system) along the lineament, which partially accounts for existing basin-range topography. The normal faults
are inferred to merge at depth into the Snowcrest-Greenhorn thrust system.
INTRODUCTION
Although a variety of criteria may be used to
document recurrent fault movement, conclusions
about reactivation are often made principally on the
observation that a clearly later structure follows the
trend of an earlier structural grain. It is usually
impossible to establish that a specific fault is really
the result of movement on an older fault or, as with
Laramide faults which follow earlier foliation
trends, that the later fault is actually significantly
influenced by metamorphic foliation (for example,
Swanson, 1986; Ratcliffe and Newell, 1986). The
likelihood that a fault will follow earlier anisotropy
depends on (among other things) the cohesion and
the continuity of the anisotropy (see Prucha, this
volume).
The tectonic evolution of southwestern Mon-
tana has been strongly influenced by recurrent
movement on at least two sets of faults: (1) a NW-
trending set which originated at least as early as
Middle Proterozoic and was recurrently active, with
reversals of throw, during Late Cretaceous (Lara-
mide) shortening and during Neogene (basin-range)
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 341-357. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
341
342 B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
110-0
t08
01'
Figure 1. Regional tectonic map showing the Great Falls tectonic zone and other major recurrent fault trends in Montana
(modified from Schmidt and Garihan, 1986). Fault trends from O'Neill and Lopez (1985), Reynolds (1979), Schmidt and
Garihan (1983), and Stone (1969). MZ, Madison mylonite zone; SGL, Snowcrest-Greenhorn lineament; SSZ, Snowy shear zone;
BSG (Fault), Bismark-Spanish Peaks-Gardiner (fault). The location of Figure 2 is indicated.
extension (Schmidt and Garihan, 1986; O'Neill and
others, 1986); and (2) a NE-trending set, which is
part of a wide, regional fault zone (Figure 1) termed
the Great Falls tectonic zone (O'Neill and Lopez,
1985). The purpose of this paper is to review the
evidence, principally from the Snowcrest Range, for
four distinct periods of movement along faults in
this zone during widely separated time intervals:
late Archean or Early Proterozoic; late Paleozoic,
Late Cretaceous, and Neogene.
The general NE trend of the Snowcrest Range,
including the Laramide folds and faults and
Neogene faults which bound the range on the
northwest (Figure 2), is especially anomalous in
terms of Laramide structural trends. Regional
syntheses either assume that (NE-SW) Laramide
shortening in Wyoming and Montana was perpen-
dicular to the major fault and fold trends of the
Rocky Mountain foreland (Hamilton, 1988; Brown,
1988), or that the shortening shifted from "'E-W to
'" N-S with time (Gries, 1983). Neither of these
schemes can account for the NE trend of structural
elements in the Snowcrest and adjacent ranges
without a significant influence of inherited struc-
tural trends.
PRECAMBRIAN STRUCTURES
Although a diversity of structural trends exist
in Archean metamorphic rocks of the principal
ranges in southwestern Montana, the regional NE
trend is most persistent (Figure 3). NE-trending
structural elements include compositional layering,
metamorphic foliation, fold hinges on several scales,
major fracture patterns, and mylonite zones
(Hadley, 1960; 1969a,b; Reid and others, 1975;
Spencer and Kozak, 1975; Duncan, 1976; Vitaliano
Multiple reactivation of a collisional boundary, southwestern Montana 343
OT
OT
BEAVERHEAD
TEATIAAYVOLCANICAOCKS

MESOZOIC ROCKS
lliJ PAl,.OZOIC ROCKS
METAMOFlPHtC
...l......L.
_u
0
-+-
-+-
IIZ'OO'
10
I
KILOMETERS
MAJOR LITHOLOGIC CONTACT
MAJOR NORMAL FAULT
MAJOR THRUST fAULT
HIGH ANGL.E fAULT SHOWING
RELAtIVE MOVEMENT
OTHER FAULTS
ANTICl.JNE
-<r
OIL WEll
SYNCLINE

MEASUREO
SECTION
OT
ENNI

N
1
20
I
Figure 2. Tectonic map of the Snowcrest and adjacent ranges. Principal sources: Brasher (1950), Flanagan (1958), Hadley
(1969), Karasevich and others (1981), Keenmon (1950), Klepper (1950), Mann (1954), McBride (1988), Monroe (1976), Schmidt
and Garihan (unpublished), Scholten and others (1955), Sheedlo (1984), Tilford (1976), Tysdal (1976), Vaughn (1948).
Structural features dashed where approximately located, dotted where concealed; blind thrusts dotted; transect line parallel to
regional Laramide shortening direction. Paleogeologic cross-sections, showing deformational sequence (Figures 5, 8, and 11)
along this transect, extend 20 km west of map boundary.
and Cordua, 1979; Fountain and Desmarais, 1980;
Karasevich and others, 1981; Erslev, 1982, 1983;
Mogk and Henry, 1988).
The trend of the regional structural grain is
broadly curved. In a general way, the more
northerly ranges (e.g., Tobacco Root Range and
344 B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
? -

....... J

PRECAMBRIAN SHEAR
.. ZONE CASH EO
... SOUTHERN WHERE INFEARED
MAJOR LARIMIOE
THRUST - DASHED
WHERE IN FERRee

D'
MAJOR l"ARIMIDE
ftEVERSE. FAULT
SHOWING RELATIVE
MOVEMENT . DASHED
LATE ARCHEAN STILLWA.TER IGNEOUS COMPL.EX
AND ASSOC lATE 0 MET ASEDIM ENT ARY
HORNFELS AuREOLE
I"ATE ARCHEAN METASEDIMENTARY SCHISTS
Af40 GNEISSES LOCALLY INTAuDED 8'1"
lATE ARCHEAN METASEDIMENTARY ROCKS
INCLUOI NG OUAATOFELOSPATHIC GNEISS,
PELITIC SCHISTS, MAR8lE., AND
AMPHIBOLITE INTRUDED 8'1' METABASITE
LATE ARCHEAN GRANITOIDS
NORTHEASTTRENDING LATE ARCHEAN
MOBILE BEL.T OF TECTONICAL.LV INTERl.EAVED
METAIGNEOuS AND METASEDIMENTARY' ROCKS
D
.: ..
.': .
LATE ARCHEAN QUARTZOFElDSPATH.IC GNEI SS
ASSEMBLAGE AND aRANITOID PLUTONS
rl
Lj
SIMI LAR TO GRAVEllY TERRAN e
EARLY PROTEROZOIC aUARTZOfELOSPATHIC
GARNET. BIOTITEONEISS, AMPHIBOLITE.
DOMINANTLY LATE ARCHEAN
CRANITOIOS AND
LATE ARCHEAN OUAATAOFEL.OSPATHIC GNeiSS
ASSEMel.AGE AND GRANITOI D PL.UTONS
AND METABASITE ASSEMBLAGE
Figure 3. Regional map of Precambrian basement terranes in southwestern Montana, known or inferred Precambrian shear
or mylonite zones, and principal Laramide faults. Terrane designations and shear zones are based, in part, on Hadley
(1969a,b), Erslev (1983), Mogk and Henry (1988). MLSZ, Mirror Lake shear zone; MMZ, Madision mylonite zone; BSG
(fault), Bismark-Spanish Peaks-Gardiner (fault); HT, Hellroaring thrust; HCT, Highland-Cabin Creek thrust.
North Snowy Block) have foliation, fold, and
mylonite trends that vary from nearly N-S to
N45E, whereas the more southerly ranges have
structural trends that vary from about N300E to
N80E. Later structures show the same variation in
trend.
Metamorphic foliation and compositional layer-
ing, although locally consistent in trend, are nearly
everywhere folded on N- to NE-trending axes. In
some cases the folds are very broad (Vitaliano and
Cordua, 1979; Spencer and Kozak, 1975; Duncan,
1976; O'Neill and others, 1988a,b), and both folia-
tion strike and dip change around folds. Therefore,
the foliation trend cannot be extrapolated to great
depths.
One type of Precambrian basement fabric that
does have a more consistent regional trend along
strike, and presumably at depth, is mylonite zones
associated with a broad late Archean and/or Early
Proterozoic compressional orogen belt (Fountain
and Desmarais, 1980; Erslev, 1982; Mogk and
Henry, 1988; Mogk and others, 1988; Mogk and
others, this volume) similar to that described by
Dubendorfer and Houston (1987) for nearly iden-
tical rocks and structural trends along the Chey-
enne belt of southeastern Wyoming. Two major NE-
trending mylonite zones in basement rocks are the
Madison mylonite zone (Erslev, 1982, 1983, this
volume) in the southern Madison Range and the
Snowy shear zone along the northwestern boundary
Multiple reactivation of a collisional boundary, southwestern Montana 345
of the Beartooth Mountains. The segments of these
zones are separated by two major Laramide faults
(Figure 3): the NW-trending Bismark-Spanish
Peaks-Gardiner fault zone and the Hilgard-Cabin
Creek fault zone in the southern Madison Range),
which have undergone several kilometers of left-
reverse movement and dip-slip thrust movement,
respectively (Schmidt and Garihan, 1983; Tysdal,
1986). These two mylonite zones may be fortui-
tously aligned by Laramide faulting and thus may
not be segments of the same zone as previously
inferred by Erslev (1982, this volume).
The Madison mylonite zone is as wide as 3 km
and separates two distinct lithostratigraphic se-
quences: a gneissic terrane (NW) and metasedi-
ments (SE), including marbles, (Erslev, 1983). It
dips 45NW and structural elements within the zone
are consistent with dip-slip reverse faulting (Erslev,
1983). Amphibolite-facies rocks within the zone
were retrograded to greenschist-facies assemblages,
accompanied by development of phyllonite. This
retrograde event is probably correlative with the
1600-1800 Ma regional retrograde metamorphism
(Giletti, 1966, 1971; Reid and others, 1975; Peter-
man, 1979; Montgomery, 1981; O'Neill and others,
1988a,b). Initial shearing responsible for the higher
grade rocks and mylonitic fabric likely was a late
Archean event; Erslev (1983) described several
other smaller mylonite zones within the southern
Madison Range. Sumner and Erslev (1988) indica-
ted that supracrustal rocks of the southernmost
Madison Range and Henry's Lake Mountains are
dominated by SE-dipping thrust faults of probable
late Archean age (2700 Ma). The Early Proterozoic
Madison mylonite zone, with SE-vergence, is
interpreted to have been superimposed (= 1900 Ma)
on this earlier thrusting (Sumner and Erslev, 1988).
The Snowy shear zone separates two disinct
lithostratigraphic sequences: (1) a high-grade silli-
manite + cordierite-bearing biotite gneiss and augen
gneiss terrane (SE), and (2) an andalusite-bearing
biotite-staurolite schist terrane (NW) (Erslev, this
volume). It dips 46NW and structural elements
within the zone indicate right-lateral normal
faulting (Erslev, this volume). Rocks within the
zone were retrograded to greenschist-facies assem-
blages during normal faulting. Evidence within the
zone suggests it follows an earlier, SE-vergent
(Archean?) thrust zone.
Another NE-trending mylonite zone (Mirror
Lake zone) is located in the Spanish Peaks of the
northern Madison Range (Figure 3) (Mogk and
Henry, 1988). According to Mogk (personal com-
mun., 1988) this steeply NW-dipping zone is =0.5
km wide and strikes N30E. It also separates two
distinctly different lithostratigraphic sequences
that have different conditions of metamorphism: a
metasedimentary, supracrustal terrane (NW; 680-
750C, 6-7 kb), and a tonalitic, migmatitic gneiss
terrane (SE; 650-700C, 7-8 kb) (Mogk and Henry,
1988). As with the Madison mylonite, the small-
scale structures within the zone indicate the north-
west side has moved up relative to the southeast
side. An Archean age for this zone is consistent with
the ages of less deformed granitic plutons (2600-
2500 Ma) that intrude the shear zone (Mogk and
others, 1988). This shear zone apparently does not
show the strong 1600-1800 Ma regional retrograde
thermal event. Extrapolation of the Mirror Lake
shear zone has not been confirmed, but a similar
lithologic contrast occurs between metasedimentary
rocks and granitic gneiss in the west-central Madi-
son Range and the east-central Gravelly Range, and
this contact may also be mylonitic. The left sepa-
ration of this possible extrapolated contact may be
due to left-reverse movement along the Spanish
Peaks fault (Garihan and others, 1983).
Although these are the only described major
mylonite zones, the juxtaposition of distinctly dif-
ferent lithostratigraphic units in other locations
(such as between the Tobacco Root Mountains on the
east and the Highland Mountains metamorphic
terrane on the west) suggests that other major NE-
trending zones of late Archean or Early Proterozoic
deformation may exist. If basement rocks in south-
western Montana were deformed in a multiple-
suture accretionary zone, as suggested for the Chey-
enne belt (Dubendorfer and Houston, 1987), and
later overprinted by ductile normal-faulting during
an Early Proterozoic extensional event, then NW-
dipping, NE-trending mylonitic shear zones should
be ubiquitous across the region. Some of these are
clearly exposed but many others may be buried
beneath Phanerozoic Rocks. Any of these zones may
have had a stronger influence on later faulting than
did ordinary foliation or compositional layering (see
also Mogk and others, this volume).
If a shear zone exists at depth, then this ob-
servable surface pattern of NE-trending, NW-
dipping regional structural grain suggests that the
shear zone would have the same attitude. Although
isolated outcrops of basement rocks on the hanging-
wall of the Laramide Snow crest Thrust do not show
a mylonitic character as in other ranges, the fact
346 B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
that structure from three Phanerozoic tectonic
events all follow this earlier structural grain along
the western boundary of the Snowcrest Range
suggests both (1) the presence of a major shear zone
within the basement structure, and (2) the likeli-
hood that this postulated basement shear zone
influenced the strike and dip of Phanerozoic struc-
tures. We have modified the earlier term for this
feature, Greenhorn lineament (Maughan and Perry,
1982), to the term Snowcrest-Greenhorn lineament,
because most of the evidence for its existence is
found in the Snowcrest Range.
SNOWCREST TROUGH (PALEOZOIC)
Assuming that the Snowcrest-Greenhorn linea-
ment is controlled by a Precambrian mylonite zone,
the first episode of recurrent activity along it was a
Paleozoic event, as indicated by thickness and facies
changes in Paleozoic rocks. The principal structural
feature is referred to as the Snowcrest trough
(Maughan, 1984) (Figure 4).
The Snowcrest trough is interpreted as a NE-
trending, NW-dipping, downthrown block (half-
graben) adjacent to the Beartooth shelf (Maughan
and Perry, 1982). It is believed to be the south-
western extension of the Big Snowy trough of
central Montana (Key, 1986; Maughan, 1984) and
may extend westward to open up into the Cordi-
Figure 4. Palinspastic isopach map of Upper Missis-
sippian and Pennsy lvanian strata in western Montana
showing inferred major tectonic elements. SGL, Snow-
crest-Greenhorn lineament. Modified from Peterson
(1981).
lleran miogeosyncline (Key, 1986). Active deposi-
tion probably occurred within the trough through-
out the Paleozoic, but because of discontinuous and
limited exposures of Devonian and Mississippian
Madi.son Group rocks, only a general trend of the
trough's geometry can be determined from isopachs
of early Paleozoic strata.
One or more NE-trending listric normal faults
probably bounded the trough on the east, as inferred
from the great change in thickness of upper Paleo-
zoic formations across the Snowcrest-Greenhorn
lineament and the gradual thinning of these forma-
tions northwest of the lineament. Because all upper
Paleozoic formations thicken toward the lineament,
a growth fault history for this boundary may also be
inferred. Growth faulting was greatest during
Meramecian/Chesterian time, with deposition of the
Late Mississippian Lombard Limestone. This is re-
flected in a W/E-transect from the western Blacktail
Range southwest of Dillon, Montana, to the eastern
Snowcrest Range (Figure 2). The Lombard Lime-
stone is absent near Dillon (Petersen, 1981). Appro-
ximately 24 km southeast in the Blacktail Range, it
has thickened to 126 m (Wardlaw and Pecora, 1985).
The Lombard continues to thicken to the southeast,
where it is >457 m along the Blacktail Road in the
Snow crest Range (Byrne, 1985; Perry, 1982). How-
ever, 2.4 km to the southeast, across the inferred
normal fault boundary, the Lombard significantly
thins to 8 m within the Marathon Cornell Camp
"-
\....,,0
o.
500
'-"'''''
(
; 160
\
(
1330
1660
500
Pennsylvanian

KILOMETERS
_500 Thickness in meters
Multiple reactivation of a collisional boundary, southwestern Montana 347
federal 1-20 well. It appears to be absent within the
Shell 34x-13 well 11 km farther to the southeast
(Perry, 1986). Lithofacies sequences within the
Lombard indicate episodic, differential subsidence
within different parts of the basin, along with basin-
wide subsidence and deposition (Byrne, 1985).
Movement on NW-trending faults of Proterozoic
origin appears to have accommodated the differen-
tial subsidence within the trough (Maughan and
Perry, 1982).
Thickness variations require at least 449 m of
throw along the trough-bounding normal fault dur-
ing Meramecian/Chesterian time. This is compar-
able to inferred displacements adjacent to the Bear-
tooth shelf within the Big Snowy trough (Maughan,
1984) during this same period. Lithofacies, within
the Blacktail Range expsosure, of the Lombard
Limestone represent a more stable platform envi-
ronment compared to the Snowcrest Range (Byrne,
1985; Key, 1985; Maughan,personal commun.,
1988) and the Blacktail Range section may have
been near the hinge zone for subsidence within the
Snow crest trough (Byrne, 1985; Key, 1985).
By incorporating all Phanerozoic formations
beneath the regional Jurassic erosional surface and
plotting their cumulative thicknesses in their ap-
propriate pre-Laramide geographical positions, a
pre-Jurassic cross-section of the Snowcrest trough
was constructed along the transect line (Figure 5).
Pre-Laramide geographic positions were estimated
by assuming 20% Late Cretaceous shortening and
15% basin-range extension. The cross-section indi-
cates rollover of the Archean basement/Phanerozoic
contact toward the east (Figure 5).
The trajectory of the normal fault with depth
can be roughly estimated from rollover curvature
and dip using the displacement vector method of
Kilsdonk and Spang (1988). Rollover is estimated
from sub-Jurassic thicknesses. The normal fault
may have been reactivated as a Late Cretaceous
thrust (discussed below). If the normal fault dip was
the same as the present Snowcrest thrust (35NW),
the depth to detachment for the normal fault is = 21
km below sea level.
Because the trend of the Snowcrest trough
parallels the structural grain within Archean base-
ment rocks, the basin-bounding normal fault system
for the Snowcrest trough is interpreted to have
developed along the pre-existing structural grain of
the basement. In addition, the Snowcrest trough
probably created new zones of weakness within the
Phanerozoic rocks which were exploited by Lara-
mide thrusting.
SNOWCREST THRUST (LATE CRETACEOUS)
The Snowcrest-Greenhorn thrust system repre-
sents the frontal thrust system of the Blacktail-
Snowcrest Rocky Mountain foreland uplift. It con-
sists of a complex package of thrust imbricates,
splays, and shear zones that are interpreted to
merge at depth within Archean basement rocks
before flattening into a regional detachment. At the
surface, the thrust system is broadly concave west-
ward, with trends of N50o-60oE in the southern
region, gradually becoming oriented N-S to the
north (Figure 6). This change in trend is nearly
MARATHON'
S HI LL
Cora .1I C ... p
T.d..nl 1-20
.34 :1 - 13
E
w
Pre-Jurassic sur.ace

Snowcrest Trou h


0
o 10 20km
______ ______
Figure 5 Inferred pre-Jurassic geometry of the Snowcrest trough across the E-W transect of Figure 2. Black squares indicate
strike-parallel extrapolation of principal measured sections; well symbols indicate strike-parallel extrapolation of principal
well data; LP, lower Paleozoic rocks; UP, upper Paleozoic rocks; LM, lower Mesozoic (Triassic) rocks.
348 B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
112.- 00'
EXPLANATION
Q!J Quaternary- Tertiary basin

Late Cretaceous
Beaverhead Group

Mesoloic rocks

Paleoloic rocks
Ftteambfion {Archean)
U " metamorphic cornpltx.
undiwided
- lmtOl..OG.1C GOuPfOARI ES
-- 'THRUST ".AULT .......
",OY
......... HOAIIIA!. fAULT
.... 'N
HIQ.K-A*lLE FAlfLT
. RELAnVIE MOVl:lillun
-
onmA
I
AHTICl'H<
fSHOWlNO
PI.UNOlil
- {- SYNCLINE
"'"
"-
f
'<
"-..
.;:
5
U'"1I(111'5.
Figure 6. Tectonic map of the Snowcrest, Greenhorn, and
Gravelly Ranges. Area of Figure 7 is shown, Principal
references are Gealy (1953), Hadley (1969a), Mann (1954),
McBride (1988a), Sheedlo (1984), Zeigler (1954), and Perry
and others (1988).
identical to the change in trend observed in (1) the
structural grain of Archean basement rocks, (2) the
trace of the Snowcrest trough, and (3) the orienta-
tion of the Snowcrest-Greenhorn lineament.
The Snowcrest-Greenhorn thrust system is
composed of at least seven distinguishable thrust
zones that progressively become more gently dip-
ping eastward. In cross-section, the thrust system is
interpreted to represent a flat-ramp-flat geometry.
The hypothetical regional detachment within base-
ment is the lower flat, the transition from the base-
ment detachment to the Paleozoic/Archean contact
is the ramp, and the progressive flattening toward
the east, once the thrusts are no longer bounded by
basement along the footwall, is the upper flat. The
progressive (W-E) decrease in the dip of thrusts at
'.
I
.0
,
EXPLANATION


i
t::=J


...........

GLiI.Cl&1. UCl'WNE

C!J

lu ....... ,nltOCl(!I

UN"""_
I ! mfl Ce r(lIb:SfCflM./lftoH

c...::J
-' 1

flJlTHEAIlI o\,lun
t
-r ........... " ....
.,.l ..... t


..
..
o
I
0 .5 t
IllOMElElS
Figure 7. Geologic map of the west-central Snowcrest
Range showing basement rocks on the hanging wall of the
Snowcrest thrust. Map location shown in Figure 6. Lower
hemisphere equal-area contoured diagrams represent
Archean (foliat ion) and Laramide (fractures and fault
plane) fabric of basement rock.
Multiple reactivation of a collisional boundary, southwestern Montana 349
the surface is interpreted to represent the transition
from the ramp domain to the upper flat domain. The
flat-ramp-flat geometry, a common model used to
describe thrusts within the thrust belt, has been
only infrequently used to describe faults associated
with Rocky Mountain foreland uplifts (Erslev, 1986;
Bruhn and others, 1986).
The most prominent fault of the Snow crest-
Greenhorn thrust system is the Snowcrest thrust,
which places Archean quartzofeldspathic gneisses
and amphibolites above overturned Late Mississip-
pian Lombard Limestone. In the central Snowcrest
Range, the trend of the Snowcrest thrust is N40o-
50
o
E, with 35-50
o
NW dips calculated by the three
point method. These orientations mimic foliation
attitudes (N54E/41 NW) within Archean basement
(Figure 7).
Shear fracture analysis of Archean basement
rocks in the central Snowcrest Range indicates the
most common shear plane is oriented N54E, 39NW
(Figure 7). Assuming that the dominant shear-
fracture orientation mimics the master thrust
plane, a strong correlation between the dominant
shear fractures, basement foliation, and the trend of
the Snowcrest thrust can be made. The undated
shear fractures are probably Laramide features,
formed during uplift of the Blacktail-Snowcrest
uplift.
o 10 20 km
L... ' __ -'- , _ _ -',
p
The great thickness variations of the Late Mis-
sissippian Lombard Limestone across the present
Laramide thrusts of the Snowcrest Range indicate a
close spatial relationship between the thrusts and
the inferred Paleozoic trough-bounding normal
fault. Reactivation is inferred to have caused juxta-
position of thicker Snowcrest-trough facies adjacent
to thinner shelf facies. Assuming Laramide defor-
mation utilized pre-existing planes of weakness to
achieve shortening, foreland deformation within the
Snowcrest Range can be modeled by progressive
shortening of the Snow crest trough cross-section
(Figure 8). Footwall and hanging-wall imbricates
were placed where they were: (1) observed in the
field, (2) calculated in cross-sections, or (3) located in
wells.
Kinematic indicators such as slickensides, ex-
tension fractures, and calcite twin lamellae indicate
a Late Cretaceous S80o-85E principal shortening
direction (McBride, 1988a,b), which requires that a
significant component of right-lateral strike-slip
occurred within the Snowcrest-Greenhorn thrust
system in the southern Snowcrest Range, with an
increasing component of dip-slip motion to the
north. If no pre-existing basement structures were
present before Laramide thrusting, the expected
trend of thrusts would be perpendicular to the prin-
cipal shortening direction, or N10E. Pre-existing
-.
Figure 8. Early and late stages of
progressive Laramide shortening of
Snowcrest trough geometry (Figure
5) to form the Blacktail-Snowcrest
uplift. Transect line and rock desig-
nations are the same as in Figure 5.
Gravetly Terrane
- O!)
-"
..
. ,
- >0
350 B. C. McBride, C. ,J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
structures, therefore, appear to have deflected the
trace of the thrust system into a more northeasterly
orientation in order to exploit the pre-existing zone
of weakness regardless of the shortening direction.
SNOWCREST-GREENHORN RANGE-FRONT
FAULT SYSTEM (NEOGENE)
The northeasternmost extent of the Basin and
Range physiographic and structural province over-
laps the part of the Rocky Mountain foreland that
contains the Snowcrest and adjacent ranges (Par-
dee, 1950; Reynolds, 1979). The structures deve-
loped by Laramide thrusting and arching strongly
influenced the Neogene basin-range structures
(Schmidt and others, 1984; Schmidt and Dresser,
1984; Schmidt and Garihan, 1986).
A major system of range-front normal faults
(Sage Creek normal fault system) extends for"" 130
km along the western flank of the Snowcrest and
Greenhorn Ranges (Figure 6). The NNE-trending
fault system closely follows the strike of the Snow-
crest thrust and bounds it on the west. Like the
Snowcrest thrust, it strikes ""N-S in the Greenhorn
Range and changes progressively to N600E in the
central and southern Snowcrest Range. Three-point
constructions of the primary normal fault system
indicate 400-80
0
NW dips. The Snowcrest thrust dips
:5 40. Because the Snowcrest and associated
thrusts dip more gently than the normal faults,
their traces merge. In some cases, the fault traces
merge gradually; in others, normal faults appear to
truncate the Snowcrest thrust.
The primary normal fault system places Oligo-
cene to Pliocene basin deposits against Archean and
Paleozoic rocks. A secondary system of synthetic
Figure 9. Two hypotheses showing the
relationship of principal Neogene range-
front fault to Snowcrest-Greenhorn thrust
system. A, Al - the normal fault inter-
sects the ramp in the Snowcrest thrust
(ST) but merges with the deeper Green-
horn thrust (GT). B, BI - normal faults
merge into a Snowcrest thrust ramp.
normal faults between the primary normal fault
system and the thrusts, places upper Paleozoic rocks
against older Paleozoic or Archean rocks (Figure 7).
Locally in the central Snow crest Range, the se-
condary synthetic faults appear to have reactivated
older thrust surfaces within Mississippian rocks.
Although the exact stratigraphic throw along the
primary system of normal faults is unknown,
gravity modeling by Burfeind (1967) along a line
through the northern part of the range at Ruby
Canyon and by Kulik and Perry (1988) along a line
through the central part of the range at the West
Fork Blacktail Creek are consistent with 1000-3000
m of Cenozoic basin deposits adjacent to the range-
bounding fault.
The range-front fault system is interpreted to
be listric. This interpretation is based (1) on the
observation, supported by gravity data, that the
Upper Ruby and Sage Creek basins are asymme-
trical, with the deepest parts on the southeast
adjacent to the range-front fault system, and (2) on
the existence of two rollover anticlines in Tertiary
rocks on the hanging wall of the principal range-
front fault (Figure 6). The Ledford Creek anticline
in the Upper Ruby basin adjacent to and parallel
with the range-front fault in the northern Snow crest
Range was described by Sheedlo (1984). It is a
broad, horizontal anticline with gentle interlimb
angles (145-160) with the eastern limb dipping
< 25 into the principal range-front fault. The
Whiskey Spring anticline in the Sage Creek basin
adjacent to and subparallel to the principal range-
front fault in the central Snowcrest Range was
described by McBride (1988a). The interlimb angle
varies from 131 to 159. The Tertiary beds, which
form the southeastern limb of this fold, dip (14-
B
Multiple reactivation ofa collisional boundary, southwestern Montana 351
35SE) toward the Snowcrest Range and into the
range-front normal fault system.
The surface expression of the range-front fault
system lies west of the Snowcrest-Greenhorn thrust
system; therefore, either the two fault systems
merge at depth or they intersect and continue to
follow separate trajectories. Because the rollover of
these two anticlines is gradual and the dips are
gentle, the principal range-front fault must curve
gently (in the subsurface), maintaining a fairly
steep dip (40_50) where the two fault systems
merge or intersect. The range-front fault probably
either truncates the Snowcrest thrust and merges
with the deeper Greenhorn thrust (Figure 9A), or
the Snowcrest thrust locally steepens at the range
margin to form a basement ramp before flattening
and the two faults merge along this ramp (Figure
9B), similar to the localization of normal faults
along thrust ramps in the Cordilleran thrust belt
(Royse and others, 1975; Dixon, 1982).
The Sage Creek and Upper Ruby basins are
interpreted to be the result of the collapsed arch of
the Blacktail-Snowcrest uplift (Schmidt and others,
1984; Sheedlo, 1984). Collapse of the Laramide up-
lift was initiated when Cenozoic extension caused
the uplift to "slide back down" the earlier Snow-
crest-Greenhorn thrust system, forming the basins
and isolating the toes of the frontal thrust sheets.
The position of the principal range-front normal
fault may have been controlled by the position of the
basement ramp of the Snow crest-Greenhorn thrust
system and/or the position of major extensional
faults that developed on the arch during thrusting
(Figure 9).
OTHER REACTIVATED NE-TRENDING
STRUCTURES
Suturing of Archean terranes along NE-trend-
ing shear zones took place over a distance > 130 km
perpendicular to strike; thus, reactivation of the
suture zone should be widespread in basement rocks
of southwestern Montana. Erslev (1983) showed the
existence of at least two thrusts (presumably Lara-
mide) on either side of the Madison mylonite zone,
both appear to follow schistose zones in NW-dipping,
sheared Archean rocks. These zones merge with the
N-trending Hilgard thrust system of Tysdal (1986),
suggesting that they are Laramide features.
The Cenozoic Deep Creek normal fault, which
forms the western boundary of the Beartooth Range
and is the fault along which the Yellowstone Valley
down-dropped between the Gallatin and Beartooth
Ranges, follows the NW -extrapolated extension of
the Snowy shear zone (Montagne and Chadwick,
1982; Erslev, this volume). Along the N-trending
Madison Range-front fault, there are several short
NE-trending segments which clearly follow the
earlier foliation trends and form range-margin
corners with the principal range-margin fault
(Hadley, 1969a; Johns and others, 1982).
Along the western boundaries of the Tobacco
Root Mountains and Ruby Range, NE-trending
Laramide thrusts with Archean rocks on their hang-
ing walls are bounded on the west by NE-trending,
steeply dipping, range-front normal faults (Figure
10) (Samuelson and Schmidt, 1981; Schmidt and
Garihan, 1983). The geometries are nearly identical
to those we have described for the Snowcrest Range.
In the Tobacco Root Mountains, the westernmost
thrust (Hellroaring thrust) places Precambrian
marbles, schists, and quartzites on steep to over-
turned upper Paleozoic rocks. Layering in Archean
rocks strikes nearly parallel to the thrust and dips
(35_50) southeasterly into the thrust (Figure
lOA). These Archean rocks are significantly differ-
ent in composition and attitude from Archean rocks
in the core of the Tobacco Root Mountains 3.2 km to
the southeast. However, they are very similar in
composition and attitude to basement rocks in the
Highland Range across the Jefferson Valley to the
west (Duncan, 1976; O'Neill and others, 1988a,b).
Actual separation along the thrust is unknown (pro-
bably = 3-5 km). However, because restoring any
amount of thrust movement will not juxtapose
similar Archean rocks on the hanging wall and
footwall, it is likely that the Laramide Hellroaring
thrust occurred along an Archean or Early Protero-
zoic structure. This proposed structure may be the
same fault along which the 1800 Ma, thermally
metamorphosed rocks of the Highland Range were
juxtaposed against 2700 Ma rocks in the Tobacco
Root Range, largely unaffected by the Early Pro-
terozoic thermal event according to O'Neill and
others (1988a,b).
Tysdal (1976) mapped a 3.2 km-long disrupted
block of schists, amphibolites, and marbles on the
northwestern flank of the Ruby Range (Figure
lOB). Schmidt and Garihan (1983) suggested that
this block was a W-dipping, basement-involved
thrust similar to, if not ultimately connected with,
the Hellroaring thrust in the western Tobacco Root
Mountains. The thrust dips 30
0
-400NW, and steeply
352 B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
Figure 10. Maps of the western Tobacco
Root Mountains (A) and northwestern
Ruby Range (B) showing the relationship
of Laramide thrusts to Neogene normal
faults and to fabric elements in hanging-
wall basement rocks.
LEGEND
NORTHWESTERN RUBY RANGE
r 0.. =

.t..I.l $.fIllUCIU.....l'U.' .... U 00/1$1'4,,,.1_1"[
":II1II(0-
oon(Owlt(iII(CDNCfAlto
o
..
- ! ::= -,
. .,...,
. - ''!'HeLm! -
27 POLS TO
MEU. ... ORPHIC fOUAnoN
NW-dipping foliations in the hanging wall strike
parallel to the thrust (Figure lOB). Although the
composition of hanging-wall rocks is somewhat
similar to those that core the northern Ruby Range
to the southeast (Karasevich, 1980), the orienta-
tions are completely different and thicknesses (10
m) of marble units in the fault block do not compare
with those (300 m) in the core of the range at Ruby
Peak. Again, this lack of comparison in thickness
and structure suggests pre-Laramide movement
along this NE-trending zone.
The range-bounding normal faults in the west-
ern Tobacco Root Mountains and western Ruby
Range have the same strike as, but a steeper dip
than, Laramide thrusts and probably have the same
relationship to thrusts as that postulated along the
western flank of the Snowcrest Range.
SUMMARY AND CONCLUSIONS
Reactivation of pre-existing zones of weakness,
rather than formation of new faults, is a widely
recognized phenomenon of intracontinental defor-
mation (Prucha, this volume; Sykes, 1978; Jackson,
1980; Winslow, 1981; Sibson, 1985). We have pre-
sented evidence for three episodes of reactivation
along NE-trending, NW-dipping faults or zones of
weakness in southwestern Montana following their
inception during late Archean continental accre-
tion. The principal deformational episodes may be
summarized as follows:
1. NE-trending zones of mylonite and pervasive
foliation formed over a 100-150-km zone of suturing
during final amalgamation of at least three distinct
lithostratigraphic terranes in late Archean time
(2600-2500 Ma; Mogk and others, 1988) (Figure
lIA). Some of these zones were strongly modified
by Early Proterozoic (1800-1600 Ma) shearing and
retrograde metamorphism (Erslev, 1983; O'Neill
and others, 1988a,b).
2. Formation of the Snowcrest trough in late
Paleozoic time probably occurred along a NE-trend-
ing, W-dipping, listric normal fault inferred to close-
ly follow one of the principal zones of Precambrian
weakness in the Snowcrest Range (Figure lIB).
3. Regional Late Cretaceous W-E shortening ap-
parently reactivated the normal fault as a right
oblique-slip thrust (Figure lIC). This thrust and
its associated hanging wall and footwall splays
constitute the Snowcrest-Greenhorn thrust system
on the eastern margin of the Blacktail-Snowcrest
uplift.
4. Crustal extension, also along a roughly E-W-
trending line, during Neogene time caused the
Blacktail-Snowcrest uplift to collapse along a major
Multiple reactivation of a collisional boundary, southwestern Montana 353
Figure 11. Inferred sequence of events along the E-W-
transect across the Blacktail and Snowcrest Ranges
(Figure 2). Rock units are the same as in Figure 5; addi-
tional rock units: UM, upper Mesozoic rocks; Kb, Upper
Cretaceous Beaverhead Group; Ts, Tertiary sedimentary
rocks.
range-bounding normal fault system inferred to
merge at depth with the Snowcrest-Greenhorn
thrust system (Figure llD).
Reactivation of the Laramide Snowcrest-Green-
horn thrust system by Neogene normal faults is
inferred from the following: (1) the two fault sys-
tems have subparallel strikes; (2) the thrusts dip
gently northwestward; and (3) the normal faults,
although steep at the surface, have rollover anti-
clines, which unequivocally establish their listric
geometry. The exact nature of the interaction of the
two faults remains conjectural, but we suggest that
the normal faults merge into a Laramide thrust
ramp.
The hypothesis of late Paleozoic reactivation of
a Precambrian zone of weakness is also conjectural,
supported largely by regional and local coincidence
of Precambrian trends and upper Paleozoic isopachs.
However, the close association of later (Laramide)
faults and Precambrian shear zones and foliation
trends is well documented in the Snowcrest and
adjacent ranges. Also, NE-trending Laramide
faults, formed by a WoE shortening direction, sup-
port pre-existing structural control for these anoma-
lous Laramide trends.
The importance of a wide belt of NE-trending
structures in the fold and thrust belt in east-central
Idaho and west-central Montana has been recog-
nized for a long time (for example, Billingsley,
1915), and the suggestion that such structures were
recurrently active since Precambrian time was
made over forty years ago (Billingsley and Locke,
1941). O'Neill and Lopez (1985) recently docu-
mented recurrent activity along this NE structural
trend and named the feature the Great Falls tec-
tonic zone (Figure 1). They, as well as earlier
authors, did not extend the width of the zone very
far into the basement rocks of southwestern Mon-
tana. However, recent work by Erslev (1983), Mogk
and Henry (1988), and O'Neill and others (1988a,b)
on the Precambrian fabric, and the recognition that
these Precambrian trends are inherited by later
deformation, suggests that the Great Falls tectonic
zone extends southeastward at least as far as the
western Beartooth Range. This almost doubles the
width of the zone and provides strong supporting
evidence that it was active from late Precambrian
time to Neogene time, as O'Neill and Lopez (1985)
suggested. The recognition that the zone incor-
porates the Precambrian structures, particularly
mylonite zones, which are exposed in the basement
uplifts of southwestern Montana also lends credence
to the speculation made by O'Neill and Lopez (1985)
that the region is underlain by NE-trending zones of
crustal weakness that separate two Precambrian
structural terranes. It appears that this suture zone
must be at least 250 km wide.
Although Mogk and others (this volume) have
argued convincingly for the existence of amalga-
mated Archean tectonostratigraphic terranes or
domains, based on work in the Beartooth, Gallatin,
and northern Madison ranges, Erslev (1988) has
argued equally convincingly for stratigraphic proto-
lith correlations across shear zones, based on work
in the Beartooth, southern Madison, and Gravelly
ranges. If such correlations are correct, then the
existence of discrete blocks (or terranes), like we
have shown (Figure 3) is less likely. It is unlikely
that this problem will be resolved soon, because a
satisfactory resolution involves regional recognition
and documentation of the extent and signature of
the late Archean and Early Proterozoic events.
354 B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
However, our principal point and interpretation in
this paper, that Late Proterozoic, Late Cretaceous,
and Neogene structures are influenced by old
Precambrian shear zones, is not affected by either
interpretation of Precambrian history.
ACKNOWLEDGMENTS
We thank Dave Mogk and Mike O'Neill for
sharing ideas and information about Precambrian
fabrics and events in southwestern Montana, Ed
Maughan for sharing ideas and information about
the Snowcrest trough, Bill Perry for sharing ideas
about the Snowcrest-Greenhorn thrust system,
Dave Lageson and Eric Erslev for reviewing and
suggesting improvements for this paper. We are
also indebted to the late Robert Chadwick for
information and earlier references on the Great
Falls tectonic zone. McBride gratefully acknow-
ledges the support of the American Association of
Petroleum Geologists' Grants-in-Aid program,
Marathon Oil Company, and Western Michigan
University Graduate College. Schmidt gratefully
acknowledges the financial support of the Depart-
ment of Geology, Utah State University, and thanks
Jim Evans and Coleen Sheehan for providing sup-
port during manuscript preparation.
REFERENCES
BILLINGSLEY, P., 1915, The Boulder batholith of Montana:
American Institute of Mining Engineers Transactions, v.
51, p. 31-47.
__ , and A. F. LOCKE, 1941, Structures and ore districts
in the continental framework: American Institute of Min-
ing Engineers Transactions, v. 144, p. 9-59.
BRASHER, G. K., 1950, Geology of Part of the Snowcrest
Range, Beaverhead County, Montana: MS thesis, Uni-
versity of Michigan, Ann Arbor, Michigan, USA, 58 p.
BROWN, W. G., 1988, Laramide deformational style in the
Wyoming foreland; pp. 1-26 in C. J. Schmidt and W. J.
Perry (eds.), Interaction of the Rocky Mountain Fore-
land and the Cordilleran Thrust Belt: Geological Socie-
ty of America, Memoir 171,582 p.
BRUHN R. L., M. D. PICARD, and J. S. ISBY, 1986, Teconics
and sedimentology of Uinta arch, western Uinta Moun-
tains, and Uinta basin; pp. 333-352 in J. A. Peterson (ed.),
Paleotectonics and Sedimentation in the Rocky Moun-
tain Region, United States: American Association of
Petroleum Geologists, Memoir 41, 693 p.
BURFEIND, W. J., 1967, A Gravity investigation of the
Tobacco Root Mountains, Jefferson Basin, Boulder
Batholith and Adjacent Areas of Southwestern Mon-
tana: PhD dissertation, Indiana University, Bloomington,
Indiana, USA, 89 p.
BYRNE, D. J., 1985, Stratigraphy and Depositional
History of the Upper Mississippian Big Snowy Forma-
tion in the Snowcrest Range, Southwestern Montana:
MS thesis, Oregon State University, Corvallis, Oregon,
USA,417,24p.
DIXON, J. S., 1982, Regional structural synthesis, Wyom-
ing salient of western Overthrust Belt: American Associa-
tion of Petroleum Geologists Bulletin, v. 66, p. 1560-1580.
DUBENDORFER, E. M., and R. S. HOUSTON, 1987, Pro-
terozoic accretionary tectonics at the southern margin of
the Archean Wyoming craton: Geological Society of Ameri-
ca Bulletin, v. 98, p. 554-568.
DUNCAN, M. S., 1976, Structural Analysis of the Pre-
Beltian Metamorphic Rocks of the Southern Highland
Mountains, Madison and Silver Bow Counties, Mon-
tana: PhD dissertation, Indiana University, Bloomington,
Indiana, USA, 222 p.
ERSLEV, E. A., 1982, The Madison mylonite zone: A major
shear zone in the Archean basement of southwestern
Montana; pp. 213-221 in S. G. Reid (ed.), Geology of the
Yellowstone Park Area: 33rd Annual Field Confer-
ence Guidebook: Wyoming Geological Association, Cas-
per, Wyoming, USA, 221 p.
__ ,1983, Pre-Beltian Geology of the Southern Madi-
son Range, Southwestern Montana: Montana Bureau of
Mines and Geology, Memoir 55, 26 p.
__ , 1986, Basement balancing of Rocky Mountain fore-
land uplifts: Geology, v. 14, p. 259-262.
FLANAGAN, W. H., 1958, Geology of the Southern part
of the Snowcrest Range, Beaverhead County, Mon-
tana: MS thesis, Indiana University, Bloomington, In-
diana, USA, 41 p.
FOUNTAIN, D. M., and N. R. DESMARAIS, 1980, Waboden
terrane of Manitoba and the pre-Belt basement of south-
western Montana: A comparison; pp. 35-46 in M. R. Miller
(ed.), Guidebook of the Drummond-Elkhorn Areas,
West-Central Montana: Montana Bureau of Mines and
Geology, Special Publication 82, 46 p.
GARIHAN, J. M., C. J. SCHMIDT, S. W. YOUNG, and M. A.
WILLIAMS, 1983, Geology and recurrent movement history
of the Bismark-Spanish Peaks-Gardiner fault system,
southwest Montana; pp. 295-314 in J. D. Lowell and R.
Gries (eds.), Foreland Basins and Uplifts: Proceedings
Multiple reactivation of a collisional boundary, southwestern Montana 355
of the Symposium: Rocky Mountain Association of Geo-
logists, 392 p.
GEALY, W. J., 1953, Geology of the Antone Peak
Quadrangle, Southwestern Montana: PhD dissertation,
Harvard University, Cambridge, Massachusetts, USA, 143
p.
GILE'ITI, B. J., 1966, Isotopic ages from southwestern
Montana: Journal of Geophysical Research, v. 71, p. 4029-
4036.
__ , 1971, Discordant isotopic ages and excess argon in
biotites: Earth and Planetary Science Letters, v. 10, p. 157-
164.
GRIES, R, 1983, North-south compression of Rocky Moun-
tain foreland basins and uplifts; pp. 9-32 in J. D. Lowell
and R Gries ( eds.), Foreland Basins and Uplifts: Pro-
ceedings of the Symposium: Rocky Mountain Associa-
tion of Geologists, 392 p.
HADLEY, J. B., 1960, Geology of the northern part of the
Gravelly Range, Madison County, Montana: Billings Geo-
logical Society Guidebook, Eleventh Annual Field Confer-
ence, p. 149-153.
__ , 1969a, Geologic Map of the Varney Quadrangle,
Madison County, Montana: U.S. Geological Survey, Map
GQ-814, 1:62,500 scale.
__ , 1969b, Geologic Map of the Cameron Quad-
rangle, Madison County, Montana: U.S. Geological Sur-
vey, Map GQ-813, 1:62,500 scale.
HAMILTON, W., 1988, Laramide crustal, shortening; pp.
27-40 in C. J. Schmidt and W. J. Perry, Jr. (eds.), Inter-
action of the Rocky Mountain Foreland and the Cor-
dilleran Thrust Belt: Geological Society of America,
Memoir 171, 582 p.
JACKSON, J. A., 1980, Reactivation of basement faults and
crustal shortening in orogenic belts: Nature, v. 283, p. 343-
346.
JOHNS, W. M., W. T. STRAW, R N. BERGANTINO, H. W.
DRESSER, T. E. HENDRIX, H. G. McCLERNAN, J. C.
PALMQUIST, and C. J. SCHMIDT, 1982, Neotectonic Fea-
tures of Southern Montana East of 11230' West Lon-
gitude: Montana Bureau of Mines and Geology, Open File
Report 91,79 p.
KARASEVICH, L. P., 1980, Structure of the Pre-Beltian
Metamorphic Rocks of the Northern Ruby Range,
Southwestern Montana: MS thesis, Pennsylvania State
University, University Park, Pennsylvania, USA, 172 p.
__ , J. M. GARIHAN, P. S. DAHL, and A. F. OKUMA,
1981, Summary of Precambrian metamorphic and struc-
tural history, Ruby Range, southwest Montana; pp. 225-
237 in T. E. Tucker (ed.), Southwest Montana Field
Conference Guidebook: Montana Geological Society,
Billings, Montana, USA, 406 p.
KEENMON, K. A., 1950, Geology of the Blacktail-
Snowcrest Region of Beaverhead County, Montana:
PhD dissertation, University of Michigan, Ann Arbor,
Michigan, USA, 207 p.
KEY, C. F., 1986, Stratigraphy and Deposition History
of the Amsden and Lower Quadrant Formations,
Snowcrest Range, Beaverhead and Madison Counties,
Montana: MS thesis, Oregon State University, Corvallis,
Oregon, USA, 187 p.
KILSDONK, B., and J. H. SPANG, 1988, New geometrical
techniques to infer listric normal fault shapes [abstract):
American Association of Petroleum Geologists Bulletin, v.
72, no. 2, p. 206.
KLEPPER, M. R, 1950, A geologic reconnaissance of parts
of Beaverhead and Madison Counties, Montana: U.S. Geo-
logical Survey, Bulletin 969-C, p. 55-85.
KULICK, D. M., and W. J. PERRY, Jr., 1988, The Blacktail-
Snowcrest foreland uplift and its influence on structures of
the Cordilleran thrust belt - Geological and geophysical
evidence for the Overlap Province in southwestern
Montana; pp. 291-306 in C. J. Schmidt and W. J. Perry, Jr.
(eds.), Interaction of the Rocky Mountain Foreland
and the Cordilleran Thrust Belt: Geological Society of
America, Memoir 171, 582 p.
MANN, J. A., 1954, Geology of Part of the Gravelly
Range, Montana: YBRA Contribution 190, Red Lodge,
Montana, USA.
MAUGHAN, E. K., 1984, Paleogeographic setting of Penn-
sylvanian Tyler Formation and relation to underlying
Mississippian rocks in Montana and North Dakota:
American Association of Petroleum Geologists Bulletin, v.
68, p. 178-195.
__ , and W. J. PERRY, Jr., 1982, Paleozoic tectonism in
southwestern Montana [abstract): Geological Society of
America Abstracts with Programs, v. 14, no. 6, p. 341.
McBRIDE, B. C., 1988a, Geometry and Kinematics of the
Central Snowcrest Range: A Rocky Mountain Fore-
land Uplift in Southwestern Montana: MS thesis,
Western Michigan University, Kalamazoo, Michigan,
USA, 267p.
__ , 1988b, Effects of Late Cretaceous compression in
southwestern Montana: Implications of thrust motions
and Rocky Mountain foreland deformation from the
Snowcrest Range [abstract): Geological Society of America
Abstracts with Programs, v. 20, no. 6, p. 431.
MOGK, D. M., and D. J. HENRY, 1988, Metamorphic petro-
logy of the northern Archean Wyoming province, south-
western Montana: Evidence for Archean collisional tec-
tonics; pp. 363-382 in W. G. Ernst (ed.), Metamorphism
356 B. C. McBride, C. J. Schmidt, G. E. Guthrie, and M. K. Sheedlo
and Crustal Evolution of the Western United States
(Rubey Volume 7). Prentice-Hall, Englewood Cliffs, New
Jersey, USA, ll53 p.
__ , P. MUELLER, E. WEYAND, J. WOODEN, and D.
BOWES, 1988, Protracted magmatic history of an Archean
collisional zone, northern Madison Range, SW Montana
[abstract]: Geological Society of America Abstracts with
Programs, v. 20, no. 7, p. A50.
MONROE, J. S., 1976, Vertebrate Paleontology, Strati-
graphy, and Sedimentation of the Upper Ruby River
Basin, Madison County, Montana: PhD dissertation,
University of Montana, Missoula, Montana, USA, 345 p.
MONTAGNE, J., and R A. CHADWICK, 1982, Cenozoic
history of the Yellowstone Valley between Yellowstone
Park and Livingston, Montana; pp. 31-51 in S. G. Reid
(ed.), Geology of the Yellowstone Park Area: 33rd An-
nual Field Conference Guidebook: Wyoming Geologi-
cal Association, Caspar, Wyoming, USA, 221 p.
MONTGOMERY, C. W., 1981, Metamorphic history of
Archean schists from the southwest Beartooth Mountains,
Montana and Wyoming [abstract]: Geological Society of
America Abstracts with Programs, v. 13, no. 4, p. 220.
O'NEILL, J. M., and D. A. LOPEZ, 1985, Character and
regional significance of Great Falls Tectonic zone, east-
central Idaho and west-central Montana: American Asso-
ciation of Petroleum Geologists Bulletin, v. 69, p. 437-447.
__ , M. S. DUNCAN and R E. ZARTMAN, 1988a, An
Early Proterozoic gneiss dome, Highland Mountains,
Montana [abstract]: Geological Society of America Ab-
stracts with Programs, v. 20, no. 6, p. 461-462.
__ , M. S. DUNCAN and R E. ZARTMAN, 1988b, An
Early Proterozoic gneiss dome, Highland Mountains, Mon-
tana; pp. 81-88 in S. E. Lewis and R B. Berg, (eds.),
Precambrian and Mesozoic Plate Margins: Montana
Bureau of Mines and Geology, Special Publication 96, 195
p.
__ , D. C. FERRIS, C. J. SCHMIDT, and D. HANNEMAN,
1986, Recurrent movement along northwest-trending
faults at the southern margin of the Belt Basin, Highland
Mountains, southwestern Montana; pp. 209-216 in S. M.
Roberts (ed.), Belt Supergroup: A Guide to Proterozoic
Rocks of Western Montana and Adjacent Areas: Mon-
tana Bureau of Mines and Geology, Special Publication 94,
311p.
PARDEE, J. T., 1950, Late Cenozoic block faulting in
western Montana: Geological Society of America Bulletin,
v. 61, p. 359-406.
PERRY, W. J., Jr., 1982, The thrust belt in the Lima-Dell,
Montana area; pp. 69-78 in P. Beaver (ed.), The Over-
thrust Province in the Vicinity of Dillon, Montana, and
How This Structural Framework has Influenced Min-
eral and Energy Accumulation: Guidebook to the 7th
Annual Field Conference: Tobacco Root Geological
Society, 87 p.
__ , 1986, Critical deep drill holes and indicated
Paleozoic paleotectonic features north of the Snake River
downwarp in southern Beaverhead County, Montana, and
adjacent Idaho. U.S. Geological Survey, Open File Report,
p.86-413.
__ , J. C. HALEY, D. J. NICHOLS, and P. W. HAMONS,
1988, Interactions of Rocky Mountain foreland and
Cordilleran thrust belt in the Lima region, southwest
Montana; pp. 267-290 in C. J. Schmidt and W. J. Perry, Jr.
(eds.), Interaction of the Rocky Mountain Foreland
and the Cordilleran Thrust Belt: Geological Society of
America, Memoir 171,582 p.
PETERMAN, Z. E., 1979, Geochronology and the Archean of
the United States: Economic Geology, v. 74, p. 1544-1562.
PETERSON, J. A., 1981, General stratigraphy and regional
paleostructure of the western Montana overthrust belt; pp.
5-35 in T. E. Tucker (ed.), Southwest Montana Field
Conference and Symposium: Montana Geological
Society, Billings, Montana, USA, 406 p.
RATCLIFFE, N. M., and W. L. NEWELL, 1986, Comment on
"Preexisting fault control for Mesozoic basin formation in
eastern North America": Geology, v. 14, p. 973-974.
REID, R R, W. J. McMANNIS, andJ. D. PALMQUIST, 1975,
Precambrian Geology of the North Snowy Block,
Beartooth Mountains, Montana: Geological Society of
America, Special Paper 157,235 p.
REYNOLDS, M. W., 1979, Character and extent of basin-
range faulting, western Montana and east-central Idaho;
pp. 185-193 in G. Newman and H. Goode (eds.), Basin and
Range Symposium: Proceedings: Rocky Mountain As-
sociation of GeologistslUtah Geological Association, Salt
Lake City, Utah, USA, 662 p.
ROYSE, F. C., M. A. WARNER, and D. L. REESE, 1975,
Thrust belt structural geometry and related stratigraphic
problems, Wyoming-Idaho-northern Utah; pp. 41-54 in D.
W. Bolyard (ed.), Deep Drilling Frontiers of the Central
Rocky Mountains: Proceedings of the Symposium:
Rocky Mountain Association of Geologists, 354 p.
SAMUELSON, K. J., and C. J. SCHMIDT, 1981, Structural
geology of the western Tobacco Root Mountains; pp. 191-
199 in T. E. Tucker (ed.), Southwest Montana Field
Conference and Symposium Guidebook: Montana Geo-
logical Society, Billings, Montana, USA, 406 p.
SCHMIDT, C. J., and H. DRESSER, 1984, Late Cenozoic
structural inversion in a region of small extensional strain,
southwestern Montana [abstract]: Geological Society of
America Abstracts with Programs, v. 16, p. 646-647.
__ , and J. M. GARIHAN, 1983, Laramide tectonic
development of the Rocky Mountain foreland of south-
Multiple reactivation of a collisional boundary, southwestern Montana 357
western Montana; pp. 271-294 in J. D. Lowell and R. Gries
(eds.), Foreland Basins and Uplifts: Proceedings of the
Symposium: Rocky Mountain Association of Geologists,
392 p.
__ , and, J. M. GARIHAN, 1986, Role of recurrent
movement of northwest-trending basement faults in the
tectonic evolution of southwestern Montana; pp. 1-15 in M.
J. Aldrich and R. Laughlin (eds.), Basement Tectonics:
Proceedings of the 6th International Conference
(Sante Fe, New Mexico, 1985): International Basement
Tectonics Association, Salt Lake City, Utah, USA, 208 p.
__ , M. K. SHEEDLO, and M. WERKEMA, 1984, Control
of range-boundary normal faults by earlier structures,
southwestern Montana [abstract]: Geological Society of
America Abstracts with Programs, v 16, p. 253.
SCHOLTEN, R., K. A. KEENMON, and W. O. KUPSCH,
1955, Geology of the Lima region, southwestern Montana
and adjacent Idaho: Geological Society of America Bulletin,
v. 66, p. 345-403.
SHEEDLO, M. K., 1984, Structural Geology of the
Northern Snowcrest Range, Beaverhead and Madison
Counties, Montana: MS thesis, Western Michigan Uni-
versity, Kalamazoo, Michigan, USA, 131 p.
SIBSON, R. H., 1985, A note on fault reactivation: Journal
of Structural Geology, v. 7, p. 751-754.
SPENCER, E. W., and J. L. KOZAK, 1975, Precambrian
evolution of the Spanish Peaks area, Montana: Geological
Society of America Bulletin, v. 86, p. 785-792.
STONE, D. S., 1969, Wrench faulting and Rocky Mountain
tectonics: Mountain Geologist, v. 6, p. 67-79.
SUMNER, W., and E. A. ERSLEV, Late Archean thin-
skinned thrusting of the Cherry Creek metamorphic suite
in the Henry's Lake Mountains, southern Madison Range,
Montana and Idaho; pp. 69-79 in S. E. Lewis and R. B Berg
(eds.), Precambrian and Mesozoic Plate Margins: Mon-
tana Bureau of Mines and Geology, Special Publication 96,
195p.
SWANSON, M. T., 1986, Preexisting fault control for Meso-
zoic basin formation in eastern North America: Geology, v.
14, p. 419-422.
SYKES, L. R., 1978, Intraplate seismicity, reactivation of
preexisting zones of weakness, alkaline magmatism, and
other tectonism post-dating continental fragmentation:
Reuiews of Geophysics and Space Physics, v. 16, p. 621-687.
TILFORD, M. J., 1976, Structural Analysis of the South-
ern and Western Greenhorn Range, Madison County,
Montana, and Magnetic Beneficiation of Montana
Talc Ores: MS thesis, Indiana University, Bloomington,
Indiana, USA, 143 p.
TYSDAL, R. G., 1976, Geologic Map of the Northern Part
of the Ruby Range, Madison County, Montana: U.S.
Geological Suruey, Miscellaneous Geologic Investigations
Map 1-951, 1:24,000 scale.
__ , 1986, Thrust faults and back thrusts in the Madison
Range of southwestern Montana foreland: American
Association of Petroleum Geologists Bulletin, v. 70, no. 4, p.
360-376.
VAUGHAN, W. J., 1948, The Geology of the West Fork of
the Madison River Area, Montana: MS thesis, Uni-
versity of Michigan, Ann Arbor, Michigan, USA, 43 p.
VITALIANO, C. J., and W. S. CORDUA, 1979, Geologic
Map of the Southern Tobacco Root Mountains, Madi-
son County, Montana: Geological Society of America,
Map and Chart Series MC-31, 1:62,500 scale.
WARDLAW, B. R., and W. C. PECCORA, 1985, New Missis-
sippian-Pennsylvanian stratigraphic units in southwest
Montana and adjacent Idaho: U.S. Geological Survey,
Bulletin 1656-B, p. B1-B9.
WINSLOW, M. A., 1981, Mechanism for basement short-
ening in the Andean foreland fold belt of southern South
America; pp. 513-528 in K. McClay and N. J. Price (eds.),
Thrust and Nappe Tectonics: Geological Society of
London, Special Publication 9, 539 p.
ZEIGLER, J. M., 1954, Geology of the BlacktaU Area;
Beaverhead County, Montana: PhD dissertation, Har-
vard University, Cambridge, Massachusetts, USA, 147 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Long-lived basement weak zones and their role in extensional
magmatism in the Ogilvie Mountains, Yukon Territory
C. F. ROOTS and R. I. THOMPSON
Geological Survey of Canada, 100 West Pender Street, Vancouver, BC, V6B 1R8, Canada
(received August 12, 1988; revision accepted February 17, 1989)
ABSTRACT
In the Ogilvie Mountains of west-central Yukon, intrusive breccias and volcanic rocks are aligned along
faults that were active during evolution of the Cordilleran miogeocline. Character of the magmatic rocks
provide clues to changing lower crust or upper mantle conditions beneath the western margin of ancestral
North America.
The unusually long (Middle Proterozoic to Jurassic) sedimentary record was influenced by episodes of
crustal stretching. Extensional events are indicated by proximal clastic sequences deposited in half-grabens
during extensional faulting. Igneous rocks associated with each stretching event have been dated: hematitic
megabreccia (1270 Ma), bimodal tholeiitic volcanic rocks (750 Ma), Cambro-Ordovician alkalic basalt, and
mid-Cretaceous alkalic granites (90 Ma).
Although crystalline basement rocks are not exposed in the northern fold and thrust belt, the combination
of inferences from the sedimentary record, ancient structures, and igneous chemistry imply recurring zones of
weakness and periodic mantle disturbance beneath the ancient continental margin.
INTRODUCTION
The Cordilleran miogeocline has a stratigraphic
record that spans more than 1300 million years - a
remarkably long period of time. Most strata consist
either of mature terrigenous clastics or shallow
water platform and shelf carbonates, facies typical
of a passive margin setting. Elsewhere, passive
margin successions, ancient and modern, record far
less time; for example, 250 million years or less for
the Alps and the Appalachians, 200 million years or
less for today's Atlantic margins, and only 10
million years for the Proterozoic Coronation mio-
geocline (Hoffman, 1983). The process interpreted
to "drive" subsidence, conductive cooling of a ther-
mal anomaly, reaches equilibrium in "" 200 million
years or less (McKenzie, 1978); as the thermal
anomaly decays, so too does the rate of subsidence.
If we adopt this model for basin subsidence, how
does one explain the long-lived subsidence neces-
sary for the accumulation of thick, shallow-water,
siliciclastic and carbonate sequences of Middle and
Late Proterozoic and Phanerozoic ages along the
margin of ancestral North America? One possibility
(Thompson and others, 1987) is to view the miogeo-
cline as an overlapping succession of discrete pas-
sive margin successions, each recording subsidence
associated with conductive cooling of a discrete
thermal anomaly. Another possibility is to view
parts of the miogeocline as having accumulated in
episutural basins (Hoffman, 1988) or in a foreland
basin (Ross, 1988).
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 359-372. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
359
360 C. F. Roots and R. I. Thompson
We use the term "ancestral North America"
when referring to the North American craton prior
to construction of the Cordillera and accretion of
exotic terranes (Quesnellia, Stikinia, Cache Creek,
Wrangellia, Alexander; Monger and others, 1982);
the term "Cordilleran miogeocline" is used to de-
scribe the thick sedimentary terrace wedge that
accumulated along the western margin of ancestral
North America from Middle Proterozoic until
Middle Jurassic time (Gabrielse and Yorath, 1989).
The idea that the Cordilleran miogeocline
underwent multiple rift events is not new: Burch-
fiel and Davis (1975) suggested continental breakup
Ma, followed by renewed extension
Ma; backstripping models applied to the southern
Canadian Cordillera by Bond and Kominz (1984)
point to an extension 550 Ma; sedimentologic and
structural relations in the central Yukon suggest a
period of extension in the northern Canadian Cordi-
llera during the late Devonian and early Carboni-
ferous (Abbott, 1982; Gordey, in press). Thompson
and others (1987) have described evidence for
extension in the Ogilvie Mountains = 1200, 750,
and 450 Ma. In each case, the hypothesized rift
event was followed by platform and shelf sedi-
mentation typical of a passive margin setting.
Stratigraphic facies relationships and paleocurrent
data suggest that segments of the margin were
paleogeographically complex, composed of crustal
"islands" or "ribbons" separated from the main
continent by narrow? seaways or troughs (Struik,
1987).
The purpose of this paper is to describe the
stratigraphic and magmatic history for part of the
Canadian Cordilleran miogeocline in west-central
Yukon - the Ogilvie Mountains - recognizing
those time intervals when the basement was active
magmatically and tectonically. Our results suggest
that episodic activity along a long-lived basement
weak zone played an important role in the evolution
of this part of the ancestral North American margin.
REGIONAL GEOLOGIC SETTING
The Cordilleran miogeocline consists of four
major time-stratigraphic packages: (1) the Purcell
and Wernecke Supergroups, the Pinguicula Group,
and the Mackenzie Mountains Supergroup, 1500-
800 Ma; (2) the Windermere Supergroup, 800-600
Ma; (3) the Cambrian to Upper Devonian terrace
wedge, 600-370 Ma; and (4) the Upper Devonian to
Jurassic terrace wedge, 370-180 Ma. All four
stratigraphic packages are present in the Ogilvie
Mountains, where we can observe how one suc-
cession relates, stratigraphically and structurally,
to the other. We conclude that a zone of basement
weaknesses was present = 1200 Ma and that it
influenced the thickness, distribution, and facies of
strata from then until the Mesozoic. From this we
gain insight into the tectonic conditions that may
have influenced initiation of each stratigraphic
package.
Because the Cordilleran miogeocline contains
stratigraphic characteristics consistent with a pas-
sive margin depositional setting, it is tempting to
explain each of the four major depositional packages
as an outgrowth of rifting followed by continental
separation - a sequence of "rift-drift" episodes
(Thompson and others, 1987). For example, the Pur-
cell, Wernecke, Mackenzie Mountain, and Pingui-
cula successions are thick (:520 km), onlap the
continent, and are composed of fine-grained
terrigenous and carbonate rocks. Burke and Dewey
(1973), Sears and Price (1978), McMechan and Price
(1982), Gabrielse and Yorath (1989), among others,
have argued continental fragmentation and associa-
ted passive margin formation = 1500 Ma. An al-
ternative, proposed recently by Hoffman (1988), is
that the Belt-Purcell succession accumulated in an
episutural basin floored by oceanic crust, analogous
to the present-day Black Sea and south Caspian Sea.
His proposal is based on (1) Nd isotopic systematics
implying a source terrain that had Proterozoic
mantle-separation ages, and (2) regional variations
in grain size of the basin fill which suggest a
sediment source from the west rather than the east
or proximal south. Convergence with incomplete
closing of irregular shaped margins is offered as an
alternative to rifting.
Stewart (1972, 1976) and Eisbacher (1981,
1985) are among those who have proposed con-
tinental separation = 800 Ma, at the beginning of
Windermere sedimentation. The Windermere
Supergroup consists of a westward-thickening ter-
race wedge, :5 9 km thick, of turbiditic feldspathic-
sandstone, shale, and some carbonate that has a
strike length of 4000 km. The Windermere extends
from southern California to the west central Yukon;
its facies are so distinctive and consistent that they
are recognizable anywhere along this length. More
recently, Ross (1988) has argued that the vagaries of
preservation allow only the eastern side of a two-
sided basin to be observed; SW-NE paleocurrent
Long-lived basement weak zones and extensional magmatism, Ogilvie Mountains 361
trends and the tremendous thickness (:s; 7 km) of
turbiditic clastic rocks suggest to him that the
Windermere Supergroup accumulated in a trough
created during convergence and assembly of a Late
Proterozoic supercontinent.
Using results from back-stripping experiments
for part of the Cambrian to Upper Devonian terrace
wedge, Bond and Kominz (1984) argued that the
Cordilleran proto-Pacific margin was initiated in
latest Proterozoic time (600 Ma). The terrace wedge
of carbonate and mature clastic rocks is analogous
to passive margin successions accumulating now.
However, the rate of subsidence and sediment
accumulation did not decay uniformly from Cam-
brian to Devonian time, as one might predict from a
simple rift-drift model (MacKenzie, 1978); anoma-
lously thick Ordovician shales, for instance, have
been explained by Bond and others (1988) in terms
oflocal rift events superposed on the passive margin
setting. Christie-Blick and Levy (1988) pointed out
that stratigraphic evidence for latest Proterozoic to
Early Cambrian rifting is surprisingly limited when
compared with the magnitude of thermally driven
subsidence during the Palaeozoic.
The fourth major time-stratigraphic package,
the Upper Devonian to Jurassic terrace wedge,
contains many aspects of a passive margin succes-
sion in the Canadian Cordillera. In the north, fault-
bounded conglomerate and local accumulations of
alkalic volcanics occur near the base of the suc-
cession (Gordey, in press); it is overlain by a shelf-
offshelf assemblage typical of an Atlantic-type
margin. Yet 2000 km farther south, in Nevada,
strata of the same age are ascribed to a convergent
regime, the Antler Orogeny, in which sedimentation
was controlled by E-W compression, mountain
building, and foreland basin development.
Clearly, the 1400 million year stratigraphic
history of the Cordilleran miogeocline does not fit a
single-stage rift-drift depositional model.
In the Ogilvie Mountains, during Jurassic and
Cretaceous time, thin, far travelled sheets of Upper
Proterozoic and younger off-shelf strata were thrust,
from the south, onto the flank of what is now an
anticlinorium of Proterozoic and Palaeozoic shelf
strata (Figure 1) - we call the anticlinorium the
"Coal Creek Inlier" (Figure 2). The boundary
between thin, far travelled sheets, and the Coal
Creek Inlier is significant for several reasons: (1) it
parallels Middle and Late Proterozoic extension
faults and associated magmatic rocks; (2) it is the
locus of Palaeozoic facies transitions from shelf
(north) to off-shelf (south) lithotypes; (3) it is the
boundary between Late Jurassic-Early Cretaceous
thrust faulting and mid-Cretaceous and younger
thrusting; (4) it is the boundary between structural
styles; and (5) it may even have influenced mid-
Cretaceous magmatism. Basement was involved,
repeatedly, over time, influencing stratigraphic and
structural styles - hence the term "long-lived
basement weak zone."
The terms "shelf' and "off-shelf," as used above,
refer to different facies of the same time-
stratigraphic succession. The former is composed of
shallow-water shelf carbonate and terrigenous
clastic strata, and the latter represents deeper-
water shales and fine sandstones more typical of a
continental slope.
The stratigraphic nomenclature specific to the
Ogilvie Mountains is summarized in Figure 1. The
Middle Proterozoic is represented by the Wernecke
Supergroup (equivalent to the Purcell Supergroup)
and the Fifteenmile group - called Pinguicula
farther east (Eisbacher, 1981); the latter does not
have a stratigraphic correlative farther south where
there is a large gap in the stratigraphic record
between the Middle Proterozoic Purcell Supergroup
and the Late Proterozoic Windermere Supergroup.
The Upper Proterozoic is represented by the Hyland
and Harper Groups, both of which correlate with the
Windermere Supergroup. The lower Palaeozoic is
represented by shelf carbonate belonging to a map
unit called CDb and by off-shelf shales and fine
clastics belonging to the Road River Group. The
upper Palaeozoic-Mesozoic is represented by conglo-
merate and shale belonging to the Earn Group and
the Keno Hill quartzite, and to unnamed shale,
siltstone, and sandstone map units. The latter suc-
cession is not discussed further.
MIDDLE PROTEROZOIC EXTENSION
Basin Setting
The Wernecke Supergroup, deposited between
"'" 1600 Ma and 1300 Ma, consists of thick fluvial to
marine-turbiditic clastic rocks with intercalated
stromatolitic carbonate complexes (Delaney, 1981).
Locally, the succession has undergone metamor-
phism and penetrative deformation similar to the
Purcell Supergroup (McMechan, 1981), but we do
not know how regionally significant this event was.
Gabrielse (1972) termed it the Racklan Orogeny,
362 C. F. Roots and R. 1. Thompson
Figure 1. General stratigraphy of the Ogilvie
Mountains. The shelf succession underlies the
Mackenzie.Ogilvie Platform (early to middle
Palaeozoic), which covers northern Yukon and
much of the Western Northwest Territories. The
offshelf succession underlies the early Palaeozoic
Selwyn Basin. The Misty Creek Embayment
(MCE) of Selwyn Basin contains Ordovician alkalic
basalts similar to those in the southern Ogilvie
Mountains.
Figure 2. Structural elements in the
Ogilvie Mountains, Dawson map
area (modified from Anderson, 1987).
The off shelf succession is exposed in
the Chandindu, Antimony and Tomb
stone thrust sheets. The shelf succes
sion is exposed in Coal Creek Inlier, a
broad anticlinorium that has under
gone little internal deformation or
contraction. Open circles are drawn
on the downfaulted sides of Middle
Proterozoic ('" 1200 Ma) extension
faults; closed circles are drawn on the
downfaulted sides of Late Protero
zoic extension faults ('" 750 Ma);
teeth are drawn on the upthrust side
of Mesozoic thrust faults; Cretaceous
alkalic plutons shown with cross
symbol.
o 5
u
(5
N
o
'" ..J
It
151 Mo
26/18
(U'Pbl
u
(5
N
o
a:

o
a:
Q.

d,'5(:O{'It
Shelf
ROAD
RIVER Group
l-
Db doloml e
'.I delle
z"''''
:::lQ.
['000"
00:: e
i
c.
::>
o
"
W
...J

z:
w
w
t-
!:
1
.....
r
25 krn . OAW SON
Offshelf
OAD
RIVER
Group

HYLAND
Group
I"G,,' \"" ")
Cleat nol
1t'_I>OSt<l
Ootk ",,,elSlone.0'9,llIle
Sloel( thetl
mudstone
SOl'lIllSlon . ,.1 stonlt
Dolostone. Iome, lone
ConQ1omcrOle
R ),ot"e 110"' 1' onel Turl
Bosoille flow, OnCl-e>reCCICI
H .. mQhl, mOft,l bretCIO
ArQ,lIac,OU$ $lll$lone. 01'1)'1111 .
Long-lived basement weak zones and extensional magmatism, Ogilvie Mountains 363
but he has since discarded the term in a recent
synthesis of Cordilleran geology (Gabrielse and
y orath, 1989). Thickness and facies of the Wer-
necke Supergroup suggest deposition in a large,
intracratonic basin (Eisbacher, 1981).
At about 1200 Ma, the Wernecke Supergroup
underwent block faulting and intrusion. In the
Ogilvie Mountains, this was a time of growth fault-
ing and breccia intrusion; in the Mackenzie Moun-
tains, 400 km to the east, there is a spectacular
angular unconformity (Eisbacher, 1981); and on the
Canadian Shield, a major NW-trending basaltic
dike swarm (Mackenzie Dyke Swarm) accompanied
crustal extension.
Fifteenmile Half-Graben
In the Ogilvie Mountains, growth faults con-
trolled deposition of the Fifteenmile Group (1300-
1200 Ma; Figure 3a). The faults are steep, dip
northward, and are the boundaries between sedi-
mentary facies and abrupt changes in thickness of
Fifteenmile strata. Breccias, composed of altered
country rock, intrude the succession and are loca-
lized, in part, along the growth faults (Figure 4).
We suggest a genetic relationship between growth
faulting and breccia intrusion.
Two subparallel faults trend northeast across
the Coal Creek Inlier (Figure 3a). Adjacent to the
southernmost one, the downdropped Fifteenmile
half-graben is filled by >400 m of sandstone, shale,
and carbonate; on the upfaulted block is a discon-
tinuous succession, :s; 50 m thick, of shallow water
stromatolitic carbonate, maroon sandstone, and
siltstone; adjacent to the fault, blocks of the updip
stratigraphy, meters in diameter, form part of the
basin fill. Vertical displacement across the fault is
= 1 km. The northern fault is the locus of breccia
intrusion (Figure 4). If there was a half-graben of
Fifteenmile Group rocks, the evidence has since
been eroded. The presence of breccias suggests both
faults were active at or about the same time.
A second, NW-trending belt of Fifteenmile
Group strata (Figure 3a) parallels the western
flank of the Coal Creek Inlier. We infer that a NW-
trending growth fault exists to the southwest in the
subsurface.
The breccia bodies consist of sedimentary clasts
(mostly Quartet, but also Fairchild and Gillespie
Lake rocks of the Wernecke Supergroup; Figure 1),
chaotically mixed in a rock-flour matrix with abun-
dant secondary hematite and aligned along the
faults. It is very unlikely that the breccia repre-
sents paleo-talus or debris flows, because they are in
contact with every stratigraphic level of the Wer-
necke Supergroup and clearly intrude lower Fif-
teenmile strata. We consider them to be intrusive
pipes with accompanying spurs and sills along
favorable bedding planes. The breccias do not occur
in younger shelf units and were probably emplaced
during or immediately after deposition of the lower
Fifteenmile clastic sediments. The NOR breccia,
which is 200 km northeast of this area, has a similar
composition and structural relations and yielded a
U-Pb monazite age of 127040 Ma (Parrish and
Bell, 1987).
Breccia bodies typically exhibit cross-cutting
phases that can be distinguished by clast lithology
and degree of "milling." Breccia margins may be
sharply cross-cutting or gradational into intensely
fractured host strata (Figure 4). Wide zones of sodic
alteration and whitish, silicified host rock are
common, although none in the Ogilvie Mountains is
known to be mineralized (in U, Co, Cu) to the extent
of those 150 km to the east (Archer and Schmidt
1977; Laznicka and Edwards, 1979). Cross-cutting
relationships, localization along faults, and internal
fabrics suggest, to us, a hydraulic fracture mech-
anism of formation and emplacement in association
with gas streaming. An alternative hypothesis,
brecciation associated with upward mobilization of
evaporite layers, was proposed by Bell (1986) for
similar breccias in the Wernecke Mountains (300
km farther east). We emphasise that the Ogilvie
Mountains examples are distributed along growth
faults.
Extension and breccia intrusion associated with
Fifteenmile Group deposition does not account for
the development of cleavage in the underlying Wer-
necke rocks. We cannot rule out an older compres-
sional event (Mercier, 1989).
LATE PROTEROZOIC EXTENSION
Windermere Basin
In the Ogilvie Mountains, the Dawson Thrust
Fault (Figure 2) juxtaposes "off-shelf' Windermere
facies, called the Hyland Group, with "rift" facies,
informally named the Harper group (Figure 3b).
Correlation of the Harper group with the Hyland
Group and Windermere Supergroup is warranted by
7
5
0

M
a
:

R

H
.
,
p
.
r

,
r
o
i
l
l
i
p

U

d
'
l
I
I
e
"
r
.
"

I
'
o
e
t
.


H
.
,
,
,
.
,

I
J
r
o
o
l
o
t
p

I
!
.L
!
J

"
o
'
e
.
"
,
l
c
:

r
o
c
l
.

b
)

1
2
7
0

M
a
:


1
0

'

F
'
l
f
l

"
m
l
l
e

l
:
...:
:
.j

"
,
o
u
p


W
.
I
'
l
I
e
c
t

'
1
p
e


a
)

.
.

-


F
A
U
L

T

1

,
0

'
t
i
l

D
.
,
l
e
f
m
e
"


-
-
=
=

o

1
0

2
0

I
<
m

d
)

4
5
0

M
a
:


A
I
"
.
"
c

"
.

M
A
C
K
E
N
Z
I
E

,
.
L
A


,

.
.
.
.
.
.
.

.

c
l

9
0
-
1
1
0

M
a
:

1
f
,
.
S
,
.
_
"
"
.
,


P
l
u
t
o
n


:
.
.
.
;
:
t
T
o
m
O
.
r
O
l
l
e

P
l
u
l
o
n

A
n
t
l
m
o
n
,


P
l
u
t
o
(
l

F
i
g
u
r
e

3
.

D
i
s
t
r
i
b
u
t
i
o
n

o
f
e
x
t
e
n
s
i
o
n

f
a
u
l
t
s

a
n
d

a
s
s
o
c
i
a
t
e
d

g
r
a
b
e
n

d
e
p
o
s
i
t
s

i
n

a
n
d

a
r
o
u
n
d

t
h
e

C
o
a
l

C
r
e
e
k

I
n
l
i
e
r
.

a
)

D
i
s
t
r
i
b
u
t
i
o
n

o
f

1
2
0
0

M
a

F
i
f
t
e
e
n
m
i
l
e

G
r
o
u
p

s
h
o
w
n

i
n

s
t
i
p
p
l
e

p
a
t
t
e
r
n
;

d
i
s
t
r
i
b
u
t
i
o
n

o
f

W
e
r
n
e
c
k
e
-
t
y
p
e

b
r
e
c
c
i
a
s

s
h
o
w
n

i
n

t
r
i
a
n
g
l
e

p
a
t
t
e
r
n
;

t
r
a
c
e

o
f

f
a
u
l
t
s

t
h
a
t

c
o
n
t
r
o
l
l
e
d

b
r
e
c
c
i
a

e
m
p
l
a
c
e
m
e
n
t

a
n
d

s
e
d
i
m
e
n
t
a
t
i
o
n

s
h
o
w
n

w
i
t
h

d
a
s
h
e
s
;

g
e
o
l
o
g
y

a
n
d

c
r
o
s
s

s
e
c
t
i
o
n

i
n
t
e
r
p
r
e
t
a
t
i
o
n

o
f

a
r
e
a

o
u
t
l
i
n
e
d

b
y

r
e
c
t
a
n
g
l
e

s
h
o
w
n

i
n

F
i
g
u
r
e

4
.

b
)

D
i
s
t
r
i
b
u
t
i
o
n

o
f

7
5
0

M
a

M
o
u
n
t

H
a
r
p
e
r

G
r
o
u
p

(
W
i
n
d
e
r
m
e
r
e
)

s
e
d
i
m
e
n
t
a
r
y

(
s
t
i
p
p
l
e
)

a
n
d

v
o
l
c
a
n
i
c

(
d
a
s
h
e
s
)

r
o
c
k
s
;

t
r
a
c
e

o
f

s
y
n
s
e
d
i
m
e
n
t
a
r
y

e
x
t
e
n
s
i
o
n

f
a
u
l
t
s

s
h
o
w
n

w
i
t
h

h
e
a
v
y

d
a
s
h
e
s
.

c
)

D
i
s
t
r
i
b
u
t
i
o
n

o
f
4
5
0

M
a

v
o
l
c
a
n
i
c
s

(
l
i
n
e
s
)

a
d
j
a
c
e
n
t

t
o

t
h
e

s
h
e
l
f
-
o
f
f
s
h
e
l
f
t
r
a
n
s
i
t
i
o
n

(
d
a
s
h
e
d

l
i
n
e
)
.

d
)

D
i
s
t
r
i
b
u
t
i
o
n

o
f
m
i
d
-
C
r
e
t
a
c
e
o
u
s

a
l
k
a
l
i
c

p
l
u
t
o
n
s

(
c
r
o
s
s
e
s
)
.

'
"

m

*
'
"

Q


:
:
0

o

o

M
-
e
n

I
I
I

5
-
P
J

!
"
"
'
1
-
3

P
"

o

.
g

e
n

o

i:
1

Long-lived basement weak zones and extensional magmatism, Ogilvie Mountains 365
the age of volcanic rocks in the Mount Harper group,
similarity of trace fossils between the groups, and
similarities of tectonic and depositional history
(Thompson and others, 1987; Mustard and others,
1988a,b). A generalized comparison of strata from
in, and south of, the Coal Creek Inlier is presented
in stratigraphic columns (Figure 1).
The Harper group accumulated synchronously
with the formation of small half-grabens (Figure
3b). Unlike the Mackenzie Mountains where Eis-
bacher (1985) described an overall southwestward
thickening of similar facies across a hinge line
marked by extension faults, we have not observed a
regional source or thickening direction for the
Harper group. We speculate that the Harper group
accumulated in a relatively narrow (30 km?) zone of
incipient rifting and that it was separate from the
basin into which the Hyland Group was accumu-
lating farther south. We base this interpretation, in
part, on the stratigraphic character of the Hyland
Group: (1) the monotonous succession of turbiditic
sandstone and shale is quite distinctive and very
typical of Windermere-type rocks the length of the
Cordillera; (2) the Hyland Group coarsens south-
FIF TEENMILE GROUP
@ Mos'Slve doloston.e
Figure 4. Map and cross section
(A-B) illustrating structural and
stratigraphic relations associated
with Middle Proterozoic extension
and breccia emplacement.
Dolomil ie pOCkS. one
(' 's'
- If .. :I. Helerol i lhlC breCCIQ
(Wernec.ke mtlu'SlonsJ
o Shale, $llfsfOl'le,
conglomerale
o
WERNECKE SUPERGROUP
uo.rtel sandstone,

FOul ! (ball on
Ihrown Side)
km
ward into a gritty sandstone containing abundant
quartz and feldspar grains, evidence of source
proximity toward the south; and (3) paleocurrent
data, though meager, indicate paleoflow from a
southerly direction. It is unlikely that the quartz-
and feldspar-rich sands could have been transported
from the north or east, across the grabens control-
ling Mount Harper group sedimentation. Ross
(1988) has suggested the Windermere basin was
two-sided; we suggest that it was not only two-sided
but internally complex as well.
An important distinction between Windermere
strata in the Ogilvie Mountains and the Mackenzie
Mountains is the lack of glaciogenic deposits in the
Mount Harper group. In the Mackenzie Mountains,
the Rapitan Group contains abundant evidence of
proglacial and ice-marginal facies (Eisbacher, 1981,
1985; Yeo, 1981).
Harper Half-Graben
In the Coal Creek Inlier, one can map the fault
structures responsible for the initiation of sedimen-
366 C. F. Roots and R. I. Thompson
Figure 5. Diagrammatic cross section showing struc-
tural and stratigraphic relations associated with exten-
sion faulting during the Late Proterozoic. The Harper
Fault (wavy line) was active during deposition of the
Lower Harper Group; strata belonging to the Fifteen-
mile, Gillespie Lake, and Quartet groups were eroded
from the southern uplifted block and deposited as coarse
conglomerate in the adjacent half-graben.
tation and the changes in facies of sedimentary and
volcanic rocks that accumulated as faulting pro-
ceeded (Figures 3b and 5).
The lowest unit within the Harper group - the
lower Harper group - is a northward tapering
wedge of conglomerate that reaches thicknesses
> 900 m adjacent to the Harper Fault (Figure 5).
Basaltic flows overlie the conglomerate as well as
overstep the fault, indicating that faulting began
prior to volcanism; volcanic clasts within adjacent
conglomerate wedges demonstrate that volcanism
was also synchronous with faulting. South of the
Harper Fault, on the uplifted block, volcanic flows
disconformably overlie Wernecke carbonate (Gilles-
pie Lake Group), there are no Fifteenmile strata,
and there is no basal conglomerate belonging to the
Mount Harper group. Uplifted Fifteenmile Group
was the source for Mount Harper conglomerate.
Rarely is the interaction between fault displace-
ment, erosion, and sedimentation so clearly pre-
served.
The Mount Harper volcanic complex is the best
preserved and compositionally most diverse volcanic
center associated with the Windermere Supergroup.
It comprises two superimposed accumulations. Sub-
aqueous to subaerial basaltic flows and breccias (the
lower suite), :5 2400 m thick, are interpreted as a
seamount (105 km
2
). Following a period of erosion,
a much smaller volume of andesitic flood-flows and
rhyolite domes, the upper suite, was extruded from
four or more subaerial vents whose easterly trend
coincides with the Harper Fault - if it were pro-
jected beneath the volcanic complex. During and
following volcanism, the northern part of the
complex subsided so that a half-graben along the
reactivated Harper Fault was filled with volcanic
debris, siltstone, and limy mudstone here referred to
as the upper Harper group (Figure 5).
Whole-rock analysis of less altered flows and
intrusions from the Harper group reveal that the
upper suite is tholeiitic and the lower suite straddles
the calcalkalic-tholeiitic boundary (Figure 6).
First-erupted basalts were olivine-phyric (now
replaced) and compositionally homogeneous, with
Mg/Fe ratios of 45-55 and low incompatible element
contents (Table 1). The upper suite flows are
plagioclase-phyric and strongly iron-enriched.
Similar incompatible element patterns for the
rhyolite and andesite suggest that the upper suite
magma fractionated, and garnet xenocrysts in the
rhyolite suggest extrusion from middle crustal
depths (Pallister, 1987). In contrast, the lower suite
flows are relatively depleted in incompatible ele-
ments that may have extruded more directly from
upper mantle or lower crustal depths (Figure 6).
Volcanic occurrences associated with initial
Windermere sediments are scattered from east-
central Alaska to Utah. Most consist of altered
basalt with alkalic affinities (Harper and Link,
1985; Devlin and others, 1985, 1988). Their com-
positions and petrographic features are similar only
to the upper suite andesite flows of the Harper
complex. The relatively depleted lower suite and
the rhyolite are not matched elsewhere. These rock
types may be the result of an unusually large
magma pulse that could ascend through thick
continental crust rapidly enough, initially, to
extrude relatively primitive flows and to permit
later fractionation in a mid-crustal magma cham-
ber. Such a scenario might occur over a fault having
significant vertical displacement to penetrate the
lower crust.
Long-lived basement weak zones and extensional magmatism, Ogilvie Mountains 367
Member A
1000
" c:
100 :-
u
.
.
'" c
o
.,
OJ

u
o
c:
1000 Member B
100 ... -.. .. - . . ....... , .. , . ... -" . -. . .. .. . - ... ...... . . -
' .0
Vb 8a Rb it-. K Nb La Sr P Sm Zf TI Hf Y Yb
Member 0 ' 000
c)
__ __
So At> Th K Nb Lo S f P Sm Zr Ti HI Y Yb
:
u
:
...
c
.::
OJ
"
u
o
a:
Member E -
'0
d)
l ao Rb K NtJ. 1..0 Sf P Sm Zr T.
Y Yb
Figure 6. Chrondrite-normalized, (Yn=15) in compatible element abundance diagrams for some Mount Harper volcanic
rocks (from Roots, 1987). Upper Suite members (c and d) have similarly shaped diagrams and are more enriched than the
Lower Suite (a and b).
Zircons from a rhyolite sample high in the vol-
canic succession gave a U-Pb date of 751 + 26/ -18
Ma (Roots and Parrish, 1989). To our knowledge,
this is the most tightly constrained date on the start
of Windermere sedimentation. In the Mackenzie
Mountains, basaltic sills assumed to be coeval with
volcanics in the Windermere Group have been dated
at 770 Ma (Armstrong, and others, 1982); farther
south, a preliminary Sm-Nd age of 76244 Ma
(Devlin and others, 1988) is from volcanics near the
base of the Windermere. It appears that long seg-
ments of the ancestral North American margin
underwent extension at the beginning of Winder-
mere sedimentation.
368 C. F. Roots and R. I. Thompson
Table 1. Representative chemical analyses of volcanic rocks from the 780 Ma Mount Harper complex, and the
Cambro-Ordovician alkalic basalts in the Dawson map area.
Mt. Harper Complex Mt. Harper Complex Cambro-Ordovician
Lower Suite Upper Suite Basalt
Pillow Dyke Rhyolite Andesite
Si0
2
57.23 48.43 77.03 59.51 41.20 42.80
Al
2
0
3
15.40 16.46 16.85 13.25 13.60 11.90
FeO 8.90 10.34 3.37 11.22 12.00 13.00
MgO 7.19 10.83 0.60 3.57 7.49 10.90
CaO 4.74 11.50 1.13 7.51 14.60 11.40
Na20 4.24 1.16 1.23 0.96 1.07 1.13
K
2
0 4.24 0.31 4.14 2.06 0.42 1.69
Ti0
2
0.81 0.57 0.41 1.24 3.47 2.18
P
2
0
5
0.81 0.04 0.20 0.30 1.36 0.57
MnO 0.14 0.19 0.04 0.14 0.19 0.18
Cr203
0.03 0.04 0.00 0.02 0.01 0.05
Ba 100 90 1010 1970 260 570
Nb 5 2 12 17 140 60
Rb 0 7 140 347 90 40
Sr 364 63 79 182 680 210
Y 20 24 31 60 30 20
Zr 73 28 222 296 370 190
Analyses by XRF at the Geological Survey of Canada and X-Ray Assay Laboratories, Toronto. Precision is
0.10% major oxides, 10 ppm for trace elements. Full discussion in Roots (1987,1988).
EARLY PALEOZOIC EXTENSION
Basin Setting
In the Ogilvie Mountains, Windennere Group
sedimentation continued into the early Palaeozoic.
The margin appears to have taken on aspects of an
Atlantic-type passive margin, with development of a
hinge line separating shelf carbonate facies (basal
part of CBb; Figure 1) from deeper water off-shelf
shales and fine sandstones. Windennere sediment
must have continued to come from a southerly
direction, because a vast carbonate shelf extended to
the north and east. The abundance of quartz, which
was not derived from the carbonate shelf, suggests
crystalline basement was exposed farther to the
south. Detrital zircon from Windermere sediments
in the southern Yukon indicate erosion of Early
Proterozoic crystalline basement (Erdmer and
Baadsgaard, 1987) and Late Proterozoic granite
(Even chick and others, 1984). If continental frag-
mentation in the latest Proterozoic or earliest
Palaeozoic led to a proto-Pacific passive margin
(Bond and Konimz, 1984), there is little in the way
of stratigraphic, structural, or magmatic evidence in
the Ogilvie Mountains to support this idea.
Long-lived basement weak zones and extensional magmatism, Ogilvie Mountains 369
Volcanism and the Shelf-Offshelf Hinge
Windermere sedimentation ended in the Middle
Cambrian with the onset of volcanism. Belts of
basalt comprising pillows, related breccias, and
subaqueous tuff typically separate the older Hyland
Group from the younger Road River Group (Roots,
1989; Figures 1 and 3c). The volcanics occur with-
in the interval Middle Cambrian to Middle Ordo-
vician; intercalations with fossil-bearing Hyland
Group shales at the base of the succession support a
Middle Cambrian age, and the volcanics are over-
lain with apparent conformity by Middle Ordovician
chert and shale ofthe Road River Group. The belt of
lower Palaeozoic volcanics extends east into the
Mackenzie Mountains where they are Middle Ordo-
vician in age (Cecile, 1982).
The hinge separating shelf carbonate from off-
shelf shale marks the boundary between the Selwyn
Basin and the Mackenzie platform (Figure 1). In
the Ogilvie Mountains, the hinge parallels the
southern flank of the Coal Creek Inlier, close to the
older growth faults of Middle and Late Proterozoic
age. The Cambro-Ordovician volcanics accumu-
lated along the Selwyn Basin side of the hinge line.
They consist of deep-water volcanic centers. Gab-
broic dikes that fed the centers form E-trending
swarms in the Hyland Group. The abundance of
subparallel dikes and the alkalic nature of the
volcanics (Table 1; Roots, 1988) are consistent with
an extensional tectonic setting.
MESOZOIC THRUST FAULTING AND
SYNOROGENIC ALKALIC MAGMATISM
The southern flank of the Coal Creek Inlier
continued to be an important boundary during
Jurassic and Cretaceous folding and thrust faulting.
South of the inlier, thin, complexly deformed and far
travelled thrust sheets comprise the deformed
wedge of supracrustal rocks (McMechan and
Thompson, in press). There are three major thrust
sheets: Chandindu, Tombstone, and Antimony
(Figure 2). The inlier, on the other hand, is a broad
anticlinorium, cut by a few thrusts around its
margin. This change in structural style reflects an
abrupt change in the depth to detachment; strata
comprising the Coal Creek Inlier are detached at
depth and have moved as a thick coherent mass,
possibly over basement ramps(s) formed by earlier
extension. This style contrasts with off-shelf strata
south of the inlier, which are detached at shallow
levels in the Keno Hill Quartzite, the Road River
Formation, and the Hyland Group (Figures 1 and
2; McMechan and Thompson, in press). Total short-
ening south of the Coal Creek Inlier to the Tintina
Fault may exceed 100 km; shortening across the
Coal Creek Inlier is < 20 km.
In fold and thrust belts, structural style is
sensitive to composition of the stratigraphic succes-
sion. Thick competent successions will form thick,
broadly folded thrust sheets; less competent multi-
layers tend to form chervon folds and become
detached at two or more levels simultaneously
(Thompson, 1981, 1988). In the Ogilvie Mountains,
stratigraphic facies reflect the presence of reacti-
vated basement features; therefore two structural
heterogenieties are built into the region: (1) the
basement features that influenced the geometry of
facies transitions, and (2) the facies transition itself.
This does not necessarily mean, however, that the
basement was active during deformation.
A-type syenite stocks of 90-110 Ma (Figures 2
and 3d; Anderson, 1988) cross-cut and therefore
post-date movement on the Antimony, Tombstone,
and Chandindu thrust sheets but pre-date open folds
in Upper Cretaceous strata on the northern flank of
the Coal Creek Inlier. Anderson (1988) suggests
that plutonism occurred between periods of com-
pression during a brief interval of crustal relaxa-
tion. This belt of alkalic plutons continues to the
southeast, parallel to the Tintina Fault.
CONCLUSIONS
Extension faulting and magmatism record
crustal stretching in the Ogilvie Mountains. The
succession of igneous events - (1) mantle degassing
(Middle Proterozoic Wernecke breccias), (2) tholeii-
tic volcanism (Upper Proterozoic Harper volcanics),
(3) alkalic volcanism (lower Palaeozoic shelf to off-
shelf transition, and (4) intrusion of undersaturated
plutons (mid-Cretaceous "intraorogenic" plutonism)
- is compatible with a long- lived, episodic, ext-
ensional history. The zone of active stretching was
relatively narrow, "'30 km, along the southern
flank of the Coal Creek Inlier. Once basement
weaknesses had been established at '" 1200 Ma,
they remained susceptible to reactivation. The
thickness and distribution of sedimentary facies was
influenced by each extensional event; this, in turn,
370
C. F. Roots and R. I. Thompson
influenced the style of structures developed during
Mesozoic thrust faulting and folding.
The Ogilvie Mountains part of the Cordilleran
miogeocline was not a simple prograded terrace
wedge; there is ample stratigraphic evidence from
the Late Proterozoic until the Jurassic that land
masses and shallower seas existed outboard (south)
of the Ogilvies. The margin was complex and
probably consisted of more than one linear basin or
trough.
The character of magmatic products through
time is consistent with continental crust as base-
ment. There is no particular reason to propose that
the Wernecke Supergroup was deposited into an
episutural basin floored by oceanic crust (Hoffman,
1988), but there is no proof this was not the case.
Stratigraphic and structural evidence supports ex-
tension from 1200 Ma on. There is no compelling
reason to suppose the Windermere Supergroup was
deposited into a foreland basin (Ross, 1988), but
there is little doubt the basin was two-sided. If
continental fragmentation did not occur until the
latest Proterozoic (Bond and Kominz, 1987), there is
little evidence in the stratigraphic, structural, or
magmatic record to support the rift phase of such a
break-up.
REFERENCES
ABBO'IT, J. G., 1982, Structure and stratigraphy of the
MacMillan fold belt: Evidence for devonian faulting; pp.
22-33 in Yukon: Exploration and Geology, 1981: Cana-
dian Department of Indian and Northern Affairs, White-
horse, Yukon, Canada, Open File Report (May 1982).
ANDERSON, R. G., 1987, Plutonic rocks in the Dawson
map-area, Yukon Territory; pp. 689-697 in Current Re-
search, Part A: Geological Suruey of Canada, Paper 87-
lA, 946p.
ARCHER, A. R., and U. SCHMIDT, 1977, Mineralized
breccias of Early Proterozoic age, Bonnet Plume River
District, Yukon Territory: Canadian Institute of Mining
and Metallurgy Bulletin, v. 71, no. 796, p. 53-58.
ARMSTRONG, R. L., G. H. EISBACHER, and P. D. EVANS,
1982, Age and stratigraphic-tectonic significance of Pro-
terozoic diabase sheets, Mackenzie Mountains, north-
western Canda: Canadian Journal of Earth Sciences, v. 19,
p.316-323.
BELL, R. T., 1986, Megabreccias in the northeastern
Wernecke Mountains, Yukon Territory; pp. 375-384 in
Current Research, Part A: Geological Survey of Canada,
Paper 86-1A.
BOND, C., and A. KOMINZ, 1984, Construction of tectonic
subsudence curves for the early Palaeozoic miogeocline,
southern Canadian Rocky Mountains: Implications for
subsidence mechanisms, age of breakup and crustal thin-
ning: Geological Society of America Bulletin, v. 85, p. 155-
173.
M. A. KOMINZ, and W. J. DEVLIN, 1988, Com-
plexities in the evolution of the early Palaeozoic passive
margin of the southern Canadian Rockies [abstract):
Geological Society of America Abstracts with Programs, v.
20, no. 7, p. A197.
BURCHFIEL, B. C., and A. DAVIS, 1975, Nature and
controls of Cordilleran orogenesis, western United States:
Extensions of an earlier synthesis: American Journal of
Science, v. 275-A, p. 363-396.
BURKE, K., and J. F. DEWEY, 1973, Plume-generated
triple junctions: Key indicators in applying plate tectonics
to older rocks: Journal of Geology, v. 81, p. 406-433.
CECILE, M. P., 1982, The Lower Palaeozoic Misty Creek
Embayment, Selwyn Basin, Yukon and Northwest
Territories: Geological Survey of Canada Bulletin, no.
335, 78p.
CHRISTIE-BLICK, N., and M. LEVY, 1988, Late Proterozoic
and Early Cambrian evolution of the western United
States [abstract]: Geological Society of America Abstracts
with Programs, v. 20, no. 7, p. A298.
DELANEY, G. D., 1981, The mid-Proterozoic Wernecke
Supergroup, Wernecke Mountains, Yukon Territory; pp. 1-
24 in F. H. A. Campbell (ed.), Proterozoic Basins of
Canada: Geological Survey of Canada, Paper 81-10, 444 p.
__ , H. K. BRUECKNER, and G. C. BOND, 1988, New
isotopic data and a preliminary age for volcanics near the
base of the Windermere Supergroup, northeast Washing-
ton, U.S.A.: Canadian Journal of Earth Sciences, v. 25, p.
1906-1911.
EISBACHER, G. H., 1981, Sedimentology, Tectonics and
Glacial Record in the Windermere Supergroup, Mac-
kenzie Mountains, Northwestern Canada: Geological
Suruey of Canada, Paper 80-27, 40 p.
__ , 1985, Late Proterozoic rifting, glacial sedimenta-
tion, and sedimentasry cycles in the light of Windermere
deposition, western Canada. Palaeogeography, Palaeocli-
matology, Palaeoecology, v. 51, p. 231-254.
ERDMER, P., and H. BAADSGAARD, 1987, 2.2 Ga age of
zircons in three occurrences of Upper Proterozoic clastic
rocks of the northern Cassiar terrane, Yu8kon and British
Columbia: Canadian Journal of Earth Sciences, v. 24, p.
1919-1924.
EVANCHICK, C. A., R. R. PARRISH, and H. GABRIELSE,
1984, Precambrian gneiss and Late Proterozoic sedimenta-
Long-lived basement weak zones and extensional magmatism, Ogilvie Mountains 371
tion in north-central British Columbia: Geology, v. 12, p.
233-237.
GABRIELSE, H., 1972, Younger Precambrian of the Cana-
dian Cordillera: American Journal of Science, v. 272, p.
521-536.
__ , and C. J. YORATH, 1989, DNAG #4: The Cordi-
lleran orogen in Canada: Geoscience Canada, v. 16, no. 2,
p.67-83.
CORDEY, S. P., in press, Nahanni Map-Area, Yukon and
Northwest Territories: Geological Survey of Canada,
Memoir.
GREEN, L. H., 1972, Geology of Nash Creek, Larsen
Creek and Dawson Map-Areas, Yukon Territory: Geo-
logical Survey of Canada, Memoir 364,157 p.
HARPER, G. D., and P. H. LINK, 1986, Geochemistry of
Upper Proterozoic rift-related volcanics, northern Utah
and southeastern Idaho: Geology, v. 14, p. 864-867.
HOFFMAN, P. F., 1983, Three stage subsidence of the 1.9
Ga Coronation Margin [abstract]. Geological Association
of Canada Abstracts with Programs, v. 8, p. A33.
__ , 1988, Belt basin: A landlocked remnant oceanic
basin? (analogous to the south Caspian and Black Seas)
[abstract]: Geological Society of America Abstracts with
Programs, v. 20, no. 7, p. A50.
LAZNICKA, P., and R. J. EDWARDS, 1979, Dolores Creek,
Yukon - A disseminated copper mineralization in sodic
metasomatites: Economic Geology, v. 74, p. 1352-1370.
McKENZIE, D., 1978, Some remarks on the development of
sedimentary basins: Earth and Planetary Science Letters,
v. 40, p. 25-32.
McMECHAN, M. E., 1981, The Middle Proterozoic Purcell
Supergroup in the southwestern Rocky and southeastern
Purcell Mountains, British Colombia, and the initiation of
the Cordilleran miogeocline, southern Canada and adja-
cent United States: Bulletin of Candian Petroleum Geo-
logy, v. 29, p. 583-621.
__ , and R. A. PRICE, 1982, Superimposed low-grade
metamorphism in the Mount Fisher area, southeastern
British Columbia - Implications for the East Kootenay
Orogeny: Canadian Journal of Earth Sciences, v. 19, p.
476-489.
__ , and R. I. THOMPSON, The Candian Cordilleran
foreland belt south of 66N and its influence on the
Western Interior Basin; in W. G. E. Caldwell and E. G.
Kauffman (eds.), The Western Interior Basin: Geological
Association of Canada, Special Paper, in press.
MERCIER, E., 1989, Evenments tectoniques d'origine
compressive dans Ie Proterozoique du nord de la Cordillere
Candienne (montagnes Ogilvie, Yukon): Canadian Jour-
nal of Earth Sciences, v. 26, p. 199-205.
MONGER, J. W. H., R. A. PRICE, and D. J. TEMPELMAN-
KLUIT, 1982, Tectonic accretion and the origin of the two
major metamorphic and plutonic welts in the Canadian
Cordillera: Geology, v.10, p. 70-75.
MUSTARD, P. S., J. A. DONALDSON, and R. I. THOMPSON,
1988a, Trace fossils and stratigraphy of the Precambrian-
Cambrian boundary sequence, upper Harper Group, Ogil-
vie Mountains, Yukon; pp. 197-203 in Current Research,
Part E: Geological Survey of Canada, Paper 88-lE.
__ , J. A. DONALDSON, and R. I. THOMPSON, 1988b,
Precambrian-Cambrian trace fossils of the Ogilvie Moun-
tains, Yukon: Significance to regional correlations [ab-
stract]: Geological Association of Canada Abstracts wtih
Programs, v. 13, p. A90.
PALLISTER, J. S., 1987, Magmatic history of Red Sea rift-
ing: Perspective from the central Saudi Arabian coastal
plain: Geological Society of America Bulletin, v. 98, p. 400-
417.
PARRISH, R. R., and R. T. BELL, 1987, Age of the NOR
breccia pipe, Wernecke Supergroup, Yukon Territory:
Geological Survey of Canada, Paper 87-2, p. 39-43.
ROOTS, C. F., 1987, Regional Tectonic Setting and Evo-
lution of the Late Proterozoic Mount Harper Volcanic
Complex, Ogilvie Mountains, Yukon: PhD thesis,
Carleton University, Ottawa, Ontario, Canada, 219 p.
__ , 1988, Cambro-Ordovician volcanic rocks of the
eastern Dawson map-area, Ogilvie Mountains, Yukon; pp.
81-87 in Yukon Geology, Volume 2: Canadian Depart-
ment of Indian Affairs and Northern Development, Ex-
ploration and Geological Services Division, Whitehourse,
Yukon, Canada.
ROSS, G. M., 1988, The Windermere foredeep: Evidence
for Late Proterozoic plate convergenece in the western
Canadian Cordillera [abstract]: Geological Society of
America Abstracts with Programs, v. 20, no. 7, p. A297.
SEARS, J. W., and R. A. PRICE, 1978, The Siberian con-
nection: A case for Precambrian separation of the North
American and Siberian cratons: Geology, v. 6, p. 267-270.
STEWART, J. H., 1972, Initial deposits in the Cordillera
geosyncline: Evidence of a Late Precambrian 850 m.y.J
continental separation: Geological Society of America
Bulletin, v. 83, p. 1345-1360.
___ , 1976, Late Precambrian evolution of North Ameri-
ca: Plate tectonic implication: Geology, v. 4, p. 11-15.
STRUIK, L. C., 1987, The Ancient Western North Am-
erican Margin: An Alpine Rift Model for the East-
Central Canadian Cordillera: Geological Survey of
Canada, Paper 87-15,19 p.
372 c. F. Roots and R. 1. Thompson
TEMPELMAN-KLUIT, D. J., 1979, Transported Cata-
ciastie, Ophiolite and Granodiorite in Yukon: Evi-
dence of Arc-Continent Collison: Geological Survey of
Canada, Paper 79-14, 27 p.
THOMPSON, R. 1., 1981, The nature and significance of
large "blind" thrusts within the northern Rocky Moun-
tains of Canada; pp. 449-462 in K. R. McClay and N. J.
Price (eds.), Thrust and Nappe Tectonics: Geological
Society of London, Special Publication 9, 539 p.
__ , 1988, Stratigraphy, Tectonic Evolution and
Structural Analysis of the Halfway River Map Area
(94 B), Northern Rocky Mountains, British Columbia:
Geological Survey of Canada, Memoir 425, 119 p.
__ , and C. F. ROOTS, 1982, Ogilvie Mountains Project,
Yukon, Part A: A new regional mapping program; pp. 403-
414 in Current Research, Part A: Geological Survey of
Canada, PaperB1-1A.
__ , and G. H. EISBACHER, 1984, Late Proterozoic rift
assemblages, northern Canadian Cordillera [abstract]:
Geological Society of America Abstracts with Programs, v.
16, no. 5, p. 336.
__ , E. MERCIER, and C. ROOTS, 1987, Extension and its
influence on Canadian Cordilleran passive margin evolu-
tion; pp. 409-418 in M. P. Coward, J. F. Dewey, and P. L.
Hancock (eds.), Continental Extensional Tectonics:
Geological Society of London, Special Publication 28, 650 p.
THOMPSON, R. N., M. A. MORRISON, A. P. DICKIN, and G.
L. HENDRY, 1983, Continental flood basalts ... Arachnids
rule OK?; pp. 158-185 in C. J. Hawkesworth and M. J.
Norry (eds.), Continental Basalts and Mantle Xenoliths:
Shiva Publishing Ltd., Nantwich, Cheshire, UK, 272 p.
YEO, G. M., 1981, The Late Proterozoic Rapitan glaciation
in the northern Cordillera; pp. 25-46 in F. H. A. Campbell
(ed.), Proterozoic Basins of Canada: Geological Survey
of Canada, Paper 81-10, 444 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Geotectonics of the cratonic margin from paleomagnetism of the
Middle Proterozoic Aldridge (Prichard) Formation and
Moyie sills of British Columbia and Montana
DAVID T. A. SYMONS and E. ALLAN TIMMINS
Department of Geology, University of Windsor, Windsor, Ontario, N9B 3P4, Canada
(received August 12, 1988; revision accepted January 26, 1989)
ABSTRACT
The Precambrian Aldridge (Prichard) Formation crops out at or near the base of the Middle Proterozoic
Belt-Purcell Supergroup in the Purcell anticlinorium, and the Moyie sills intrude the Lower and Middle
Aldridge Formation. Paleomagnetic analysis was done on specimens from 27 sites in layered argillites of the
Aldridge and from 30 sites in the Moyie metagabbros using AF and thermal (Th) demagnetization methods.
Fold and contact tests show that no primary remanence components were isolated. However, three post-
folding metamorphic components were found: A was found in 9 argillite (AF, 22.9, 64.1, u95 = 9.8) and 11
sill sites (AF, 35.2, 62.0, U95 = 7.1); B was found in 8 argillite (AF, 320.4, 52.0, U95 = 11.8) and 10 sill
sites (Th, 303.6, 56.9, U95 = 7.8); and C was found in 19 argillite sites (AF, 209.0, 3.1, U95 = 6.0). All three
components include normal and reversed sites. The A, B, and C poles are discordant with the positions
expected for known metamorphic or intrusive events, but correction for a simple 37 clockwise rotation leads to
concordancy of the A pole with the"" 175 Ma position, B with the"" 810 Ma position, and C with the"" 1400 Ma
position on the APWP for the North American craton. These ages coincide with the known Laramide, Goat
River, and East Kootenay orogenies. Evidently, the terrane housing the Purcell Anticlinorium: (1) has
always been part of the North American craton with little or no translation; (2) was folded and
metamorphosed in the East Kootenay orogeny prior to 1400 Ma; (3) was thermally overprinted in the Goat
River orogeny; and (4) was thermally overprinted and then rotated"" 37 12 clockwise as Terrane I of the
Cordillera impacted from the southwest against the craton during the mid-Mesozoic Laramide orogeny.
INTRODUCTION
The Middle Proterozoic Belt-Purcell Super-
group is exposed in southeastern British Columbia
and in contiguous parts of Alberta, Montana, Idaho,
and Washington (Figure 1). The Aldridge Forma-
tion in Canada and the equivalent Prichard Forma-
tion in the United States are stratigraphically near
the base of the supergroup. The Moyie sills intrude
the lower part of the Aldridge (Prichard) Formation.
The study area is in the Purcell anticlinorium where
paleomagnetic fold tests can indicate if a given
remanence component was formed before or after
folding. The object of the study was to determine if
this Precambrian terrane was rotated or translated
relative to the North American craton, as was found
in Mesozoic rocks of Terrane I of the Cordillera to
the immediate west (Symons and Litallien, 1984).
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 373-384. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
373
374
D. T. A. Symons and E. A. Timmins
Q
o Km 100
L I
Wash I

) Idaho
study
area
Figure 1. Location of the study area in the Belt-Purcell
Supergroup (B, P) outcrop area. The arrows denote the
structural trends. The boundaries of the Rocky Mountain
Belt (R) and the Quesnellia (Q) and Eastern (E) sub-
terranes of Terrane I are also shown.
Geology
The Belt-Purcell Supergroup rests unconform-
ably along its eastern margin on a plutonic-
metamorphic basement that yields radiometric ages
of ""1700100 Ma (Burwash and others, 1962;
Giletti, 1968) and probably of "" 1540 Ma (Reid and
others, 1973; Armstrong, 1975; Evans, 1986). This
sets a maximum age for the supergroup. In Canada
the Aldridge is comprised of "" 5000 m of wackes and
argillites in a flysch sequence that overlies "" 2000
m of argillites and quartzites of the Fort Steele
Formation at the base of the Purcell Supergroup
(Hamilton and others, 1982). The stratigraphic
equivalent of the Aldridge in Montana is the 7500-
m-thick Prichard Formation at the base of the Belt
Supergroup (Cressman, 1984). The Aldridge (Prich-
ard) is overlain in turn by several kilometers of
wackes, argillites, and quartzites with minor coarse
clastic, carbonate, and basalt units in the super-
group, and above these by younger Proterozoic and
Paleozoic formations.
The Moyie sills are metagabbroic in composi-
tion and up to 350 m thick. They are found in only
the Aldridge (Prichard) Formation. They show
minor crosscutting relationships in the Lower Ald-
ridge division but are essentially stratabound in the
Middle Aldridge division. One sill, the Crossport C
sill near Eastport, Idaho, gives a U-Pb zircon age of
143310 Ma which dates intrusion, as well as Rb-
Sr and K-Ar ages of 1285 165 and 850 45 Ma,
respectively, which reflect subsequent metamor-
phism (Zartmann and others, 1982). Altered wall
rock from the Sullivan Pb-Zn-Ag deposit in the
Lower Aldridge near Kimberly, B.C., gives a K-Ar
biotite age of 1436 Ma (LeCouteur, 1979). Thus,
1435 Ma is likely both the actual age of the sills and
the minimum age for the Aldridge Formation. This
result is consistent with the U-Pb zircon age of
1370 110 Ma from porphyritic granite plutons in
the Salmon-Leesburg area of Idaho and the Rb-Sr
isochron age of 1305 152 Ma found for the
Hellroaring Creek granodiorite stock, because these
felsic plutons cut Prichard sediments and Moyie
sills and post-date much of the deformation and
regional metamorphism in the Purcell anti-
clinorium (Ryan and Blenkinsop, 1971, recalculated
using A = 1.42X10-
11
year-\ Evans, 1986). From
the metamorphic mineralogy and isotope data,
McMechan and Price (1982) concluded that regional
metamorphism reached lower middle greenschist
facies temperatures of about 375C in the Belt-
Purcell rocks of the study area. They also identified
three metamorphic episodes: (1) the East Kootenay
orogeny from 1350 to 1300 Ma, which produced
most of the folding; (2) the Goat River orogeny from
900 to 800 Ma; and (3) the Laramide orogeny from
Jurassic to early Tertiary time. The Laramide oro-
geny signals the accretion of allochthonous Terrane
I onto the craton in the Early Jurassic (Monger and
others, 1982; Brown and others, 1986). Mesozoic
rocks in the Quesnellia subterrane of Terrane I to
the immediate west of the study area were rotated
clockwise "" 43 and translated northward "" 13
relative to the craton (Symons and Litalien, 1984).
Cressman (1984) reported that sole marks on
turbidites in the Aldridge (Prichard) Formation to
the northwest have been rotated "" 45 clockwise on
Laramide thrust faults relative to transport direc-
tions found to the southeast. Archibald and others
(1983, 1984) and Price (1986) consider that rocks of
the Purcell anticlinorium were deeply buried in
Jurassic time, were rapidly uplifted in Early Creta-
ceous time and overthrust into a supracrustal
position, and were then intruded by Late Cretaceous
to early Tertiary granitic intrusions. Lithoprobe
seismic reflection data indicate that a series of W-
dipping thrust faults is present in the supracrustal
Geotectonics of the cratonic margin, British Columbia and Montana 375
Proterozoic rocks of the anticlinorium, which sit on
an autochthonous basement at a depth of 9-12 km,
and that a major E-dipping fault zone appears to
truncate basement on the west side of the anti-
clinorium where Terrane I abuts (Cook and others,
1987).
Sampling
The 27 Aldridge Formation sites were collected
from layered pelagic interturbidite zones or argil-
lites (Figure 2A). The 30 Moyie sill sites include
two baked contact sites (sites 503 and 542) (Figure
2B). All sites had well-defined attitudes based on
sediment bedding or an exposed contact. In general,
most of the sites have intermediate easterly or
westerly dips (Figures 3A and 4A). Four to six 2.5-
cm-diameter drill cores were collected from each site
after orienting them in situ with a sun compass.
Two or more specimens were sliced from each core
for study.
MAGNETIZATION
Natural Remanence
The natural remanent magnetization (NRM) of
each specimen was measured on a Schonstedt SSM-
1A magnetometer. The sediments and sills have
weak median NRM intensities of ::.8XlO-
3
Am-
1
and Am-I, respectively. Both have a
normal-to-reverse polarity ratio (N:R) of about 2:1.
After excluding a few inhomogeneous NRM direc-
tions from specimens with a within-specimen angu-
lar standard deviation of >35
0
(Harrison, 1980), the
remaining vectors were plotted on an equal-area
stereonet, smoothed with a 1% areal window, and
contoured in percentage density (Kamb, 1959;
Stupavsky and Symons, 1982). Vectors in both sedi-
ments and sills form highly significant
antiparallel normal (N) and reverse (R) populations
which, when combined, yield a single directional
anomaly for each (Figures 3B and 4B). Their unit-
directions, calculated by averaging the specimen
directions to get the site mean-directions and in
turn averaging these to get the unit means (Fisher,
1953), are essentially identical (Table 1). Because
the cones of 95% confidence about the NRM unit
means encompass the present Earth's magnetic field
(PEMF) direction, it is likely that the NRM induces
530
/ 509
KIMBERLEY ... /508
531/1512
.... 511 514
.546
541

CRANBROOK
502-.1
534
501 533
547"548
,551 BRITISH
z 549 COLUMBIA
---T-------
IDAHO i MONTANA
I
CROSSPORT I .41
I .42
BONNERS I 26
FERRY
A
I "1'-48
I 47
I
1I6W

CRANBROOK
o Km 20
503
5361532
538!e"535
539 .540
5i
o
BRITISH
COLUMBIA
(\! ----r-------
IDAHO 1 MONTANA
3r]&)9
1
BONNERS 14649
FERRY _, 50 25

144
1 33
116W
B
Figure 2. Location of the sampling sites in: A) Aldridge
(Purcell) sediment; B) Moyie sills.
a substantial modern viscous remanent magnetiza-
tion (VRM) component. However, the presence of
both Nand R populations also suggests a more
stable underlying component for both sediments and
sills.
Sill Demagnetization
Initially one specimen per site was AF demag-
nitized in 15 steps up to 100 mT using a Schonstedt
GSD-IA AF demagnetizer. When the remanence
directions for the sill specimens are grouped into 0-
15, 20-50, and 60-100 mT populations without tec-
tonic correction for dip (Figure 4C,D,E), the ano-
maly peak direction shallows slightly as VRM is
removed at low mT fields. The peak density in-
creases from 9% for the NRM to 13% as VRM is first
preferentially removed and then decreases progres-
sively to 10% and then 8% as the A component is
isolated and itself demagnetized. Also, the N:R
ratio for the directions decreases from 2:1 to 1:1,
indicating that the A component was acquired over
a period of time during which the Earth's field un-
derwent one or more polarity reversals. Tectonic
correction (e.g., Figure 4F for the 60-100 mT popu-
376
D. T. A. Symons and E. A. Timmins
Figure 3. Lower hemispheres of equal-area stereo-
nets showing for the Aldridge (Prichard) sedi-
ments.
A) Poles to bedding planes for each site.
B-E) Vector percentage density stereoplots for N
specimens with a set at the 95% confidence
level and without tectonic correction:
B) NRM;
C) 0-15 mT AF cleaning;
D) 20-50 mT AF cleaning;
E) 60-100 mT AF cleaning.
F) C component remagnetization circles
without tectonic correction.
The triangle in B denotes the Earth's present
magnetic field direction.
lation) causes the A component anomaly to split into
two anomalies of much lower significance that
reflect the intermediate easterly (Ae) and westerly
(Aw) dips found in the Purcell anticlinorium. This
shows that the A component is postfolding in origin
and, therefore, is likely a metamorphic overprint.
Figure 4C,D,E also shows a SW-extending lobe off
the main A-component anomaly. It suggests the
presence of a second less-common but stable com-
ponent (B). Where present, the A and B components
are readily identified using Zijderveld plots, in both
Nand R directions (Figures 5 and 6, respectively).
Initial removal of VRM with fairly rapid initial
intensity decay is a common feature (Figure 5,
#536; Figure 6, #503, #513). A few specimens
show very stable directions and slow intensity decay
(Figure 5, #049; Figure 6, #043). For most speci-
mens, the coherent movement is away from the A
direction above 50 mT; however, no end points were
270
270
270
0
+
() + C?
0
00
+
00
A AF20-5OmT
0
0
-
0
.+ -+
+ ++
()
+ + 90
0
00
0
0
00
+
+
+
+
+

+ + ++ + + + + 90
ALDRIDGE ALDRIDGE
NRM
+ +
AF !SO-tOO mT
N 215

F
90
Ocr
ALDRIDGE
180 180
reached and the trend of movement was inconsistent
so that the remagnetization circle method of Halls
(1976) was unable to isolate an underlying compon-
ent (Figure 5, #532, #550, #035). AF demagne-
tization of all the remaining sill specimens in the
2010 mT range duplicated the anomaly pattern
found for the 20-50 mT range (Figure 4D) but with
a slightly lower peak density - about 8%.
Two specimens per site were thermally demag-
netized using a Schonstedt TSD-1 demagnetizer in
11 steps up to 650C. A few specimens gave co-
herent demagnetization tracks up to about 525C
(Figure 7, #550), but most have rapid intensity
drops to much lower blocking temperatures followed
by erratic unstable directions (Figure 7, #536,
#541). From Pullaiah and others (1975), it is evi-
dent that virtually none of the sill remanence is
likely to predate a ""375C regional metamorphic
event.
Geotectonics of the cratonic margin, British Columbia and Montana 377
0
0
Figure 4. Lower hemisphere equal-area stereonets
A 0
showing for the Moyie sills.
+

A) Poles to contact planes for each site.

+

.. --",- . B-F) Vector percentage density stereoplots for:
.oo
10
270
..: ,.,+
+ + + + + + 90

AF 20-50 mT
B) NRM;
N -152


C) 0-15 mT AF cleaning;
.9
+
+ 6
D) 10-50 mT AF cleaning;

8
+ + + + 90
E) 60-100 mT AF cleaning;
AF 60-100 mT
N-116

F) 60-100 mT AF cleaning with
270
+

tectonic correction.
4
6
G) 525C thermal cleaning. +
+ + +
90
AF 60-100 mT
N-116
Conventions as in Figure 3. The plots are without +
+CORRECTED
6*

tectonic correction except for E.
10
12
270
+
+ + + + + +
90
THER. 525"C
AF 0-15 mT N-143
N- 172 + +
+ +
180 180
Table 1. Component mean directions.
Unit
Number of Site
Mean Remanence Direction
Component
Treatment
Means (Specimens) Decl Inc! R K a95
Aldridge Sediments
A 10-20mT 9 (29) 22.9 64.1 8.640 23 9.8
B 10-20mT 8 (21) 320.4 52.1 7.601 18 11.8
C 50-100mT 19 (82) 209.0 3.1 18.437 32 6.0
Moyie Sills
A 20-50mT, 11 (57) 18.2 47.5 10.842 63 5.8
525C
B 525C 10 (28) 303.6 56.9 9.764 38 7.9
The mean remanence direction is given by its declination (0 Decl), inclination (0 Inc!), vector resultant (R), precision
parameter (K), and radius of cone of 95% confidence (0 u95)' From Fisher (1953).
378
Figure 5. Orthogonal vector de-
magnetization plots for example Moyie
sill specimens with the A component.
The horizontal plane (north, east,
south, west) is shown as circles and the
vertical plane (up, down) as triangles.
The axial intensity lengths are ex-
pressed as a ratio of the NRM intensity
for each specimen. The AF cleaning
field is given in mT from 0 for the
NRM.
Figure 6. Orthogonal vector for de-
magnetization plots for example Moyie
sill specimens with the B component on
AF demagnetization. Conventions as
in Figure 5.
D. T. A. Symons and E. A. Timmins
wu wu
0.4
S ---;,------1 N
1.0
ED
10 50
10
o
...
1.0
ED 550
wu
1.0
wu
25
1.0
I--_-__ -:t- N
532
ED
wu
0.8


1.0
503
049
N U
0.2
1.0
so
ED
o
NU
0.4
W---+----"fo-E
0.4
513
0.8
SO
536
ED
WU
1.0
S --+---+-...""..lI-N
o
043
0.2
EO
Geotectonics of the cratonic margin, British Columbia and Montana 379
400N U
NU
500
w 0
0.4
w 0.6
W
0 0
510
SD
550 541
In a final attempt to isolate a primary reman-
ence, 6 fresh specimens from each of the 24 most
stable sites were thermally demagnetized at 525C.
Before tectonic correction, two anomalies of > 99%
significance were isolated (Figure 4G), whereas
after correction none of > 95% significance was
found. Thus, both anomalies record a post-folding
remanence and reflect the most stable record of the
residual A and B components. The B component is
likely the older because A dominates the lower
blocking-temperature and coercivity spectrums,
whereas both are about equally represented in the
higher blocking-temperature domains. Demagne-
tization characteristics are consistent with coarse-
grained magnetite being the main remanence
carrier. Also both contact test sites gave negative
results, indicating that the A and B components are
post-intrusive.
Sediment Demagnetization
As before, one specimen per sediment site was
AF demagnetized in 15 steps to 100 mT. When the
0-15 mT population is grouped, two anomalies of
>99% significance are found that correspond to A
and B component directions seen in the sills.
Several anomalies of marginal (= 95%) significance
are directed towards the northeast and southwest
perimeter (Figure 3C). On tectonic tilt correction, 8
anomalies of only marginal significance remain,
demonstrating again that both A and B components
are post-folding in origin, as found in the sill
specimens. When the 20-50 mT vector population
for the sediments is formed without tilt correction,
the B component disappears, leaving only the A
component (Figure 3D). The A component is re-
N U
Figure 7. Orthogonal vector demagne-
tization plots for example Moyie sill
200 specimens on thermal demagnetiza-
400
tion. Conventions as in Figure 5, ex-
E
cept that the temperature steps are in
200
C.
0.8
S D
moved in turn when the 60-100 mT population is
formed without tilt correction; however, a signifi-
cant anomlay appears at the SW and NE perimeter
that is termed the C component (Figure 3E). It is
about equally represented by Nand R specimens,
indicating acquisition during one or more reversals
of the EMF. Individual specimens provide examples
of several types of demagnetization, such as: (1) the
isolation of the A component only (Figure 8, #533);
(2) the isolation of A, then movement to C (#041);
(3) the isolation of B, then movement to C (#045);
(4) the movement towards and then isolation of C
(#544); and (5) the simple isolation ofC with both N
and R polarities (#508,531).
Fourteen fresh specimens from 14 sediment
sites were thermally demagnetized in 11 steps up to
650C. All specimens showed erratic intensity and
directional behaviour, even at low temperatures.
This indicates that Aldridge argillites are not amen-
able to thermal demagnetization because of rapid
chemical reactions within them upon heating.
All remaining sediment specimens were first
AF cleaned at 10 to 25 mT. Only one site had a cone
of 95% confidence (Fisher, 1953) of <32 for the C
component, indicating very poor within-site consis-
tency. Therefore, two approaches were adopted.
First, the remagnetization circle method of
Halls (1976) was used on the AF step data. The 10
best-defined demagnetization tracks heading in the
C component direction give a mean direction of 219,
9 (u95 = 15) before tilt correction (Figure 3F) and
218, 1 (U95 = 28) after tilt correction. The former
is significantly better grouped at the 95% confidence
level, which indicates that the C component is also
post-folding in origin.
The second approach was to use population
level analysis on all > 50 mT AF data without tilt
380 D. T. A. Symons and E. A. Timmins
wu wu
100
Figure.8. Orthogonal vector demagnetiza
tion plots for example Aldridge (Purcell)
sediment specimens on AF demagnetiza
tion. Conventions as in Figure 5.
S --4r--__ --+
S
Figure 9. Pole positions for the average A,
B, and C components (Table 2) showing
their uncorrected (triangle) position and
corrected (circle with oval of 95% confi
dence) position after 37 of clockwise rota
tion relative to segments of the APWP
corresponding to the Laramide 200 Ma;
Irving, 1979), the Goat River (900-800 Ma;
Dunlop, 1984), and the East Kootenay
(1450-1300 Ma; Irving, 1979) orogenies. P
is the location of the unaltered Prichard
Formation pole of Elston and Bressler
(1980).
0
1.0 1.0
ED
533
WU
0.8
0
10
100
S 0 ..... 6----..... t--N
ED
531
ED
o
0
041
0
wu
o
0.4
100
S N
0.8
E D
508
045
WU
0.4
02
E D
Geotectonics of the cratonic margin, British Columbia and Montana 381
correction. This defines one anomaly (Figure 3E)
with a mean direction of 209, 3 (U95 = 6, k = 32)
before tilt correction and of 209, 6 (U95 = 10, k =
6) after correction. With the C component being
significantly better grouped at the 95% confidence
level before rather than after tilt correction, it is
evident from this fold test that the C component is
post-folding in origin. The C component also likely
pre-dates both A and B because it is preferentially
found in high-coercivity domains. Finally, the AF
and thermal demagnetization characteristics are
consistent with the remanence being carried by fine-
grained magnetite and probably some pyrrhotite.
DISCUSSION OF POLE POSITIONS
Pole positions found for the A, B, and C
components in Moyie sills and Aldridge sediments
(Table 2) are shown in Figure 9. Also shown are
segments of the apparent polar wander path
(APWP) for the North American craton that cor-
respond to the East Kootenay, Goat River, and
Laramide orogenies (Irving, 1979; Dunlop and
Stirling, 1985). All poles in the sediments, except
perhaps for C, are discordant with respect both to
the APWP at these times of known regional meta-
morphism and to the pole determined for the
unmetamorphosed Prichard Formation (Elston and
Bressler, 1980). This discordancy implies that the
terrane housing the Purcell anticlinorium has
undergone significant geotectonic translation
and/or rotation. If simple translation is invoked,
then the terrane would have had to originate ==: 30
to the south or southwest to achieve concordancy.
This is unlikely because the most distant known
outcrop that might have been deposited in the Belt-
Purcell basin is in the south-central part of the
U.S.A. - or ==:15 away. More cogently, the Belt-
Purcell rocks in the study area extend, with
apparent stratigraphic continuity, to the east and
the southeast into relatively undeformed areas of
the Cordillera where they give concordant pole
positions (Evans and others, 1975; Vitorello and van
der Voo, 1977; Elston and Bressler, 1980). This
suggests that tectonic rotation about a local vertical
axis is a more probable cause of discordancy. Noting
that the regional WNW-NW structural trends found
in the Belt-Purcell to the southeast swing to NNW-
NNE trends in the study area, clockwise rotation of
==:45 is an obvious possibility (Hoy, 1984; Reynolds,
1984; Schmidt and Garihan, 1986). A 45 clockwise
rotation is consistent with the rotated sole marks or
transport direction observed by Cressman (1984) in
the study area. Also, paleomagnetic data from the
Quesnellia subterrane of Terrane I, which abuts on
the western side of the study area, indicate an ==: 43
clockwise rotation (Symons and Litallien, 1984).
Significantly, correcting the A and B poles in both
sills and sediments for an ==: 37 12 clockwise
rotation shifts both poles onto the APWP at ==: 180
Ma in Laramide time and at ==:810 Ma in Goat River
Table 2. Pole positions.
Component/Unit
Rotated
LongoE Lat. ON
SO
m
So
p
A Aldridge No 337.3 74.2 15.6 12.5
Moyie Sill No 23.2 65.2 7.5 4.9
Average Yes 115.0 73.0 12.0 9.0
B Aldridge No 139.7 56.3 16.1 11.0
Moyie Sill No 159.9 48.4 11.5 8.4
Average Yes 176.0 29.0 14.0 10.0
C Aldridge No 208.6 -33.5 6.0 3.0
Aldridge Yes 248.0 -38.0 11.0 6.0
Aldridge Yes 248.0 -38.0 11.0 6.0
Pole positions calculated using a mean site location of 244.1E, 49.1N. The rotation
correction is 37 clockwise. The pole positions are by easterly longitude and northerly lati-
tude, with the semi-axes ofthe oval of 95% confidence 0;). From Fisher (1953).
382
D. T. A. Symons and E. A. Timmins
time (Figure 9). This rotation simply moves the C
pole from == 1370 to 1400 Ma on the path but still
within the available time frame from 1436 to 1370
Ma for the East Kootenay orogeny. These age
assignments for the A, B, and C poles are also con-
sistent with their relative ages of apparent
remanence acquisition based on their demagne-
tizing characteristics. Further, they are consistent
for the sill A component where the harder > 525C
direction gives an earlier Early Jurassic pole,
whereas the softer 20-50 mT direction gives an later
Middle Jurassic pole. This is the expected pattern
for a terrane that was being uplifted during the
Laramide orogeny.
GEOTECTONIC SUMMARY
The paleomagnetic results from Aldridge
sediments and associated Moyie sills indicate the
following sequence of events for the study area when
integrated with existing geochronologic data.
Deposition of the Aldridge Formation began at
== 1540 Ma, based on the basement age of Reid and
others (1973) and Evans (1986) and on the pole
position found for unmetamorphosed Prichard
Formation by Elston and Bressler (1980) (Figure 9).
Intrusion of Moyie sills and formation of the
Sullivan ore body took place at about 1435 Ma
(LeCouteur, 1979; Zartmann and others, 1982).
Compression in the East Kootenay orogeny
deformed and uplifted the rocks of the Purcell anti-
clinorium to produce the post-folding C component
in the Aldridge sediments at == 1400 Ma, based on
its rotated pole position. This implies an earlier
time for the orogeny than McMechan and Price
(1982) proposed, by == 50 Ma, but it is still well
within the available time window.
Intrusion of granitic plutons at == 1370 Ma
(Evans, 1986) occurred after deformation in the East
Kootenay orogeny (McMechan and Price, 1982).
The B remanence component found in both
Aldridge sediments and Moyie sills was acquired
during the Goat River orogeny at ==810 Ma, based
on their rotated pole positions.
The A remanence component found in both the
Aldridge sediments and the Moyie sills was ac-
quired during the Laramide orogeny, based on their
rotated pole positions. As noted for the sills, the
most stable, or "hardest," A domains were magne-
tized at == 200 Ma in the earlier Early Jurassic, with
progressively less stable, or "softer," A domains
being magnetized through to == 170 Ma in Middle
Jurassic time. This corresponds to uplift following
the == 20-km-deep burial event postulated by
Archibald and others (1983) for the Purcell anti-
clinorium at this time, as the eastern side of Terrane
I was obducted eastwards onto the continental
margin (Monger and others, 1982; Brown and
others, 1986).
After mid-Middle Jurassic time, the terrane
housing the Purcell anticlinorium was rotated
clockwise == 37 12 about a local vertical axis as a
fairly coherent block. Archibald and others (1984)
suggest that these Proterozoic rocks were raised to
supracrustal levels by mid-Cretaceous time and
then subjected to latest Cretaceous to earliest
Tertiary thrusting, as the accretion of Terrane I was
terminated with the outboard docking of Terrane II.
Price (1986) and Cook and others (1987), citing
geological and seismic evidence, followed the same
timetable for thrusting supracrustal rocks along
west-dipping thrust faults over the autochthonous
basement of the craton. The clockwise sense of
rotation is attributed to: (l)the impact of Terrane I
from the southwest (Irving and others, 1985);
(2) ''ball bearing-type" rotations associated with
dextral transcurrent faulting (Beck, 1976; Gab-
rielse, 1987); or (3) rotation of the southern end of
Terrane I only around the end of the Coast Plutonic
Complex as it collided from the west (Symons, 1971).
Of these alternatives, (1) and (3) fit best with known
Mesozoic-Cenozoic offshore plate motions for west-
ern North America (Engebretson and others, 1985).
Finally, these paleomagnetic results are equivocal
with respect to the "enclosed lacustrine" and "ocean-
margin marine" hypotheses for the origin of the Belt
basin (Winston, 1986).
ACKNOWLEDGMENTS
The authors are grateful to F. R. Edmunds and
J. M. Hamilton (Cominco Ltd.) for their help in
many ways throughout this study, to V. E. Symons
who did most of the measuring, and to Cominco Ltd.
for funding this study.
REFERENCES
ARCHIBALD, D. A., J. K. GLOVER, R. A. PRICE, E.
FARRAR, and D. M. CARMICHAEL, 1983, Geochronology
and tectonic implications of magmatism and metamor-
phism, southern Kootenay Arc and neighbouring regions,
Geotectonics of the cratonic margin, British Columbia and Montana 383
southeastern British Columbia, Part I: Jurassic to mid-
Cretaceous: Canadian Journal of Earth Sciences, v. 20, p.
1891-1913.
__ , T. E. KROGH, R. L. ARMSTRONG, and E. FARRAR,
1984, Geochronology and tectonic implications of magma-
tism and metamorphism, southern Kootenay Arc and
neighbouring regions, southeastern British Columbia,
Part II: Mid-Cretaceous to Eocene: Canadian Journal of
Earth Sciences, v. 21, p. 567-583.
ARMSTRONG, R. L., 1975, Precambrian (1500 m.y. old)
rocks of central Idaho - The Salmon River arch and its
role in Cordillera sedimentation and tectonics: American
Journal of Science, v. 275-A, p. 437-467.
BECK, M. E., 1976, Discordant paleomagnetic pole posi-
tions as evidence of regional shear in the western Cordi-
llera of North America: American Journal of Science, v.
276, p. 694-712.
BROWN, R. L., J. M. JOURNEAY, L. S. LANE, D. C.
MURPHY, and C. J. REES, 1986, Obduction, backfolding
and piggyback thrusting in the metamorphic hinterland of
the southeastern Canadian Cordillera: Journal of Struc-
tural Geology, v. 8, p. 255-268.
BURWASH, R. A.,H. BAADSGAARD, and Z. E. PETERMAN,
1962, Precambrian K-Ar dates from the western Canadian
sedimentary basin: Journal of Geophysical Research, v. 67,
p. 1617-1625.
COOK, F. A.,P. S. SIMONY, K. C. COFLIN, A. G. GREEN, B.
MILKEREIT, R. A. PRICE, R. PARRISH, C. PATENAUDE, P.
L. GORDY, and R. L. BROWN, 1987, Lithoprobe southern
Canadian Cordillera transect: Rocky Mountain thrust belt
to Valhalla gneiss complex: Geophysical Journal of the
Royal Astronomical Society, v. 89, p. 91-98.
CRESSMAN, E. R., 1984, Paleogeography and paleotectonic
setting of the Pritchard Formation - A preliminary inter-
pretation; pp. 18-22 in S. W. Hobbs (ed.), The Belt: Belt
Symposium II (1983), Abstracts with Summaries:
Montana Bureau of Mines and Geology, Special Publication
90,1l7p.
DUNLOP, D. J., and J. M. STIRLING, 1985, Post-tectonic
magnetizations from the Cordova gabbro, Ontario, and
Palaeozoic reactivation in the Grenville Province: Geo-
physical Journal of the Royal Astronomical Society, v. 85,
p.521-550.
ELSTON, D. P., and S. L. BRESSLER, 1980, Paleomagnetic
poles and polarity zonation from the Middle Proterozoic
Belt Supergroup, Montana and Idaho: Journal of Geophy-
sical Research, v. 85, p. 339-359.
ENGEBRETSON, D. C., A. COX, and R. G. GORDON (eds.),
1985, Relative Motions Between Oceanic and Contin-
ental Plates in the Pacific Basin: Geological Society of
America, Special Paper 206, 59 p.
EVANS, M. E., D. K. BINGHAM, and E. W. McMURRAY,
1975, New paleomagnetic results from the Upper Belt-
Purcell Supergroup of Alberta: Canadian Journal of Earth
Sciences, v. 12, p. 52-61.
FISHER,R. A., 1953, Dispersion on a sphere: Proceedings of
the Royal Society of London, v. 217, p. 295-305.
GABRIELSE, H., 1987, Regional transcurrent faults in the
Canadian Cordillera [abstract): XIX General Assembly of
the IUGG, Abstracts, v. 1, p. Ill.
GILE'ITI, B. J., 1968, Isotopic geochronology of Montana
and Wyoming; pp.1ll-146 in E. J. Hamilton and R. M.
Farquhar (eds.), Radiometric Dating for Geologists:
Interscience, London, England, UK, 506 p.
HALLS, H. C., 1976, A least-squares method to find a
remanence direction from converging remagnetization
circles: Geophysical Journal of the Royal Astronomical
Society, v. 45, p. 297-304.
HAMILTON, J. M., D. T. BISHOP, H. C. MORRIS, and O. E.
OWENS, 1982, Geology of the Sullivan orebody, Kimberly,
B.C., Canada: Geological Association of Canada Papers, v.
25, p. 597-665.
HARRISON, C. G. A., 1980, Analysis of the magnetic vector
in a single specimen: Geophysical Journal of the Royal
Astronomical Society, v. 60, p. 489-492.
HOY, T., 1984, The Purcell Supergroup near the Rocky
Mountain Trench, southeastern British Columbia; pp. 36-
38 in S. W. Hobbs (ed.), The Belt: Belt Symposium II
(1983), Abstracts with Summaries: Montana Bureau of
Mines and Geology, Special Publication 90, 117 p.
IRVING, E., 1979, Paleopoles and paleolatitudes of North
America and speculations about displaced terrains: Cana-
dian Journal of Earth Sciences, v. 16, p. 669-694.
KAME, W. B., 1959, Ice petrofabric observations from Blue
Glacier, Washington, in relation to theory and experiment:
Journal of Geophysical Research, v. 64, p. 1891-1909.
LECOUTEUR, P. C., 1979, Age of the Sullivan lead-zinc
deposit: Geological Association of Canada (Cordilleran
Section), Program with Abstracts, p. 19.
McMECHAN, M. E., and R. A. PRICE, 1979, Superimposed
low-grade metamorphism in the Mount Fisher area, south-
eastern British Columbia - Implications for the East
Kootenay orogeny: Canadian Journal of Earth Sciences, v.
19, p. 476-489.
MONGER, J. W. H., R. A. PRICE, and D. J. TEMPLEMAN-
KLUIT, 1982, Tectonic accretion and the origin of two
major metamorphic and plutonic welts in the Canadian
Cordillera: Geology, v. 10, p. 70-75.
PRICE, R. A., 1986, The southeastern Canadian Cordillera:
Thrust faulting, tectonic wedging, and delamination of the
384 D. T. A. Symons and E. A. Timmins
lithosphere: Journal of Structural Geology, v. 8, p. 239-
254.
PULLAIAH, G., E. IRVING, K. L. BUCHAN, and D. J.
DUNLOP, 1975, Magnetization changes caused by burial
and uplift: Earth and Planetary Science Letters, v. 28, p.
133-143.
REID, R. R, D. A. MORRISON, and W. R GREENWOOD,
1973, The Clearwater orogenic zone, a relict of Proterozoic
orogeny; pp. 10-56 in Belt Symposium 1973, Volume I:
Idaho Bureau of Mines and Geology and the Department of
Geology, University ofIdaho, Moscow, Idaho, USA, 322 p.
REYNOLDS, M. W., 1984, Tectonic setting and develop-
ment of the Belt Basin, northwestern United States; pp.
44-46 in S. W. Hobbs (ed.), The Belt: Belt Symposium II
(1983), Abstracts with Summaries: Montana Bureau of
Mines and Geology, Special Publication 90, 117 p.
RYAN, B. D., and J. BLENKINSOP, 1971, Geology and
geochronology of the Hellroaring Creek stock, British
Columbia: Canadian Journal of Earth Sciences, v. 8, p. 85-
95.
SCHMIDT, C. J., and J. M. GARIHAN, 1986, Middle
Proterozoic and Laramide tectonic activity along the
southern margin of the Belt Basin; pp. 217-235 in S. M.
Roberts (ed.), Belt Supergroup: A Guide to Proterozoic
Rocks of Western Montana and Adjacent Areas: Mon-
tana Bureau of Mines and Geology, Special Publication 94,
311p.
STUPAVSKY, M., and D. T. A. SYMONS, 1982, Isolation of
early Paleohelikian remanence in Grenville anorthosites
of the French River area, Ontario: Canadian Journal of
Earth Sciences, v. 19, p. 819-828.
SYMONS, D. T. A., 1971, Paleomagnetism of the Triassic
Guichon batholith and rotation in the Interior Plateau,
British Columbia: Canadian Journal of Earth Sciences, v.
8, p. 1388-1396.
__ , and C. R LITALIEN, 1984, Paleomagnetism of the
Lower Jurassic Copper Mountain intrusions and geotec-
tonics of Terrane I, British Columbia: Geophysical Re-
search Letters, v. 11, p. 685-688.
VITORELLO, 1., and R. VAN DER VOO, 1977, Late Hadry-
nian and Helikian pole positions from the Spokane Forma-
tion, Montana: Canadian Journal of Earth Sciences, v. 14,
p.67-73.
WINSTON, D., 1986, Sedimentology of the Ravalli Group,
Middle Belt carbonate and Missoula Group, Middle
Proterozoic Belt Supergroup, Montana, Idaho and Wash-
ington; pp. 85-124 in S. M. Roberts (ed.), Belt Super-
group: A Guide to Proterozoic Rocks of Western Mon-
tana and Adjacent Areas: Montana Bureau of Mines and
Geology, Special Publication 94, 311 p.
ZARTMAN, R E., Z. E PETERMAN, J. D. OBRADOVICH, M.
D. GALLEGO, and D. T. BISHOP, 1982, Age of the Crossport
C sill near Eastport, Idaho: Idaho Bureau of Mines and
Geology, Bulletin, v. 24, p. 61-69.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Impact origin of large intracratonic basins, the stationary
Proterozoic crust, and the transition to modern plate tectonics
J. W. SEARS and D. ALT
Department of Geology, University of Montana, Missoula, Montana 59812, USA
(received August 12, 1988; revision accepted March 1, 1989)
ABSTRACT
Impacts oflarge meteorites or comets into Proterozoic continental crust may have opened basins as much
as hundreds of kilometers in diameter. We suggest that basalt magma generated through pressure-relief,
partial melting flooded the craters to form lava lakes, which evolved into lopoliths capped with flood-basalt
flows.
Thermal-contraction subsidence of these heavy igneous rocks created basins that continued to subside
under sedimentary loading. Meanwhile, continued partial melting in the upper mantle generated more basalt
magma that entered the basin fill as either flows or diabase sills. Continuing igneous activity within the
basin, instead of along a hotspot track, implies that the Proterozoic lithosphere was stationary. Sharply
limited rifting associated with large Proterozoic basins similarly suggests a nearly stationary crust. Perhaps
very large areas of relatively thin continental crust blocked sweeping plate movement.
Thrusting associated with subduction within extremely large impact basins of generally oceanic
character may have thickened the Proterozoic crust while diminishing its area. Relatively straight and long
mobile belts appeared near the end of Proterozoic time, apparently marking the time when ocean basins
became large enough to permit the transition to modern plate tectonic settings.
INTRODUCTION
Why do deep sedimentary basins form within
continents in the absence of apparent plate tectonic
context? And why are they so commonly circular?
It is difficult to imagine the essentially linear
processes of plate tectonics creating isolated and
nearly circular basins in places remote from plate
boundaries.
Rampino and Stothers (1988) suggested that
flood-basalt plateaus originate through large im-
pacts. Alt and others (1988) argued that impacts of
very large meteorites during Mesozoic and Tertiary
time created a large lava plateau, a hotspot track,
and a spreading ridge. That array may be the
distinctive tectonic signature of an impact explosion
that opens a crater significantly > 100 km in dia-
meter - large enough to prompt significant partial
melting and upwelling in the upper mantle. In our
model (A It and others, 1988), molten basalt then
floods the crater to create a lava lake, which
develops into a flood-basalt plateau that covers a
layered gabbro and granophyre lopolith. Very large
impacts fracture stressed lithosphere. If the litho-
sphere is free to move, the fracture may become a
spreading ridge. Continued volcanic activity above
the mantle site of the impact generates a migrating
hotspot track in the moving lithosphere. Both the
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 385-392. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
385
386 J. W. Sears and D. Alt
spreading ridge and the hotspot may remain active
for a very long time if magma generated through
pressure-relief melting continues to erupt and drift
off the site, thus maintaining a low pressure cell in
the mantle.
This paper extends that argument to very large
impacts during Proterozoic time, when the tectonic
signature appears to have been different. We sug-
gest that Proterozoic impacts opened large craters
that evolved into more or less circular intracratonic
sedimentary basins. But rift development was
minor, and the hotspot remained in place, injecting
mafic magmas into the accumulating basin-fill
sediments instead of distributing them along a
track. Development of this argument leads us to
suggest that Proterozoic time was a period of
sharply restricted plate movement.
Geologic maps of large shield areas, such as the
geologic maps of Canada (1969) and Australia
(Palfreyman and others 1976), and the tectonic map
of North America (King, 1969) show that Early
Proterozoic fold belts tend to be shorter and more
arcuate than those formed since. The transition to
long and relatively straight mobile belts appears in
Late Proterozoic time. We suggest that thrusting
along the short Proterozoic fold belts thickened and
aggregated the continental crust formed during
Archean time, causing it to cover progressively less
of the Earth's surface. Large areas were thus freed
of continental crust, creating the large ocean basins
necessary for the sweeping plate movements
characteristic of modern plate tectonic processes.
ANATOMY OF A LARGE INTRACRATONIC
SEDIMENTARY BASIN: A COMPOSITE MODEL
Our idealized composite basin section (Figure
1) progresses from Paleozoic and Mesozoic sedi-
mentary rocks at the top downward through Pro-
terozoic sedimentary rocks inter layered with basalt
flows and diabase sills. Those lie on a flood-basalt
plateau, which rests on a layered gabbro and grano-
phyre lopolith. Basement rock beneath the lopolith
is brecciated and shock metamorphosed.
Mesozoic and Paleozoic Basin Fill
The upper part of our section shows Paleozoic
basin-fill sediments resting on a deeper fill of Pro-
terozoic sediments interlayered with diabase sills
and basalt flows. An example of that situation
exists in the Michigan basin, where Sleep and Sloss
(1978) and Fowler and Kuenzi (1978) described a
deep borehole that penetrated Proterozoic sedimen-
tary rocks and gabbro sills beneath the Paleozoic
section. The geologic map of Canada (1969) shows
further examples in the Amundsen and Hudson Bay
basins, where Paleozoic sedimentary rocks lie on a
deeper fill of Proterozoic formations. The geologic
maps of Montana (Ross and others, 1958) and Idaho
(Bond, 1978) show Paleozoic formations resting on
the Proterozoic section in the Belt sedimentary
basin.
Protoerozoic Sedimentary Formations, Basalt
Flows, and Diabase Sills
Proterozoic basin-fill sequences typically con-
tain inter layered basalt flows and diabase sills.
Young (1981) and Dostal and others (1986), for
example, described diabase sills and thick sections
of basalt flows inter layered in the Proterozoic
sections along the northern and southern flanks of
the Amundsen basin of northwestern Canada.
The Proterozoic Belt basin of the northern
Rocky Mountains contains an enormously thick sec-
tion of sedimentary rocks interlayered with numer-
ous large diabase sills. Harrison (1972) estimated
the total thickness of the Belt section in one area of
western Montana at "" 20 km, and that in several
other areas at > 10 km. The Arco Paul Gibbs #1
well west of Kalispell, Montana, penetrated 11
diabase sills, that ranged in thickness from 7.5 to
408 m, in drilling some 5280 m of the Prichard For-
mation, the lowest formation in the Belt section.
The aggregate thickness of sills, some 790 m, ac-
counted for "" 15% of the section (Warren Shepard,
pers. commun., 1988). Bishop (1973) reported a
total of 2 km of sills in the Prichard Formation near
Crossport, in northern Idaho. One of those was 600
m thick. Although no comprehensive formal studies
of the diabase sills exist, we observe that they be-
come notably less abundant and generally thinner
upward.
Daniel and Berg (1981) cited radiometric dates
on diabase sills in the Belt basin that range from
1433 to 826 Ma, a span that must approximately
correspond to the period of Belt deposition. Plumb
and others (1981) described a similarly long period
of igneous activity in the Proterozoic McArthur
basin of northern Australia. The time span is
Impact origin oflarge intracratonic basins 387
radUl rift and dike . v ar.
PlIle:o!:o 1 c scd ll11entl
plateau b .... lt
ttenuated continental crust
Keeweenawan basin and Duluth lopolith
gabbl"o/granophyre lopolith
brecciated .and shock.ed blillellent rocks
Figure 1. Schematic composite section through an idealized intracratonic sedimentary basin. Vertical bars indicate control
from selected basins.
similarly long in the Amundsen basin, where the
Coppermine River basalts at the base of the Pro-
terozoic section erupted == 1200 Ma and some of the
sills within the section are as young as 600 Ma.
The Hudson Bay basin appears to consist of a
cluster of three slightly overlapping circles, each a
separate basin. Ricketts and Donaldson (1980) de-
scribed a sequence of Early Proterozoic sediments
and flood-basalt flows, 7-9 km thick, in the Belcher
Islands near the eastern side of Hudson Bay.
Moores (1918) described thick diabase sills in the
same section. Again, the large volume of flows with-
in the section shows that igneous activity accom-
panied sedimentation.
Plumb and others (1981) briefly summarized
the nearly circular Kimberley basin of northern
Australia, which is == 400 km in diameter. It con-
tains an accumulation == 5 km thick of Proterozoic
sedimentary rocks, flood basalts, and diabase sills,
some with granophyre caps. The aggregate thick-
ness of the sills is as much as 3 km, and at least one
near the base of the section appears to lie beneath
the entire basin.
A number of authors, including Grout and
others (1959), Halls (1966, 1978), and Green (1972),
broadly described the Middle Proterozoic Keweena-
wan basin of the Lake Superior region. It contains a
thick section of Proterozoic sedimentary rocks inter-
layered with diabase sills and basalt and grano-
phyric rhyolite flows. Like that in the Amundsen
basin, the Proterozoic section in the Keweenawan
basin lies on a thick section of basalt flows.
Flood Basalts and Lopoliths
Baragar (1977) interpreted the Coppermine
River basalts at the base of the Proterozoic section
in the Amundsen basin as a flood-basalt plateau.
These overlie the Muskox layered mafic intrusion.
White (1968), Annells (1973, 1974), and Green
(1977) interpreted the Keweenawan flows as flood
basalts. Phinney (1970) provided compositional evi-
dence that the Keweenawan lavas erupted from a
magma chamber in the immediately underlying
Duluth lopolith.
388 J. W. Sears and D. Alt
IMPACT ORIGIN OF SEDIMENTARY BASINS
Alt and others (1988) argued that the coinci-
dence in timing between deposition of the terminal
Cretaceous boundary clay and eruption of the
Deccan flood basalts is circumstantial evidence for
an impact origin of the Deccan plateau. The pre-
sence of the Duluth lopolith directly beneath the
Keweenawan flood basalts suggests a link between
flood basalts and lopoliths. Rocks surrounding the
Sudbury lopolith of Ontario provide independent
evidence of an impact origin.
Dietz (1964) showed that the apices of shatter
cones in the country rock around the Sudbury lopo-
lith consistently point toward the center of the
structure and concluded that the Sudbury basin
formed through meteorite impact. He interpreted
the Sudbury gabbro and granophyre lopolith as a
crystallized lava lake and the Whitewater series
above it as sediments deposited in an impact basin.
Dence (1972) showed that the country rock around
the Sudbury structure contains evidence of shock
metamorphism, and French (1968, 1970) showed
that the Onaping tuff, at the base of the Whitewater
series, contains shock-metamorphosed debris.
The Stillwater lopolith of Montana provides
further independent evidence of an origin as a lava
lake flooding a basin. Page and Koski (1973) and
Page and Zientek (1985) described a sequence of
sediments beneath the Stillwater complex that
includes layers of ultramafic diamictites, which
they interpreted as tillite. We suggest an alterna-
tive interpretation, as impact rubble in the floor of a
crater, should be considered. Fragmental rocks
beneath the Stillwater complex should be examined
for any evidence of shock metamorphism that may
have survived contact metamorphism.
Labotka (1985) concluded that hornfels meta-
morphism of rocks beneath the Stillwater complex
proceeded at a pressure between 2 and 4 kb, and
country rock baked while it was at an ambient tem-
perature of The pressure closely corres-
ponds to the probable original thickness of the
lopolith. That, along with the evidence of high
ambient temperature in the country rock, fits our
hypothesis of contact metamorphism beneath a lava
lake filling a deep crater.
Hamilton (1970) interpreted the Bushveld
lopolith and the nearby Vredefort Ring as the result
of simultaneous impact of several large meteorites.
Rhodes (1975) interpreted the lower part of the
Rooiberg felsites above the Bushveld lopolith as
partially melted impact rubble comparable to the
Onaping tuff of the Sudbury complex, and he sug-
gested that chaotic diamictites in the basal part of
the Waterberg sedimentary series are crater ejecta
dumped into annular synclines.
If some lopoliths originated through meteorite
impact, then it seems reasonable to apply the same
interpretation generally. We therefore suggest that
the Duluth lopolith formed as a lava lake flooding a
large impact crater, as did the large layered gabbro
and granophyre complex that Ham and others
(1964) described in southwestern Oklahoma.
Hamilton (1956) had earlier interpreted that
complex as a lopolith.
THE LAHOOD DIAMICTITE
The LaHood diamictite of western Montana
may contain direct evidence that the Belt sedi-
mentary basin formed through meteorite impact.
The formation is locally exposed along an E-
trending line that separates terranes to the south,
where the Middle Cambrian Flathead sandstone lies
on Archean basement, from those to the north where
a thick veneer of Proterozoic Belt formations covers
the basement. McMannis (1963) interpreted the
LaHood Formation as a basin marginal facies of
coarse sediment shed across an active fault scarp
from a rising highland in the south into a subsiding
basin in the north. In a tabulation of the history of
interpretation of the LaHood Formation, Schmidt
and Garihan (1986) showed that many later authors
broadly agreed with that view. Our own observa-
tions of the LaHood Formation lead us to suggest
otherwise.
The type LaHood Formation exposed near the
head of Jefferson Canyon, just east of Cardwell,
Montana, is a dark greenish-gray deposit of frag-
ments of a wide variety of igneous and metamorphic
rock types. Bedding is extremely crude, sorting
almost absent, the general aspect chaotic. Frag-
ments smaller than about a centimeter across are
generally sharply angular; those larger than that
size are to some extent rounded. Rock fragments
float in a matrix of sharply angular crystals of
quartz, feldspar, amphibole, pyroxene, and mica -
no clay fraction.
Thin-sections of LaHood rocks reveal numerous
examples of structures and textures that resemble
those other authors have interpreted as evidence of
shock metamorphism. In particular, we find:
Impact origin oflarge intracratonic basins 389
quartz grains that exhibit parallel micro-
lamellae;
strongly kink-banded grains of biotite and al-
tered biotite;
plagioclase crystals that show multiple offsets
of albite twins;
mosaic deformation of hornblende; and
shattered rock fragments full of microbreccia
dikes that do not extend into the matrix.
If the LaHood Formation does indeed contain
shock-metamorphosed material, then it may well be
an apron of explosion rubble ejected from the crater
that eventually developed into the Belt sedimentary
basin. If so, then transport rounding of the larger
fragments occurred within the rapidly moving cloud
of crater ejecta. We suggest that the Proterozoic
fault scarp bounding the southern margin of the
LaHood exposure may be the southern boundary of a
radial rift within which the ejecta was preserved.
A SCENARIO OF BASIN EVOLUTION
If it were possible to make sedimentary rocks
magically transparent and then view the Earth
from space, the dark igneous rocks within intra-
cratonic basins would appear as round patches of
dark rock set within the lighter rocks of the
continental crust. Or, if the Earth were as devoid of
sedimentary rocks as the Moon, the intracratonic
Proterozoic basins would be dozens of round patches
of dark igneous rocks set within the continents.
These terrestrial mafic basins so analogous to lunar
maria in form and composition may also be
analogous in origin.
We propose a model for basin formation that
starts with the impact of a very large meteorite
blasting an initial crater some tens of kilometers in
diameter. The walls of the original crater collapse
almost immediately along concentric listric faults,
broadening it to a much shallower final basin, > 100
km in diameter, floored with impact and collapse
debris. We suggest that the strength of the litho-
sphere probably limits the depths of final craters to
something between 3 and 5 km but does not restrict
their widths.
Basalt magma derived from pressure-relief
partial melting of the upper mantle would soon flood
the crater. The expansion that accompanies melting
would raise the top of the magma column to a height
that depends upon the depth to which melting
penetrates. A relatively small and shallow crater
would relieve pressure enough to cause partial
melting only to shallow depths in the mantle and
would therefore fill with a shallow pool of magma;
Sudbury is probably an example. Deeper craters
that cause partial melting to much greater depths
may completely fill with basalt and overflow into
surrounding lowlands.
The cooling lava lake erupts floods of basalt
and, in some cases, rhyolite across its surface to
create a lava plateau, while the deeper parts cool to
become a gabbro and granophyre lopolith. The
result at this stage is an essentially circular area in
which substantially denser gabbro and basalt re-
place much of the original continental crust. As the
heated lithosphere beneath the impact site contracts
upon cooling, its surface sags to form a basin that
collects sediments, which cause further subsidence.
Inversion of gabbro to eclogite may also cause the
basin to sink. Turcotte (1980) briefly reviewed these
mechanisms of basin subsidence.
Still larger craters may deplete the mantle
before they fill with basalt and thus remain basins
capable of receiving sediment from the time of their
formation. We suggest that some extremely large
craters may so completely remove the continental
crust that they develop floors of essentially oceanic
aspect in which basalt flows cover layered gabbros
that rest directly on depleted mantle. Such a crater
floor would presumably differ from ordinary oceanic
crust mainly in lacking a sheeted-dike horizon.
Furthermore, the crater rim, with its pattern of
concentric listric faults, has the structural aspect of
a rifted continental margin. Deposition of sedi-
ments along such a rim would produce an array of
structural and depositional features nearly indis-
tinguishable from those that form along a modern
trailing continental margin. Assemblages common-
ly interpreted as Proterozoic continental margins
could equally well be regarded as rims of extremely
large craters.
MAFIC IGNEOUS ACTIVITY WITHIN BASINS
Meanwhile, the lithosphere beneath the basin
bleeds basalt magma into its accumulating fill.
Why?
It is easy to imagine that residual magma may
continue to rise from the lava lake during the early
stages of basin filling. That may explain the ap-
parent tendency for diabase sills and sections of
basalt flows to be thickest and most abundant low in
390 J. W. Sears and D. Alt
the Proterozoic basin-fill sequence. But it is difficult
to suppose that the lava lake could continue to
supply basalt magma to the basin for hundreds of
millions of years. However, if the crust were sta-
tionary, then the mantle hotspot would remain
beneath the basin.
A persistent mantle hotspot beneath a basin
might provide a continuing source of magma for a
very long time, presumably until the accumulating
weight of the basin fill stopped pressure-relief
partial melting. The volume of magma in the basin
fill, although impressive, is not remotely compar-
able with the total of that in a modem oceanic
hotspot track, although it may match that in a
single large island.
THE STATIONARY PROTEROZOIC CRUST
Proterozoic basins typically contain sediments
deposited on and inter layered with rocks of basaltic
composition, but they lack the hotspot track and
spreading ridge characteristic of analogous Meso-
zoic and Tertiary mafic basins. It also seems that
our proposed Proterozoic impacts tended to open
narrow rifts or dike swarms instead of oceanic
spreading ridges. If general movement of continen-
tal lithosphere did not begin until Late Proterozoic
time, it would explain why igneous rocks that would
now be distributed along a hotspot track in a moving
plate instead became part of the basin fill in a
stationary plate. It would also explain why litho-
spheric fractures that would now develop into
spreading ridges became narrow rifts or dike
swarms.
We suggest that, during much of Proterozoic
time, continental crust covered enough of the
Earth's surface to inhibit plate movement through a
sort of planetary gridlock. The major exceptions lay
within impact basins substantially larger than
those now preserved as intracratonic sedimentary
basins - presumably in basins large enough to have
no continental crust in their floors. Such craters
would be generally oceanic in character and may
have collapsed through subduction of their floors to
create the distinctively short and arcuate Protero-
zoic mobile belts. We suggest that thrusting
associated with subduction of those large basin
floors telescoped Archean crust enough to transform
the Earth from a planet covered mostly with rela-
tively thin continental crust to one with smaller but
thicker continents. That change would have created
a situation in which very long spreading ridges
could drive the more open plate movements neces-
sary to develop long mobile belts.
CONCLUSIONS
If Proterozoic time actually was, as we contend,
a period of heavy meteorite bombardment, then
blankets of impact ejecta could survive in many
localities. Such deposits could, quite readily, be
misinterpreted as tillite. We therefore suggest
examining Proterozoic diamictites, especially basal
diamictites that resemble tillite, for evidence of
shock metamorphism.
Obviously, a scenario of prolonged pressure-
relief partial melting beneath the subsiding basin
requires an assumption of a nearly stationary
lithosphere. Continued partial melting in the upper
mantle would otherwise generate a migrating
hotspot track as the lithosphere moves across the
impact wound in the mantle. We suggest that the
first appearance of migrating hotspot tracks may
record the beginning of lithospheric plate move-
ment. Within continents, those may appear as
linear, anorogenic granites that progress in age
along their length.
We further suggest that the first widespread
appearance in the geologic record of ophiolite
complexes with sheeted-dike horizons should mark
the time when spreading ridges began to generate
essentially modern oceanic crust - in effect, the
time of transition to essentially modem plate
tectonic processes.
ACKNOWLEDGMENTS
D. W. Hyndman, S. D. Sheriff, and A. E. Engel
helped us develop these alarming ideas through
many properly skeptical and entirely stimulating
conversations. K. Frye and M. R. Rampino contri-
buted many helpful suggestions in their reviews of
the manuscript. D. Mogk provided valuable edi-
torial help.
REFERENCES
A1T, D., J. W. SEARS, and D. W. HYNDMAN, 1988, Terres-
trial Maria: The origins of large basalt plateaus, hotspot
Impact origin oflarge intracratonic basins 391
tracks, and spreading ridges: Journal of Geology, v. 96, p.
647-662.
ANNELLS, R. N., 1973, Proterozoic Flood Basalts of
Eastern Lake Superior: The Keweenawan Volcanic
Rocks of the Mamainse Point Area, Ontario: Geological
Survey of Canada, Paper 77-10,51 p.
__ , 1974, Keweenawan Volcanic Rocks of Michip-
icoten Island, Lake Superior, An Eruptive Center of
Proterozoic Age: Geological Survey of Canada, Bulletin
218,141 p., and Map 1353A.
BARAGAR, W. R. A., 1977, Volcanism of the stable crust;
pp. 377-405 in W. R. A. Baragar, L. C. Coleman, and J. M.
Hall (eds.), Volcanic Regimes in Canada: Geological
Association of Canada, Special Paper 16, 476 p.
BISHOP, D. T., 1973, Petrology and geochemistry of the
Purcell sills in Bondary County, Idaho; pp. 15-66 in Belt
Symposium 1973, Volume 2: Idaho Bureau of Mines and
Geolog and the Department of Geology, University of
Idaho, Moscow, Idaho, USA, 138 p.
BOND, J. (compiler), 1978, Geologic Map of Idaho: Idaho
Bureau of Mines and Geology, Moscow, Idaho, USA,
1:250,000 scale
DANIEL, F., and R. B. BERG, 1981, Radiometric Dates of
Rocks in Montana: Montana Bureau of Mines and Geo-
logy, Bulletin 114, 136 p.
DENCE, M. R., 1972, Meteorite impact craters and the
structure of the Sudbury basin; pp. 7-18 in J. V. Guy-Bray
(ed.), New Developments in Sudbury Geology: Geologi-
cal Association of Canada, Special Paper 10,124 p.
DIETZ, R. L., 1964, Sudbury structure as an astrobleme:
Journal of Geology, v. 72, p. 412-434.
DOSTAL, J., W. J. R. BARAGAR, and C. DUpuy, 1986,
Petrogenesis of the Natkusiak continental basalts, Vic-
toria Island, Northwest Territories, Canada: Canadian
Journal of Earth Sciences, v. 23, p. 622-632.
FOWLER, J. H., and W. D. KUENZI, 1978, Keweenawan
turbidites in Michigan (deep borehole red beds): A foun-
dered basin sequence developed during evolution of pro-
toceanmic rift system: Journal of Geophysical Research, v.
83, p. 5833-5843
FRENCH, B. M., 1968, Sudbury structure, Ontario: Some
petrographic evidence for an origin by meteorite impact;
pp. 383-412 in B. M. French and N. M. Short (eds.), Shock
Metamorphism of Natural Materials: Mono Book Cor-
poration, Baltimore, Maryland, USA, 644 p.
__ , 1970, Possible relations between meteorite impact
and igneous petrogenesis as indicated by the Sudbury
structure, Ontario, Canada: Bulletin of Volcanology, v. 34,
p.466-517.
GEOLOGICAL SURVEY OF CANADA, 1969, Geological
Map of Canada, Map 1250A, 1:5,000,000 scale.
GREEN, J. C., 1972, General geology, northeastern
Minnesota; pp. 292-317 in Geology of Minnesota: A Cen-
tennial Volume: Minnesota Geological Survey, St. Paul,
Minnesota, USA, 632 p.
__ ,1977, Keweenawan plateau volcanism in the Lake
Superior region; pp. 407-422 in W. R. A. Baragar, L. C.
Coleman, and J. M. Hall (eds.), Volcanic Regimes in
Canada: Geological Association of Canada, Special Paper
16,476p.
GROUT, F. F., R. P. SHARP, and G. M. SCHWARTZ, 1959,
The Geology of Cook County Minnesota: Minnesota
Geological Survey, Bulletin 39, 163 p.
HALLS, H. C., 1966, A review of the Keweenawan geology
ofthe Lake Superior region; pp. 3-27 in J. S. Steinhart and
T. J. Smith (eds.), The Earth Beneath the Continents:
American Geophysical Union, Geophysical Monograph 10,
663p.
__ , 1978, The late Precambrian central North American
rift system: A survey of recent geological and geophysical
investigations; pp. 111-123 in 1. B. Ramberg and E. R. Neu-
mann (eds.), Tectonics and Geophysics of Continental
Rifts: Reidel Publishing Company, Hingham, Massa-
chusetts, USA, 444 p.
HAM, W. E., R. E. DENISON, and C. A. MERRITT, 1964,
Basement Rocks and Structural Evolution of South-
ern Oklahoma: Oklahoma Geological Survey, Bulletin 94,
302p.
HAMILTON, W. B., 1956, Precambrian rocks of Wichita
and Arbuckle Mountains, Oklahoma: Geological Society of
America Bulletin, v. 67, p. 1319-1330.
__ , 1970, Bushveld complex - Product of impacts?:
Geological Society of South Africa, Special Publication 1, p.
367-379.
KING, P. B. (compiler), 1969, Tectonic Map of North
America: U.S. Geological Survey, 1:5,000,000-scale.
LABOTKA, T. C., 1985, Petrogenesis of the metamorphic
rocks beneath the Stillwater complex: Assemblages and
conditions of metamorphism; pp. 70-76 in G. K. Czamanske
and M. L. Zientek (eds.), The Stillwater Complex,
Montana - Geology and Guide: Montana Bureau of
Mines and Geology, Special Publication 92, 396 p.
McMANNIS, W. J., 1963, LaHood Formation - A coarse
facies of the Belt series in southwestern Montana: U.S.
Geological Survey Bulletin, v. 74, p. 407-436.
MOORES, E. S., 1918, The iron formation on Belcher
Islands, Hudson Bay, with special references to its origin
and its associated algal limestones: Journal of Geology, v.
26, p. 412-438.
392 J. W. Sears and D. Alt
PAGE, N. J., and R A. KOSKI, 1973, Precambrian
diamicite below the base of Stillwater Complex, Montana:
U.S. Geological Survey Journal of Research, v. 1, p. 403-
414.
__ , and M. L. ZIENTEK, 1985, Petrogenesis of the
metamorphic rocks beneath the Stillwater Complex:
Lithologies and structures; pp. 55-69 in G. K. Czamanske
and M. L. Zientek (eds.), The StiUwater Complex,
Montana - Geology and Guide: Montana Bureau of
Mines and Geology, Special Publication 92, 396 p.
PALFREYMAN, W. D., G. W. D'ADDARIO, R. A. SWOBODA,
J. M. BULTITUDE, and I. T. LAMBERTS (compilers), 1976,
Geology of Australia: Australian Department of Natural
Resources, 1:2,400,000-scale.
PHINNEY, W. C., 1970, Chemical relations between Ke-
weenawan lavas and the Duluth complex, Minnesota:
Geological Society of America Bulletin, v. 81, p. 2487-2496.
PLUMB K. A, G. M. DERRICK, R S. NEEDHAM, and RD.
SHAW, 1981, The Proterozoic of northern Australia; pp.
205-308 in D. R Hunter (ed.), Precambrian of the
Southern Hemisphere: Elsevier, New York, New York,
USA, 882p.
RAMPINO, M. R, and R B. STOTHERS, 1988, Flood basalt
volcanism during the past 250 million years: Science, v.
241, p. 663-668.
RHODES, R C., 1975, New evidence for impact origin ofthe
Bushveld complex: Geology, v. 3, p. 549-554.
RICKETI'S, B. D., and J. A. DONALDSON, 1980, Sedi-
mentary history of the Belcher Group of Hudson Bay; pp.
235-254 in F. H. A. Campbell (ed.), Proterozoic Basins of
Canada: Geological Survey of Canada, Paper 81-10, 444 p.
ROSS, C. P., D. A. ANDREWS, and I. J. WITKIND (com-
pilers), 1958, Geologic Map of Montana: Montana Bur-
eau of Mines and Geology, 1:500,000 scale.
SCHMIDT, C. J., and J. M. GARIHAN, 1986, Middle Pro-
terozoic and Laramide tectonic activity along the southern
margin of the Belt basin; pp. 217-236 in S. M. Roberts (ed.),
Belt Supergroup: A Guide to Proterozoic Rocks of
Western Montana and Adjacent Areas: Montana Bur-
eau of Mines and Geology, Special Publication 94, 311 p.
SLEEP, N. H., and L. L. SLOSS, 1978, A deep borehole in
the Michigan basin: Journal of Geophysical Research, v.
83, p. 5815-5819.
TURCOTTE, D. L., 1980, Models for the evolution of
sedimentary basins; pp. 21-26 in A. W. Bally, P. L. Bender,
T. R McGetchin, and R 1. Walcott (eds.), Dynamics of
Plate Interiors: Geodynamics Series, Volume 1: Am-
erican Geophysical Union, Washington, DC, and the Geo-
logical Society of America, Boulder, Colorado, 162 p.
WHITE, W. S., 1968, The Keweenawan lavas of Lake
Superior: An example of flood basalts: American Journal
of Science, v. 258-A (Bradley Volume), p. 367-374.
YOUNG, G. M., 1981, The Amundsen embayment, North-
west Territories: Relevance to the Upper Proterozoic
evolution of North America; pp. 203 -218 in F. H. A. Camp-
bell (ed.), Proterozoic Basins of Canada: Geological
Survey of Canada, Paper 81-10, 444 p.
CHAPTER 4
ApPALACHIAN MARGINS
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Late Precambrian tectonism - the opening of the Iapetus Ocean
NICHOLAS RAST
Department of Geological Sciences, University of Kentucky, Lexington, Kentucky 40506, USA
(received May 15, 1989; revision accepted September 18, 1989)
ABSTRACT
At the end of the Proterozoiclbeginning of the Paleozoic, the Laurentian craton formed a passive
continental margin facing the Iapetus Ocean. The evidence for this paleogeographic reconstruction, in North
America, is as follows: (1) disparity in thicknesses and lithologies of lower Paleozoic sediments of the cratonic
cover and the Appalachian orogenic belt; (2) the existance of graben-like Late Proterozoic basins at the edge of
the craton; (3) the presence of the Late Proterozoic volcanic deposits in and at the edges of the grabens,
implying extensional tectonics; and (4) the widespread occurrence of Late Proterozoic dikes that stretch
parallel to the Laurentian margin. Cambrian sedimentary rocks rest unconformably to paraconformably on
all these rocks and represent the shoreline facies of the Iapetus Ocean.
On the southeastern side of the northern Appalachian chain, the rocks of the Avalon composite terrane
show some of the same characteristics - including the Late Proterozoic grabens; even more extensive volcanic
rocks; dikes; and the lower Paleozoic sedimentary cover of moderate thickness. However, the detailed
successional and faunal differences between the Avalon and Laurentian rocks are such that the Iapetus Ocean
must have originally separated the two sides of the Appalachian system. It is suggested that this ocean began
as a series of rifts.
INTRODUCTION
In eastern North America, the western margin
of the Appalachian orogenic belt (Figure 1) trends
approximately parallel to the trend of the Late
Proterozoic Grenville belt that was formed = 1000-
1100 Ma (Easton, 1986). Aspects of this belt have
recently been reviewed in a Geological Association
of Canada Special Paper (Moore and others, 1986)
and by Rivers and others (1988), and McLelland and
others (1988). Within the Appalachian mountain
belt, Grenville inliers and slices have been widely
identified in both the southern Appalachians (Bar-
tholomew, 1984) and the northern Appalachians
(Ratcliffe and others, 1988). Although strong de-
formation affects the Paleozoic rocks that overlie the
Grenville basement, there are numerous exposures
where the unconformity between lower Paleozoic
rocks and the underlying metamorphosed Grenville
formations (Rodgers, 1971) is preserved. In other
places, however, thick but local sequences of Late
Proterozoic post-Grenville metasedimentary rocks
intervene between the Grenville and the Paleozoic
rocks and are interpreted as deposits of grabens
(Schwab, 1986; Schwab and others, 1988), the for-
mation of which preceded Cambrian transgression
and the formation of a Cambro-Ordovician car-
bonate bank (Rodgers, 1968) that accumulated at a
passive continental margin of the Laurentian craton
facing the Iapetus (proto-Atlantic) Ocean.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 395-406. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
395
396 N. Rast
The other side of the ocean is usually placed in
the Avalon composite terrane, disjointed fragments
of which have been recognized at the eastern margin
of the Appalachian orogenic belt (Figure 1) in New-
foundland (Williams, 1964; Williams and others,
1972), New Brunswick and Nova Scotia (Rast and
others, 1976), eastern Massachusetts (Skehan and
Murray, 1978; Zartman and Naylor, 1984), and the
Carolinas (Williams and Hatcher, 1983; Secor and
others, 1984). The broadly accepted current inter-
pretation of the composite terrane is that, in Late
Proterozoic time, it was formed as a volcanic island
arc.
In between lie the deformed remnants of Paleo-
zoic pericontinental metasedimentary rocks, asso-
ciated intercalations, volcanic island arcs, and
oceanic crust (ophiolites) intruded by granites and
other plutons. This complicated assemblage of gen-
erally deformed rocks is known as the Central
Mobile Belt (Williams, 1964). Consideration of this
complex belt is beyond this paper, which is princi-
pally concerned with the development and transfor-
mation of the Avalon side of the Iapetus Ocean and
its relationship with the Laurentian Border.
THE LAURENTIAN BORDER
The Laurentian Border of the Appalachian
orogen in North America (senus stricto) throughout
its length is represented by Grenville amphibolite
and granultic gneisses, which are unconformably
covered by Cambrian strata that are much less de-
formed and metamorphosed. The Grenville gneisses
have a complex geologic history that lasted several
hundred million years (Rivers and Chown, 1986;
McLelland and Isachsen, 1986). Although the age of
the Grenville orogeny is often quoted as 1100-900
Ma, it is quite certain that the orogeny, as well as
the associated metamorphism, affected much older
rocks ranging back into the Archean. In addition to
the lithologically dominant high-grade gneisses,
supracrustal metavolcanic and metasedimentary
schists and gneisses, both pre- and post-orogenic
granitic plutons, and intrusions of anorthosites are
abundant and are considered as typical of the
Grenville belt (Figure 1). In both Canada and part
of the northeastern USA (the Adirondacks), Gren-
ville rocks crop out on the surface, but farther south
on the craton they are hidden under a thick cover of
Phanerozoic strata, although they crop out in the
central and southern portions of the Appalachian
Figure 1. Relationships of the Grenville orogenic belt and
the Appalachians. Key: triangle teeth, Grenville Front;
tick marks, Appalachian Front; black areas, Grenville
outcrops within the Appalachian orogen; stippled areas,
Avalon superterrane massifs; A, Avalong Peninsula;
B, Burin Peninsula; CB, Cape Breton; NS, Nova Scotia;
NB, New Brunswick; S, St. John; CpB, Coastal Plain
Boundary, AD, Adirondacks; M, Massachusetts; R, Ra-
leigh, CS, Carolinas.
Late Precambrian tectonism - The opening of the Iapetus Ocean 397
orogen. Nevertheless, boreholes and geophysical
work have led to their recognition below the surface
and the charting of the so-called Grenville front
(Figure 1) at least as far south as Alabama (Muehl-
berger, 1980; Denison and others, 1984). Grenville-
age rocks are found as far south as southern Mexico.
The recently much investigated Grenville front
tectonic zone is usually interpreted to be the result
of a continental collision (Green and others, 1988).
The unconformable sedimentary rocks resting
on Grenviiie gneisses can be divided into Late
Proterozoic and succeeding Cambro-Ordovician
deposits (Schwab and others, 1988; Skehan, 1988).
The former are generally thinner in the north (New
England and New York) and much thicker and
localized in the south (Virginia, North Carolina,
Georgia), where they are interpreted as infillings of
grabens (Schwab and others, 1988) among Gren-
villian massifs (Bartholomew and Lewis, 1988).
However, in western Newfoundland, the Fleur de
Lys Supergroup, which is questionably partly Pro-
terozoic, reaches some 10 km in thickness (Kennedy,
1975). In general, the Cambro-Ordovician sequence
starts from a clastic succession, such as the Lower
Cambrian Chilhowee Group in the south and the
Upper Cambrian Chesire Quartzite in Quebec and
New Emgland, and equivalents in New York. These
basal clastics rocks, in both the northern and south-
ern Appalachian sectors of the adjacent craton, are
succeeded by a carbonate bank, implying formation
of a passive continental border, in the aftermath of
late Precambrian rifting. In places (e.g., Virginia)
the Cambrian transgression, which marked the
opening of the Iapetus Ocean, was preceded by vol-
canism - as, for instance, the Catoctin sequence of
Virginia (Badger and Sinha, 1988). The overlying
Chilhowee sediments of Cambrian age have been
interpreted, on a paleostratigraphic basis, as partly
concurrent with Iapetan rifting and partly post-
rifting (Simpson and Eriksson, 1989).
Until recently, the Iapetus rifting was assumed
to be a single episode, and the generation of the rift
volcanics was accompanied by the injection of a
single set of dikes that, in the central and southern
Appalachians, were explained as feeder dikes to the
Catoctin and Mount Rogers lavas. For instance, on
deducing an age of =600 Ma for rift-related dikes
cutting the Grenville gneisses of western Newfound-
land, Williams and others (1985) inferred that older
dikes at the edge of southern Appalachians reflected
the diachronous nature of rifting. Badger and Sinha
(1988), however, indicated that in Virginia and
North Carolina there are two rifting events, = 700
Ma and = 570 Ma; this has been independently
confirmed by Aleinkoff and others (1991). Thus the
= 730 Ma Bakersville dikes (Goldberg and others,
1986) and the =760 Ma Mount Rogers Volcanics
(Aleinkoff and others, 1991) ofthe southern Appala-
chians represent a different but broadly related
event to those =600 Ma dikes of western New-
foundland and the = 570-600 Ma Catoctin volcanics.
It thus can be suggested that dike swarms, while
reflecting rifting in general, can be injected in a
series of discrete episodes. In the southern Ap-
palachians, the presence of (1) Bakersville dikes and
Mount Rogers volcanics (=730-760 Ma), (2) =680-
730 Ma meta luminous to perakaline granites of the
CrossnorelRobertson River-type (Odom and Fulla-
gar, 1984; Tollo and others, 1991), which chemically
suggest a rift setting (Tollo and Arav, this volume),
and (3) still later 570-600 Ma tholeiitic rift-related
Catoctin basalts imply the possibility of several
separate periods of rifting.
In the northern Appalachian sector, so far only
the =600 Ma (=615 Ma in Labrador; Kamo and
others, 1989) event has been definitely documented
for Late Proterozoic dikes, although there is a
variety of Middle Proterozoic dikes within the
Grenville terrane, varying in age from 1250 to 1140
Ma. According to Gorbatschev and others (1987),
these older dikes are related to the rifting that
preceded the initiation of the Grenvillian orogeny.
In addition, =750 Ma Franklin dikes have been
reported from the Labrador Sea region (Fahrig,
1987). Piper (1988) reported the chronology of
minor intrusions associated with Grenville-age
rocks of Scandinavia that vary in age from = 900 to
544 Ma. These dikes may reflect the onset of frag-
menta ton of the Proterozoic supercontinent. The
apparent differences in dike ages between the north-
ern and southern Appalachians partly depend on
the fact that data from the southern Appalachian
are mainly derived from Grenville blocks in the
interior of the orogenic belt (Blue Ridge), where
these blocks were brought up to the surface by
Alleghanian (Late Carboniferous-Permian) thrusts.
In the northern Appalachians, however, the Alle-
ghanian thrust-front lies well to the east of the main
Appalachian Central Mobile Belt (Rast, 1989).
Therefore, in the northern Appalachians there are
fewer Grenville massifs within the orogenic belt,
and the true relationships at the edge of the Iapetus
Ocean are hidden under the cover of lower Paleozoic
overthrust sediments.
398 N. Rast
It is, however, significant that ""680 Ma dikes
are recorded from the Davis Strait separating
Greenland from Canada (Nielsen, 1987), indicating
an episode of rifting of the same age as the Cross-
nore plutonic series.
An appreciable spread of isotopic ages of dikes
has been recorded in western Newfoundland.
Pringle and others (1971), on the basis of whole-rock
K/Ar dating, reported an ",,805 Ma age. This was
then reinterpreted by Stukas and Reynolds (1974)
who, using 4oAr/39Ar data from a nearby area, found
dikes to be "" 605 Ma and suggested that the greater
age found by Pringle and others arose from pre-
ferential absorption of argon by mafic minerals.
From yet another locality, Williams and others
(1985) determined a U-Pb age of 60210 Ma on a
granite that is cut by mafic dikes. But the dikes do
not penetrate the overlying latest Precambrian
metasedimentary rocks, thus constraining the age
of the dikes as ""600 Ma. The diversity of measured
ages from different localities permits a possibility of
the presence of dikes of several generations that
penetrate the Grenville gneisses of western New-
foundland and reflect several episodes of rifting.
Throughout the Laurentian margin of the Iape-
tus Ocean, therefore, the existing evidence points to
a prolonged series of riftogenic movements lasting
almost 200 million years, from 760 Ma to the Cam-
brian.
THE A V ALONIAN BORDER
The Avalonian border of the Appalachians has
been subject to extensive investigations summari-
zed by Rast and others (1976), Rast and Skehan
(1983), Skehan (1988), Rankin and others (1989),
and Lefort (1989). Diagnostic features of the Ava-
lonian border are late Precambrian, mainly felsic,
volcanic, and volcaniclastic rocks, intermingled
with siliciclastic deposits and intruded by "" 600 Ma
granitoid bodies. These rocks are paraconformally
to uncomformably covered by Cambrian-age marine
strata. In places there are indications of contem-
poraneous intrusives of mixed mafic to intermediate
and felsic magma of Late Proterozoic age (Rast and
Skehan, 1983) and coexisting dioritic to tonalitic
plutons, as well as mafic dike swarms (Rast, 1979).
Late Proterozoic sedimentary-metasedimentary
strata are known from most major outcrops of the
Avalon blocks and are frequently interbedded or
partially interbedded with volcanic rocks. In a few
places, as in southeastern New England, sedimen-
tary formations fill deep graben-like basins of de-
position with some strata interpreted as glacial
tillites (reviewed by Socci and Smith, 1987). Glacio-
genic deposits are also known from Newfoundland
(Bruckner and Anderson, 1971).
Sedimentary beds also occur below the base of
the volcanogenic succession. In New Brunswick and
Nova Scotia, ",,800 Ma quartzites, carbonates, and
argillites have been reported to overlie the com-
posite, partly dioritic Brookfield gneiss (originally
unconformably?). The gneiss has yielded a U-Pb age
of 83026 Ma (Olszewski and Gaudette, 1982),
although now this age is rejected because the gneiss
is, in all probability, intrusive into the metasedi-
ments (Rast and Skehan, 1991). In Nova Scotia,
gneisses of 950 Ma have been dated by Barr and
others (1987), but these gneisses have been sepa-
rated from the Avalon terrane and included in the
Laurentian craton. This interpretation is disputed
by Murphy and others (1989). I interpret the base-
ment of the Avalonian succession as Grenville
gneiss, admittedly separated from the main Gren-
ville platform of the Laurentian continent by the
Central Mobile Belt of the Appalachian orogen.
In the southern Appalachians again, the Ava-
lon equivalents (Carolina Slate Belt and East
Carolina Slate Belt) are also closely associated with
the Grenville rocks (Farrar, 1984) of the Goochland-
Raleigh region, where they appear to be in tectonic
contact (Horton and others, 1986) with an interven-
ing ophiolitic melange. The Eastern Slate Belt is
also in tectonic contact with the Grenville rocks of
the Raleigh region. This relationship I consider as
reflecting the collision of the Grenville terrane to
the east with the Avalonian Carolina Slate terrane
to the west, and the melange represents the squee-
zed out intervening oceanic crust. Even if the Caro-
lina Slate Belt is locally allochthonous, the fact that
the melange lies west of the Raleigh (Goochland)
terrane carries the probability that this Grenville
block was separated from the Laurentian craton, as
also independently suggested by Bartholomew and
Lewis (1988).
Like at the Laurentian side of the Iapetus
Ocean, late Precambrian diabase dikes and associa-
ted late Precambrian basaltic flows were emplaced
along the Avalonian border. In Newfoundland,
similar volcanicity and associated intrusive activity
was described from the Burin Pennisula by Strong
and others (1978) and recently dated (U-Pb method)
by Krogh and others (1988) as "" 763 Ma.
Late Precambrian tectonism - The opening of the Iapetus Ocean 399
In New Brunswick, Rast (1979) inferred mul-
tiple intrusion of metadiabase dikes, the earliest of
which were strongly deformed and the latest,
although metamorphosed, were not appreciably
deformed. At the time, this was interpreted as a
rapid alternation of tensile (emplacement of dikes)
and compressive (deformation) regimes and was
assumed to have occurred within a short period of
time. It seems now possible that a much longer
interval of time between successive stages of
emplacement and deformation may have existed. A
few late and little deformed dikes cut even Cam-
brian rocks. In the city of St. John (New Bruns-
wick), these late dikes intrude strongly deformed
gneisses and amphibolites of the Brookfield gneiss.
Thus, the multiple nature of these dikes, analyzed
in detail by Matthews (1985), may have arisen
during several successive episodes of distension. An
extra complication is that Siluro-Devonian dikes,
only slightly metamorphosed, are found to coexist
with the Precambrian intrusions (McLeod and Rast,
1988).
Hadley (1973) reported that earlier gabbros and
mafic volcanics near Oxford, North Carolina, are cut
by numerous metadiabase dikes, indicating at least
two distensional episodes. The earlier mafic rocks
are alkaline and presumably rift-related. The later
dikes are continental tholeiites. These rocks are
now being investigated at the University of Ken-
tucky. Though limited, the present information
suggests that, as in New Brunswick and Newfound-
land, the distensional activity was probably epi-
sodic. This activity is here attributed, as suggested
Pre-1050
by Piper (1987), to the Late Proterozoic breakup of a
supercontinent which included Laurentia and
Baltica and had a margin formed by the Grenville
orogenic belt (Figure 2).
The breakup of the supercontinent determined
the margins of the Iapetus Ocean that were formed
at the beginning of Cambrian time. In the ensuing
oceanic stretch, an island arc or coterminous arcs
originated, ultimately forming the Avalon super-
terrane. The superterrane has, in part, a basement
of Grenville gneiss that is, in part, overlain by the
Green Head supercrustal rocks. Subsequent early
Paleozoic events involved complex interactions of
island-arc and oceanic and continental terranes that
existed in the Iapetus Ocean (Rast, 1989).
RELATIONSHIPS WITH GONDWANA
Both the Grenville belt and the Avalon compo-
site terrane are represented across the Atlantic
Ocean (Figure 1) and across the trace of the Iapetus
Ocean, which also continues into Europe. Further-
more, rocks analogous to the Avalon terrane have
been detected in Morocco (Schenk, 1971; Hughes,
1972). The ""680-600 Ma deformation of these
strata, referred to as "Pan-African" by French geo-
logists (Choubert and Faure-Muret, 1980; Fabre,
1983), is of a distinct orogenic type. On the basis of
chronologic comparisons, Rast and Skehan (1983)
have correlated the later episodes of these move-
ments with the Cadomian of France. Recently,
Cadomian orogenic events were equated with the
Figure 2. Part of the Proterozoic supercontinent, during and after the Grenville orogeny (after Piper, 1987). Ornamented,
Grenville orogenic belt; B, Baltica; G, Greenland; L, Laurentia.
400
N. Rast
Avalonian of North America (Nance, 1987). Al-
though there is little doubt of the broad comtem-
poraneity of Cadomian and Avalonian movements,
it seems desirable to differentiate between their
regional positions and causes. Avalonian orogenic
effects influence the Avalon superterrane in North
America and in Midlands of England and they occur
both west and north of it. The Cadomian orogenic
belt occurs in northern France (Cabanis and others,
1987), to the southeast of the Avalon terrane of the
Midlands of England, and is suggested to extend
into Central Europe and northern Spain. It is on
this basis that Rast and Skehan (1983) connected it
with the Pan-African structures that lie to the
northeast and east of the West African craton.
Consequently, the Late Proterozoic succession to the
west and northwest of this craton (Figure 3) has
been suggested to correlate with the Avalon super-
terrane, which must have been, in part, an epi-
continental volcanic arc, although other parts of it
may have been oceanic. This type of tectonic unit
would be analogous to the present-day Kamchatka-
Kuril Island arc of the northwestern Pacific Ocean,
where the Kamchatka part is epicontinental and the
Kurils are partly oceanic. Thus, at the end of the
Proterozoic, the Avalon arc was associated in part
with the West African craton. On the other hand, if
the basement of the arc, in places such as New
Brunswick or Nova Scotia, is Grenville then the
position of the West African craton must have been
proximate to the Grenville belt (Figure 4a).
Hagstrum and others (1980) suggested, based
on paleomagnetic studies, that the Armorican mas-
sif as well as the Avalonian terrane of southern
England formed, in the Late Proterozoic and Early
Cambrian, a single continetal mass related to the
West African craton. However, whether the West
African craton at that time was a part of Gondwana
is less certain (Perrin and Prevot, 1988). There is
appreciable evidence that the West African craton
joined the rest of proto-Africa in stages, possibly
beginning at Hoggar along the eastern edge of the
West African craton at 870 Ma (Caby, 1987). The
730 Ma U/Pb event in eastern Hoggar was particu-
larly important, as is the 720 Ma event in the Ardar
des Iforas (Mali). Structural evidence for 620 Ma
compressional movements is also present here. The
sediments, ophiolites, etc., were thrust over the
eastern margin of the West African craton, sug-
gesting that prior to 730 Ma, the West African
craton was separated from the other African shields
and that its suturing to some of these took place in
\
\
I
\
I
I
\
J __ _
I
I
\
I
I
I

I
I
I
I
I
\20
Figure 3. Distribution of the Avalon composite terrane
and associated tectonic elements in Europe and North
Africa. Europe: A, Avalon composite terrane (stippled);
CD, Cadomian orogen (dashed). African cratons: N, Ni-
geria, T, Tuareg, WA, West African. Fragments of oro-
genic belts in Africa (dashed): HG, Hogger Iforas;
AA, Anti-Atlas; DH, Dahomeyides. Note that all these
belts are ""600 Ma, indicating that, prior to then, the West
African craton was separated from Gondwana.
stages between 730 and 610 Ma. Late Proterozoic
granitoids that cut this sequence of deformed rocks
are dated as "'" 615-580 Ma.
Along the northern margin of the West African
craton (Figure 3), within the Anti-Atlas and High
Atlas, a similar sequence of events has been
suggested by Brabers (1988). His model starts with
a stage of rifting at "'" 788 Ma and separation of
blocks, and develops a first major compressive stage
Late Precambrian tectonism - The opening of the Iapetus Ocean 401
Figure 4. Suggested kinematics of continental relationships of the supercontinent (Laurentia-Greenland-Baltical and the
West Mrican craton. Key: flowlines, Grenville rocks; crosses, West Mrican craton; solid triangles, volcanic arcs;
A, Armorica; B, Baltica; G, Greenland; GD, Gondwana; N, Nigerian craton; T, Tuareg craton. Arrows indicate relative
direction of movement of the West African craton. a) ==1050 Ma - formation of the Grenville orogenic belt; b) ==800 Ma -
beginning of fragmentation of the supercontinent; c) ==750 Ma - growth of volcanic arcs and collision of the West Mrican
craton with other Gondwana shields, formation of the Cadomian ocean; d) ==600-650 Ma - Cadomian and Avalonian
orogenies, continentalization of the Amorican plate; adhering blocks of volcanic arc, shown in solid triangles, represent the
Avalon Superterrane; e) ==580 Ma - beginning of post-Cadomian extension, formation of grabens in Amorica, shown as
parallel lines with stippling inside; f) == 520 Ma - Cambrian relationships, extensive Iapetus Ocean.
of obduction at 724 50 Ma, followed by a still later
collision and development of post-collisional Late
Proterozoic volcanism. This volcanism is of the
same character as that of the same age in the A va-
Ion superterrane in Newfoundland (Hughes and
Bruckner, 1971). It is inferred that at this time
Baltica was separated from the West African craton
by a wide oceanic stretch with continental frag-
ments, some of which contained relicts of Archean
minerals and themselves may have been fragments
of the West African craton (Guerrot and others,
1989). The Cadomian orogeny is interpreted to be
the result of the collision of these blocks with each
other, with the West African craton, and with the
Avalonian volcanic are, leading to formation of the
Armorican plate (Hagstrum and others, 1980). The
distance of this plate from Baltica at the end of the
Proterozoic is difficult to estimate, especially
because both of these masses were undergoing ex-
tensional movements at about the same time as the
West African craton underwent peripheral com-
pression. Ages of .., 730 Ma indicate periods of dike
injection in Laurentia, including Bakersville gabbro
dikes in North Carolina-Tennessee at ..,734 Ma
(Goldberg and others, 1986); ..,750 Ma Franklin
dikes on Baffin Island in the Labrador Sea (Tahrig,
1987), and Baltior-Egersund dolerites (850-663 Ma;
Piper, 1988) and 745 Ma metadiabases (Krill, 1983)
in Norway. These were succeeded by later dikes
with ages :5 600 Ma and probably younger.
402 N. Rast
KINEMATIC RECONSTRUCTION
The events following the Grenville orogenic
cycle are concerned with (1) the formation of a Late
Proterozoic supercontinent, (2) its subsequent frag-
mentation and the interplay of the resultant frag-
ments, (3) the formation of at least one new contin-
ental mass, and (4) the ultimate opening of the Iape-
tus Ocean, partly in Late Proterozoic and partly in
Early Cambrian time. The model proposed here
covers North America, western Europe, and north-
western Africa. It is suggested that until "'" 760 Ma,
Baltica, Laurentia, and the West African craton
were all part of a supercontinent along with the
Grenville orogenic belt that had been formed by the
collision of the West African craton and Laurentia-
Baltica at "'" 1000 Ma when the supercontinent
originated (Figure 4a). At ""'800-730 Ma, fragmen-
tation of the supercontinent, marked by injection of
some diabase dikes, began with Baltica and Lauren-
tia drifting away. In particular, the West African
craton moved away from Baltica and between the
two cratons a series of isolated blocks remained
within the Cadomian Ocean (Figure 4b), within
which the Avalon volcanic arc was formed, in part
Figure 5. Early Mesozoic configuration of the
North Atlantic region. Key: flowline ornament,
Grenville rocks; shaded areas, Caledonian Acadian
belt; solid triangles, Avalon terrane; A, Armorica;
B, Baltica; G, Greenland; L, Laurentia; WAC, West
African craton.
WAC
overlapping the West African craton. Meanwhile,
the West African craton and the Tuareg craton
collided (Figure 4b) with each other and also maybe
almagamated with volcanic arcs that then existed in
the Cadomian Ocean. One island arc (Figure 4c),
however, was partly oceanic and partly continental
and later became the Avalon superterrane. When
the Cadomian orogeny took place at "'" 620-580 Ma,
the Avalon arc was deformed and adhered to the
continentalized Cadomian orogenic belt, incor-
porating fragments of the West African craton and
other island arcs (Schermerhorn and others, 1986).
In the process of deformation related to orogeny, the
Avalon island arc was probably broken up into dis-
tinct blocks that moved against each other, giving
rise to the Avalon superterrane (Figure 4d), which
mayor may not have collided with Laurentia. From
580 Ma to the beginning of the Cambrian, exten-
sional movements caused the formation of grabens
and injection of dikes in the Avalon superterrane
(Figure 4e). Although, as mentioned before, a
whole series of compressional and extensional epi-
sodes probably occurred, depending not only on
regional plate interactions but also on local plate
interactions between smaller blocks. As a result,

,
I
I
I
,
\
\
\
,
\
\
,
"
Late Precambrian tectonism - The opening of the Iapetus Ocean 403
age determinations of different episodes varies from
place to place, and the system as a whole is remark-
ably complex.
The opening of the Iapetus Ocean sensu stricto
began really only a short while before the Cambrian
and continued into Cambrian time. This opening
marks the termination of the Cadomian orogenic
events that led to continental growth and amal-
gamation of the Armorican plate. Continental
masses at this time were Laurentia (including
Greenland), Baltica, Armorica, and Gondwana, the
latter by then incorporating the West African
craton. The edges of Gondwana, Armorica, and Bal-
tica were invaded by Cambrian seas, and the frag-
ments of the Avalon superterrane thus covered by
Paleozoic sediments became distributed between the
Armorican and Gondwana continents. They ul-
timately became blocks of the composite Avalon
terrane of North America and Europe, forming parts
of a complex jigsaw puzzle by the end of Paleozoic
time (Figure 5), which then underwent Mesozoic
fragmentation. Although the relative distances of
the main continental cratons (Laurentia, Baltica,
West African craton, Gondwana) varied through
geologic time, it was the interaction of all four in the
last 1000 million years that probably governed the
pattern of Atlantic geology.
ACKNOWLEDGMENTS
I wish to acknowledge discussions with J. W.
Skehan, SJ, M. J. Bartholomew, and Kieran
O'Hara. Jeanie Thompson typed the paper and
Steve Greb helped with figures.
REFERENCES
ALEINIKOFF, J. N., R E. ZARTMAN, D. W. RANKIN, P. T.
LYTTLE, W. C. BURTON, and R C. McDOWELL, 1991, New
U-Pb zircon ages for rhyolite of the Catoctin and Mount
Rogers formations - More evidence for two pulses of lape-
tan rifting in the central and southern Appalachians [ab-
stract]: Geological Society of America Abstracts with Pro-
grams, v. 23, no. 1, p. 2.
BADGER, RL., and A. K. SINHA, 1988, Age and Sr isotopic
signature of the Catoctin volcanic province: Implications
for subcrustal mantle evolution: Geology, v. 16, p. 692-695.
BARR, S. M., R P. RAESIDE, and O. VAN BREEMAN, 1987,
Grenvillian basement in the northern Cape Breton High-
lands, Nova Scotia: Canadian Journal of Earth Sciences, v.
24, p. 992-997.
BARTHOLOMEW, M. J. (ed.), 1984, The GrenvUle Event
in the Appalachians and the Related Topics: Geo-
logical Society of America, Special Paper 194,287 p.
___ , and S. E. LEWIS, 1988, Perigrination of Middle
Proterozoic massifs and terranes within the Appalachian
orogen, eastern U.S.A.: Universidad de Oviedo (Spain),
Trabajos de Geologia, v. 17, p. 155-165.
BRABERS, P. M., 1988, A plate tectonic model for the
Panafrican Orogeny in the Anti-Atlas, Morocco; pp. 61-80
in V. H. Jacobshagen (ed.), The Atlas System of Moroc-
co: Studies on its Geodynamic Evolution: Lecture Notes
in Earth Sciences 15, Springer-Verlag, Berlin, Germany,
499p.
BRUCKNER, W. D., and M. M. ANDERSON, 1971, Precam-
brian glacial deposits in southeastern Newfoundland: A
preliminary note: Geological Association of Canada, Pro-
ceedings, v. 24, p. 95-102.
CABANIS, B., J. CHANTRAINE, and D. RABU, 1987, Geo-
chemical study of the Brioverian (Late Proterozoic) vol-
canic rocks in the northern Armorican Massif (France):
Implications for geodynamic evolution during the Cadom-
ian; pp. 525-539 in T. C. Pharoah, R D. Beckinsale, and D.
T. Rickard (eds.), Geochemistry and Mineralization of
Proterozoic Volcanic Suites: Geological Society of Lon-
don, Special Publication 33, 575 p.
CABY, R, 1987, The Pan-African belt of west Africa from
the Sahara Desert to the Gulf of Benin; pp. 125-170 in J. P.
Schaer and J. Rodgers (eds.), The Anatomy of Mountain
Ranges: Princeton University Press, Princeton, New
Jersey, USA, 298 p.
CHOUBERT, G., and A. FAURE-MURET (eds.), 1980, The
Precambrian in mobile zones (Final report, rGCP Project
No.2): Earth-Science Reviews, v. 16, no. 2/3, p. 85-316.
DENISON, R E., E. G. LIDIAK, M. E. BICKFORD, and E. B.
KISVARSANYI, 1984, Geology and geochronology of Pre-
cambrian rocks in the central interior region of the United
States; pp. CI-C20 in J. E. Harrison and Z. E. Peterman
(eds.), Correlation of Precambrian Rocks of the United
States and Mexico: U.S. Geological Survey, Professional
Paper 1241-C, 20p.
EASTON, R M., 1986, Geochronolgy of the Grenville Pro-
vince; pp. 127-137 in J. M. Moore, A. Davidson, and A. J.
Baer (eds.), The GrenvUle Province: Geological Associa-
tion of Canada, Special Paper 31, 358 p.
FABRE, J. (ed.), 1983, Afrique de l'Ouest: Introduction
Geologique et Termes Stratigraphiques: Pergamon
Press, Oxford, England, UK, 396 p.
404 N. Rast
FAHRIG, W.F., 1987, The tectonic setting of continental
mafic dyke swarms: Failed arm and early passive margin;
pp. 331-348 in H. C. Halls and W. F. Fahrig (eds.), Mafic
Dyke Swarms: Geological Association of Canada, Special
Paper 34, 503 p.
FARRAR, S. S., 1984, The Goochland granulite terrane:
Remobilized Grenville basement in the eastern Virginia
Piedmont; pp. 215-227 in M. J. Bartholomew (ed.), The
Grenville Event in the Appalachians and Related
Topics: Geological Society of America, Special Paper 194,
287p.
GOLDBERG, S.A., J. R. BUTLER, and P. D. FULLAGAR,
1986, The Bakersville dike swarm: Geochronology and
petrogenesis of Late Proterozoic basaltic magmatism in the
southern Appalachian Blue Ridge: American Journal of
Science, v. 286, p. 403-430.
GORBATSCHEV, R., A. LINDH, Z. SOLYOM, 1. LAlTAKARl,
K. ARO, S. B. LOBACH-ZHUCKENKO, M. S. MARKOV, A. I.
IVLIER, and 1. BRYNHI, 1987, Mafic dyke swarms of the
Baltic Shield; pp. 361-372 in H. C. Halls and W. F. Fahrig
(eds.), Mafic Dyke Swarms: Geological Association of
Canada, Special Paper 34, 503 p.
GREEN, A. G., B. MILKERElT, A. DAVIDSON, C. SPENCER,
D. R. HUTCHINSON, W. F. CANNON, M. W. LEE, W. F.
AGENA, J. C. BEHRENDT, and W. J. HINZE, 1988, Crustal
structure of the Grenville front and adjacent terranes:
Geology, v. 16, p. 788-792.
GUERROT, C., J. J. PEUCAT, R. CAPDEVlLA, and L.
DOSSO, 1989, Archean protoliths within Early Proterozoic
granulitic crust of the west European Hercynian belt:
Possible relics of the West African craton: Geology, v. 17,
p.241-244.
HADLEY, J. B., 1973, Igneous rocks of the Oxford area,
Granville County, North Carolina: American Journal of
Science, v. 273-A, p. 217-233.
HAGSTRUM, J. T., R. VAN DER VOO, B. AUVRAY, and N.
BONHOMMET, 1980, Eocambrian-Cambrian palaeomagne-
tism of the Armorican Massif, France: Geophysical Jour-
nal of Royal Astronomical Society, v. 61, p. 489-517.
HORTON, J. W., Jr., D. E. BLAKE, A. S. WYLIE, and E. F.
STODDARD, 1986, Metamorphosed melange terrane in the
eastern Piedmont of North Carolina: Geology, v. 14, p.
551-555.
HUGHES, C., 1972, Geology of the Avalon Peninsula, New-
foundland, and its possible correspondence with Morocco:
Notes et Memoires du Service Geologique, v. 236, p. 265-
275.
HUGHES, C. J., and W. D. BRUCKER, 1971, Late Precam-
brian rocks of eastern Avalon Peninsula, Newfoundland:
A volcanic island complex: Canadian Journal of Earth
Sciences, v. 8, p. 899-915.
KAMO, S. L., C. F. GOWER, and T. E. KROGH, 1979,
Birthdate for the Iapetus Ocean? A precise U-Pb zircon
and baddleyite age for the Long Range dikes, southeast
Labrador: Geology, v. 17, p. 602-605.
KENNEDY, M. J., 1975, Repetitive orogeny in the north-
eastern Appalachians - New plate models based upon
Newfoundland examples: Tectonophysiccs, v. 28, p. 39-87.
KRILL, A. G., 1983, Rb-Sr study of dolerite dikes and psam-
mite from the Western Gneiss region of Norway: Lithos, v.
16, p. 85-93.
KROGH, T. E., D. F. STRONG, S. J. O'BRIEN, and V. S.
PAPEZlK, 1988, Precise U-Pb zircon dates from the Avalon
terrane in Newfoundland: Canadian Journal of Earth
Sciences, v. 25, p. 442-453.
LEFORT, J. L., 1989, Basement Correlation Across the
North Atlantic: Springer Verlag, Berlin, Germany, 148p.
McLELLAND, J., and Y. W. ISACHSEN, 1986, Synthesis of
geology of the Adirondack Mountains, New York, and their
tectonic setting within the southwestern Grenville Pro-
vince; pp. 31-50 in J. M. Moore, A. Davidson, and A. J. Baer
(eds.), The Grenville Province: Geological Association of
Canada, Special Paper 31,358 p.
__ , J. CHIARENZELLI, P. WHITNEY, and Y. ISACHSEN,
1988, U-Pb zircon geochronology of the Adirondack
Mountains and implications for their geologic evolution:
Geology, v. 10, p. 920-924.
McLEOD, M. J., and N. RAST, 1988, Correlations and fault
systematics in the Passamaquoddy Bay area, southwestern
New Brunswick: Maritime Sediments and Atlantic Geo-
logy, v. 24, p. 289-300.
MATTHEWS, N. A., 1985, A Petrographic and Chemical
Study of a Late Precambrian Dike Swarm in South-
eastern New Brunswick, Canada: MS thesis, Univer-
sity of Kentucky, Lexington, Kentucky, USA, 116 p.
MOORE, J. M., A. DAVIDSON, and A. J. BAER (eds.), 1986,
The Grenville Province: Geological Association of Cana-
da, Special Paper 31, 358 p.
MUEHLBERGER, W. R., 1980, The shape of North America
during the Precambrian; pp. 175-183 in Continental Tec-
tonics: National Academy of Sciences, Washington DC,
USA, 197p.
MURPHY, J. B., J. D. KEPPlE, R. D. NANCE, and J.
DOSTAL, 1989, Reassessment of terranes in the Avalon
composite terrane of Atlantic Canada [abstract]: Geo-
logical Society of America Abstracts with Programs, v. 21,
no. 2,p. 54.
NANCE, R. D., 1987, Model for the Precambrian evolution
of the Avalon terrane in southern New Brunswick, Cana-
da: Geology, v.15, p. 753-756.
Late Precambrian tectonism - The opening of the Iapetus Ocean 405
NIELSEN, T. F. D., 1987, Mafic dyke swarms in Greenland:
a review; pp. 349-360 in H. C. Halls and W. F. Fahrig
(eds.), Mafic Dyke Swarms: Geological Association of
Canada, Special Paper 34, 503 p.
ODOM, A.L., and P. D., FULLAGAR, 1984, Rb-Sr whole-
rock and inherited zircon ages of the plutonic suite of the
Crossnore complex, southern Appalachians, and their
implications regarding the time of opening of the Iapetus
Ocean; pp. 255-280 in M. J. Bartholomew (ed.), The Gren-
ville Event in the Appalachians and Related Topics:
Geological Society of America, Special Paper 194, 287 p.
OLSZEWSKI, W. J., Jr., and H. E., GAUDE'ITE, 1982, Age of
the Brookville gneiss and associated rocks, southeastern
New Brunswick: Canadian Journal of Earth Sciences, v.
19, p. 2158-2166.
PERRIN, M., and M., PREVOT, 1988, Uncertainties about
the Proterozoic and Paleozoic polar wander path of the
West Mrican craton and Gondwana: Evidence for succes-
sive remagnetization events: Earth and Planetary Science
Letters, v. 88, p. 337-347.
PIPER, J. D. A., 1987, Paleomagnetism and the Con-
tinental Crust: John Wiley and Sons, New York, New
York, USA, 434 p.
__ , 1988, Paleomagnetism of (late Vendian-earliest
Cambrian) minor alkaline intrusions, Fen Complex, south-
east Norway: Earth and Planetary Science Letters, v. 90, p.
1325-1330.
PRINGLE,!' R., J. A. MILLER, and D. M. WARRELL, 1971,
Radiometric age determinations from the Long Range
Mountains, Newfoundland: Canadian Journal of Earth
Sciences, v. 8, p. 1325-1330.
RANKIN, D. W., A. A. DRAKE Jr., L. GLOVER III, R.
GoLDSMITH, L. J. HALL, D. P. MURRAY, N. M.
RATCLIFFE, J. F. READ, D. T. SECOR Jr., and R. S.
STANLEY, 1989, Pre-orogenic terranes; pp. 7-100 in R. D.
Hatcher Jr., W. A. Thomas, and G. W. Viele (eds.), The
Geology of North America, Volume F2: The Appala-
chian-Ouachita Orogen in the United States: Geo-
logical Society of America, Boulder, Colorado, USA, 767 p.
RAST, N., 1979, Precambrian meta-diabases of southern
New Brunswick: The opening of the Iapetus Ocean?: Tec-
tonophysics, v. 59, p.127-137.
__ , 1989, The evolution of the Appalachian Chain; pp.
323-348 in A. W. Bally and A. R. Palmer (eds.), The Geo-
logy of North America, Volume A: An Overview: Geo-
logical Society of America, Boulder, Colorado, USA, 619 p.
__ , and J. W. SKEHAN, 1983, The evolution of the
Avalonian plate: Tectonophysics, v. 100, p. 257-286.
__ , and J. W. SKEHAN, SJ, 1991, Tectonic relation-
ships of Carboniferous and Precambrian rocks in south-
westwestern New Brunswick; pp. 209-231 in A. Ludman
(ed.), Geology of the Coastal Lithotectonic Block and
Neighboring Terranes, Eastern Maine and Southern
New Brunswick: Proceedings of the B3rd Annual New
England Intercollegiate Geological Conference:
Princeton, Maine, USA, 400 p.
__ , B. H. O'BRIEN, and R. J. WARDLE, 1976, Rela-
tionships between Precambrian and lower Paleozoic rocks
of the "Avalon Platform" in New Brunswick, the northeast
Appalachians and the British Isles: Tectonophysics, v. 30,
p.315-338.
RATCLIFFE, N. M., W. C. BURTON,J. N. SU'ITER, andS.A.
MUKASA, 1988, Stratigraphy, structural geology, and
thermochronology of the northern Berkshire Massif and
the southern Green Mountains, Part I - Pittsfield, MA, to
Stamford, VT; pp. 1-31 in W. A. Bothner (ed.), Guidebook
for Field Trips in Southwestern New Hampshire,
Southeastern Vermont, and North-Central Massachu-
setts: BOth Annual New England Intercollegiate Geo-
logical Conference, 372 p.
RIVERS, T., and E. H. CHOWN, 1986, The Grenville orogen
in eastern Quebec and western Labrador - Definition,
identification and tectonometamorphic relationships of
autochthonons, parantochthonous and allochthonous ter-
ranes; pp. 31-50 in J. M. Moore, A. Davidson, and A. J.
Baer (eds.), The Grenville Province: Geological Associa-
tion of Canada, Special Paper 31,358 p.
__ , J. MARTIGNOLE, A. DAVIDSON, and C. F. GOWER,
1988, On the tectonic analysis of the Grenville Orogen [ab-
stract): Geological Association of Canada, Mineralogical
Association of Canada, and Canadian Society of Petroleum
Geologists Joint Annual Meeting, Program with Abstracts,
v. 13, p. A104.
RODGERS, J., 1971, The Tectonics of the Appalachians:
Wiley Interscience, New York, New York, USA, 271 p.
__ , 1968, The eastern edge of the North American
continent during the Cambrian and Early Ordovician; pp.
141-149 in E-An Zen, W. S. White, J. B. Hadley, and J. B.
Thompson, Jr. (eds.), Studies of Appalachian Geology:
Northern and Maritime (BilUngs Volume): Inter-
science, New York, New York, USA, 475 p.
SCHENK, P. E., 1971, Southeastern Atlantic Canada,
northwestern Mrica and continental drift: Canadian
Journal of Earth Sciences, v. 8, p. 1218-125l.
SCHERMERHORN, L. J. G., E. WALLBRECHER, and M.
HUCH, 1986, Der Subduktionskomplex, granitplutonisms
and Schertektonik in grundgebirge des Sirwa-Doms (Anti-
Atlas, Marokko): Berliner Geowissenschaft liche Abhand-
lungen, v. 66A, p. 301-322.
SCHWAB, F. L., 1986, Latest Precambrian-earliest Paleo-
zoic sedimentation, Appalachian Blue Ridge and adjacent
areas: Review and speculation; pp. 115-137 in L. Glover III
and R. C. McDowell (eds.), Studies in Appalachian Geo-
406 N. Rast
logy (Lowry Volume): Virginia Polytechnic Institute and
State University, Department of Geological Sciences, Mem-
oir3,137p.
__ , J. P. NYSTUEN, and L. C. GUNDERSON, 1988, Pre-
Arenig evolution of the Appalachian-Caledonide orogen:
Sedimentation and stratigraphy; pp. 75-91 in A. L. Harris
and D. J. Feltes (eds.), The Caledonian-Appalachian
Orogen (IGCP Project 233): Geological Society of Lon-
don, Special Publication 38,643 p.
SECOR, D. T., Jr., S. L. SAMSON, A. W. SNOKE, and A. R.
PALMER, 1983, Confirmation of the Carolina Slate Belt as
an exotic terrane: Science, v. 221, p. 649-650.
SIMPSON, E. L., and K. A. ERIKSSON, 1989, Sedimentology
of the Unicoi Formation in southern and central Virginia:
Evidence for Late Proterozoic to Early Cambrian rift-to-
passive margin transition: Geological Society of America
Bulletin, v. 101, p. 42-54.
SKEHAN, J. W., 1988, Evolution of the Iapetus Ocean and
its borders in pre-Arenig times: A synthesis; pp. 185-229
in A. L. Harris and D. J. Fettes (eds.), The Caledonian-
Appalachian Orogen (IGCP Project 233): Geological
Society of London, Special Publication 38, 643 p.
__ , and D. P. MURRAY, 1980, Geologic profiles across
southeastern New England: Tectonophysics, v. 69, p. 285-
319.
SOCCI, A. D., and G. W. SMITH, 1987, Evolution of the
Boston basin: A sedimentological perspective; pp. 87-100
in C. Beaumont and A. J. Tankard (eds.), Sedimentary
Basins and Basinforming Mechanisms: Canadian So-
ciety of Petroleum Geologists, Memoir 12, 527 p.
STRONG, D. F., S. J. O'BRIEN, S. W. TAYLOR, P. G.
STRONG, and D. H. WILTON, 1978, Geology of the Marys-
town and St_ Lawrence Map Areas, Newfoundland:
Newfoundland Department of Mines and Energy, Report
77-8,33p.
STUKAS, V., and P. H. REYNOLDS, 1974, 4oAr/Ar39 dating
of the Long Range dikes, Newfoundland: Earth and Plane-
tary Science Letters, v. 22, p. 256-266.
TOLLO, R. P., J. N. ALEINIKOFF, and K. J. GRAY, 1991,
New U/Pb zircon isotopic data from the Robertson River
Igneous Suite, Virginia Blue Ridge: Implications for the
duration of Late Proterozoic anorogenic magmatism [ab-
stract]: Geological Society of America Abstracts with
Programs, v. 23, no. 1, p. 139.
WILLIAMS, H., 1964, The Appalachians in northeastern
Newfoundland, a two-sided symmetrical system: Ameri-
can Journal of Science, v. 262, p. 1137-1158.
__ , and R. D. HATCHER Jr., 1983, Appalachian suspect
terranes; pp. 33-54 in R. D. Hatcher Jr., H. Williams, and 1.
Zietz (eds.), Contribution to the Tectonics and Geo-
physics of Mountain Chains: Geological Society of Am-
erica, Memoir 158,223 p.
__ , R. T. GILLESPIE, and O. VAN BREEMAN, 1985, A
late Precambrian rift-related igneous suite in western
Newfoundland: Canadian Journal of Earth Sciences, v. 22,
p. 1727-1735.
__ , M. J. KENNEDY, and E. R. W. NEALE, 1972, The
Appalachians structural province; pp. 181-261 in R. A.
Price and R. J. W. Douglas (eds.), Variations in Tectonic
Styles in Canada: Geological Association of Canada,
Special Paper 11,688 p.
ZARTMAN, R. E., and R. S. NAYLOR, 1984, Structural
implications of some radiometric ages of igneous rocks in
southeastern New England: Geological Society of America
Bulletin, v. 95, p. 522-539.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Transition from alluvial to deep-water sedimentation in the
lower Lynchburg Group (Upper Proterozoic), Virginia*
FREDERICK WEHR
Exxon Production Research Company, P.O. Box 2189, Houston, Texas 77001, USA
(received August 5, 1988; revision accepted February 6, 1989)
*Publication ofthe Orogenic Studies Laboratory, Virginia Polytechnic Institute and State University,
Blacksburg, Virginia, USA
ABSTRACT
Two distinct assemblages of Upper Proterozoic rift-related metamorphic rocks can be delineated in the
southern Appalachian Blue Ridge: (1) areally discontinuous, alluvial and shallow-water clastic and volcanic
deposits to the north and west; and (2) a more continuous belt of deep-water clastic, mafic and ultramafic rocks
to the south and east. An apparently conformable transition between these two assemblages is preserved in
the lower Lynchburg Group in central Virginia.
The lower Lynchburg Group is composed of three lithostratigraphic units: (1) a basal unit of very coarse-
grained, cross-stratified feldspathic arenite; (2) a middle unit of fine- to very fine-grained, thin-bedded
sandstone and laminated argillite; and (3) an upper unit of very coarse-grained, thick-bedded to massive
feldspathic sandstone, pebbly sandstone, and dark argillite. The lower Lynchburg Group is interpreted as a
retrogradational succession of braided-alluvial through delta-front and slope into deep-water fan deposits.
The vertical arrangement and the thick succession of deep-water deposits preserved in the upper unit and
above suggest accumulation near the margin of a rapidly subsiding rift basin.
The transition preserved in these rift deposits may have been controlled by a crustal hinge zone,
separating slightly thinned continental crust to the north and west from highly thinned crust to the south and
east. However, there is no evidence in the northern and central Virginia Blue Ridge for oceanic crust
formation during Late Proterozoic extension. The rift succession in this area therefore appears to be para-
authochthonous with respect to North America.
INTRODUCTION
Among the oldest stratified rocks of the
southern Appalachians are the Upper Proterozoic
clastic and volcanic strata of the Blue Ridge pro-
vince. They nonconformably overlie 1.0 Ga (Gren-
ville) basement and are thought to reflect crustal
extension that preceded formation of an early Paleo-
zoic passive margin (Brown, 1970; Hatcher, 1972,
1978; Rodgers, 1972; Rankin, 1975; Wehr and
Glover, 1985; Schwab, 1986). Paleozoic metamor-
phism and deformation have obscured the original
relationships among these nonfossiliferous rocks;
consequently, their sedimentology and stratigraphy
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 407-423. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
407
408 F.Wehr
are poorly known. However, sedimentologic inves-
tigations of these rocks can yield significant results
in unravelling the earliest history of the southern
Appalachian orogen and in understanding sedimen-
tation patterns along rifted margins.
The focus of this paper is the relationship
between alluvial and deep-water deposits within the
Upper Proterozoic Lynchburg Group near Culpeper,
Virginia, =:100 km southwest of Washington, DC.
The lower three formations of the Lynchburg Group
are interpreted as transitional from alluvial
through delta front-slope into submarine fan envi-
ronments. This transition provides an important
link between two contrasting types of Upper Pro-
terozoic rift deposits that occur throughout the
central and southern Blue Ridge (Schwab, 1986). It
also indicates deposition in a rift system formed
adjacent to now exposed Grenville basement and
attached to the ancestral North America craton.
GEOLOGIC SETTING
In northern Virginia, the Blue Ridge geologic
province is a NE-trending anticlinorium cored by 1
Ga basement rocks (Figure 1). Upper Proterozoic
sedimentary and volcanic rocks occur nonconform-
ably on both limbs and in a narrow belt within the
core of the anticlinorium <Wehr and Glover, 1985).
On the northwest limb, basement is overlain by up
to 3 km of dominantly mafic volcanic rock (Catoctin
Formation). Locally, lenses of metamorphosed clas-
tic rocks as much as 450 m thick (Swift Run
Formation) occur at the base of the Catoctin. In the
core of the Blue Ridge is a 1-3 km wide belt of
metasedimentary rocks (Mechum River Formation)
that extends =: 100 km along strike. The thickest
accumulations of Upper Proterozoic strata occur
beneath the Catoctin on the southeast limb of the
Blue Ridge. In northern Virginia, they are called
the Fauquier Formation (Espenshade and Clarke,
1976) and consist of metamorphosed feldspathic
sandstone, conglomerate, argillite, and minor
carbonate. The Fauquier Formation interfingers
southward with a much thicker succession of
metamorphosed wacke, argillite, and conglomerate
called the Lynchburg Formation (Jonas, 1927;
Brown, 1958; Allen, 1963) or Group (Furcron, 1969;
Wehr, 1983, 1985).
Upper Proterozoic rocks to the north and west
appear to record mostly alluvial and shallow water
deposition. The Swift Run Formation is interpreted
as a valley-fill sequence deposited on a surface of
considerable relief (King, 1950; Reed, 1969; Gath-
right, 1976). Alluvial deposition was also suggested
for the Mechum River Formation (Schwab, 1974),
although Brown (1973) considered it to be a western
inlier of Lynchburg Group rocks. In addition, non-
marine deposition was suggested for the Fauquier
Formation on the basis of abundance of cross-
stratification (Espenshade and Clarke, 1976). All of
the above units differ markedly from the Lynchburg
Group to the southeast, which consists of a thick
succession of deep-water, resedimented deposits
(Brown, 1970; Wehr, 1981, 1983, 1985; Schwab,
1982).
The pattern of thin, shallow-water deposits to
the west and thick, deep-water deposits to the east
persists throughout the southern Appalachian Blue
Ridge <Wehr and Glover, 1985; Schwab, 1986), but
nearly everywhere the two sequences are separated
by thrust faults. Thus, the transition preserved in
the Culpeper area provides a sedimentary link of
regional significance to Appalachian geology.
STRATIGRAPHY
The Lynchburg Group in the Culpeper area is
divisible into six formations (Wehr, 1985). The
transition from alluvial to deep-water deposition
occurs among the lower three units: the Bunker
Hill, Monumental Mills, and Thorofare Mountain
Formations. They occupy a steeply SE-dipping to
overturned belt adjacent to basement rocks (Figure
2). Metamorphism is upper greenschist (almandine)
facies, and the rocks were penetratively deformed at
least twice. Outcrops are small and widely scat-
tered, and none of the contacts between formations
are exposed. Despite these limitations, a coherent
lithostratigraphy can be traced throughout the
study area and beyond (Wehr, 1985), and sedimen-
tary structures are remarkably well preserved.
Figure 3 summarizes the lithostratigraphy of
the lower Lynchburg Group. The lowermost unit
(Bunker Hill Formation) consists of coarse-grained,
cross-stratified feldspathic arenite. It is thickest
north of the Hazel River and pinches out to the
south around the nose of a major NE-plunging anti-
cline (Figure 2). Predominance oflarge-scale cross-
stratification, coarse grain size, and arenitic texture
suggest an alluvial origin for the Bunker Hill For-
mation. Sparse paleocurrent data suggest domin-
ance of NNW -transport during Bunker Hill depo-
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 409
Figure 1. Geologic map of the Blue Ridge of northern
Virginia: Be, Grenville basement complex and Upper
Proterozoic intrusives, undivided; LY. Lynchburg Group
(includes "Fauquier formation" in northern Virginia);
MR, Mechums River Formation; CV, Catoctin and Swift
Run Formations (both Upper Proterozoic); EV, Evington
Group; CH, Chilhowee Group (Upper Proterozoic to early
Paleozoic); S, sedimentary rocks (Mesozoic).
sition. The Bunker Hill Formation is overlain by
the Monumental Mills Formation, which also
pinches out southward. The Monumental Mills
Formation consists of a lower sandstone member
and an upper laminated-siltstone member. It is
relatively fine grained, contains abundant evidence
of slumping and dewatering, and is interpreted as a
deepening-upward succession of delta front and
slope deposits. Overlying the Monumental Mills
siltstone member is a thick succession of coarse-
grained sandstone, argillite, and conglomerate
called the Thorofare Mountain Formation. It thick-
VA.
MD.
s
:--STUDY
AREA
0 10 20

30
N -
KM
0 10 20
30

MI
ens southward and is interpreted as a resedimented
deep-water deposit.
Sandstones of the lower Lynchburg Group are
uniformly feldspathic, consisting primarily of rutil-
lated blue quartz and perthitic feldspar in a recry-
stallized matrix. Plagioclase is a minor detrital
component but is abundant as metamorphic albite
in the matrix. Coarse fraction petrography of the
Lynchburg Group indicates derivation from a source
terrane very similar to the now exposed Blue Ridge
basement complex. Lynchburg Group petrography
is presented in detail elsewhere (Wehr, 1985).
410
Figure 2. Geologic map of the lower
Lynchburg Group in the Culpeper
area; mapping by Wehr (1983), U.S.
Geological Survey 7.5' Quadrangles:
JEF, J effersonton; CAS, Castleton;
BRA, Brandy Station; BRT, Bright-
wood; CLW, Culpeper West; CLE.
Culpeper East; MAD, Madison Mills;
RAP, Rapidan.
Figure 3. Summary of lithostrati-
graphy for the lower three forma-
tions of the Lynchburg Group in the
Culpeper area. Paleocurrent data
from the Bunker Hill Formation
have been rotated around a regional,
N-plunging fold axis. Thicknesses
are structural.
F. Wehr
THOROFARE MOUNTAIN FM.
MONUMENTAL MILLS FM.
BUNKER HILL FM. JEF
BRA
CLE

- H-
R
1 0
CLW
0
__ -==--===-.. 5 MI
RAP
MASSIVE AND GRADED BEDS OF
FELDSPATHIC WACKE, COARSE -
GRAINED TO PEBBLY; LAMINATED
SANDSTONE-MUDSTONE AND
COBBLE CONGLOMERATE
RI PPLED SANOSTONE, ABUNDANT
LI QUEFACTION AND SLUMP
FEATURES, FINE- TO VERY
TABULAR AND TROUGH CROSS-
STRATIFIED ARENITE, COARSE -
GRAINED TO PEBBLY
BLUE RIDGE BASEMENT COMPLEX
SUBMARINE
FAN
DELTA FRONT
BRAIDED
ALLUVIAL
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 411
BUNKER HILL FORMATION
The Bunker Hill Formation consists of as much
as 1 km of metamorphosed feldspathic arenite and
argillite at the base of the Lynchburg Group, resting
nonconformably on granitic basement rocks. It is
dominated by coarse-grained, poorly sorted sand-
stone containing trough and tabular cross-strati-
fication, with minor interbedded fine-grained
sandstone and argillite. The Bunker Hill Formation
is divided into four major lithofacies and is inter-
preted as a braided alluvial deposit.
Planar and trough cross-stratified sandstones
are the two prevalent lithofacies in the Bunker Hill
Formation. Planar sets are lenticular or wedge-
shaped beds 5-100 cm thick, typically 5-20 cm
(Figure 4A). Foresets are defined by crude segre-
gation of coarser and finer material and are com-
monly graded. Reactivation surfaces (Collinson,
1970) are locally well-developed as low-angle,
concave-upward discordance marked by thin, silty
drapes separating packages of fore sets (Figure 4B).
Lenses of trough cross-stratified sandstone as much
as 2 m thick occur interbedded with planar beds
(Figure 4C). Exposures of planar fore sets in the
Bunker Hill Formation are adequate for paleo-
current measurements. A total of 21 fore sets scatter
around a vector mean of 341
0
(Figure 3). Visual
inspection of many outcrops where measurement
was impossible indicates a prevalence of NNW-
transport.
Massive to crudely stratified sandstone is pre-
sent, mainly in the lower parts of the Bunker Hill
Formation (Figure 4D). This lithofacies occurs in
beds 0.5-2 m thick, ranging from fine grained to
pebbly. Stratification is defined by cm-scale grain-
size segregation, with coarse-tail grading present in
some of the thicker strata. The coarsest deposits of
the Bunker Hill Formation are found in this litho-
facies includes scattered pebbles as large as 2 cm.
Laminated fine-grained sandstone and argillite
occur as interbeds within this lithofacies, generally
in discontinuous bedsets < 20 cm thick. Although
planar-parallel and current-ripple laminae have
been recognized, internal structures are generally
obscured by deformation and metamorphism.
The Bunker Hill Formation is interpreted as a
braided alluvial deposit, based on the abundance of
planar and trough cross-stratified sandstone, the
arenitic texture of the sandstones, the general lack
of fine-grained material, and the roughly unimodal
paleocurrent indicators. No turbidites or other fea-
tures indicative of mass-flow deposition were recog-
nized, and the light-colored, arenitic sandstones
contrast markedly with the dark, micaceous wackes
that characterize the Lynchburg Group further up-
section.
The predominance of planar cross-stratification
(facies Sp of Miall, 1977) suggests an alluvial
system dominated by transverse or linguoid bar
forms. Reactivation surfaces reflect partial erosion
and local mud draping of accretion surfaces during
low water. Lenses of trough cross-stratified sand-
stone (facies St of Miall, 1977) are interpreted as
intrabar channel fill. Fine-grained sandstone and
argillite represent low-energy deposition, probably
as infill of minor channels and scour hollows on bar
surfaces during low water (Miall, 1977, p. 36). Mas-
sive to horizontally stratified sandstone (facies Gm
of Miall, 1977).is interpreted as longitudinal bar
deposits, formed by aggradation of gravel sheets
during low water. (Hein and Walker, 1977) The
concentration of this coarse facies in the lower parts
of the Bunker Hill Formation may be significant,
given that longitudinal bars are more characteristic
of the proximal reaches of modern alluvial systems
than linguoid or transverse bars (Smith, 1970;
Boothroyd and Ashley, 1975). With the overall
retrogradational trends in the lower Lynchburg
Group, it seems reasonable that the Bunker Hill
Formation section represents a smaller-scale retro-
gradation from proximal to distal alluvial environ-
ments.
MONUMENTAL MILLS FORMATION
The Monumental Mills Formation comprises a
belt as much as 1.5 km wide between the alluvial
Bunker Hill Formation and the deep-water Thoro-
fare Mountain Formation (Figure 2). It is infor-
mally divided into two members (Wehr, 1985): (1) a
lower sandstone member dominated by fine-
grained, thin-bedded sandstone and argillite, and
(2) an upper siltstone member composed of dark,
thinly laminated siltstone and argillite. The silt-
stone member is almost certainly equivalent to
rocks exposed north of the study area that were
interpreted by Theismeyer (1939) as "varved lacu-
strine deposits." The Monumental Mills Formation
is distinguished by the preponderance of fine-
grained deposits relative to the rest of the Lynch-
burg Group and by the abundance of soft-sediment
412 F.Wehr
Figure 4. Bunker Hill Fonnation lithofacies: A) tabu-
lar cross-stratified bed showing foreset grading; B) re-
activation surface (arrow) within a large cross-stratified
bed, marked by a partially eroded argillaceous drape;
C) trough cross-stratified sandstone; D) crudely hori-
zontally stratified sandstone.
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 413
Figure 5. Monumental Mills Formation, sandstone
member lithofacies: A) thin-bedded sandstone erosional-
ly overlain by current-rippled sandstone; B) massive to
horizontally stratified sand.stone interbedded with dark,
current-rippled argillite; C) polished slab of current-
rippled sandstone showing lamIna set geometry; D) de-
positional-stoss climbing current ripples; E) lens of
coarse-grained sandstone interbedded within thin-
bedded sandstone facies.
414 F. Wehr
defonnation structures. It is interpreted as a delta
front and slope deposit.
Sandstone Member
The sandstone member makes up the bulk of
the Monumental Mills Formation, comprising fine-
to medium-grained, light gray feldspathic sandstone
and dark argillite. Sandstone occurs principally in
thin planar beds 1-4 cm thick (locally up to 40 cm)
separated by 1-2 cm beds of dark, slaty argillite
(Figure 5B). Sandstone beds are sharp-based and
generally massive, although horizontal stratifica-
tion and weak grading are evident locally. Current-
rippled sandstone occurs in lamina sets as thick as
60 cm (Figure 5A,C), locally filling broad, shallow
scours. Climbing ripples are common and are pre-
sent in both erosional-stoss and depositional-stoss
forms (Figure 5D) (Harms and others, 1982). Com-
monly, planar-bedded and current-rippled sand-
stone occur in bedsets a few decimeters thick,
separated by low-angle discordances (Figures 6C
and 7).
2
I I I II
Figure 6. Soft-sediment deformation and liquefaction structures in the Monumental Mills Formation, sandstone member:
A) pillar structures resulting from in situ liquefaction; B) large-scale pillar in thin-bedded sandstone facies; C) discordant
surfaces above large-scale cross-stratified(?) sandstone - note onlap onto upper surface; D) detail of pillar structure from
polished slab of current-rippled sandstone.
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 415
Minor lithofacies in the sandstone member
include large-scale cross-stratified sandstone and
thin lenses of massive, very coarse-grained sand-
stone. Large-scale cross-stratification was recog-
nized in three localities: twice as an isolated, sharp-
ly bounded wedge planar set up to 45 cm thick
(Figure 7c), and once as a l.3-m-thick bedset of
trough cross-strata separated by thin, argillaceous
interbeds. Transport direction for these isolated
beds is to the NNW, similar to the fore sets in the
underlying Bunker Hill Formation. Massive, very
coarse-grained sandstone facies was found in two
outcrops. It fills small lenticular scours as much as
40 cm thick and is interbedded with rippled and
planar-bedded sandstone (Figure 5E).
The sandstone member of the Monumental
Mills Formation contains many features interpreted
as products of deformation during compaction and
dewatering. These features are best developed in
fine-grained sandstone facies, reflecting suscepti-
bility to sediment liquefaction (Lowe, 1975). They
include pillar structures, soft-sediment folds, and
erosional-depositional discordances (Figure 6).
Pillar structures (Lowe and LoPiccolo, 1974) are
irregular zones of disruption that cut bedding at a
high angle. They result from water escape during
compaction and are indicative of relatively high
rates of deposition. Soft-sediment folds are typi-
cally developed as rootless intrafolial isoclines with
amplitudes of a few dm and wavelengths < 10 cm.
They presumably reflect shear strain accommoda-
tion along bedding planes, probably associated with
small-scale slumping. Low-angle erosional-deposi-
Figure 7. Detailed sections through
the Monumental Mills, sandstone
member, illustrating the vertical ar-
rangement of facies. The most con-
sistent facies succession within a single
bed is shown in <D. These deposits are
interpreted as delta-front turbidites.

tiona I discordances are common bedset bounding
surfaces in the sandstone member. They occur as
gently undulating or irregular surfaces with as
much as 1.5 m of relief (Figure 6C).
The origin of these discordances is problematic:
the gently undulating surfaces may relate to de-
formation associated with slumping, with the
discordance representing sediment drape of a slump
scar. More irregular surfaces (Figure 5A) appear to
be cut-and-fill structures, perhaps scoured by non-
depositional sediment gravity-flows and subse-
quently draped by finer grained material.
The sandstone member of the Monumental
Mills Formation is interpreted as a type of delta
front to prodeltaic deposit that accumulated basin-
ward of the Bunker Hill alluvial system. It was
deposited principally from sediment gravity-flows
and suspension settling. It is, however, distinct
from deep-water resedimented deposits found
elsewhere in the Lynchburg Group in that: (1) it is
much finer grained and more thinly bedded; (2) soft-
sediment folds, dewatering structures, and ero-
sional-depositional discordances are pervasive; and
(3) current-ripple lamination is abundant, notably
as high-aggradation climbing ripples (Figure 7).
These features, combined with its position sand-
wiched between two much coarser units interpreted
as alluvial (Bunker Hill) and deep water (Thorofare
Mountain), point to deposition in a transitional
setting bypassed by the coarsest sediment. Coarse
sediment may have been transported across the
delta front in shallow chutes (Buch and Bottje,
1985), which are preserved only as draped low-angle
FACI ES
THINsEOO.O SANOSTONE
o MASSIVE SANOSTONE
RIPPLED SANOSTONE
d
CllM8tNG RIPPLED SANDSTONE
LAMINATED SILTSTONE
b

CROSS.STAATIFIED SANDSTONE
i
e

LIOUEFACTION
I
B
EROSIONAL. SURfACE
a c
416
F.Wehr
discordances and as thin lenses of very coarse-
grained sandstone. The overall lack of coarse ma-
terial in this member suggests efficient transport
across the delta front, perhaps reflecting a relatively
steep slope. Abundant liquefaction and soft-sedi-
ment deformation features provide additional evi-
dence for accumulation in an unstable setting.
Fallout from episodic sediment gravity-flows
appears to be the main mechanism for sandstone
member deposition. Sharp-based, planar-bedded
sandstone beds are interpreted as turbidites. The
presence of grading and horizontal stratification is
indicative of upper flow regime conditions typical of
delta-front sandstones (for example, Cotter, 1975).
Interbedded current-rippled sandstones represent
sediment fallout from lower velocity flows, and the
predominance of climbing ripples indicates rapid
sediment fallout from suspension. Both planar-
bedded and rippled sandstones may represent
deposition from the dilute tails of large, erosive
sediment gravity-flows that deposited the bulk of
their load downslope.
Siltstone Member
The siltstone member of the Monumental Mills
Formation overlies the sandstone member in the
southern half of the Castleton quadrangle (Figure
2). It consists principally of laminated fine-grained
sandstone, siltstone, and argillite. Beds are gen-
erally < 2 cm thick, sharp-based (locally with flame
structures), and consist of very fine-grained sand-
stone or siltstone bases grading up into argillaceous
tops (Figure B). Horizontal stratification and small-
scale starved current ripples are common. The
single most distinguishing characteristic of the silt-
stone member is the abundance of synsedimentary
deformation features, including intrafolial folds,
cm-scale faults, and erosional-depositional discor-
dances (Figure BCE).
Thicker beds of fine- to medium-grained sand-
stone 5-20 cm thick occur in several outcrops of the
siltstone member. These beds are Tabc turbidites:
massive or weakly graded with sharp, erosive bases
and horizontal to current-ripple lamination in the
upper portions (Figure BB).
The siltstone member is interpreted as a slope
deposit, based on the dominance of fine-grained
facies, the prevalence of synsedimentary deforma-
tion features, and its stratigraphic position sand-
wiched between rocks interpreted as delta-front and
deep-water fan deposits. Thin, graded sand-argillite
couplets are the product of low-density turbidity
currents and hemipelagic sedimentation. The per-
vasive deformation features are typical of unstable
slope settings, particularly the erosional-deposi-
tional discordances (cr. Pickering, 1982). The lack of
sandstone in this unit suggests either a sand-poor
depositional system or highly efficient sand trans-
port across the slope. Given the overall coarse-
grained and sand-rich nature of the lower Lynch-
burg Group, efficient sand transport seems a more
likely explanation. In such systems, slope channels
are abruptly abandoned and infilled with fine-
grained sediment (Dott and Bird, 1979).
THOROFARE MOUNTAIN FORMATION
The Thorofare Mountain Formation occurs in a
belt 1-7 km wide in the Culpeper area (Figure 2)
and can be traced southwestward along strike for
over 100 km (Wehr, 1985). The formation consists of
coarse-grained to pebbly feldspathic wacke, argil-
lite, and conglomerate sharply overlying the Monu-
mental Mills Formation. Sandstones are dark gray
in color and are composed of subangular grains of
bluish quartz and feldspar in a recrystallized meta-
morphic matrix (Figure 9A). The Thorofare Moun-
tain Formation is typical of the coarse-grained
wackes that make up the bulk of the lower Lynch-
burg Group in central Virginia (Conley, 1978; Wehr,
1985). It is interpreted as a deep-water fan deposit.
The dominant lithofacies exposed in the Thoro-
fare Mountain Formation is coarse-grained to
pebbly, massive sandstone in beds from 15 cm to >8
m thick (averaging 1 m). Beds range from tabular
to broadly lenticular in shape, locally with deeply
scoured bases (Figure 9B). Massive sandstone beds
are amalgamated or interbedded with thin, discon-
tinuous beds of laminated sandstone and argillite,
which are partially scoured out by overlying sand-
stone beds (Figure 10). Argillite-draped scour
surfaces are present in several outcrops. Horizons
rich in intraclasts of argillite locally define amal-
gamation surfaces. Graded beds are common and
particularly well-developed in beds with deeply
scoured bases. Internally, massive sandstones are
structureless or may contain a crude horizontal
stratification, generally in the basal portions. Rare
lenses of cross-stratified sandstone as much as 10 cm
thick, are interbedded in a few localities.
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 417
Figure 8. Monumental Mills, siltstone member
lithofacies: A) polished slab of laminated siltstone-
argillite facies - light flecks are garnet porphyro-
blasts, darker spots are biotites; B) thin-bedded Tab<:
turbidites, possibly filling an upper slope channel;
C) soft-sediment fold in laminated sandstone-argillite
facies; D) detail of synsedimentary normal fault
draped by overlying beds; E) erosional-deposition.al
discordances (arrow) in laminated siltstone-argillite
facies.
418
F.Wehr
Figure 9. Thorofare Mountain Formation
lithofacies: A) slab of massive sandstone con-
taining intraclasts of laminated argillite, scale
is 1 cm; B) scour surface with coarse-tail grad-
ing at the base of a 1-m-thick bed; C) large
flame (or "flap") of mudstone injected into over-
lying sandstone bed.
Other lithofacies in the Thorofare Mountain
Fonnation include tabular-bedded sandstone,
laminated sandstone-argillite, and conglomerate.
Tabular beds of sandstone as much as 40 cm thick
are interbedded with massive sandstone and argil-
lite. These are "classical" turbidites (facies C of
Walker and Mutti, 1973) - sharp-based, commonly
graded beds with local development of Bouma
sequences Tabe and Tace. Laminated very fine-
grained sandstone and argillite are present but are
poorly exposed and strongly defonned. Where pre-
served, this lithofacies consists of horizontally
stratified or current-rippled sandstone laminae
interbedded with black slaty argillite. Cobble
conglomerate is rare in the Thorofare Mountain
Fonnation, found associated with thick beds of
massive sandstone. It consists of rounded clasts of
granitoid and granitoid gneiss in a coarse-grained
sandstone matrix. Beds are matrix-supported, are
1-2 m thick, and may grade upward into massive
sandstone.
The Thorofare Mountain Fonnation appears to
be a deep-water, resedimented deposit that accumu-
lated from sediment gravity-flows and suspension
sedimentation. The abundance of coarse-grained
sandstone, the presence of cross-stratification and
mud-draped scours, and the tectonic setting are all
consistent with deposition on a sand-rich, "poorly-
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 419
Wl-3
" ,
...
FACIES
(==:J MASSiVE SAHO$TONE
G ll8UlAR-8EDO[D SA"'DSTON'E
FilIPPI..ED SANDSTONE
LAMINATED SANDSTONE-ARGIL.L I TE
_ ARGILL ITE
COvERED
STRUCTURES
I...: .... -I RIPPED-uP CLASTS
1- :::j HORIZONTAL STRATIFI CATION
FLAME STR-ueIURES
5a LOAD CASTING
CONvOLUTE LAMINATION
1- ---- ] AMALGAMATION
CROSS- STRATIFIED LENSES
t,;RAVEL LENSES
MULTIPLE GRADING
Figure 10. Two measured sections through sand-rich Thorofare Mountain exposures, illustrating the dominance of thick-
bedded, erosionally based beds. Section WI-3 (left) is the type section from Thorofare Mountain.
efficient" submarine fan (Mutti, 1979; Nilsen, 1980;
"type II" of Mutti and Normark, 1987). Although
the original thickness of the Thorofare Mountain
Formation is unknown, its regional extent and
structural thickness suggest deposition in a deep,
rapidly subsiding and laterally extensive basin.
Lithofacies associations in the Thorofare Moun-
tain Formation are typical of sand-rich submarine
fan deposits. Coarse-grained massive sandstones
and cobble conglomerates represent deposition from
high-density turbidity currents (Lowe, 1982) -
when a high-density sediment gravity-flow decele-
rapidly and loses its capacity to transport
sediment, and they are characteristic of base-of-
slope settings (for example, Hiscott and Middleton,
1979). Cross-stratified sandstones are similar to
facies reported from other deep-water settings (Cas,
1979; Hiscott and Middleton, 1979; Hein and
Walker, 1982; Mutti, 1979; Mutti and Normark,
1987) and may reflect tractive processes associated
with a zone of "hydraulic jump" at the base of slope.
Mud-draped scours are further evidence for local
erosion and sediment bypass in a hydraulic jump
zone (Mutti and Normark, 1987).
"Poorly efficient" or "type II" submarine fan
systems are found in a number of ancient rift basins
(Surlyk, 1978; Mutti and others, 1981; Stow, 1984;
Devlin,1989). In these settings, the combination of
high relief and a narrow shelf results in rapid intro-
duction of sand into deep water where it accumu-
lates in a thick wedge at or near the base of slope.
DISCUSSION
Depositional Model
The preservation of sedimentary structures in
Lynchburg Group rocks is remarkable and provides
good data for facies interpretation at the outcrop
scale. However, the original depositional strati-
graphy is uncertain because of the structural com-
plexity, widely spaced outcrop, and lack of age
control. The following stratigraphic model assumes
relative conformity of the Bunker Hill-Thorofore
Mountain succession and is based on the following
factors: (1) a logical succession of deepening-upward
depositional environments, (2) a consistent pattern
of sedimentary younging toward the southeast, (3) a
420
F. Wehr
consistent petrographic character (Wehr, 1985), and
(4) the absence of field evidence for major faults or
unconformities within the succession.
Figure 11 shows a cross-section of the central
Virginia Blue Ridge near the end of lower Lynch-
burg deposition. The Bunker Hill Formation ac-
cumulated as braided alluvium along the margins of
a major graben system. It interfingered basinward
with the delta front and slope deposits of the Monu-
mental Mills Formation, which in turn inter-
fingered with the much thicker turbidites and
related deposits of the Thorofare Mountain Forma-
tion. The amount of synchronicity among these
units is conjectural, but the factors listed above are
consistent with accumulation as part of a single
phase of basin fill. High rates of sediment input and
subsidence are indicated by the abundance of coarse,
poorly sorted sandstones and the overall thickness
of the section. Sediment influx may have been aug-
mented by glaciation in adjacent highlands (Wehr,
1986).
It is unknown whether the deep-water facies of
the Lynchburg Group are marine or lacustrine in
origin. The regional extent of Upper Proterozoic
deep-water deposits in the eastern Blue Ridge
require at least a very large lake. Marine conditions
were clearly established in the western Blue Ridge
by Early Cambrian time (Simpson, 1987).
Paleocurrent Data
Paleocurrent data from Bunker Hill alluvial
deposits, although sparse, yield a consistent NNW
NW
NORTHERN
VA
direction of transport. These data conflict with the
regional model (Figure 11), which predicts trans-
port toward the southeast. A number of explana-
tions are possible to explain these data.
First, the Bunker Hill (and perhaps the Monu-
mental Mills) deposits may record a somewhat
earlier phase of basin fill prior to establishment of
dominant subsidence gradients. In this case, the
overlying deep-water deposits are unconformable
upon the Monumental Mills Formation. This
possibility cannot be discounted with the available
data and would require a more complex stratigra-
phic model than shown in Figure 11. Second,
fluvial transport may have been predominantly
parallel to the axis of a NNW-trending graben,
rather than normal as implied (Figure 11). This
possibility still results in transport away from the
main depocenter to the southeast. Third, deposi-
tional gradients may have been controlled by
rotation of listric fault blocks rather than by fault
scarps. Rotation of fault blocks along NE-trending
normal faults could cause regional tilting to the
northwest. This possibility would require that the
entire Lynchburg succession formed on the down-
thrown side of the graben system. It would there-
fore not account for the regional lithofacies
variations in the Virginia Blue Ridge. A more radi-
cal approach is to interpret the entire Lynchburg
succession as allochthonous and accreted to the
North American craton. This possibility is unlikely,
as it requires an unrecognized orogenic event
between Lynchburg deposition and deposition of the
overlying Catoctin Formation, which can be traced
around the nose of the Blue Ridge Anticlinorium
I CRUSTAL 1
- HINGE ZONE--
SE
CULPEPER ROCKFISH LYNCHBURG
AREA RI VER AREA AREA
Figure 11. Generalized model for
lower Lynchburg Group stratigraphy;
BH, Bunker Hill Formation (alluvia!);
" ,
\ -
MM, Monumental Mills Formation (delta front-
slope); TM, Thorofare Mountain Formation (deep-
water fan) . Dark pattern in Lynchburg area represents
mafic igneous rocks intercalated with Lynchburg deep-wa ter
deposits south of Culpeper (after Brown, 1970). Locations labeled
/'
/
/
as above refer to sections across the present-day outcrop belt (see Figure 12).
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 421
(Figure 1). Finally, the data simply may not reflect
regional paleotransport, in view of the small (n=21)
sample population. A more regional study of Upper
Proterozoic fluvial rocks in northern Virginia may
help to clarify this problem.
Regional Paleogeography
In a paleogeographic model for northern and
central Virginia during Lynchburg Group deposi-
tion (Figure 12), the present orientation of the
Lynchburg outcrop belt can be interpreted as a
slightly oblique section through a rifted margin,
separating thinned continental crust to the east
from normal crust to the west (Wehr and Glover,
1985; Wehr, 1985). To the north and west, rela-
tively thin discontinuous alluvial deposits accumu-
lated in isolated grabens, roughly analogous to the
early Mesozoic graben system exposed in the
Appalachian Piedmont. To the south and east,
thicker accumulations of turbidites record a deeper,
rapidly subsiding basin formed on actively thinning
continental crust. Intercalated mafic volcanic rocks
become more abundant toward the southwest
(Brown, 1970; Conley, 1978), perhaps related to
more extensive crustal thinning and possible ocea-
nic crust formation (Hatcher, 1978).
The transition preserved in the Culpeper area is
significant in that it provides a stratigraphic link
across a major paleotectonic hinge zone. This hinge
was a long-lived feature: it marked the eastern
boundary of early Paleozoic miogeoclinal deposits
and, to the southwest, appears to have been re-
activated as a zone of thrusting during the Taconic
orogeny (Wehr and Glover, 1985). Recognition of
stratal continuity across this zone supports a para-
autochthonous model for the deep-water clastic
rocks of latest Proterozoic age in the eastern Blue
Ridge, at least as far south as central Virginia.
ACKNOWLEDGMENTS
This paper is the result of dissertation research
at Virginia Polytechnic Institute and State Univer-
sity. The advice and support of Lynn Glover (chair-
man), Ken Eriksson, and Fred Read are gratefully
acknowledged. This research was supported pri-
marilyby National Science Foundation Grant EAR-
8009549-02 and Nuclear Regulatory Commission
Grant NRC-04-75-237 to Lynn Glover and John
MD.
/
VA.
......
-1'
THIN ALLUVIAL .
DEPOSITS . ..... . .... /
.. .., .. .
.. r: . - ; .
00 0 __
A oc!:-=- .
Yftj:::/: .. : ..
: : . .- - - DEPOSITS
.:. oRe-
.: _ ORIENTATION
_ t- _ - OF EBR OUTCROP BELT
-= = -t; - 0 50 KM

Figure 12. Regional paleogeography of northern Virginia
during Lynchburg time. The present trend of the Blue
Ridge outcrop belt is interpreted as an oblique section
across a zone of crustal thinning. Northward termination
of the rift is after Fisher and others (1979).
Costain. A dissertation support grant from ARCO
provided partial funding for field work. Permission
to publish was granted by Exxon Production Re-
search Company. Esso Australia Ltd provided assis-
tance with drafting and typing. Two anonymous
reviewers improved the manuscript.
REFERENCES
ALLEN, R. M., Jr., 1963, Geology and Mineral Re-
sources of Greene and Madison Counties: Virginia
Division of Mineral Resources, Bulletin 78, 102 p.
BOOTHROYD, J. C., and G. M ASHLEY, 1975, Processes,
geomorphology, and sedimentary structures on braided
outwash fans, northeastern Gulf of Alaska; pp.193-222 in
A. V. Jopling and B. C. McDonald (eds.), Glaciofluvial
and Glaciolacustrine Sedimentation: Society of Econo-
mic Paleontologists and Mineralogists, Special Publication
23,320 p.
BROWN. W. R., 1958, Geology and Mineral Resources of
the Lynchburg Quadrangle, Virginia: Virginia Divi-
sion of Mineral Resources, Bulletin 74,99 p.
422
F. Wehr
__ , 1970, Investigations of the sedimentary record in
the Piedmont and Blue Ridge of Virginia; pp. 335-349 in G.
W. Fischer ,F. J. Pettijohn, J. C. Reed, and K. N. Weaver
(eds.), Studies of Appalachian Geology: Central and
Southern: Wiley-Interscience, New York, New York,
USA,460p.
__ , 1973, Evidence in Virginia for Precambrian
separation of North America from Africa [abstract):
Geological Society of America Abstracts with Programs, v.
5, p. 381-382.
BUCH, S. P., and D. J. BOTI'JE, 1985, Continental slope
deposits from Late Cretaceous tectonically active margin,
southern California: Journal of Sedimentary Petrology, v.
55, p. 843-855.
CAS, R, 1979, Mass-flow arenites from a Paleozoic interarc
basin, New South Wales, Australia: Mode and environ-
ment of emplacement: Journal of Sedimentary Petrology,
v. 49, p. 29-44.
COLLINSON, J. D., 1970, Bedforms of the Tana River,
Norway: Geografiska Annaler, v. 52, p. 31-56.
CONLEY, J. F., 1978, Geology of the Piedmont of Virginia
- Interpretations and problems; pp. 115-149 in Contribu-
tions to Virginia Geology - III: Virginia Division of
Mineral Resources, Publication 7, 154 p.
CO'ITER, E., 1975, Deltaic deposits in the Upper
Cretaceous Ferron Sandstone, Utah; pp. 471-484 in M. L.
Broussard (ed.), Deltas: Models for Exploration: Hous-
ton Geological Society, Houston, Texas, USA, 555 p.
DEVLIN, W. J.,1989, Stratigraphy and sedimentology of
the Hamill Group in the northern Selkirk Mountains
British Columbia: Evidence of latest Proterozoic-earl;
Precambrian extensional tectonism: Canadian Journal of
Earth Sciences, v. 26, p. 515-533.
DOTI', R H., Jr., and K. J. BIRD, 1979, Sand transport
through channels across an Eocene shelf and slope in
southwestern Oregon; pp. 327-342 in L. J. Doyle and O. H.
Pilkey (eds.), Geology of Continental Slopes: Society of
Economic Paleontologists and Mineralogists, Special Pub-
lication 27, 374 p.
ESPENSHADE, G. H., and J. W. CLARKE, 1976, Geology of
the Blue Ridge Anticlinorium in Northern Virginia:
Geological Society of America, Northeast-Southeast Sec-
tions Joint Meeting, Field Trip Guidebook 5, 26 p.
FISHER, G. W., M. W. HIGGINS, and I. ZEITZ, 1979, Geo-
logical Interpretations of Aeromagnetic Maps of the
Crystalline Rocks in the Appalachians, Northern Vir-
ginia to New Jersey: Maryland Geological Survey, Re-
ports of Investigations 32, 43 p.
FURCRON, A. S., 1969, Late Precambrian and early
Paleozoic erosional and depositional sequences of northern
and central Virginia; pp. 57-88 in Precambrian-Paleo-
zoic Appalachian Problems: Georgia Geological Survey,
Bulletin 80, 139 p.
GATHRIGHT, T. M., II, 1976, Geology of the Shenandoah
National Park, Virginia: Virginia Division of Mineral
Resources, Bulletin 86, 93 p.
HARMS, J. C., J. B. SOUTHARD, and R G. WALKER, 1982,
Structures and Sequences in Clastic Rocks: Society of
Economic Paleontologists and Mineralogists, Short Course
9,161 p.
HATCHER, R D., Jr., 1972, Developmental model for the
southern Appalachians: Geological Society of America
Bulletin, v. 83, p. 2735-2760.
__ , 1978, Tectonics of the western Piedmont and Blue
Ridge, southern Appalachians: Review and speculation:
American Journal of Science, v. 278, p. 276-304.
HEIN, F. J., 1982, The Cambro-Ordovician Cap Enrage
Formation, Quebec, Canada: Conglomeratic deposits of a
braided submarine channel with terraces: Sedimentology,
v. 29, p. 309-329.
__ , and R G. WALKER, 1977, Bar evolution and
development of stratification in the gravelly, braided Kick-
ing Horse River, British Columbia: Canadian Journal of
Earth Science, v. 14, p. 562-570.
HISCOTI', R N., and G. V. MIDDLETON, 1979, Depositional
mechanics of thick-bedded sandstones at the base of a
submarine slope, Tourelle Formation (Lower Ordovician),
Quebec, Canada; pp. 307-326 in L. J. Doyle and O. H. Pil-
key (eds.), Geology of Continental Slopes: Society of Eco-
nomic Paleontologists and Mineralogists, Special Publica-
tion 27,374 p.
JONAS, A. 1., 1927, Geologic reconnaissance in the Pied-
mont of Virginia: Geological Society of America Bulletin, v.
38, p. 837-846.
KING, P. B., 1950, Geology of the Elkton Area, Virginia:
U.S. Geological Survey, Professional Paper 230,82 p.
LOWE, D. R, 1975, Water-escape structures in coarse-
grained sediments: Sedimentology. v. 22, p. 157-204.
__ , 1982, Sediment gravity flows, II: Depositional
models with special reference to the deposits of high-
density turbidity currents: Journal of Sedimentary Petr-
ology, v. 52, p. 279-297.
__ , and LOPICCOLO, RD., 1974, The characteristics
and origins of dish and pillar structures: Journal of Sedi-
mentary Petrology, v. 44, p. 484-50l.
MIALL, A. D., 1977, A review of the braided river depo-
sitional environment: Earth-Science Reviews, v. 13, p.1-
62.
MUTI'I, E. 1979, Tubidites et cones sous-marines profonds;
pp. 353-419 in P. Homewood (ed.), Sedimentation Detri-
Transition from alluvial to deep-water sedimentation in the lower Lynchburg Group 423
tique (FluviatUe, Littoralet Marine): Institute Geologie,
Universite de Fribourg, Fribourg, Switzerland, 511 p.
__ , and W. R. NORMARK, 1987, Comparing examples of
modern and ancient turbidite systems: Problems and con-
cepts; pp. 1-38 in J. K. Leggett and G. G. Zuffa (eds.), Mar-
ine Clastic Sedimentology: Concepts and Case Stu-
dies. Grahamm and Trotman, London, England, UK.
__ , C. CAZZOLA, F. FONNESU, G. RAMPONE, M.
SONNINO, and B. VIGNA, 1981, Geometry and facies of
small, fault-controlled deep-sea fan systems in a trans-
gressive depositional setting (Tertiary Piedmont Basin,
northwestern Italy); pp. 1-56 in F. Ricci-Lucchi (ed.), Ex-
cursion Guidebook with Contributions on Sedimento-
logy of some Italian Basins: 2nd European Regional
Meeting of the International Association of Sedimen-
tologists (Bologna, Italy): Oxford University Press,
Oxford, England, UK, 342 p.
NILSEN, T. R., 1980, Modern and ancient submarine fans:
Discussion of papers by R. G. Walker and W. R. Normark:
American Association of Petroleum Geologists Bulletin, v.
64, p. 1094-1099.
PICKERING, K. T., 1982, A Precambrian upper basin-slope
and prodelta in northeast Finnmark, north Norway - A
possible ancient upper continental slope: Journal of Sedi-
mentary Petrology, v. 52, p. 171-186.
RANKIN, D. W., 1975, The continental margin of eastern
North America in the southern Appalachians, the opening
and closing ofthe proto-Atlantic Ocean: American Journal
of Science, v. 273-A, p.298-336.
REED. J. C., Jr., 1969, Ancient lavas in Shenandoah
National Park near Luray, Virginia: U.S. Geological Sur-
vey Bulletin, v. 1265, p. 1-42.
RODGERS, J., 1972, Latest Precambrian (post-Grenville)
rocks ofthe Appalachian region: American Journal of Sci-
ence,v. 272,p. 507-520.
SCHWAB, F. L., 1974, Mechum River Formation: Late Pre-
cambrian(?) alluvium in the Blue Ridge province of Vir-
ginia: Journal of Sedimentary Petrology, v. 44, p. 862-87l.
__ , 1982, Late Precambrian-early Paleozoic sedimen-
tary tectonic framework in the central and southern
Appalachians [abstract]: Geological Society of America
Abstracts with Programs, v. 14, p. 80-8l.
__ , 1986, Latest Precambrian-earliest Cambrian sedi-
mentation, Appalachian Blue Ridge and adjacent areas:
Review and speculation; pp.115-137 in R. C. McDowell and
L. Glover III (eds.), Studies in Appalachian Geology
(Lowry Volume): Virginia Polytechnic Institute and State
University, Department of Geological Sciences, Memoir 3,
137p.
SIMPSON, E. L., 1987, Sedimentology and Tectonic Im-
plications of the Late Proterozoic to Early Cambrian
ChUhowee Group in Southern and Central Virginia:
PhD dissertation, Virginia Polytechnic Institute and State
University, Blacksburg, Virginia, USA, 298 p.
SMITH, N. D., 1970, The braided stream depositional
environment: Comparison of the Platte River with some
Silurian clastic rocks, north-central Appalachians: Geo-
logical Society of America Bulletin, v. 81, p. 2993-3014.
STOW, D. A. V., 1984, Upper Jurassic overlapping-fans
slope-apron system: Brae oilfield, North Sea: Geomarine
Letters, v. 3, no. 4, p. 217-222.
SURLYK, F., 1978, Submarine fan sedimentation along
fault scarps on tilted fault blocks (Jurassic-Cretaceous
boundary, East Greenland): Bulletin Gronlands Geol.
Unders., v. 128, p. 1-108.
THIESMEYER, L. R., 1939, Varved slates in Fauquier
County, Virginia: Virginia Geological Survey, Bulletin 51-
D, p. 105-118.
WALKER, R. G., and MUTTI, E., 1973, Turbidite facies and
facies associations; pp. 119-157 in R. G. Walker and E.
Mutti (eds.), Turbidites and Deep Water Sedimenta-
tion: Society of Economic Paleontologists and Minera-
logists, Pacific Section, Short Course, 157 p.
WEHR, F. L., II, 1981, Coarse-grained, resedimented
deposits of the Lynchburg Group, late Precambrian,
central Virginia [abstract]: Geological Society of America
Abstracts with Programs v. 13, p. 577-578.
__ , 1983, Geology of the Lynchburg Group in the
Culpeper and Rockfish River Areas, Virginia: PhD
dissertation, Virginia Polytechnic Institute and State
University, Blacksburg, Virginia, USA, 254 p.
__ , 1985, Stratigraphy of the Lynchburg Group and
Swift Run Formation, Late Proterozoic (730-570 Ma),
central Virginia: Southeastern Geology, v. 25, p. 225-239.
__ , 1986, A proglacial origin for the Upper Proterozoic
Rockfish Conglomerate, central Virginia, U.S.A.: Pre-
cambrian Research, v. 34, p.157-174.
__ , and L. GLOVER III, 1985, Stratigraphy and tec-
tonics of the Virginia-North Carolina Blue Ridge: Evolu-
tion of a Late Proterozoic-early Paleozoic hinge zone:
Geological Society of America Bulletin, v. 96, p. 285-295.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
The Robertson River Igneous Suite (Blue Ridge Province,
Virginia) - Late Proterozoic anorogenic (A-type)
granitoids of unique petrochemical affinity
RICHARD P. TOLL0
1
AND SARA ARAV2
lDepartmentofGeology, George Washington University, Washington, DC 20052, USA
2U.S. Geological Survey, 5 Aerial Way, Syosset, New York 11791, USA
(received August 5, 1988; revision accepted February 6,1989)
ABSTRACT
Detailed field mapping and petrochemical analysis indicate that the previously defined Robertson River
Formation encompasses a suite of A-type granitoids and felsites that intruded Middle Proterozoic gneisses in
the core of the Blue Ridge anticlinorium in northern Virginia. Constituent granitoids in the northern portion
of the suite include the herein named (from oldest to youngest): Laurel Mills Granite (LMG) , Cobbler
Mountain Alkali Feldspar Quartz Syenite (CMAS), Amissville Alkali Feldspar Granite (AAG), and Battle
Mountain Complex (BMC). The Robertson River Igneous Suite includes both subsolvus (LMG) and hyper-
solvus assemblages, ranges from predominantly metaluminous (LMG, CMAS) to peralkaline (AAG, BMC) in
composition, and displays mineralogic and bulk chemical affinities to A-type granites described from other
localities. The suite exhibits a broad range in GaIAI, Nb, Y, Zr, and Zn that corresponds closely to published
values for the Younger Granites of Nigeria, which have been shown to represent anorogenic magmatic
activity related to extensional tectonics. The extreme enrichment in Ga, Zr, Nb, and Zn in both suites is
probably indicative of a petrologic history involving extensive F complexing, consistent with the presence of
abundant fluorite and F -bearing amphiboles. Marked depletion in Sr and variable RblSr ratios suggest
extensive plagioclase fractionation, which is consistent with the presence of this mineral as a liquidus phase
in the earliest intrusive unit of the suite. Although the generation of A-type melts does not necessarily
indicate an anorogenic or rifting tectonic environment, the unique geochemical characteristics and marked
similarity to the demonstrably anorogenic Younger Granites of Nigeria are consistent with the interpretation
that the Robertson River Igneous Suite records an extensive period of anorogenic magmatism associated with
the early phase of Late Proterozoic rifting of Laurentia.
INTRODUCTION
The Blue Ridge anticlinorium in northern Vir-
ginia (Figure 1) includes a complex series of high-
grade gneissic lithologies that are Middle Protero-
zoic in age and are overlain by a sequence of Late
Proterozoic to Early Cambrian low-grade metasedi-
mentary and metavolcanic rocks (Espanshade,
1970; Bartholomew and Lewis, 1984; Clarke, 1984).
The gneiss assemblage is interpreted to represent
Laurentian basement which experienced intense
Grenville-age orogenesis involving abundant plu-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 425-441. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
425
426
R. P. Tollo and S. Arav
Lower Jurassic
to
Upper Triassic
[]
clastic sequence, basalt, and diabase (rift-
related) : Newark Supergroup
Figure 1. Generalized geologic map
of the Blue Ridge anticlinorium from
the vicinity of Carlisle (CA), Penn-
sylvania, to Charlottesville (CH),
Virginia, modified after Rankin and
others (1989). VA, Virginia; W.VA,
West Virginia; MD, Maryland; PA,
Pennsylvania.
pre-Middle
Ordovician
clast ic sequence and basal t (accreted
terranes) rocks: Jefferson terrane
Middle Ordovician
to
platform/shelf sequence : Canasauga
and Knox Groups
lower Cambrian
lower Cambrian
D
fW
basal clastic seq uence: Chilhowee Group
Upper Proterozoic
1BJ
clastic sequences, basalt, and rhyolite
(rift-rel ated) : Swift Run, Mechum River,
Fauquier, and Catoctin Formations
anorogenic grani te: Robertson River Igneous
Suite
Middle Proterozoic
o
granitiC orthogneiss: Blue Ridge
Basement Complex
tonism and regional metamorphism during the
period 950-1150 Ma (Bartholomew and Lewis,
1984). This basement terrane was subdivided by
Bartholomew and others (1981) into the Pedlar
(west) and Lovingston (east) massifs. The Pedlar
massif is characterized by limited lithologic diver-
sity and high T/high P granulte-facies metamorphic
mineral assemblages, whereas the Lovingston mas-
sif contains a more complex association of rock types
and evidence for high Tllower P granulite-facies
metamorphism (Sinha and Bartholomew, 1984). In
northern Virginia, the Pedlar massif is unconform-
ably overlain by terrestrial clastic deposits of the
Swift Run Formation and geographically extensive
greenstones of the latest Proterozoic Catoctin
Formation, which pinches out toward the southwest
in central Virginia. The Lovingston massif is over-
lain by Catoctin metavolcanic rocks in northern-

o CI ==::::::JI __ 50 mi
most Virginia and Maryland and is overlain farther
south along the eastern flank of the anticlinorium
by a thick sequence of terrestrial to marine clastics
assigned to the Lynchburg Formation (Bloomer and
Werner, 1955; Bartholomew and others, 1981; Bar-
tholomew and Lewis, 1984; Wehr and Glover, 1985).
Relatively undeformed and unmetamorphosed
granitoids of the Robertson River Igneous Suite
intrude gneissic units of the Lovingston massif in
northern Virginia. Field evidence indicating that
Robertson River magmatism was completed prior to
latest Proterozoic time includes greenstone dikes
(which may have served as feeders for the Catoctin
Formation; Clarke, 1984) observed to cut the older,
demonstrably intrusive units of the Robertson River
Suite, and granitic cobbles of probable Robertson
River affinity in the Mechum River Formation
(Schwab, 1974; Hutson and Tollo, 1991a,b).
Robertson River Igneous Suite - Late Proterozoic anorognic granitoids 427
Rankin (1975) noted the occurrence of peralka-
line rhyolite within the dominantly basaltic Catoc-
tin Formation at South Mountain, Pennsylvania
and Maryland, and in the Mt. Rogers Formation in
southern Virginia. The peralkaline rhyolite is simi-
lar to specific units of the Robertson River Igneous
Suite and also to possibly correlative granitoids of
the Crossnore Suite in North Carolina. This basalt-
peralkaline rhyolite petrologic assemblage and the
association with terrestrial clastics led Rankin
(1975, 1976) to suggest that both the volcanics and
compositionally similar granitoids were produced
during Late Proterozoic rifting of Laurentian base-
ment. New geochronological ages of "" 700-760 Ma
(Odom and Fullagar, 1984; Aleinikoff and others,
1991) for these volcanic and intrusive rocks of North
Carolina and adjacent Virginia are consistent with
their formation during the early stage (Badger and
Sinha, 1988; Aleinikoff and others, 1991) of Late
Proterozoic rifting of Laurentian basement. How-
ever, new data indicating an age of "" 570-600 Ma
for the Catoctin Formation (Badger and Sinha,
1988; Aleinikoff and others, 1991) and an age of
"" 700-730 Ma (Tollo and others, 1991; Tollo and
Aleinikoff, 1992) for the Robertson River Igneous
Suite indicate that Robertson River magmatism and
peralkaline volcanism of the Catoctin were dia-
chronous, with a significant time interval ("" 100
million years) separating these magmatic events.
This paper presents the results of detailed field
mapping, petrographic description, and geochemical
analysis of the Robertson River Igneous Suite in
northern Virginia. The data indicate that the pre-
viously defined Robertson River Formation (Allen,
1963) encompasses a suite of alkaline granitoids and
felsites representing a magmatic event that occur-
red over a significant interval of time. Trace ele-
ment data from the Robertson River Igneous Suite
exhibit enrichments comparable to those characteri-
zing complexes of the Nigerian Younger Granites.
These data support previously proposed models for
emplacement of the Robertson River Igneous Suite
in a similar anorogenic tectonic regime associated
with continental rifting.
FIELD RELATIONS
Previous mapping (mostly 1:24,000 scale; Allen,
1963; Lukert and Nuckols, 1976; Lukert and Halla-
day, 1980; Clarke, 1984) has established that the
Robertson River Igneous Suite is post-Grenville and
pre-Catoctin in age, but the internal complexities of
this formation have received little attention. Initial
field studies by Lukert and Halladay (1980) sug-
gested that the observed textural and mineralogic
differences within the Robertson River are grada-
tional, leading some authors to suggest that the
suite represents a single pluton (Mose and Nagel,
1984). However, recent field studies (Tollo, 1986;
Wallace and Tollo, 1986; Tollo and Arav, 1987;
Arav, 1989) have demonstrated that the northern
portion of the Robertson River Igneous Suite is
composed of at least four mappable petrologic units
with contrasting mineralogic and petrologic charac-
teristics and discernible mutual field relations.
Rocks assigned to the Robertson River Igneous
Suite define a generally elongate outcrop belt with-
in the core ofthe Blue Ridge anticlinorium in north-
ern Virginia (Figure 1). The suite is bordered by
basement gneisses and associated intrusive units
throughout the area, except where in fault contact
with the Mechum River Formation in the Massies
Corner quadrangle (Lukert and Halladay, 1980),
and is overlain with apparent unconformity by
metasedimentary strata of the Fauquier Formation
in the south (Figure 1). Field evidence for the
intrusive nature of the Robertson River into the
basement terrane includes: (1) undeformed granitic
dikes cutting the country rock, (2) xenoliths of
gneiss included within the granitoids, and (3) the
presence of inliers (possible screens and roof pen-
dants) of gneiss mapped within the principal outcrop
belt of the granite (Lukert and Nuckols, 1976;
Lukert and Banks, 1984; Clarke, 1984; this study).
Constituent granitoids of the Robertson River
Igneous Suite range from coarse-grained, 2-feldspar
granite to hypersolvus felsite and preserve textural
and mineralogic evidence for progressively shal-
lower levels of intrusion. The Laurel Mills Granite
(LMG) occurs along the western margin of the
outcrop belt, except near the northern termination
where it is absent (Figure 2). This granite is the
oldest unit of the suite, as evidenced by dikes of all
other Robertson River lithologies that cut this rock
type. Unique characteristics of the LMG include
consistent coarse grain size, subequal amounts of
two primary feldspars, and the presence of numer-
ous, outcrop-scale, anastomosing ductile deforma-
tion zones. The petrogenetic relation of this
demonstrably older unit to the rest of the suite is
unclear. The aforementioned characteristics argue
for a petrologic history distinct from the younger
granitoids of the suite. However, the presence of
428
R. P. Tollo and S. Arav
u
'0
N
CI) =
.. ...
., CI)
-l'"
= ...

u
o
CI) N
contect. locetion epproxlmete
contect, locet lon inferred
__ D_ feult. locetlon epproximete
D down-feulted block
__ Q __ feult, locetlon inferred
0= down-feulted block
D Swift Run Formetion
Mechum River Formetion
Bettie Mountein Complex
Amissville Alkeli Gronit.
f:=.:;':1 Cobbler Mountein Alkali Syenite
Leurel Mills Grenl te

Robertson River Sui te
-=
i: :;; Blue Ridge Besement Complex (gnei ss end grenitoi d)
.- ...
l: =
...

Figure 2. Generalized geologic map of the
Robertson River Igneous Suite. AG, Ashby
Gap; U, Upperville; FH, Flint Hill; 0, Orlean;
Me, Massies Corner; J, Jeffersonton; C, Castle-
ton.
hastingsitic hornblende, allanite, and fluorite (see
Table 1) indicates a mineralogic similarity to other
lithologic units of the suite and suggests that the
LMG represents an initial phase of the Robertson
River magmatic event.
The Cobbler Mountain Alkali Feldspar Quartz
Syenite (CMAS) occurs along the eastern margin of
AG
o
J
5mi

the Robertson River outcrop belt, except in the
northern terminus where it is the sole lithology
present (Figure 2). Consistent porphyritic textures,
variable grain size, and a high K-feldspar/quartz
ratio (Figure 3) are distinguishing field characteris-
tics of this unit. Quartz syenite dikes and pegmatite
veins of Cobbler Mountain affinity that cut the
Robertson River Igneous Suite - Late Proterozoic anorognic granitoids 429
o AAG
(MAS
A dike
Q plex suggest that these units may represent the
upper crustal roots of a Late Proterozoic peralkaline
volcanic center.
The Battle Mountain Complex (BMC) occupies
an area of approximately 16 km
2
in the Massies Cor-
ner and Castleton quadrangles (Figure 2). Prelimi-
nary mapping indicates that the complex includes
both massive and flow-banded felsite in association
with aegirine-bearing alkali feldspar granite (D.
Hawkins, MS thesis, in preparation). The felsite,
which is younger than the granitic rocks, exhibits a
variety of features consistent with high-level in-
trusion, or possibly extrusion, including: (1) ubiqui-
tous fine grain size, (2) flow banding, (3) autolithic
inclusions, (4) miarolitic cavities (more common in
granite), and (5) lithophysae. These features indi-
cate that the BMC represents the only known
sizable body of rift-related subvolcanic rocks of
Crossnore affinity in the central Appalachians be-
A =-_..L.--____ --'-_____ ----L ____ tween South Mountain (PennsylvanialMaryland)
Figure 3. Modal mineralogy of samples of the Amissville and Mt. Rogers (Virginia). The BMC is associated
Alkali Feldspar Granite (AAG, open circles) and Cobbler with compositionally similar equivalent volcanic
Mountain Alkali Feldspar Quartz Syenite (CMAS, solid rocks of the nearby Mechum River Formation (Hut-
circles) units of the Robertson River Igneous Suite plotted son and Tollo, 1991a,b, 1992).
in terms of quartz-alkali feldspar-plagioclase (QAP). Data
for a sample of a riebeckite-aegirine dike (solid triangle) of
presumed AAG affinity, which intrudes the CMAS in the
Orlean quadrangle, is included for comparison. After
Streckiesen (1973). Fields: 1, alkali feldspar syenite; 2, al-
kali feldspar quartz syenite; 3, alkali feldspar granite;
4, granite.
adjacent gneisses and gneissic inlier preserve other-
wise uncommon evidence for the intrusive nature of
the Robertson River Igneous Suite in the Blue Ridge
anticlinorium core terrane. The relative age of the
CMAS within the suite is bracketed by syenitic
dikes cutting the LMG and by aegirine/ riebeckite
dikes of Amissville Alkali Feldspar Granite affinity
that intrude the CMAS.
The Amissville Alkali Feldspar Granite (AAG)
has been identified primarily in the Orlean and
Massies Corner quadrangles (Figure 2), occurring
in close association with rocks of the Battle Moun-
tain Complex throughout most of the area. This
lithology is characterized by prominent quartz
phenocrysts, consistent grain size throughout the
outcrop belt, variable K-feldspar/plagioclase ratio,
and abundant fluorite. The hypabyssal textures,
modal mineralogy, and spatial occurrence of this
lithology with rocks of the Battle Mountain Com-
PETROGRAPHY
Petrologic characteristics of the constituent
lithologies of the Robertson River Igneous Suite are
summarized in Table 1. Primary igneous textures
are preserved throughout the suite. Textural
evidence of metamorphism is generally limited and
includes fine-grained aggregates of recrystallized
quartz, retrograde assemblages formed after pri-
mary ferromagnesian phases (described below), and
localized recrystallization along ductile deformation
zones.
Distinguishing features of the LMG include
subsolvus feldspar assemblage, lack of pyroxene,
and locally abundant stilpnomelane + biotite as
retrograde reaction products after primary amphi-
bole. Electron microproble anayses (Tollo, unpub.
data) indicate that the amphibole ranges in com-
position from ferro-edenitic hornblende to hasting-
sitic hornblende (nomenclature after Leake, 1978).
The CMAS is characterized by abundant K-feld-
spar as an early liquidus phase, interstitial quartz,
and amphibole showing no clear evidence of retro-
grade reactions. The primary amphibole is consis-
tently hastingsitic in composition (Arav, 1989).
Primary aegirine-augite found in one outcrop pre-
430
R. P. Tollo and S. Arav
Table 1. Petrographic characteristics of Robertson River lithologies in the study area.
LMG CMAS AAG BMC
Rock Type granite alkali feldspar quartz alkali feldspar felsite, alkali feldspar
granite syenite
Texture, Grain Size hypidiomorphic, porphyritic,
coarse medium
Kspar/Plag Ratio 1:1 to 2:1 10:1 to 8:1
granite
porphyritic,
medium
10:1 to 9:1
aphanitic, fine (fel-
site); equigranular,
medium (granite)
10:1 to 9:1
Ferromagnesian ferro-edenitic hastingsite locally riebeckite typically aegirine + biotite
(most units); some
units locally bear
hastingsitic horn-
blende
Phase(s) hornblende to preceded by aegirine- preceded by aegirine
hastingsitic augite
hornblende
Accessory Phases zircon, allanite,
stilpnomelane
zircon, allanite,
fluorite
zircon, fluorite
zircon, fluorite
Distinguishing
Features
ubiquitous; coarse
grain size; 2-feldspar
assemblage; gray-
blue quartz
mesoperthite pheno-
crysts; paucity of
quartz
quartz phenocrysts;
prismatic riebeckite
with associated
aegirine; abundant
fluorite
fine grain size and
local flow-banding
(felsites); low color
index and hypidio-
morphic texture
(granites); abundant
fluorite
LMG, Laurel Mills Granite; CMAS, Cobbler Mountain Alkali Feldspar Quartz Syenite;
AAG, Amissville Alkali Feldspar Granite; BMC, Battle Mountain Complex.
ceded the amphibole in the crystallization history of
that particular specimen.
Petrographic features of the AAG include vari-
able K-feldspar/plagioclase ratio (Figure 3), initial
sodic pyroxene followed by sodic amphibole, and a
texturally diverse population of zircons. Microprobe
analyses (Arav, 1989) indicate that the pyroxene
and amphibole are near-end-member aegirine and
riebeckite, respectively.
The BMC includes multiple lithologies which
exhibit a range of petrographic features. The felsite
units are fine grained to aphanitic and typically
contain aegirine and/or primary biotite and abun-
dant fluorite. The granitic units are equigranular to
porphyritic and include hastingsitic hornblende or
aegirine.
Modal assemblages and compositions of consti-
tuent pyroxenes and amphiboles serve to distin-
guish the different lithologies of the Robertson
River Igneous Suite and provide insight into mag-
matic chemical evolution. The suite can be divided
into two series: (1) an amphibole-bearing, 2-feld-
spar (subsolvus) lithology (LMG), and (2) pyrox-
ene+amphibole and/or biotite, I-feldspar (hyper-
solvus) lithologies (CMAS, AAG, and BMC). The
fact that Ca abundances, normalized to constant
Si0
2
values, are higher for the CMAS than for the
LMG, which contains similar amphiboles, indicates
that the observed differences in the feldspar assem-
blages are not a function of composition. This ar-
gues strongly that the rocks comprising the two
series crystallized under different f
H20
.
Primary pyroxene occurs only in the hyper-
solvus feldspar lithologies (CMAS, AAG, BMC).
Pyroxene was followed by crystallization of am-
phibole, presumably as the water content of the melt
increased in each case. Pyroxene is conspicuously
absent, however, in the subsolvus feldspar assem-
blage of the LMG, consistent with initially more
hydrous conditions during crystallization of the
assemblage. The mineral chemical characteristics
of the constituent pyriboles therefore reflect bulk
compositional differences between the initial melts
(e.g., Si0
2
content, alumina saturation) and sub-
stantiate field-based lithologic distinctions.
Robertson River Igneous Suite - Late Proterozoic anorognic granitoids 431
Near-end-member aegirine and riebeckite occur
in the high Si0
2
, low A1
2
0
3
/(CaO + N a20 + K
2
0)
lithologies (AAG, BMC), whereas aegirine-augite
and hastingsitic amphibole occur only in low Si02,
high A1
2
0
3
/(CaO + N a20 + K
2
0) lithologies (LMG,
CMAS). Extremely fine-grained, acicular riebeck-
ite, locally forming veins crosscutting primary feld-
spar and quartz, occurs in both the CMAS and the
AAG and is clearly subsolidus in origin, contrasting
with the prismatic, medium-grained riebeckite
occurring throughout the AAG. Textural relations
are insufficient, however, to uniquely document the
paragenesis of this coarse-grained riebeckite, which
may also be largely sub solidus in origin.
GEOCHEMISTRY AND PETROGENESIS
Representative geochemical data (Table 2) for
the constituent granitoids of the Robertson River
Igneous Suite define a range of compositional char-
acteristics that supports the field-based lithologic
distinctions and preserve important evidence bear-
ing on the tectonic environment of emplacement.
Silica values for the AAG are predictably high (74-
77 wt%), whereas the CMAS is characterized by a
range (67-73 wt%) of generally lower values. Felsite
of the BMC is similar in major element composition
to the AAG, whereas alkali feldspar granite from
the complex is typically lower in silica (avg 72 wt%).
The available data indicate, however, that both the
felsite and associated granite from the complex are
characterized by a relatively large range in silica
values (69-76 wt%). Limited data for the LMG
indicate relatively lower, more silica
values (69-72 wt%) for this coarse-grained lithologic
unit. All units are metaluminous to peralkaline,
with agpaitic ratios ranging from 0.9 to 1.1. Ag-
paitic ratios < 1.0 for riebeckite-bearing AAG are
inconsistent with the observed mineralogy, sug-
gesting post-crystallization loss of alkali elements.
The hypersolvus lithologies (CMAS, AAG, BMC)
are characterized by high (Na + K)/Ca, low absolute-
Ca abundances, and high Fe/Mg. The subsolvus
LMG is similar to the CMAS in terms of silica
content ("" 70 wt%), high Fe/Mg, and total alkali
elements, suggesting that the Laurel Mills lithology
may share a common petrogenetic history with at
least the Cobbler Mountain portion of the Robertson
River Igneous Suite.
Trace element concentrations for the suite
(Table 2) show a broad range of values, with mar-
ked enrichment in Nb, Zr, Zn, Y, and Ga relative to
most other low-Ca granites and with pronounced
depletion in Ni, V, and Sr. Distinct elemental
trends are not apparent for any of the units. Nb and
Zr concentrations are high in all units (Figure
4a,b), with values ranging by more than an order of
magnitude for specific lithologic units. Both Zn and
Yare variably enriched throughout the suite (Fig-
ure 4c,d), with the highest Zn values in low-silica
lithologies. Ga concentrations, which are extra-
ordinarily high in nearly all samples (30-62 ppm),
represent the most diagnostic geochemical para-
meter of the suite.
The mineralogic and major and trace element
features described above for the Robertson River
Igneous Suite are considered essential characteris-
tics of A-type granites (Loiselle and Wones, 1979;
Pitcher, 1983; Whalen and others, 1987a,b). The
suite displays trace element enrichments compar-
able to those A-type suites desribed by Whalen and
others (1987a,b). However, the typically high Nb
contents, high absolute-Ga concentrations and, par-
ticularly, the high Ga/Al ratios serve to distinguish
the Robertson River Igneous Suite from most well-
characterized A-type suites except for some com-
plexes of the Nigerian Younger Granites (Figures
4a-d and 5).
The unique petrochemical characteristics of the
Robertson River Igneous Suite are indicative of
several important aspects of the petrogenetic
history of the melts which produced these A-type
granites. The late crystallization of hydrous ferro-
magnesian silicates in most of the lithologic units of
the Robertson River Suite (CMAS, AAG, BMC)
suggests that f
H20
of these magmas was initially
relatively low, increasing sufficiently for amphibole
formation only in the latter portion of the crystal-
lization sequences. The presence of ubiquitous
fluorite and fluorine-bearing amphiboles further
suggests that fHF/fH
2
0 was relatively high compared
to other granitic melts, as originally noted by
Loiselle and Wones (1979) for A-type granites in
general. Such volatile conditions may be a conse-
quence of the type of high-temperature partial
melting of previously melt-depleted I-type source
rock proposed for the generation of A-type magmas
by Collins and others (1982).
The fluorine content of a granitic melt coexist-
ing with an aqueous fluid phase may be enhanced by
the relatively low solubility of fluorine in the fluid
(Bards, 1976). One effect of such fluorine in alkali-
rich systems is the formation of ion complexes that
432
R. P. Tollo and S. Arav
Table 2. Average geochemical compositions and range of values for each of the constituent lithologies
of the Robertson River Igneous Suite in the study area compared to average compositions
of felsic 1- and S-type granites from the Lachlan Fold Belt, Australia
(data ofB. W. Chapell as reported in Whalen and others, 1987a).
LMG CMAS AAG
Avg Max Min Avg Max Min Avg Max Min
Si0
2
70.96 72.92 68.67 70.06 72.84 66.89 75.79 76.51 74.16
Ti0
2
0.32 0.38 0.27 0.31 0.41 0.22 0.09 0.18 0.06
Al
2
0
3
14.04 15.03 13.18 14.00 15.02 13.53 11.87 12.81 11.03
Fe203
3.57 4.25 2.99
4.41 6.50 2.91 3.53 5.12 2.60
MnO 0.06 0.08 0.04
0.09 0.14 0.05 0.04 0.06 0.02
MgO 0.20 0.32 0.08 0.07 0.16 0.03 0.05 0.07 0.02
CaO 0.92 1.43 0.57
0.79 1.52 0.23 0.11 0.36 0.01
Na20 4.00 4.27 3.73
5.00 5.19 4.64 4.48 5.14 3.47
K
2
0 5.70 5.94 5.39
5.03 5.45 4.73 4.35 5.62 3.77
P
2
0
5
0.06 0.09 0.02
0.03 0.05 0.01 0.01 0.01 0.01
Total 99.82 99.78 100.30
Samples 4 5 8
Y 75 84 59
115 271 19 51 151 8
Rb 90 106 76
110 164 70 192 309 119
U 2 3 1
2 4 1 2 6 n.d.
Th 11 16 6
8 18 2 6 22 n.d.
Pb 18 20 15 21 63 3 23 111 4
Ga 32 34 30 42 50 38 51 61 47
Nb 45 47 42
67 119 35 78 182 15
Zr 452 573 353
748 1073 213 523 1403 93
Sr 92 102 73
26 63 7 4 7 1
Zn 160 191 100
194 295 129 176 355 77
Ni 7 9 6
7 13 4 5 12 1
Cr 17 21 15
13 16 10 4 12 n.d.
V 5 6 5
1 2 n.d. 1 2 n.d.
Ce 264 284 244
362 707 187 73 130 16
Ba 580 803 415
197 665 45 29 106 11
10
4
x Ga/Al 4.31 4.69 3.94
5.71 6.86 4.75 8.19 10.42 7.06
A.I. 0.91 0.93 0.89
0.98 1.01 0.94 1.02 1.07 0.96
LMG, Laurel Mills Granite, CMAS, Cobbler Mountain Alkali Feldspar Quartz Syenite; AAG, Amissville
Alkali Feldspar Granite; BMC, Battle Mountain Complex; LFB, Lachlan Fold Belt. Total iron expressed as
Fe203; A.I., agpaitic ratio = (Na + K)/AI; n.d., not detected; n.r., not reported.
may prohibit specific cations from entering mineral and Ga in the Robertson River Igneous Suite may,
structures and which, therefore, may result in at least in part, be a result of such complexing. The
marked enrichment of certain components in the concentration of Zr in felsic melts is directly related
melt phase. The high concentrations of Zr, Nb, Zn, to increased (Na20+ K
2
0)/AI
2
0
3
molar ratio (Wat-
Robertson River Igneous Suite - Late Proterozoic anorognic granitoids 433
Table 2
(continued)
BMC - Felsite BMC - Granite LFB-Felsic
Avg Max Min Avg Max Min I-Type S-Type
SiO
z
73.93 76.40 69.71
71.55 75.56 69.15 73.39 73.39
TiO
z
0.21 0.27 0.16
0.27 0.35 0.22 0.26 0.28
Al
2
0
a
12.15 15.27 10.70 13.11 14.33 11.54 13.43 13.45
Fe20a
4.07 4.43 3.43 4.29 4.93 3.05 2.07 2.28
MnO 0.06 0.08 0.03
0.05 0.11 0.02 0.05 0.04
MgO 0.13 0.39 0.05
0.17 0.19 0.16 0.55 0.58
CaO 0.24 0.58 0.01
0.94 1.22 0.68 1.71 1.28
Na20 4.55 6.44 3.20 4.91 5.32 4.37 3.33 2.81
K
2
0 4.28 4.74 3.79
4.54 4.98 4.02 4.13 4.56
P
2
0
5
0.02 0.05 0.01
0.03 0.05 0.01 0.07 0.14
Total 99.63
99.87 98.99 98.81
Samples 9 3 421 205
Y 115 88 61
149 170 128 34 33
Rb 281 489 206 244 274 214 194 277
U 9 13 7
5 5 4 5 6
Th 20 33 17
26 26 26 22 18
Pb 33 51 39
22 25 18 23 28
Ga 53 62 51
48 49 46 16 17
Nb 229 354 22 259 275 232 12 13
Zr 1741 2224 76 1136 1440 860 144 136
Sr 6 12 1
13 26 5 143 81
Zn 200 378 66
391 510 264 35 44
Ni 7 11 1
11 14 7 2 4
Cr 6 11 4
14 15 14 n.r. n.r.
V 1 2 n.d.
3 4 2 22 23
Ce 204 341 136
301 367 255 68 53
Ba 19 59 1
42 122 1 510 388
10
4
X Ga/AI 8.05 9.02 6.39
6.91 7.77 6.47 2.25 2.39
A.!, 1.00 1.09 0.89
0.99 1.00 0.99 0.74 0.71
LMG, Laurel Mills Granite, CMAS, Cobbler Mountain Alkali Feldspar Quartz Syenite; AAG, Amissville
Alkali Feldspar Granite; BMC, Battle Mountain Complex; LFB, Lachlan Fold Belt. Total iron expressed as
FezOa; A.!., agpaitic ratio = (N a + K)/AI; n.d., not detected; n.r., not reported.
son, 1979) and, in fluorine-rich alkaline systems, remained stable in melts of the Robertson River
the formation of alkali-zircono-fluoride complexes Igneous Suite until the fluorine content of the
(e.g., Na2ZrFs or Na2ZrF7) may be enhanced (Col- liquids was reduced by the relatively late-stage
lins and others, 1982). Such complexes could have precipitation of fluorine-bearing riebeckite (AAG),
434 R. P. Tollo and S. Arav
E
Co
E:
..0
z
E
Co
E:
C
N

o.
B
\00
0
0
0 0
"' , 0
-
. -
....
Whal.n and olhers (1987bl
4000
3soo
3000
2soo
E
Co
E: 2000
N
''''''
1000
soo
Whalen and olhers (1987b)
...
)C. )(
+
.
.
+ ..... ... Of> 0
o
10
__ __
2 10
"
12
10000GaiAI
1000
900
800
700
600
E
Co
500
E:
>
400 0
Whalen and others (,g.67b)
300
200
.
0
CO
100
0
C 0
,
0
2 10
"
12
10000GaiAI
Figure 4. Plots of 104 x Ga/AI us Nb (a), Zr (b), Zn (e), and Y (d) for
the Robertson River Igneous Suite, a reference suite of A-type
granites from Newfoundland, New Brunswick, Corsica, and
Australia (Whalen and others, 1987b), the Afu Complex, Nigeria
(lmeokparia, 1982), and the Shira Complex, Nigeria (Whalen and
others, 1987b).
2 10
"
12
10000GaJAI

250
200
150
100
5()
+ +
WI'I31en ar.d OlnerS 11987b)

.... . :
,: .: '. '.
" .. ::
o
o

+
x
,p
o
10
"
10000GaiAI
Robertson River Igneous Suite - BMC
Robertson River Igneous Suite - AAG
Robertson River Igneous Suite - CMAS
Afu - Nigeria
Shira - Nigeria
A-type Reference Suite
12
hastingsitie amphibole (CMAS), biotite (BMC) or,
eventually, fluorite (all). Zinc also forms complexes
with fluorine (e.g., ZnF
2
) at magmatic temperatures
and, in alkaline melts similar to those represented
by the lithologic units of the Robertson River Ig-
neous Suite, such structures may be further stabi-
lized as alkali-zincate-fluoride complexes (Aylett,
1973; Cotton and Wilkinson, 1980; Collins and
Robertson River Igneous Suite -- Late Proterozoic anorognic granitoids 435
70 ,-----------------------------------------____
Figure 5. Plot of l04X GalAI us Ga for the
Robertson River Igneous Suite and A-t ype
granites.
o
o

+
x
Robertson River Igneous Suite - BMC
Robertson River Igneous Suite -AAG
Robertson River Igneous Suite - (MAS
Afu - Nigeria
Shira - Nigeria
A-type Reference Suite
65
60
55
50
45
40
35
30
25
20
15
3 5
[]
o
ot-e +
[]
o
o '"
[]

o
"
"
[]
"
Whalen and ol hers (1967b)
6 8
10000GalAI
+
o
o
x
10 11 12
400 ;-------------------------------------------__
Figure 6. Plot of l04xGalAI us Sr for the
Robertson River Igneous Suite and A-type
granites.
o
o

+
x
Robertson River Igneous Suite - BMC
Robertson River Igneous Suite - AAG
Robertson River Igneous Suite - CMAS
Afu - Nigeria
Shira - Nigeria
A-type Reference Suite
350
300
250

E
c.
e 200
Whalen and olhers (1987b)
....
en
150
100

50
x
I ", '
. ..... :: ," '\j

"
"
others, 1982). Similar complexing of Nb may also
occur under magmatic conditions, but there is no
evidence available to indicate that magmatic com-
2 3 8 10 1 1 12
10000GalAI
plexing is a process by which Y is concentrated
(Cotton and Wilkinson, 1980).
436 R. P. Tollo and S. Arav
Gallium concentrations for the Robertson River
Igneous Suite (Figure 5) are typically higher than
those reported for A-type suites by Whalen and
others (1987a,b), except for specific intrusive units
of the Nigerian Younger Granites. This extreme
enrichment and high GalAl ratio may also result
from fluorine complexing. Both Al and Ga form
complexes with fluorine under magmatic conditions.
However, in magmas characterized by relatively
low fH 0, GaF
6
complexes appear to be more stable
than AIF 6 (see discussion in Collins and others,
1982), thus concentrating Ga in the melt phase of
silicate magmas such as those represented by the
lithologic units of the Robertson River Igneous
Suite. Furthermore, detailed studies of coexisting
phenocrysts and groundmass in volcanic rocks
(Goodman, 1972) indicate that Ga is preferentially
excluded, relative to aluminum, from the tetrahe-
dral sites in anorthitic feldspar. This suggests that
plagioclase fractionation may be an important
mechanism contributing to the characteristically
high GalAl ratio of A-type melts (Whalen and
others, 1987a). The variable RblSr ratios and the
extreme depletion in Sr shown by the Robertson
River Igneous Suite (Table 2) and the inverse
relationship of Sr with GalAl ratio for A-type
granites in general (Figure 6) are consistent with
such a fractionation model. The marked negative
Eu-anomaly characterizing the hypersolvus litho-
logies of the Robertson River Igneous Suite (Table 3
and Figure 7), and A-type granites in general,
further suggests that such fractionation may be of
fundamental importance in the petrogenesis of most
such suites.
TECTONIC IMPLICATONS
The trace element data indicate a strong geo-
chemical correspondence between the Robertson
River Igneous Suite and at least some of the
Nigerian Younger Granites, including high Ga, Nb,
and GalAl ratio (Figures 4a-d and 5). These two
suites, markedly different from previously docu-
mented A-type granitoids from Newfoundland, New
Brunswick, Australia, and Corsica (Whalen and
others, 1987a, b), appear to preserve evidence of
generation in similar tectonic environments and
may form a subset of A-type granitoids.
Both suites were intruded into high-grade base-
ment terranes in extensional tectonic settings. Both
of these anorogenic suites were emplaced prior to
the initiation of associated rifts. Emplacement of
the Robertson River followed the widespread
Grenville orogeny (1100-1000 Ma), occurred over an
extended interval ("" 730-700 Ma; Tollo and others,
Table 3. REE concentrations in selected samples of the Robertson River Igneous Suite.
An analysis of standard SDC-l is presented for comparison with the data of Abbey (1983).
SDC-I SDC-I
RR-85-20
RR-NG-400 RR-SA-74 RR-SA-78 RR-SA-6 RR-85-22 This Abbey
LMG LMG CMAS CMAS AAG BMC Study (1983)
La
236
123 141 128 47 98 42 42
Ce
290
272 325 297 102 131 96 92
Nd
130
135 129 119 66 78 44 38
Sm 26 24 17 22 15 20 8 8
Eu
2
5 1 1 1 1 2 2
Cd
5
5 2 4 3 17 3 3
Tb
3
2 1 2 2 3 1 1
Hb
3
2 1 2 2 3 1
Yb
7
7 3 6 8 10 5 4
Lu
1
1 0.5 1 1 1 1
All data expressed in ppm; LMG, Laurel Mills Granite; CMAS, Cobbler Mountain Alkali Feldspar Quartz Syenite;
AAG, Amissville Alkali Feldspar Granite; BMC, Battle Mountain Complex.
Robertson River Igneous Suite - Late Proterozoic anorognic granitoids 437
1991; Tollo and Aleinikoff, 1992), that is approxi-
mately correlative with emplacement of the Cross-
nore plutonic suite (e.g., Odom and Fullagar, 1984),
and preceded latest ("'" 600 Ma) Late Proterozoic
rifting of Laurentia (Rankin, 1975, 1976; Wehr and
Glover, 1985). The Nigerian Younger Granites
were intruded after the Pan-African orogeny ( "'" 600
Ma; Jacobson and others, 1964; Grant, 1969, 1970),
were emplaced over a time span ranging from "'" 174
to 154 Ma (van Breeman and others, 1975), and
preceded the Early Cretaceous rifting that formed
the present South Atlantic Ocean (Grant, 1971).
Both the Robertson River Igneous Suite and the
Nigerian Younger Granites were emplaced after a
long period (hundreds of millions of years) of crustal
2

u
c
o
.r::.
o
'-....
w
a::
a::
10
Battle Mountain Felsite
Amissville Alkali Granite
Cobbler Mountain Syenite
Laurel Mills Granite
La Ce Pr Nd PmSm Eu Gd Tb Dy Ho Er Tm Yb Lu
Figure 7. Plot of chrondrite-normalized REE data for
representative smaples of the Robertson River Igneous
Suite. Data normalized to the values of Anders and Ebi-
hara (1982).
quiescence, in contrast to A-type granites from
Corsica and Saudi Arabia, where intusion closely
followed the close of a major orogenic event (Bonin
and others, 1978; Harris, 1985). This distinction
was noted by other workers (e.g., Rogers and
Greenberg, 1988), who indicated that the A-type
classification encompasses petrochemical and
tectonic subtypes. Examined in this context, the
Robertson River Igneous Suite and the Nigerian
Younger Granites, both of which intruded intra-
plate continental crust after an extended period of
tectonic quiescence and were associated with large-
scale extension, are typical examples of an an-
orogenic subtype. In contract, A-type suites from
Corsica (Bonin and others, 1978) and Saudi Arabia
(Harris, 1985; Harris and others, 1986) intruded
calcalkaline terranes closely following the termina-
tion of orogenic events and represent examples of a
post-orogenic subtype. Such granites may show
elevated GalAI ratios; however, the data from Cor-
sica (Whalen and others, 1987a; Bonin and others,
1978) suggest that these granites are characterized
by lower absolute-Ga concentrations relative to the
anorogenic subtype. The Silurian-Jurassic Ras ed
Dom ring complex, which intruded basement gneiss
terranes affected by the Pan-African orogeny in
eastern Sudan (O'Halloran, 1985), is characterized
by trace element enrichments (including Ga) com-
parable to the Robertson River Igneous Suite and
the Nigerian Younger Granites; it probably repre-
sents another example of the anorogenic subtype.
However, the lack of evidence for contemporaneous
rifting in this part of the African craton suggests
that broad-scale crustal extension, not necessarily
active rifting, is an important control in the produc-
tion of such anorogenic lithologies .
Although both the Robertson River Igneous
Suite and the Nigerian Younger Granites were
associated with major continential rifting events,
neither was apparently directly, spatially related to
the actual ocean-forming rift. The Nigerian Youn-
ger Granites are located within the Jos plateau
adjacent to the Benue trough. The granites were
emplaced slightly prior to development of the
trough, which represents a mid-Cretaceous failed
arm of the ridge-ridge-transform fault triple
junction that eventually led to separation of Africa
from South America (Grant, 1971). In the central
Appalachians, stratigraphic evidence (Schwab,
1974) indicates that intrusion of the Robertson
River Igneous Suite preceded development of an
encratonic basin(s?) which did not subsequently
438 R. P. Tollo and S. Arav
evolve into the Iapetus seaway (cr Rankin, 1976;
Wehr and Glover, 1985). The southwestward (youn-
ging) progressive sequence of ages for the Nigerian
Younger Granites was interpreted by some authors
(Rhodes, 1971; van Breeman and others, 1975) to
represent evidence for magma generation resulting
from mantle hotspot migration. A similar age
sequence for the Robertson River (Tollo and others,
1991; Tollo and Aleinikoff, 1992) and A-type
granitoids of the Appalachians in general is not
apparent (but see Williams and others, 1985).
Although Late Proterozoic A-type volcanic rocks of
the Appalachians may be associated with triple
junctions (Rankin, 1976), the granitoids of the
Robertson River Igneous Suite show no such geo-
graphic restriction. As a result, the Robertson River
and associated A-type granites of the Appalachians
appear to be related to regional crustal extension
accompanying the early stages of the Late Protero-
zoic rifting of Laurentia, as inherent in the models
of Rankin (1976) and Wehr and Glover (1985). We
suggest that the distinctive compositional charac-
teristics of the Robertson River Igneous Suite and
the Nigerian Younger Granites are a consequence of
similar: (1) deep-seated source materials, (2) exten-
ded time intervals following previous orogenies, and
(3) association with large-scale continental rifting.
Our data support the conclusions of Whalen and
others (1987a) that A-type granites are characteri-
zed by a broad range in trace element composition
and that the GalAI ratio is probably the most sen-
sitive discriminant between these and other types of
granites. These new data further support their ob-
servations that A-type granites have been genera-
ted in a number of diverse tectonic settings and that
the presence of such granites is not a priori evidence
for rifting. However, the trace element signature of
the Robertson River Igneous Suite and the Nigerian
Younger Granites suggests that these represent a
subset of A-type granites produced after an ex-
tended post-orogenic interval and are directly
related to thermal anomalies associated with
penecontemporaneous continental rifting or epeiro-
genic uplift. We suggest that the characteristic
trace element compositions of these granitoids may
be a consequence of a unique combination of tectonic
variables and thus may potentially serve as a
discriminant by which such anorogenic granitoids
occurring in highly deformed basement terranes
may be recognized.
CONCLUSIONS
The data presented in this study indicate the
following:
The Robertson River Igneous Suite includes at
least four lithologic units with distinct field, min-
eralogic, and petrochemical characteristics.
Most melts of the Robertson River Igneous
Suite were initially relatively anhydrous, increased
in f
H20
with progressive crystallization, and main-
tained relatively high fHF/fH
2
0 for at least part of
the crystallization sequence.
The Robertson River Igneous Suite exhibits a
broad range in Zr, Nb, Y, Zn, Ga, and GalAI ratio
and correlates geochemically with A-type granites
compiled by Whalen and others (1987a,b).
Relative to most other A-type granite suites,
the Robertson River shows marked enrichment in
Nb, Ga, and GalAI ratio, displaying a similarity to
data from two complexes of the Nigerian Younger
Granites and the Ras ed Dom ring complex in Su-
dan, among others.
The Robertson River Igneous Suite and Niger-
ian Younger Granites form a subset of A-type gra-
nites emplaced in anorogenic terranes after an
extended post-orogenic interval and associated with
nearly contemporaneous continental rifting or epei-
rogenic uplift.
NOTE IN PROOF
After submission of this paper, the Battle
Mountain Complex was formally designated the
Battle Mountain Alkali Feldspar Granite and
defined as including both alkali feldspar granite and
(younger) subvolcanic felsite (Tollo, in press).
ACKNOWLEDGMENTS
Major and trace element data were obtained by
x-ray fluorescence spectrometry using automated
Siemens equipment at the Analytical Geochemistry
Facility, Department of Geology and Geography at
the University of Massachusetts-Amherst. REE
data were obtained by instrumental neutron acti-
vation analysis performed at the Department of
Geology and Geophysics at Boston College.
Robertson River Igneous Suite - Late Proterozoic anorognic granitoids 439
We would like to acknowledge K. Schultz, M. C.
Gilbert, D. Rankin, and two anonymous reviewers
for useful comments on an earlier draft of this
manuscript. We also acknowledge the George
Washington University Graduate School of Arts and
Sciences and the Department of Geology for finan-
cial support ofthis project. We thank D. Hawkins of
George Washington University for sharing unpub-
lished data from his MS thesis, J. Sparks of the
University of Massachusetts for assistance with
XRF analyses, S. Flagel of Boston College for
performing INAA analyses, and I. Kelly for meticu-
lous editing. Finally, we would like to thank S. J.
Kreitman for her help and support in all aspects of
this project.
REFERENCES
ABBEY, S., 1983, Studies in "Standard Samples" of
Silicate Rocks and Minerals 1969-1982: Geological
SurueyofCanada, Paper 83-15, 114p.
ALEINIKOFF, J. N., R. E. ZARTMAN, D. W. RANKIN, P. T.
LYTILE, W. C. BURTON, and R. C. McDOWELL, 1991, New
U-Pb zircon ages for rhyolite of the Catoctin and Mount
Rogers formations - More evidence for two pulses of
Iapetan rifting in the central and southern Appalachians
[abstract]: Geological Society of America Abstracts with
Programs, v. 23, no. 1, p. 2.
ALLEN, R. M., Jr., 1963, Geology and Mineral Re-
sources of Greene and Madison Counties: Virginia
Division of Mineral Resources, Bulletin 78,98 p.
ANDERS, E., and M. EBIHARA, 1982, Solar-system abun-
dances of the elements: Geochimica et Cosmochimica Acta,
v.46,p.2363-2380.
ARAV, S., 1989, Geology, Geochemistry, and Relative
Age of Two Plutons from the Robertson River Forma-
tion in Northern Virginia: MS thesis, George Washing-
ton Univeristy, Washington, DC, USA, 153 p.
AYLETI, B. J., 1973, Group lIB; pp. 187-328 in A. F.
Trotman-Dickenson (ed.), Comprehensive Inorganic
Chemistry, Volume 3: Pergamon Press, Oxford, England,
UK, 1387p.
BADGER, R. L., and A. K. SINHA, 1988, Age and Sr isotopic
signature of the Catoctin volcanic province: Implications
for subcrustal mantle evolution: Geology, v. 16, no. 8, p.
692-695.
BARTHOLOMEW, M. J., and S. E. LEWIS, 1984, Evolution
of Grenville massifs in the Blue Ridge geologic province,
southern and central Appalachians; pp. 229-254 in M. J.
Bartholomew (ed.), The Grenville Event in the Appala-
chians and Related Topics: Geological Society of Ameri-
ca, Special Paper 194, 287 p.
__ , T. M. GATHRIGHT II, and W. S. HENIKA, 1981, A
tectonic model for the Blue Ridge in central Virginia:
American Journal of Science, v. 281, no. 9, p. 1164-1183.
BONIN, B., C. GRELOU-ORSINI, and Y. VIALETTE, 1978,
Age, origin, and evolution of the anorogenic complex of
Evisa (Corsica): A K-Li-Rb-Sr study: Contributions to
Mineralogy and Petrology, v. 65, p. 425-432.
BLOOMER, R. 0., and H. J. WERNER, 1955, Geology of the
Blue Ridge region in central Virginia: Geological Society
of America Bulletin, v. 66, no. 5, p. 579-606.
CLARKE, J. W., 1984, The core of the Blue Ridge
anticlinorium in northern Virginia; pp. 153-160 in M. J.
Bartholomew (ed.), The Grenville Event in the Appala-
chians and Related Topics: Geological Society of Ameri-
ca, Special Paper 194, 287 p.
CLEMENS, J. D., J. R. HOLLOWAY, and A. J. R. WHITE,
1986, Origin of an A-type granite: Experimental con-
straints: American Mineralogist, v. 71, p. 317-324.
COLLINS, W. J., S. D. BEAMS, A. J. R. WHITE, and B. W.
CHAPELL, 1982, Nature and origin of A-type granites with
particular reference to southeastern Australia: Contribu-
tions to Mineralogy and Petrology, v.80, p. 189-200.
COTION, F. A., and G. WILKINSON, 1980, Advanced
Inorganic Chemistry: John Wiley-Interscience, New
York, New York, USA, 1136 p.
ESPANSHADE, G., 1970, Geology of the northern part of
the Blue Ridge anticlinorium; pp. 199-211 in G. W. Fisher,
F. J. Pettjohn, J. C. Reed, Jr., and K. N. Weaver (eds.),
Studies in Appalachian Geology: Central and South-
ern: Interscience, New York, New York, USA, 460 p.
GRANT, N. K., 1969, The late Precambrian to early Paleo-
zoic Pan-African orogeny in Ghana, Togo, Dahomey, and
Nigeria: Geological Society of America Bulletin, v. 80, p.
45-56.
__ , 1970, Geochronology of Precambrian basement
rocks from Ibadan, southwestern Nigeria: Earth and
Planetary Science Letters, v. 10, no. 1, p. 29-38.
__ , 1971, South Atlantic, Benue trough, and Gulf of
Guinea Cretaceous triple junction: Geological Society of
America Bulletin, v. 82, p. 2295-2298.
GOODMAN, R. J., 1972, The distribution of Ga and Rb in
coexisting groundmass and phenocryst phases of some
basic volcanic rocks: Geochimica et Cosmochimica Acta, v.
36, p. 303-317.
HARDS, N. J., 1976, Distribution of the elements between
the fluid phase and silicate melt phase of granites and
440
R. P. Tallo and S. Arav
nepheline syenites: Progress in Experimental Petrology, v.
3, p. 88-90.
HARRIS, N. B. W., 1985. Alkaline complexes from the Ara-
bian Shield: Journal of African Earth Sciences, v. 3, p. 83-
88.
__ , F. M. H. MARZOUKI, and S. ALI, 1986, The Jabel
Sayid complex, Arabian Shield: Geochemical constraints
on the origin of peralkaline and related granites. Journal
of the Geological Society of London, v. 143, p. 287-295.
HUTSON, F. E., and R. P. TOLLO, 1991a, Relationship of
Late Proterozoic felsic magmatism and terrestrial sedi-
mentation, Blue Ridge Province, Virginia [abstract): Geo-
logical Society of America Abstracts with Programs, v. 23,
no. 1, p.48.
__ , and R. P. TOLLO, 1991b, Reconstruction of Late Pro-
terozoic crustal history using exposed volcanic and meta-
sedimentary units, Blue Ridge anticlinorium, Virginia
[abstract): Geological Society of America Abstracts with
Programs, v. 23, no. 5, p. A69.
__ , and R. P. TOLLO, 1992, Mechum River Formation:
Evidence for the evolution of a Late Proterozoic rift basin,
Blue Ridge anticlinorium, Virginia [abstract): Geological
Society of America Abstracts with Programs, v. 24, no. 2,
p.22.
IMEOKPARIA, E. G., 1982, Geochemistry and relationship
to mineralization of granitic rocks from the Mu Younger
Granitic Complex, central Nigeria: Geological Magazine,
v. 119, p. 39-56.
JACOBSON, R. R. E., W. N. MacLEOD, and R. BLACK,
1958, Ring-Complexes in the Younger Granites of
Northern Nigeria: Geological Society of London, Memoir
1,72p.
__ , N. J. SNELLING, and J. F. TRUSWELL, 1964, Age
determinations in the geology of Nigeria, with special
reference to the Older and Younger Granites: Overseas
Geology and Mineral Resources, v. 9, p. 168-182.
LEAKE, B. E., 1978, Nomenclature of amphiboles: Ameri-
can Mineralogist, v. 63, p. 1023-1052.
LOISELLE, M. C., and D. R. WONES, 1979, Characteristics
and origin of anorogenic granites [abstract): Geological
Society of America Abstracts wtih Programs, v. 11, p. 468.
LUKERT, M. T." and P. O. BANKS, 1984, Geology and age
of the Robertson River Pluton; pp. 161-166 in M. J. Bar-
tholomew (ed.), The GrenviUe Event in the Appala-
chians and Related Topics: Geological Society of Ameri-
ca, Special Paper 194, 287 p.
__ , and C. R. HALLADAY, 1980, Geology of the Mas-
sies Corner Quadrangle, Virginia: Virginia Division of
Mineral Resources, Publication 17 (Geologic Map 197 A),
scale 1:24,000.
__ , and E. B. NUCKOLS III, 1976, Geology of the Lin-
den and Flint Hill Quadrangles, Virginia: Virginia
Division of Mineral Resources, Report of Investigation 44,
83p.
MOSE, D. G., and S. NAGLE, 1984, Rb-Sr age for the
Robertson River pluton in Virginia and its implication on
the age of the Catoctin Formation; pp. 167-173 in M. J.
Bartholomew (ed.), The Grenville Event in the Appala-
chians and Related Topics: Geological Society of Ameri-
ca, Special Paper 194, 287 p.
ODOM, A. L., and P. D. FULLAGAR, 1984, Rb-Sr whole-rock
and inherited zircon ages of the plutonic suite of the
Crossnore Complex, southern Appalachians, and their
implications regarding the time of opening the Iapetus
Ocean; pp. 255-261 in M. J. Bartholomew (ed.), The Gren-
ville Event in the Appalachians and Related Topics:
Geological Society of America, Special Paper 194,287 p.
O'HALLORAN, D. A., 1985, Ras ed Dom migrating ring
complex: A-type granites and syenites from the Bayuda
Desert, Sudan: Journal of African Earth Sciences, v. 3, p.
61-75.
PITCHER, W. S., 1983, Granite type and tectonic environ-
ment: pp. 19-40 in K. HSll (ed.), Mountain Building Pro-
cesses: Academic Press, London, England, UK, 263 p.
RANKIN, D. W., 1975, The continental margin of eastern
North America in the southern Appalachians: The open-
ing and closing of the proto-Atlantic Ocean: American
Journal of Science, v. 275-A, p. 298- 336.
__ , 1976, Appalachian salients and recesses: Late Pre-
cambrian continental breakup and the opening of the Iape-
tus Ocean: Journal of Geophysical Research, v. 81, p. 5605-
5619.
__ , A. A. DRAKE, Jr., and N. M. RATCLIFFE, 1989,
Geologic map of the U.S. Apalachians showing the Lau-
rentian margin and the Taconic orogen; Plate 3 (scale
1:1,500,000) in R. D. Hatcher, Jr., W. A. Thomas, and G. W.
Viele (eds.), The Geology of North America, Volume F2:
The Appalachian-Ouachita Orogen in the United
States: Geological Society of America, Boulder, Colorado,
USA,767p.
RHODES, R. C., 1971, Structural geometry of subvolcanic
ring complexes as related to pre-Cenozoic motions of
continental plates: Tectonophysics, v. 12, p. 111-117.
ROGERS, J. J. W., and J. K. GREENBERG, 1988, Post-
orogenic vs. anorogenic granites: Two lithologically dif-
ferent suites [abstract): Geological Society of America Ab-
stracts with Programs, v. 20, p. A203.
SCHWAB, F. L., 1974, Mechum River Formation: Late Pre-
cambrian(?) alluvium in the Blue Ridge Province of Vir-
ginia: Journal of Sedimentary Petrology, v. 44, p. 862-871.
Robertson River Igneous Suite - Late Proterozoic anorognic granitoids 441
SINHA, A. K., and M. J. BARTHOLOMEW, 1984, Evolution
of the Grenville terrane in the central Virginia Appala-
chians; pp. 175-186 in M. J. Bartholomew (ed.), The Gren-
ville Event in the Appalachians and Related Topics:
Geological Society of America, Special Paper 194, 287 p.
STRECKIESEN, A. L., 1973, Plutonic rocks classification
and nomenclature recommended by the lUGS Subcommis-
sion on the Systematics of Igneous Rocks: Geotimes, Octo-
ber, p. 26-30.
TOLLO, R. P., 1986, Constituent granitoids of the Robert-
son River Formation in the northern virginia Blue Ridge
[abstract]: Geological Society of America Abstracts with
Programs, v. 18, p. 269.
__ (in press), Definition and nomenclature of the
Robertson River Igneous Suite, Blue Ridge Province, Vir-
ginia: U.S. Geological Survey, Stratigraphic Notes, in
press.
__ , and J. N. ALEINIKOFF, 1992, The Robertson River
Igneous Suite, Virginia Blue Ridge: A case study in mul-
tiple-stage magmatism associated with the early stages of
lapetan rifting [abstract]: Geological Society of America
Abstracts with Programs, v. 24, no. 2, p. 70.
__ , and S. ARAV, 1987, Petrochemical comparison of
two plutons from the Robertson River Formation in
northern Virginia [abstract]: Geological Society of Ameri-
caAbstracts with Programs, v. 19, p. 133.
__ , J. N. ALEINIKOFF, and K. J. GRAY, 1991, New UlPb
zircon isotopic data from the Robertson River Igneous
Suite, Virginia Blue Ridge: Implications for the duration
of Late Proterozoic anorogenic magmatism [abstract]:
Geological Society of America Abstracts with Programs, v.
23, no. 1, p. 139.
VAN BREEMAN, 0., J. HUTCHINSON, and P. BOWDEN,
1975, Age and origin ofthe Nigerian Mesozoic granites: A
Rb-Sr study: Contributions to Mineralogy and Petrology, v.
50, p. 157-172.
WALLACE, P. J., and R. P. TOLLO, 1986, The Battle Moun-
tain felsite: Evidence of an anorogenic (A-type) subvol-
canic complex in the core of the Blue Ridge anticlinorium,
northern Virginia [abstract]: EOS, Transactions of the
American Geophysical Union, v. 67, p. 385.
WATSON, E. B., 1979, Zircon saturation in felsic liquids:
Experimental results and applications to trace element
geochemistry: Contributions to Mineralogy and Petrology,
v. 70, p. 407-419.
WEHR, F., and L. GLOVER III, 1985, Stratigraphy and
tectonics of the Virginia-North Carolina Blue Ridge: Evo-
lution of a Late Proterozoic-early Paleozoic hinge zone:
Geological Society of America Bulletin, v. 96, p. 285-295.
WHALEN, J. B., K. L. CURRIE, and B. W. CHAPELL, 1987a,
A-type granites: Geochemical characteristics, discrimina-
tion and petrogenesis: Contributions to Mineralogy and
Petrology, v. 95, p. 407-419.
__ , K. L. CURRIE, and B. W. CHAPELL, 1987b, A-Type
granites: Descriptive and Geochemical Data: Geologi-
cal Survey of Canada, Open-File Report 1411,33 p.
WILLlAMS, H., R. T. GILLESPIE, and O. VAN BREEMAN,
1985, A late Precambrian rift-related igneous suite in
western Newfoundland: Canadian Journal of Earth Sci-
ences,v. 22,p. 1727-1735.
INTERN A TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Structural characteristics of the Late Proterozoic
(post-Grenville) continental margin of the Laurentian craton
MERVIN J. BARTHOLOMEW
Earth Sciences and Resources Institute, University of South Carolina, Columbia, South Carolina 29208, USA
(received December 31, 1988; revision accepted June 1, 1989)
ABSTRACT
Distribution and orientation data of 2676 segments of 2119 Late Proterozoic to Early Cambrian felsic and
mafic dikes were compared with other Late Proterozoic to Early Cambrian features associated with two stages
of Iapetan extension, including: (1) the trend of the Robertson River rift zone; (2) the distribution of granitoid
plutons; (3) the distribution of volcanic fields; (4) trends of previously recognized surface and subsurface
extensional faults; (5) kinematic indicators from previous studies; and (6) some Paleozoic foliation data. This
comparison suggests the following:
Felsic dikes in northern Virginia are spatially related to the N25E-trending rift zone filled with
Robertson River granitoids and have a principal relict trend ofN25E.
Mafic dikes, identified as Catoctin dikes in the same area, have a principal relict orientation ofN35E but
exhibit a wider range of orientations than the dikes associated with the rift - many locally bimodal or
trimodal.
Other mafic dikes in the same area have a N25E principal trend. Previous studies indicate chemical
similarity between these dikes and Catoctin dikes. These other mafic dikes could be either associated with the
earlier stage of extension, when the Robertson River rift zone developed but are, coincidently, from chemically
similar magmas as the Catoctin volcanic rocks, or they could be associated with the younger stage of Catoctin
volcanism but intruded along fractures developed during the earlier stage.
Many amphibolite dikes in the eastern part of the Blue Ridge where Paleozoic metamorphic grade was
higher (commonly garnet) have preferred orientations parallel to regional Paleozoic foliation trends and
generally lack diverse secondary or minor trends. These amphibolites are interpreted to have been partially
transposed (.., 10_20 rotation) during Paleozoic orogenesis, but they could reflect original trends that
coincidentally are parallel to younger foliation.
Overall, preferred relict orientations for all dikes suggest Late Proterozoic-Cambrian extension was
primarily along an axis oriented ..,N55-85"W - nearly perpendicular to documented Late Proterozoic
subsurface faults in the southern Appalachians and to the reconstructed Laurentian margin during Iapetan
extension.
(continued overleaf)
In: Basement Tectonics 8: Characterization and Comparison 'of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 443-467. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
443
444
M. J. Bartholomew
A bstract Continued
Data from uppermost Late Proterozoic (post-Grenville) to Early Cambrian units of the Blue Ridge, when
plotted in possible pre-Paleozoic palinspastic locations, suggest that several centers of felsic and mafic
volcanism/plutonism developed along the southeastern portion of Laurentia during the initial stage of rifting
( "" 760- 650 Ma) associated with Iapetus. In the northern Virginia Blue Ridge, both granitoids of the
Robertson River rift zone and associated dikes were emplaced subparallel to the carton margin, derived in
previous studies, but oblique to the trend of the nonmarine/marine hinge line previously recognized from Late
Proterozoic sedimentation patterns. During the subsequent stage of rifting (= 570-600 Ma), when an
extensive mafic volcanic field (Catoctin Formation) occupied the northern Virginia Blue Ridge, the associated
dike trend was subparallel to the hinge zone. The ten degree difference in dike orientations between the early
and late stages of Iapetan rifting may reflect a similar shift in the local stress-field orientation with time
during the Iapetan extensional event.
INTRODUCTION
Previous work (e.g., Rankin, 1976; Thomas,
1977, 1983) has indicated a relationship between
Appalachian salients and recesses and late Precam-
brian tectonic features. In particular, Rankin (1970,
1975, 1976) recognized the bimodal nature of Late
Proterozoic volcanic and plutonic rocks in the Blue
Ridge geologic province and their genetic relation-
ships to Late Proterozoic rifting. Based on major
and trace element geochemical data (Rankin, 1975)
and the then-known distribution of these units,
Rankin (1976) suggested a tectonic relationship of
the plutonic suite of the Crossnore Complex to
salients during the opening of the Iapetus Ocean.
Among other points, he suggested that the Roanoke
recess was a triple junction whose failed arm was
carried away. At the time of Rankin's (1976) land-
mark paper, neither volcanic nor plutonic rocks of
this age were known within the Roanoke recess.
Although the age of the Crossnore Complex was
originally placed at about 820 Ma (Rankin and
others, 1969), extensive recent work has established
a maximum age range between 760 and 650 Ma for
most granitoid plutons like those of the Crossnore
and Robertson River plutonic suites (Aleinikoff and
others, 1991; Fullagar and Butler, 1980; Herz and
others, 1981; Hudson and Dallmeyer, 1981; Lukert
and Banks, 1984; Mose and Nagel, 1984; Odom and
Fullagar, 1984; Sinha and Bartholomew, 1984; Tollo
and others, 1991; Tollo and Aleinikoff, 1992).
Rankin (1975) designated the Crossnore Com-
plex to include plutonic rocks of both the Bakersville
Gabbro and the Crossnore-type granitoids, in-
cluding peralkaline granites of North Carolina.
Although metagabbros are rare in Virginia, per-
alkaline rocks also make up a small part of the
slightly younger (730-700 Ma) Robertson River
suite (of Virginia) which is predominantly meta-
luminous (Tollo and others, 1991; Tollo and
Aleinikoff, 1992; Tollo and Arav, this volume). The
different types and ages of granitoids rocks of these
suites reflect complex histories with compositional
variations related to source depth, crustal con-
tamination, and/or magma evolution with time
(Rankin and Tollo, 1987; Tollo and others, 1991;
Tollo and Aleinikoff, 1992). Continuing work on the
composition and age of all of the suspected Late
Proterozoic granitoids of the Blue Ridge is likely to
show that they are all part of a supersuite of tem-
poral and geographically related plutons that are
tectonically linked to Iapetan rifting (Rankin, 1975,
1976; Rankin and Tollo, 1987; Tollo and Aleinikoff,
1992). Intrusion of all granitoid plutons (Crossnore-
type) ( = 730-700 Ma) along with extrusion of Mount
Rogers ("" 760 Ma) and Grandfather Mountain vol-
canics, and intrusion of Bakersville Gabbro (= 735
Mal all characterized the earlier stage (= 760-700
Mal of Iapetan extension. Deposition of Lynchburg
synrift deposits (Wehr and Glover, 1985) followed by
extrusion of Catoctin and Chilhowee volcanics,
accompanied by sedimentation, characterized the
later stage (= 600-570 Mal of rifting in the two-
stage model proposed by Badger and Sinha (1988).
Since publication of Rankin's pioneering work,
new mapping in the Roanoke recess has demon-
strated the existence of numerous granitic plutons
whose crosscutting relationships and/or mineral
assemblages indicate that they are probably part of
the Late Proterozoic plutonic suites of granitoids
(Bartholomew, 1981; Bartholomew and Lewis, 1984;
Bartholomew and others, 1992b). Additionally,
some metavolcanic rocks also were mapped strati-
graphically between the basal metasedimentary
Late Proterozoic continental margin of the Laurentian craton 445
rocks of the Weverton Formation (basal Chilhowee
Group) and the underlying Middle Proterozoic
(Grenville) granulite facies gneisses of the basement
(Henika, 1981).
Recognition of a two-stage model of Iapetan
extension (Badger and Sinha, 1988), along with the
following, provides a new framework for evaluation
of the tectonic setting in which Late Proterozoic
igneous rocks may have evolved:
Orientation diagrams (presented herein) for
2676 segments of 2119 Late Proterozoic-Cambrian
dikes from published and unpublished maps of parts
of the Blue Ridge in Virginia and North Carolina, as
well as additional orientation data for 181 dikes
measured by Espenshade (1986).
Recent work on extensional faults related to
Iapetan rifting (Bailey and Simpson, 1991; Favret
and Williams, 1988; Simpson and Kalaghan, 1989).;
The presence of both volcanic and plutonic rocks
in the Roanoke recess.
A Late Proterozoic, palinspastic restoration of
the eastern margin of Laurentia (Bartholomew and
Lewis, 1988, 1992).
LATE PROTEROZOIC TO EARLY CAMBRIAN
DIKE ORIENTATIONS
Bartholomew and Lewis (1984) illustrated the
distribution of the plutonic suites of Late Protero-
zoic granitoids in the Blue Ridge geologic province
at the 1:500,000 scale. Sufficient numbers of dikes
have now been mapped in parts of the Blue Ridge
(Figure 1) to allow preliminary comparison of dike-
orientation data from many areas containing these
granitoids. Selected maps (Figure 1) are presented
to show spatial relationships of different types of
dikes, plutons, and major Grenvillian and Paleozoic
tectonic features in northern Virginia, the Roanoke
recess, the Gossan Lead area of southwestern Vir-
ginia, and the Boone area of North Carolina. Dike-
orientation data are presented (Figures 2 and 3) for
these and other areas in the Virginia Blue Ridge
(Figure 1).
Considerations Used for Analyzing Dike
Orientations
Dikes on many of the published maps (both
1:24,000- and 1:62,500-scale) are commonly shown
at exaggerated width and length. Moreover, dikes
shown on more recent (1:24,000) maps generally are
significantly shorter and have more diverse orienta-
tions than those shown on older (1:62,500) maps,
perhaps indicating that different methods of map-
ping and portraying dikes were utilized for different
map scales. Thus, only the trend was considered
reasonably reliable when measured directly from a
map, and data from areas where the primary source
of data was from an older (1:62,500) map were
plotted as separate diagrams (Figure 2).
Changes in trend of individual dikes were
considered significant if on the order of 5 or more,
and individual segments were measured and recor-
ded separately regardless of length. Conversely,
even long large dikes (generally on older maps) were
recorded by a single measurement if no significant
change in trend occurred. Approximately 20% of the
data (Figure 2) are from multi-segmented dikes. A
total of 2676 segments were measured from 2119
dikes.
Low-grade Paleozoic metamorphism affected all
of the region studied. The western part of the Blue
Ridge was generally at chlorite-to-biotite grade,
whereas the eastern part of the Blue Ridge was
generally at biotite-to-garnet grade. Generally
dikes within the basement massifs are steeply dip-
ping to vertical throughout the lower grade area.
Only where extensive mylonitic gneiss developed
has the dip of dikes been significantly transposed
subparallel to foliation. Five different types of
dikes, which represent variations of age, grade of
Paleozoic metamorphism, and original mineralogy
and chemistry, may be recognized in the field:
(1) light to medium gray, fine-grained felsic dikes;
(2) greenish gray, fine- to medium-grained green-
stones; (3) medium to dark gray, fine to medium-
grained biotite- or actinolite-rich mafic dikes;
(4) black, medium- to coarse-grained Bakersville
metagabbro; and (5) green to black, medium- to
coarse-grained amphibolites.
The amphibolites, of course, are generally found
in the eastern part of the Blue Ridge and, in
particular, were plotted separately (Figures 1 and
2; Table 1) to evaluate effects of Paleozoic meta-
morphism on dike orientations (discussed below).
Greenstone dikes generally contain ab + ch + ep +
act at lower grades and, with an increase in ac-
tinolite content with increasing metamorphic grade,
gradually become actinolite-bearing mafic dikes.
Other mafic dikes at lower greenschist-facies con-
ditions, such as some near Roanoke, Virginia (e.g.,
Bartholomew, 1981), may still consist largely of
446
,--
r
j
.--'-
- -':::::jENNESSEE-'-'--
g,.
M. J. Bartholomew
Figure 1. Map showing distribution of Grenvillian basement massifs (stippled areas) and location of areas, for which dike-
orientation data were obtained, relative to major features of Appalachians. Area designations and corresponding figures:
1, Figure 3; C1, C2, C3, areas of Catoctin dikes (Figure 2A,B,C,D); 2, eastern Greene and Madison Counties (Figure 2L,M);
3, Figure 2Q,R; 4, Lynchburg quadrangle (Figure 2N,P); 5, Figure 4 and Figure 2E,S; 6, Figure 5 and Figure 2F,T; 7, Figure 6
and Figure 2G,H,U; 8, Figure 7 and Figure 2J,K,V. Key: W, Washington, D.C.; RI, Richmond; RO, Roanoke; RA, Raleigh;
BR, Brevard; BZ, Brevard zone; HF, Hayesville fault; BRFS, Blue Ridge fault system; GLF, Gossan Lead fault; FFS, Fries
fault system; SRA, Smith River allochthon; KMB, Kings Mountain Belt; RB, Raleigh Belt; ESB, Eastern Slate Belt.
Figure 2. Dike-orientation diagrams with numbers of dike segments actually plotted indicated by _S, total number of
dikes indicated by (_D), and types of dikes indicated by: (C, Catoctin dikes; MD, other fine-grained (actinolite- or biotite-
bearing) mafic dikes; F, felsic dikes; G, greenstone dikes; AB and AL, amphibolites in Grenvillian basement (AB) and
Lynchburg Formation (AL); BG, Bakersville Gabbro dikes.
A. Catoctin dikes in the Pedlar and Lovingston massifs north of 3045' latitude (Figure 1, area Cl); from Gathright (1976),
Lukert and Nuckols (1976), Lukert and Halladay (1980), and Rader and others (1975).
B. Catoctin dikes in the Pedlar massif between 3845'N and 3820'N latitude (Figure 1, area C2); from Allen (1963), Brent
(1960), Gathright (1976), Lukert (1973), and Reed (1955).
C. Catoctin dikes in the Pedlar massif south of 3820'N latitude but north of the James River (Figure 1, area C3); from
Bailey (1983), Bartholomew (1977), Hudson (1981), Werner (1966) and Bloomer and Werner (1955).
D. Composite diagram of all Catoctin dikes north of the James River (Figure 1, areas Cl, C2, C3).
E. Metagabbros and other mafic dikes (MD, top) and felsic dikes (F, bottom) in the Pedlar massif near Roanoke (Figure 4);
from Bartholomew (1981) and Henika (1981).
Late Proterozoic continental margin of the Laurentian craton 447
A
C1
965 (820)
20
-1..----''---'-' _ J - E
c 0
w-
10
- --'---'---'----' - E
N
12 .
E 10
MO
575 (560)
G
F
165$ (500)
"II' -
J
F
475(430)
N
J 12
6 10

N8$'W
w-
F
G
445 (260)
H
MD
2065 (820)
K
20

30
Figure 2 (continued)
F. Greenstone and other
mafic dikes in the Pedlar
massif within the Radford
quadrangle (Figure 5); from
Bartholomew and others
(1992b).
G. Felsic dikes in the
Watauga massif, Gossan Lead
area (Figure 6); from Stose
and Stose (1957).
H. Mafic dikes in the
Watauga massif, Gossan Lead
area (Figure 6); from Stose
and Stose (1957) and Riecken
(1966).
J. Felsic dikes in the
Watauga massif, Boone area
(Figure 7); from Bartholomew
and Gryta (1980) and
unpublished map of Elk Park
quadrangle by S. E. Lewis and
M. J. Bartholomew.
K. Other mafic dikes in the
Watauga massif, Boone area
(Figure 7); from Bartholomew
and Gryta (1980), Bartholo-
mew (1983a), and unpublished
map of Elk Park quadrangle
by S. E. Lewis and M. J.
Bartholomew.
(continued overleaf)
448
Figure 2 (continued)
L. Amphibolites in the Lynch-
burg Formation, eastern Greene
and Madison Counties (Figure 1,
area 2); from Allen (1963).
M. Amphibolite dikes in the
Lovingston massif, eastern
Greene and Madison Counties
(Figure 1, area 2); from Allen
(1963) and Lukert (1973).
N. Amphibolites in the Lynch-
burg Formation within the
Lynchburg quadrangle (Figure 1,
area 4); from Brown (1958).
P. Amphibolite dikes in the
Lovingston massif within the
Lynchburg quadrangle (Figure 1,
area 4); from Brown (1958).
Q. Other mafic dikes in the
Lovingston massif, Nelson
County (Figure 1, area 3); from
Bailey (1983), Bartholomew
(1977), Evans (1984), and Herz
and Force (1987).
R. Amphibolite dikes in the
Lovingston massif, Nelson
County (Figure 1, area 3); from
Bartholomew (1977), Bloomer
and Werner (1955), and Evans
(1984).
S. Felsic dikes in the Lovings-
ton massif near Roanoke (Figure
4); from Bartholomew (1981).
T. Amphibolite dikes in the
Lovingston massif within the
Radford quadrangle (Figure 5);
from Bartholomew and others
(1992b).
U. Mafic dikes in the Elk River
massif, Gossan Lead area (Figure
6); from Stose and Stose (1957) .
V. Dikes of Bakersville Gabbro
(BG, top) and other mafic dikes
(MD, bottom) in the Elk River
massif, Boone area (Figure 7);
from Bartholomew (1983a),
Bartholomew and Gryta (1980),
and unpublished map of the
Zionville quadrangle by M. J.
Bartholomew and J. R. Wilson.
M. J. Bartholomew
L AL
78S (470)

N
AL
294S
,
100
a
MO
(190)
N/I<I

4_E
0
S
F
(80)
J<,
H

o 2
U
AB N
185 (90)
'14
10
M
AS
84S (580)

P
AS
220S (1920)
N
' 100
R
AS
- E
545 (470
IL
2 12 20
w- - E
0
T
AS
(220)
4
V
BG

:J
MO
(3D)
5
- E
Late Proterozoic continental margin of the Laurentian craton 449
Table 1. Orientation data for Late ProterozoicCambrian dikes.
Area Designation
Principal Trend (Range) Minor
Fig.l Fig. 2 (from Figures 2 and 3) Secondary Trend (Range) Trends
Western Blue Cl ACl N35E
Ridge Massifs C2 BC2 N35E
(Pedlar and C3 CC3 N35E
Watauga) ClC3 DCT N35E
5 EMD N25E
6 FG N85W
7 GF NI5E
7 GMD
N200E
8 JF N45E
8 KMD N45W
Eastern Blue 1 F N25E
Ridge Massifs 1 M N25E
(Lovingston
and Elk River) Eastofl 181MD N25E
2 LAL N35E
2 MAB N45E
4 NAL
N400E
4 PAB N45E
3 QMD
3 RAB N45E
5 SF
6 TAB
N500E
7 UAB N55E
8 VBG N25E
relict minerals and texture. Commonly, mafic dikes
at chloritetobiotite grade contain actinolite and
chlorite at lower grades but many have a significant
biotite content and typically have a well developed
biotite foliation at garnet grade.
It is not currently certain that metamorphic
grade is the sole discriminant among the mafic
dikes (meta diabases and metagabbros, greenstones,
actinolite or biotitebearing mafic dikes, and am
phibolites). Bakersville metagabbro dikes, which
are significantly older than Catoctin dikes, occur in
higher metamorphic zones yet commonly have
abundant original minerals and texture preserved.
Moreover, greenstones and other mafic dikes occur
together spatially in northern and central Virginia
(e.g., Figure 3) and in North Carolina (area 8 on
(NlO500E) N35W
(NI5500E) N5E (NI00W NI5E) N500W
(N25500E) N5E (NI00W N25E) N35W
(NW500E) N5E (NI00WNWE) N400W
(N20500E) N75W
(N70900W) NI5E
(NO400E)
(NO300E)
(N30600E) N35W
(N20500W)
(NI0400E)
(NI0500E)
(N838E) [from Espenshade, 1986]
(N20500E)
(NI0WE)
(N20600E)
(N30600E)
(NO55E)
(N30600E)
(NlO500E)
(N40600E) N55W
(N 50 60
0
E)
N45W
Figure 1), where Bartholomew and others (1983)
illustrated a biotitebearing dike cutting coarse
grained green amphibolite. Because metamor
phism, age, and chemical composition are all rela
tively unknown now, but may be important factors
in future attempts to define groups of related dikes,
I have plotted the data by field rock types and geo
graphic areas (Figure 2).
Early and LateStage Igneous Rocks, Central
and Northern Virginia Blue Ridge
The band across the Blue Ridge geologic
province included in the Massies Corner, Flint Hill,
and Linden quadrangles has been mapped at the
450 M. J. Bartholomew
1:24,000 scale (Lukert and Halladay, 1980; Lukert
and Nuckols, 1976) and represents a particularly
appropriate area (area 1 on Figure 1) in which to
compare distribution and orientation data for differ-
ent types of dikes for inferences regarding the stress
field associated with the two known stages of Late
Proterozoic to Early Cambrian Iapetan extension.
In the Flint Hill region (Figure 3), felsic dikes,
Catoctin greenstone dikes, and other (generally
actinolite-bearing) mafic dikes are all spatially as-
sociated with Robertson River granitoids.
Robertson River Rift Zone. The Robertson
River suite crops out across the Grenvillian
Lovingston massif in the core of the Blue Ridge
anticlinorium in the northern Virginia Blue Ridge
(Figures 1 and 3), and it encompasses the largest
areal extent of Crossnore-type granitoid plutons
Figure 3. Distribution maps and
orientation diagrams offelsic (F),
Catoctin (C), and other mafic (M) dikes
in the northern Virginia Blue Ridge
(from Lukert and Nuckols, 1976;
Lukert and Halladay, 1980). Quad-
rangles: L, Linden; FH, Flint Hill;
MC, Massies Corner; RRS, Robertson
River suite; CG, Chilhowee Group.
Grenvillian rocks are not patterned;
dikes are shown as solid lines.
Me
F
340
(Bartholomew and Lewis, 1984). It may be much
more extensive, but Late Proterozoic Lynchburg
metasedimentary rocks unconformably overlie the
granitoids along the eastern flank of the anticlinor-
ium (Wehr, 1985; Schwab, 1986). The Robertson
River plutons are aligned along a narrow, linear
zone (Tollo and Arav, this volume) that overall
trends"'" N25E, is "'" 120 km long but generally < 5
km wide, and has a length:width ratio >20:1. It
appears to be a large rift zone filled with elongated
granitoid plutons. Moreover, detailed mapping of
both individual plutons, which are generally elon-
gated parallel to the zone (Tollo and Arav, this
volume), and the surrounding country rocks (Allen,
1963; Lukert and Halladay, 1980; Lukert and Nuc-
kols, 1976) depict geometric relationships which
suggest that the overall shape of the zone did not
result from Paleozoic folding or thrusting. Thus,
C
575
(SOD)
w12 B 4
Late Proterozoic continental margin of the Laurentian craton 451
like most dikes, this rift zone likely developed perp-
endicular to the Late Proterozoic axis of extension
which is inferred (from the rift-zone orientation) to
be ""N65W.
Relict Dike Trends. The Catoctin Formation
(and by inference associated greenstone dikes) is
known to be "" 100 millian years younger than the
Robertson River suite (Badger and Sinha, 1988;
Aleinikoff and others, 1991; Tollo and Aleinikoff,
1992). Felsic dikes in the same area (Figure 3) are
more likely related to the older Robertson River
suite. Other mafic dikes, affected by increasing
Paleozoic metamorphic grade, may represent either
stage of intrusion. In northern and central Virginia,
a limited number of analyses suggest that Catoctin
greenstone dikes, other mafic dikes, and amphibo-
lites all have many chemical similarities (e.g., all
are high in P
2
0
5
and Ti0
2
), consistent with an
extensional environment (Reed and Morgan, 1971;
Espenshade, 1986; Ratcliffe, 1987; Bartholomew
and others, 1992a). However, chemical similarity
need not preclude significantly different times of
intrusion (e.g., comparison of Catoctin dikes and
Bakersville metagabbro; Ratcliffe, 1987; Goldberg
and others, 1986).
For the Flint Hill area (Figure 3), three
catagories of dikes - felsic (F), greenstone (G), and
other mafic (M) dikes - are plotted separately to
determine their degree of similarity in geographic
distribution and preferred orientation. All three
types of dikes occur throughout the area (Figure 3),
and all three types intrude the Robertson River
suite of granitoids (""730-700 Ma) but are not
known to intrude the Catoctin Formation or over-
lying Chilhowee Group of Early Cambrian age
(Simpson and Sundberg, 1987). Mafic dikes are 15-
20 times more abundant than either Catoctin or
felsic dikes, and they intrude the Mechum River
metasedimentary rocks, which contain detrital
clasts derived from Robertson River granitoids
(Lukert and Banks, 1984). Thus, crosscutting or
bounding relationships restrict all three types of
dikes to only a Late Proterozoic to Early Cambrian
age range (730-570 Ma).
Orientation data (Figure 3F; Table 1) for felsic
dikes (F) in northern Virginia suggest they are
closely related to the Robertson River suite. These
felsic dikes have a principal trend ofN25E, which is
also the overall trend of the Robertson River rift
zone. Orientation data for other mafic dikes (Fig-
ure 3M) have a greater range of preferred orienta-
tions than that for felsic dikes but, like felsic dikes,
are nearly symmetrical about the N25E Robertson
River trend. Thus, unless Paleozoic metamorphism
and deformation are major factors affecting dike
orientations in this area, both felsic (F) and mafic
(M) dikes were intruded under stress field condi-
tions similar to those existing during development
of the Robertson River rift zone.
In the Rectorville and Marshall quadrangles
just east of this area (in the direction of increasing
metamorphic grade), Espenshade (1986) reported a
principal trend of N25E (Table 1) for 181 meta-
diabase dikes which also have a range of preferred
orientations similar to that of the other mafic dikes
(Figure 3M; Table 1).
Orientation data for Catoctin greenstone dikes
(C) also show a strong NE trend but have a wider
range of preferred orientations. Greenstone dikes
have a principal orientation of N35E - 10 differ-
ent from the N25E trend common to the Robertson
River rift zone, felsic dikes, other mafic dikes in the
Flint Hill area, and metadiabases farther east
(Table 1). Greenstone-dike trends for three areas
(C1, C2, and C3 on Figure 1) were plotted and
compared to determine if this N35E Catoctin trend
may be significant. Catoctin dikes in the Flint Hill
region of northern Virginia (Figure 3) represent
approximately half of the database for Catoctin
dikes in the northernmost Blue Ridge (Figures 1
and 2A) north of latitude 3845'N but only 17% of
the total Catoctin database (Figure 2D).
Orientation data for Catoctin dikes in all three
areas (C1, C2, C3 on Figure 1; Figure 2A,B,C,D)
have a N35E principal trend that is consistently 10
different from the N25E Robertson River rift-zone
trend. Thus the 10 difference between the earlier
rift-zone trend and the later Catoctin greenstone
dikes may reflect a slight shift in regional stress-
field orientation during the "" 100 million years
separating these stages.
Comparison of orientation data of Catoctin
dikes and other mafic dikes is possible for each of
the three Catoctin segments (C1, C2, C3). As stated
previously, mafic dikes in the northernmost seg-
ment (Figure 3M) and farther east (Espenshade,
1986) have a symmetrical distribution about a
N25E principal trend that is parallel to the Robert-
son River rift zone. Both locally and regionally,
Catoctin greenstone dikes have a N35E principal
trend (Figures 2A and 3C). However, in area C1
(Figure 2A) and particularly in area C2 (Figure
2B), the range of preferred orientations is skewed
452 M. J. Bartholomew
toward the older N25E trend. The apparent
influence of the N25E older trend on younger
Catoctin greenstone dikes suggests that fractures
formed earlier may have influenced Catoctin dike
trends as well as the trends of other mafic dikes,
regardless of the latter's age.
Farther south (area C2 on Figure 1; Figure
2B), Catoctin dikes have both a N35E principal
N
37'22'30'
RS
37'15'
trend and a N5E secondary trend. By comparison,
in eastern Greene and Madison Counties, Virginia
(area 2 on Figure 1), amphibolites that intrude both
basement (Figure 2M) and Lynchburg metasedi-
mentary rocks (Figure 2L) in this portion of the
Blue Ridge geologic province have principal trends
of N45E and N35E, respectively. Although their
ranges are skewed somewhat to the northeast,
io
:-
'" 3r lS" ...
"- -<
'>-- -
/
.4
/
,/
/'
/'
A
/
,/
.4
'"
r:o
;;;
37'22'30-
/
.4
/'
Figure 4, Geologic map of Early Cambrian and Late Proterozoic rocks near Roanoke (Figure 1, area 5), modified after
Bartholomew (1981) and Henika (1981). Grenvillian rocks of the Blue Ridge are not patterned; black areas and lines mark
metagabbros and other mafic dikes; dotted lines denote felsic dikes; v, basal volcanic unit of Weverton Formation;
CG, Chilhowee Group. Quadrangles: V, Villamont; M, Montvale; R, Roanoke; S, Stewartsville.
Late Proterozoic continental margin of the Laurentian craton 453
neither of these mafic dike sets show either the
N5E secondary trend or the minor NW trend ex-
hibited by the Catoctin dikes, and they may princi-
pally reflect effects of Paleozoic metamorphism.
From this region southward, more highly meta-
morphosed dikes (amphibolites) are generally trans-
posed subparallel to the regional foliation (as
discussed below), but dikes subjected to lower grades
of metamorphism, at least partially retain relict
orientations inherited from the time of intrusion
during the Late Proterozoic. In the Nelson County
region of Virginia (area C3 on Figure 1; Figure
2C), Catoctin dikes in the Pedlar massif have a
principal trend of N35E - with the preferred range
skewed to the east, a N5E secondary trend, and
minor E and NW trends. In the adjacent Lovingston
massif, amphibolites (Figure 2R) have a
symmetrical orientation about a N45E principal
trend similar to the skewed orientation range of the
Catoctin dikes; this trend probably also reflects the
effects of Paleozoic metamorphism. The limited
number of dikes mapped here (area 3 on Figure 1)
as "other" (actinolite- and biotite-bearing) mafic
dikes (Figure 2Q) partially reflects the foliation
trend as well. However, this minor dike set also has
a relict N5E trend similar to the nearby Catoctin
dikes (Figure 2C).
Similar inferences appear valid farther south.
About 40-60 km west of Lynchburg, near Roanoke
(area 5 on Figure 1; Figure 4), numerous metamor-
phosed fine-grained mafic dikes, as well as a few
coarse-grained metagabbros, a metadiorite, and
several fine-grained felsic dikes, intruded the
Pedlar massif. These mafic dikes (Figure 2E, top)
have a principal N25E trend, with the distribution
slightly skewed more easterly. The few felsic dikes
in this area (Figure 2E, bottom) have a similar
range. Likewise, minor Nand NW mafic dike
trends (Figure 2E) also are interpreted as reflecting
relict Late Proterozoic trends similar to secondary
and minor dike trends farther north (Figures 1 and
2; Table 1). Within the structurally superjacent
Lovingston massif, the trend of granitic bodies pro-
bably reflects superimposed Paleozoic trends, but
the few felsic dikes (Figure 2S) retain a relict N-NE
orientation similar to the mafic dikes (Figure 2E)
in the adjacent Pedlar massif.
Southwest of Roanoke (area 6 on Figure 1; Fig-
ure 5), the N85W principal trend (Figure 2F) is
similar to the minor N75W trend of mafic dikes just
northeast of Roanoke (Figure 2E) but both are are
significantly different from other relict trends. A
minor NI5E trend is present here as well. The only
mafic volcanics found near these NW -trending dikes
of the Pedlar massif are a few flows in the basal
Unicoi Formation (Figure 5). Such orientations
could reflect stress-field changes with time (i.e.,
"" 570 Ma Chilhowee dike trends versus older
trends) or changes related to Iapetan paleogeo-
graphic features (i.e., triple junction or transform
fault at the recess; Rankin, 1976; Thomas, 1991).
The very limited number of known dikes currently
precludes anything but speculation.
Early- and Late-Stage Igneous Rocks,
Southwestern Virginia-North Carolina
Blue Ridge
Gossan Lead Area. In the Gossan Lead area
(area 6 on Figure 1; Figure 6), Stose and Stose
(1957) mapped the easternmost exposures of early-
stage Mount Rogers metavolcanic rocks "" 10 km
northwest of the elongated (5 X 16 km) Striped Rock
Pluton dated at 690 10 Ma (Odom and Fullagar,
1984). The pluton changes trend eastward from E-
W to N60E. They also mapped five smaller granitic
plutons north and west of the main body and an
array of NNE-trending felsic dikes extending
northeastward from the main pluton. They mapped
scattered felsic dikes north of the main pluton, but
none south of it - that is, none south of the Fries
fault as shown by Bartholomew and Lewis (1984).
Stose and Stose (1957) showed numerous plutons of
Cattron diorite near the Striped Rock Pluton, as
well as a mafic dike array subparallel to but slightly
west of the felsic dike array. Both felsic and mafic
dikes cut both Striped Rock granite and Cattron
diorite and are terminated by the unconformity with
the overlying, Cambrian age (Simpson and Sund-
berg, 1987), Chilhowee Group (Rankin, 1970;
Rankin and others, 1973). Thus, like elsewhere in
the Virginia Blue Ridge, both felsic and mafic dikes
are constrained to be of Late Proterozoic age (760-
570 Ma) from crosscutting and bounding relation-
ships. Stose and Stose (1957) mapped a limited
number of "altered diorite" bodies (which they
interpreted as Cattron diorite) intruded into the
Shoals Gneiss between their Gossan Lead fault and
the Fries fault, as subsequently shown by Bartholo-
mew and Lewis (1984).
Boone Area. In the Boone area (area 8 on
Figure 1; Figure 7), Bartholomew and others
F
i
g
u
r
e

5
.

G
e
o
l
o
g
i
c

m
a
p

o
f
E
a
r
l
y

C
a
m
b
r
i
a
n

a
n
d

L
a
t
e

P
r
o
t
e
r
o
z
o
i
c

r
o
c
k
s

o
f

t
h
e

R
a
d
f
o
r
d

3
0
'

X

6
0
'

q
u
a
d
r
a
n
g
l
e

(
F
i
g
u
r
e

1
,

a
r
e
a

6
)
,

m
o
d
i
f
i
e
d

a
f
t
e
r

B
a
r
t
h
o
l
o
m
e
w

a
n
d

o
t
h
e
r
s

(
l
9
9
2
b
)
.

G
r
e
n
v
i
l
l
i
a
n

r
o
c
k
s

o
f

t
h
e

B
l
u
e

R
i
d
g
e

a
r
e

n
o
t

p
a
t
t
e
r
n
e
d
;

b
l
a
c
k

a
r
e
a
s

a
n
d

l
i
n
e
s

d
e
n
o
t
e

a
m
p
h
i
b
o
l
i
t
e
s

a
n
d

a
l
t
e
r
e
d

m
a
f
i
c

d
i
k
e
s
;

v
,

m
a
f
i
c

v
o
l
c
a
n
i
c

u
n
i
t
s

o
f

b
a
s
a
l

U
n
i
c
o
i

F
o
r
m
a
t
i
o
n
;

C
G
,

C
h
i
l
h
o
w
e
e

G
r
o
u
p
;

L
F
,

L
y
n
c
h
b
u
r
g

F
o
r
m
a
t
i
o
n
.

Q
u
a
d
r
a
n
g
l
e
s
:

E
,

E
l
l
i
s
t
o
n
;

B
M
,

B
e
n
t

M
o
u
n
t
a
i
n
;

R
,

R
i
n
e
r
;

P
,

P
i
l
o
t
;

C
H
,

C
h
e
c
k
;

C
A
,

C
a
l
l
a
w
a
y
.


c
\
)
V

'
.

s
.
.
.

.
.
,


.
.
.
.

.

f
"

.
.
.
.
.
.

L
F
-
.

.
.
.

4
!


L
F

o

1

2
K
M

'
=
'
=
=
'

L
F

.
I
.

N

t

,


'
Q
\
v

C
G


(
S
J

,

,

/
'

/

,


\
v
0


"
"
"

0
1

"
"
"


.
:
.
.
.

t
;
:
j


:
+

:
:
r

o

0
"

S

(
1
)


Late Proterozoic continental margin of the Laurentian craton 455
Figure 6. Geologic map (Figure 1,
area 7) of the Gossan Lead area of
Virginia (modified after Stose and
Stose, 1957), showing felsic (dotted
lines) and mafic (short dashed lines)
dikes mapped by Stose and Stose
(1957) and Riecken (1966); Striped
Rock Pluton as mapped by Riechen
(1966); Independence (extensional)
fault from Simpson and Kalaghan
(1989) , and Kalaghan (1987); modified
trace of Fries Fault from Bartholomew
and Lewis, 1984); and Ashe Formation
of Rankin and others (1972) . Gren-
villian rocks are not patterned; CG,
Chilhowee Group; MR, Mount Rogers
Formation; F, town of Fries.
(1983) illustrated the distribution of Bakersville,
felsic, and mafic (biotite-bearing) dikes, as well as
folded Crossnore granitoid plutons, within portions
of the Watauga and Elk River massifs north of the
Grandfather Mountain window. They showed the
distribution of abundant medium- to coarse-grained,
green amphibolite dikes (but not individual dikes)
that occur within the Elk River massif near but
structurally below the Gossan Lead fault (patterned
area west offault on Figure 7).
Two major granitoid plutons (Buckeye Knob
and Leander Mountain) intrude basement gneisses
of the Elk River massif, which is structurally
superjacent to the Watauga massif containing the
Beech Pluton in the lower, recumbent limb of a
large fold (Bartholomew and others, 1983). Further-
more, within the Elk River massif, mafic, biotite-
bearing dikes are generally transposed subparallel
to foliation. Likewise, felsic and mafic dikes asso-
ciated with the Beech Pluton also trend parallel to
W -NW -trending Paleozoic fold axes.
Relict Dike Trends. In the Gossan Lead area
(Figure 6), dike orientation data for all felsic and
mafic dikes north of the Fries thrust (Figure 2G,H;
Table 1) have a N200E principal trend. The few
altered dikes between the Gossan Lead and Fries
faults (Figure 2U) have a N45E trend. The mean
trend of foliation within the Striped Rock Pluton is
N55E, 60
0
SE (Riechen, 1966; Table 2), but Late
Proterozoic fabrics associated with the Indepen-
dence fault were not distinguished from Paleozoic
fabrics. The 35 difference between the principal
dike trend and the foliation trend suggests that Late
Proterozoic dikes in this region (except between the
Fries and Gossan Lead faults) were not significantly
transposed during Paleozoic deformation; hence, the
N200E dike trend may reflect relict original orienta-
tions. This trend is consistent with results from
dike-orientation data in the central and northern
Virginia Blue Ridge, but it is less consistent with
the interpretation of Simpson and Kalaghan (1989)
- that downdip extension along the Independence
fault (and perhaps the Fries fault) produced its
characteristic mylonite zone. Their data are consis-
tent with an axis of extension of "" N35W, and dikes
intruded in such a stress field should be oriented
""N55E (more like the trend of the Striped Rock
456 M. J. Bartholomew
Figure 7. Geologic map of the Boone area
(Figure 1, area 8), modified after Bartholo-
mew and others (1983), showing felsic
(dotted lines) and other mafic dikes (short
dashed lines); trace of Forked Ridge fault
modified after Adams (1990); and trace of
Gossan Lead fault from Bartholomew and
Lewis (1988) . Grenvillian rocks are not
patterned; small letters: D, diorite; B,
Bakersville Gabbro; U, ultramafic bodies;
EK, Elk Knob sulfides; BKP, Buckeye
Knob Pluton; LMP, Leander Mountain
Pluton. Towns (largest letters): B, Boone,
EP, Elk Park, MC, Mountain City. Quad-
rangles: MC, Mountain City; BG, Baldwin
Gap; EM, Elk Mills; S, Sherwood; Z, Zion-
ville; EP, Elk Park; V, Valle Crucis;
B, Boone.
Pluton) - 35 different from the N200E trend for
dikes portrayed on the Stose and Stose (1957) map.
According to Rankin (personal communication,
1989), his observations in the region suggest that
the Stoses' mapping of the dikes is partly dia-
grammatic; hence, the Gossan Lead dike trend may
not be reliable. On the other hand, the extensional
fabric Simpson and Kalaghan (1989) examined
necessarily post-dates emplacement of the Striped
Rock Pluton and thus cannot be used as an indicator
of the stress field at the actual time of emplacement
of the pluton (i.e., at 690 10 Ma) near the end of the
earlier stage of Iapetan extension. Thus it is even
possible that the dike trends and the fault data
record changes in the extension direction with time.
Within the Watauga massif of the Boone area
(Figure 7), felsic dikes have a N45E principal trend
and a N35W secondary trend (Figure 2J; Table 1).
Mafic (biotite-rich) dikes have a N450W principal
trend (Figure 2K; Table 1). Thirteen scattered,
relatively unaltered Bakersville gabbro dikes have
a N25E principal trend (Figure 2V; Table 1) and a
N-NE overall trend across the map (Figure 7). This
N25E Bakersville dike trend is similar to the
questionable N200E trend in the nearby Gossan
Lead area and to the early-stage relict N25E trend
in the Blue Ridge north of the Roanoke recess.
Similarity of dike orientations for Bakersville dikes,
early-stage felsic dikes, and the Robertson River rift
zone is possible because the Bakersville metagabbro
has a 734 Ma Rb-Sr age (Goldberg and others, 1986)
which is younger than the zircon age (760 Ma) of the
oldest known rocks of the Crossnore Complex
(Aleinikoff and others, 1991) but older than many of
the granitoid plutons of the complex (Odom and
Fullagar, 1984).
Late Proterozoic continental margin of the Laurentian craton 457
Table 2. Paleozoic and Grenvillian foliation data.
Area
Strike and Dip
[from Fig. 1] Unit (Reference) of Mean Foliation
1 Lovingston massif (1) N35E,25SE
C3 Chilhowee Group (2) N44E,36SE
C3 Catoctin Formation (3) N42E,35SE
C3 Pedlar River Suite (3) N46E,68SE
C3 Mylonitic rocks (4) N42E,24SE
3 Archer Mountain Suite (3) N43E,24SE
3 Stage Road Layered Gneiss (3) N48E,84SE
3 Hills Mountain Gneiss (3) N40oE,71SE
N35W, 57NW*
3 Border Gneiss (3) N42E,61SE
N58E, 57NW*
5 Chilhowee Group (5) N50oE,34SE
5 Pedlar massif (5) N53E,40
o
SE
5 Stewartsville Pluton (5) N56E,40oSE
5 Lovingston massif (5) N34E,81SE
5 Mylonitic rocks (5) N54E,34SE
6 Basement and cover (6) N50o-55E,40oSE
7 Striped Rock Pluton (7) N55E, 60
0
SE
*Grenvillian segregation layering. References: 1) Lukert and Nuckols, 1976; 2) Griffin, 1971;
3) Gryta and Bartholomew, 1989; 4) Bartholomew and others, 1981; 5) Bartholomew and others,
1982; 6) Kaygi, 1979; 7) Riecken, 1966.
EFFECTS OF PALEOZOIC TECTONISM ON DIKE
ORIENTATIONS
Most dikes in the Blue Ridge are short, steeply
dipping and poorly exposed, hence folding of dikes is
generally not observed. Therefore, comparison of
dike trends with Paleozoic fabric data (Table 2) is
essential to determine the influence of Paleozoic
tectonism on relict dike trends. Comparison of dike-
orientation data for amphibolites (in higher grade
areas) with other mafic dikes (in lower grade areas)
also permits qualitative assessment of the influence
of Paleozoic metamorphism on dike orientations.
This can be done in central and northern Virginia
where more than 1700 dikes of Late Proterozoic-
Cambrian age have been mapped by numerous
workers within both the Pedlar and Lovingston
massifs and the overlying Lynchburg metasedi-
mentary rocks.
Comparison of Dikes from Areas of Different
Metamorphic Grade
The Pedlar massif of the Blue Ridge was near
the western limit of the Catoctin volcanic field.
Northward from the James River, Bartholomew and
others (1981) showed the westward pinch out of
Catoctin metavolcanic rocks over the Pedlar massif
as derived from cross sections of numerous workers.
Greenstone dikes and Catoctin flows both diminish
in abundance southward toward the James River
(Bloomer and Werner, 1955). Greenstone dikes,
which Reed and Morgan (1971) demonstrated were
chemically similar to Catoctin flows, are concen-
trated in the Pedlar massif - although scattered
greenstone dikes are also found in the Lovingston
massif (e.g., Bailey, 1983; Bartholomew, 1977;
Lukert and Nuckols, 1976). Metabasalt flows of the
Catoctin Formation do continue somewhat farther
458 M. J. Bartholomew
south along the eastern flank of the Lovingston
massif, but they thin and pinch out as a continuous
stratigraphic unit near Lynchburg, Virginia, where
Brown (1958) mapped the older metamorphosed
Moneta Gneiss metavolcanics.
Amphibolites and other mafic dikes, which are
chemically similar to Catoctin metabasalts (Espen-
shade, 1986; Ratcliffe, 1987; Bartholomew and
others, 1992a), occur abundantly throughout the
Lovingston massif, particularly where Catoctin
greenstone dikes are concentrated. They also occur
much farther south of the James River. Amphibo-
lites are generally found where the grade of Paleo-
zoic metamorphism was higher (typically garnet
grade) in the easternmost parts of the Lovingston
massif or in the unconformably overlying Lynch-
burg Formation.
From a regional perspective, dikes that in-
truded the western (Pedlar massif) and northern
parts of the Blue Ridge geologic province where
Paleozoic metamorphic grade was lower: (1) have
diverse orientation ranges, (2) have prominent se-
condary and minor trends, and (3) are characterized
by principal trends (N25-35E) subparallel to the
N25E trend of the Robertson River rift zone (Fig-
ures 2AF and 3; Table 1). In contrast, dikes that
intruded the eastern part of the Blue Ridge geologic
province (Lovingston massif and Lynchburg Forma-
tion): (1) virtually all trend NE, (2) have less
diverse orientation ranges, and (3) have principal
trends (N35-55E) that are subparallel to the strike
of regional Paleozoic foliation (Table 2).
Comparison of Dike Trends and Paleozoic
Foliation
In the Flint Hill region (Figure 3), the principal
strike (N35E) of Paleozoic fabric (Table 2) differs
by 10 from the dominant dike trend (N25E).
Although similar, the 10 difference in overall trend
of Paleozoic fabric and Proterozoic dikes may
indicate little influence of Paleozoic tectonism there
on relict dike trends because the principal dike
trends are the same as the Robertson River rift-zone
trend and because the geometry of that 120 X 5 km
rift zone is not the result of Paleozoic deformation.
The N35E Catoctin dike trend is the same as
the principal strike of Paleozoic foliation (Table 2)
in the Flint Hill area. Catoctin greenstone dikes are
generally in the western part of the Blue Ridge
where metamorphic grade is lower (chlorite-biotite
grade); hence, it seems less likely that reorientation
during Paleozoic metamorphism and deformation
would be more pronounced there than farther east
where metamorphic grade was higher. Moreover,
both the central (C2) and southern (C3) Catoctin
segments have prominent but secondary N5E
trends, and all three segments exhibit minor but
persistent NW trends (Figure 2A,B,C,D; Table 1).
Like NW-trending Grenvillian fabric (Table 2),
which lies nearly perpindicular to the Paleozoic
foliation (e.g., Bartholomew and others, 1991), NW
or N dike trends are significantly less likely to be
transposed during Paleozoic deformation and hence
reflect relict Late Proterozoic emplacement orienta-
tions.
Only within the Nelson County area (areas C3
and 3 on Figure 1) have sufficient structural ana-
lyses (Griffin, 1971; Bartholomew and others, 1981;
Gryta and Bartholomew, 1989) been performed to
determine orientation vectors for Paleozoic fabrics
within most basement units and the cover rocks.
The regional foliation determined for different rocks
of the Pedlar massif and overlying Catoctin Forma-
tion and Chilhowee Group in this region is =N45E
5 (Table 2). Thus, the skewed distribution range
(N25-500E) of C3 Catoctin dikes probably reflects
an original NE dike trend (=N25-35E) partially
transposed toward the regional Paleozoic foliation
(N45E) during Paleozoic deformation. Within the
adjacent Lovingston massif (area 3 on Figure 1), the
N45-55E dike trends (Figure 2Q,R; Table 1) is
subparallel to the mean regional Paleozoic foliation,
for several rock units hosting amphibolites and
other mafic dikes, which is also = N 45E 5 (Table
2). The N45E trend (Figure 2R) of virtually all
amphibolite dikes closely reflects the regional
Paleozoic fabric. Thus, either original principal
trends (=N25-35E) were significantly transposed
(10_20 rotation) subparallel to Paleozoic foliation,
or original dike trends had a N45E principal trend
coincidentally parallel with the younger Paleozoic
foliation.
Moreover, amphibolite dikes found = 50-80 km
farther north in Greene and Madison Counties (Fig.
ure 12; Figure 2M,L) and south in the Lynchburg
quadrangle (Figure 14; Figure 2N,P) have similar
symmetrical distributions about principal trends
that appear to parallel regional Paleozoic foliation.
By analogy, therefore, these dikes sets are also
likely to reflect significant rotation (= 10-20)
during Paleozoic orogenesis.
Late Proterozoic continental margin of the Laurentian craton 459
Near Roanoke (Figure 4), the mean regional
Paleozoic foliation in the Pedlar massif and over-
lying Chilhowee Group is N500-55E, 35-40
0
SE
(Table 2). Because of the large angular difference
(25_30) between both the N25E principal mafic
dike trend (Figure 2E; Table 1) and minor NW
trends and the regional foliation, the dike trends
appear to be relict trends little affected by Paleozoic
tectonism. Within the adjacent Lovingston massif,
the largest body of the Stewartsville Pluton (Figure
4) changes trend toward the northeast, from N700E
to N45E and then back to N70E. The mean folia-
tion within both the Stewartsville Pluton and the
adjacent Lovingston massif is N45-55E, 35-60
0
SE
(Table 2), suggesting Paleozoic modification of the
pluton's trend.
Southwest of Roanoke (Figure 5), the NW and
N-NE dike trends (Figure 2F; Table 1) in the Ped-
lar massif also have large angular variations with
Paleozoic foliation (N500-55E, 40
0
SE) in this area
and hence are likely to reflect relict dike trends. In
contrast, amphibolite dikes in the Lovingston mas-
sif (Figure 2T; Table 1) have a N500E principal
trend and a range suggestive of re-orientation
during Paleozoic tectonism.
TECTONIC SETTING OF IAPETAN RIFTING
The Robertson River rift zone and other Late
Proterozoic granitoid plutons, the northern part of
the Bakersville dike swarm, and mafic/felsic vol-
canic fields, which collectively span the Iapetan
extensional event (760-570 Ma), are plotted on a
modified version (Figure 8A) of a Late Proterozoic
palinspastic reconstruction of Grenvillian massifs
(Batholomew, 1983b) which Bartholomew and
Lewis (1988, 1992) place near the Laurentian mar-
gin (Figure 9). Part of a Late Proterozoic non-
marine/marine hinge zone, recognized by Wehr and
Glover (1985), lies cratonward of the Laurentian
margin (Figures 8A and 9).
Southwestward of the hinge trace shown on
Figure 8, Wehr and Glover (1985) followed Rankin
and others' (1972) interpretation and merged the
Rockfish Valley-Fries fault system with the Hayes-
ville/Gossan Lead fault system and interpreted the
hinge zone to lie along the Gossan Lead fault.
However, Stose and Stose (1957) originally mapped
the Fries and Gossan Lead faults as two separate,
distinct thrusts with Grenvillian rocks in between
(Figure 6). Bartholomew and Lewis (1988, 1992)
followed this distinction and showed the Hayes-
ville/Gossan LeadlMartic Line fault zone (Figures
6, 7, 8A) as marking the western boundary of the
accreted Piedmont terrane of Williams and Hatcher
(1983), the western part of which is approximately
correlative with the Jefferson terrane of Rankin
(1988).
The Fries fault at Fries, Virginia (Figure 6),
was linked through reconnaissance mapping with
the Fork Ridge fault (Figure 7) near the Grand-
father Mountain window in North Carolina by litho-
logic contrasts (similar to those farther north) in
basement massifs on opposing sides of this fault
(Bartholomew and Lewis, 1984), which may extend
as far southwestward as the Hot Springs window
(Rast and Kohles, 1986). Thus, the FrieslFork Ridge
fault south of Fries, Virginia, is analogous to the
Rockfish ValleylFries fault (north of Fries, Virginia)
in that both are inboard from the Laurentian
margin (Figures 8A and 9) and not bounding it like
the Hayesville/Gossan Lead/Martic Line fault zone
(Bartholomew and Lewis, 1988, 1992).
Badger and Sinha (1988) suggested that a
significant tectonic hiatus (= 700-600 Ma) existed
between the two stages of Iapetan extension. New
ages elsewhere (e.g., a 615Ma U-Pb age for a Long
Range dike, southeast Laborador; Kamo and others,
1989) may show this hiatus to be of more local
(Virginia Blue Ridge) significance than applicable
to the Laurentian margin (discussed by Rast, this
volume) as a whole. In Virginia a hiatus is evident.
The Robertson River suite and equivalent units are
truncated by the unconformity and/or served as a
source for Lynchburg (or equivalent) detrital
material (Allen, 1963; Schwab, 1974; Lukert and
Nuckols, 1976; Mitra and Lukert, 1982; Lukert and
Banks, 1984). The Catoctin Formation overlies the
Lynchburg. In North Carolina, Bryant and Reed
(1970) indicated that the Brown Mountain Pluton
was a likely source for detrital material in the
Grandfather Mountain Formation which contains
volcanic units.
Fichter and Diecchio (1986) suggested the
entire rift-to-drift transition (analogous to Badger
and Sinha's younger phase) took place within a 60
million year interval (630-570 Ma) based on com-
parison with other examples of rifting. Simpson and
Eriksson (1989) noted that this transition reflected
"a continuum from fault-influenced to thermotec-
tonic subsidence" control on Late Proterozoic to
Early Cambrian sedimentation. Wehr (this volume)
also favors early tectonic control on sedimentation.
460 M. J. Bartholomew
In fact, most models (Rast and Kohles, 1986;
Schwab, 1986; Thomas, 1986; Wehr and Glover,
1985) portray listric normal faulting in the base-
ment during the Late Proterozoic.
Herz and Force (1984) first postulated late
Precambrian extensional movement on the Rockfish
Valley fault. Although documentation of actual
Cambrian or Late Proterozoic displacement of con-
tacts on any exposed Blue Ridge faults is still lack-
ing, a few Blue Ridge surface faults (e.g., Indepen-
dence fault; Figure 7) are currently known to have
kinematic indicators indicative of Late Proterozoic
to Cambrian extension (Simpson and Kalaghan,
1989; Bailey and Simpson, 1991). Beneath north-
western North Carolina, Harris and others (1981)
interpreted Late Proterozoic extensional displace-
ments on subsurface faults from seismic reflection
data along a single line. To the southwest, in the
subsurface of northeastern Georgia and adjacent
South Carolina (Figure 9), interpretation of seismic
data yielded actual orientations of Late Proterozoic-
Cambrian faults. There, normal displacement is in-
ferred for N35E-trending, NW- and SE-dipping
faults and for N100W-trending, E-dipping faults
(Favret and Williams, 1988). Farther southwest-
ward, other NE-trending subsurface faults are also
known in Alabama and northwestern Georgia (e.g.,
Thomas, 1991). Overall, relict Late Proterozoic ex-
tensional axes inferred from relict dike orientations
(Figure 9), are reasonably compatible with exten-
sional surface faults and with subsurface fault
trends identified by Favret and Williams (1988).
The inferred distribution (Figures 8e and 9) of
known granitoid plutons and major volcanic fields of
the earlier stage Ma) of extension con-
trasts somewhat with the known distribution of
mafic rocks of the later Ma) stage.
Whereas some zones of felsic plutonism are sub-
parallel to the Laurentian margin and some appear
to have a large angular discordance to it (Figures
BC and 9), the younger mafic volcanic fields appear
to be distributed along the hinge zone (Figure BB).
Such changes in regional patterns could be
indicative of significant stress-field changes over
time.
Changes in stress-field orientation with time
might be expected in the two-phase model (Badger
and Sinha, 1988) and may be reflected in dike
Figure 8. Distribution of paleogeographic features related to Late Proterozoicl Early Cambrian Iapetan extensional event
on palinspastic base modified after Bartholomew (1983b).
A. Palinspastic map of Grenvillian basement (1150-1000 Ma) massifs (stippled areas) showing the Fries/Rockfish Valley
(RFVF) fault system of Bartholomew and Lewis (1984), the terrane-bounding Hayesville-Gossan Lead-Martic Line fault
system of Bartholomew and Lewis (1988,1992), and part of the hinge zone of Wehr and Glover (1985). LFF, Linville Falls
fault; GMW, Grandfather Mountain window; NMIM, nonmarine/marine; ANAC, ancestral North American craton;
AT, accreted terranes.
B. Palinspastic distribution of latest Proterozoic-Early Cambrian volcanic fields within and adjacent to Grenvillian
basement massifs of the Blue Ridge province, with extensional directions (arrows) inferred from dike-orientation data
(Figures 2 and 3) and distribution and orientation of zones of granitic intrusions (Figure 8C). BUV/BWV, Early Cambrian
basal Unicoi (BUV) and basal Weverton (BWV) volcanic field; CV, latest Proterozoic-earliest Cambrian Catoctin volcanic
field, with NW pinchout line (hatchured) delineated by Bartholomew and others (1981) and metagabbro (g) along Rockfish
River - age range from Badger and Sinha (1988) and Aleinikoff and others (1991); MV, Late Proterozoic metavolcanic rocks
of the Moneta Gneiss of Brown (1958) and Conley (1978); GMV/MRV, Late Proterozoic Grandfather Mountain (GMV) and
Mount Rogers (MRV) volcanic field, age from Aleinikoff and others (1991). The zone of Bakersville Gabbro intrusions
extends SE beyond map area; age from Goldberg and others (1986).
C. Palinspastic distribution of zones (heavy dashes) of Late Proterozoic granitic plutons (black) within and adjacent to
Grenvillian basement massifs of the Blue Ridge province, with principal extensional directions (arrows) inferred from
distribution and orientation of zones and from dike-orientation data (Figures 2 and 3), modified from Bartholomew and
Lewis (1984). Geochronological data (UlPb not underlined; Rb/Sr underlined) from Fullagar and Butler (1980); Hudson and
Dallmeyer (1981); Herz and others (1981); Lukert and Banks (1984); Mose and Nagel (1984); Odom and Fullagar (1984);
Rankin (1976); Rankin and others (1969); Sinha and Bartholomew (1984); Tollo and others (1991). RRS, Robertson River
suite of granites; RRP, Rockfish River Pluton; MMP, Mobley Mountain Pluton; IC, Irish Creek stock; SMP, Suck Mountain
Pluton; SP, Stewartsville Pluton; DMP, Dillons Mill Pluton; SRP, Striped Rock Pluton; CD, Cattron diorite; BMP, Brown
Mountain Pluton; LP, Lansing Pluton; LMP, Leander Mountain Pluton; BKP, Buckeye Knob Pluton; BP, Beech Pluton;
C, Crossnore Pluton; ES, Elkin Stock; DP, Danbury Pluton; CP, Capella Pluton; RHP, Rock House Pluton.
Late Proterozoic continental margin of the Laurentian craton 461
trends. In the northern Blue Ridge, the Robertson
River rift-zone trend (N25E) and the Catoctin trend
(N35E) differ by 10, and these trends represent (at
least partially) the earlier and later, respectively,
stages of the Iapetan extensional event. However,
these trends are not mutually exclusive:
considerable overlap exists of principal prefer-
red orientation ranges (Figures 2 and 3; Table 1);
relict secondary and minor trends of Catoctin
and other mafic dikes indicate considerable local
B
c
and regional variablity in extensional patterns
(perhaps some areas were transtensional); and
during the later stage, Catoctin and other
equivalent mafic dikes may have intruded along
fractues developed during the earlier phase.
Obviously, considerably more work on geo-
chemistry, geochronology, crosscutting relation-
ships, and kinematic indicators of different dike sets
is needed before relict dike patterns and relict
stress-field patterns can be related in detail.
ANAC
462 M. J. Bartholomew
Overall, dike trends are consistent with ex-
tension primarily along an axis in the range of
N55-75W (""'N65E). Although approached from
different lines of evidence, Thomas' (1977) recon-
struction of the rifted margin through Virginia also
has an extension direction of N65OW. However,
even if a 10 shift (reflecting the change in overall
principal dike trends from N25E to N35E) did
occur, the Robertson River rift zone and Catoctin
trends are still so similar that major changes in
stress-field orientation probably did not occur
during the time interval involved ("'" 730-570 Ma).
CONCLUSIONS
1) The Robertson River zone is a 120-km-long,
N25E-trending zone of granitic plutons with a
length:width ratio >20:1; it is interpreted as an
extensional rift zone developed perpendicular to a
Late Proterozoic (730-700 Ma) N65W axis of exten-
sion.
2) In northern Virginia, both felsic dikes and other
mafic dikes have the same principal trend (N25E)
as the rift zone, indicating that: (a) they are approxi-
mately coeval with the rift zone and intruded under
the same stress field; (b) they are younger but a
similar stress field existed at the time of their
intrusion; or (c) they are younger but were intruded
along fractures developed during the earlier stage.
3) Both locally (near the Robertson River rift zone)
and regionally, Catoctin dikes in the western Blue
Ridge exhibit a principal N35E trend with a
secondary N5E trend and a minor NW trend. These
trends, when compared with the N25E rift-zone
trend and other dike trends, suggest either some
change ("'" 10) in the stress field with time or
geographic variations in the stress field. Regionally
though, these Catoctin trends are similar enough to
both the rift-zone trend and felsic and other mafic
dike trends, that no significant change in the Iape-
tan stress field is inferred between the earlier
(""'730-700 Ma) stage (rift-zone stage) and the later
(""'600-570 Ma) stage (Catoctin stage).
3) In southwestern Virginia and northwestern
North Carolina, dike trends are less definitive than
farther north in Virginia. Although consistent with
those farther north, felsic and mafic dike trends in
the Gossan Lead area, where kinematic indicators
indicate down-dip (S35E) extension on the Indepen-
dence fault, may not be reliable. A few Bakersville
dikes in nearby North Carolina are also consistent
with an extension axis ofN65W. Other mafic and
felsic dike trends are, however, much more diverse
- in part reflecting the effects of Paleozoic tec-
tonism.
4) Extension directions inferred from Late Pro-
terozoic dikes throughout the Blue Ridge geologic
province (N55-75OW principal, N85W secondary)
are generally consistent with extension directions
from (a) the Robertson River rift zone - N65W
axis, (b) data from kinematic indicators (e.g., Inde-
pendence fault; Simpson and Kalaghan, 1989) -
N35W axis, and (c) trends of Late Proterozoic
subsurface faults (Favret and Williams, 1988) much
farther south in the autochthonous Grenvillian
basement - N550W and N85E axes.
5) Dikes located in regions that were metamor-
phosed to chlorite-to-biotite grade, reflect orienta-
tions likely acquired at the time of emplacement. In
regions that reached garnet grade, on the other
hand, dikes are generally complexly recrystallized
and are structurally subparallel to the regional
foliation so that relict trends versus transposed or
rotated trends are not distinguishable from orien-
tation diagrams alone.
ACKNOWLEDGMENTS
Both R. P. Tollo and D. W. Rankin spent a num-
ber of days in the field with me sampling suspected
Figure 9. Paleogeographic distribution of principal Late Proterozoic-Early Cambrian structural features within and
adjacent to palinspastically restored Grenvillian basement massifs (of the Blue Ridge geologic province) along the Late
Proterozoic Laurentian margin (EL) of Bartholomew and Lewis (1988, 1992). FW, subsurface faults delineated by Favret
and Williams (1988) with inferred principal (heavy arrows) and secondary (light arrows) extensional directions; WG, hinge
line of Wehr and Glover (1985); BG, area of extensive Bakersville Gabbro intrusions; stippled areas, zones of granitic
intrusions (Figure 8C); SK, Independence fault with extensional directions determined by Simpson and Kalaghan (1989);
RR, Robertson River rift zone with inferred principal extension direction (heavy arrows). Principal (longer light arrows) and
secondary (shorter light arrows) extensional directions inferred from dike-orientation data (Figures 2 and 3)); note that
diagram "north" has been palinspastic ally rotated for dikes in eastern Blue Ridge massifs.
Late Proterozoic continental margin of the Laurentian craton 463
Late Proterozoic granitic plutons and discussing
many aspects of these granites, related dikes, and
the Iapetan extensional event, Their contributions
to this paper through our discussions and by review
of early versions of this manuscript are sincerely
appreciated_ Comments and reviews from S, Farrar,
B, Kulander, S, E. Lewis, M, T, Lukert, N, Rast, C,
Simpson, and W, A, Thomas also significantly im-
/",,- -
L, _' - ' "7"""' - - ' - , - ' - ' -, -' - -, - , ("
I /', " """"', \)
)
.r
--
,.J
, ../
,,J
./
/\..,/
/ .'
.>
-'"\ \
/ :./"
\
,)
./
/
(
-.../ - , ) (' '\ l,
I ,-,,, ;\....... , \ ""-,----"
/ ' ( \ t,
./ " '" 5 j J\l '\ \
.r ',,;> 'J
)
,v C
./ \ <..,':""
n i
( '" ''-../.\ ,
\ ,J \.., ''\ \./1:",
/ " ....... .r.' 0 /.
j
' ''' ...J 0
'", ()
\: ,

---L ____ , ___ ,..r '; _, _ _ '_' _,_, _, _
\
464
M. J. Bartholomew
proved this paper and are most appreciated. The
thoughtful, and often provocative, questions posed
by all of these reviewers compelled me to sharpen
the focus of this paper and shift its emphasis. My
thanks to each of you.
REFERENCES
ADAMS, M. G., 1990, The Geology of the Valle Crucis
Area, Northwestern North Carolina: MS thesis, Uni-
versity of North Carolina, Chapel Hill, North Carolina,
USA, 95p.
ALEINIKOFF, J. N., R. E. ZARTMAN, D. W. RANKIN, P. T.
LYTTLE, W. C. BURTON, and R. C. McDOWELL, 1991, New
U-Pb zircon ages for rhyolite of the Catoctin and Mount
Rogers fonnations - More evidence for two pulses of
Iapetan rifting in the central and southern Appalachians
[abstract]: Geological Society of America Abstracts with
Programs, v. 23, no. 1, p. 2.
ALLEN, R. M., Jr., 1963, Geology and Mineral Re-
sources of Greene and Madison Counties: Virginia
Division of Mineral Resources, Bulletin 78, 102 p.
BADGER, R. L., and A. K. SINHA, 1988, Age and Sr isotopic
signature of the Catoctin volcanic province: Implications
for subcrustal mantle evolution: Geology, v. 16, no. 8, p.
692-695.
BAILEY, C. M., and C. SIMPSON, 1991, Deformation of the
basement in the Blue Ridge anticlinorium, Virginia:
Kinematic and temporal evolution [abstract]: Geological
Society of America Abstracts with Programs, v. 23, no. 1, p.
5.
BAILEY, W. M., 1983, Geology of the Northern Half of
the Horseshoe Mountain Quadrangle, Nelson County,
Virginia: MS thesis, University of Georgia, Athens,
Georgia, USA, 100 p.
BARTHOLOMEW, M. J., 1977, Geology of the Greenfield
and Sherando Quadrangles, Virginia: Virginia Divi-
sion of Mineral Resources, Publication 4, 43 p.
__ ,1981, Geology of the Roanoke and Stewartsville
Quadrangles, Virginia: Virginia Division of Mineral
Resources, Publication 34, 23 p.
__ , 1983a, Geologic Map and Mineral Resources
Summary of the Baldwin Gap Quadrangle, North
Carolina and Tennessee: North Carolina Division of
Mineral Resources, GM 220-NW and MRS 220-NW, 5 p.
__ , 1983b, Palinspastic reconstruction of the Grenville
terrane in the Blue Ridge geologic province, southern and
central Appalachians, U.S.A.: Geological Journal, v. 18,
pt. 3, p. 241- 253.
__ , and J. J. GRYTA, 1980, Geologic Map and Mineral
Resources Summary of the Sherwood Quadrangle,
North Carolina and Tennessee: North Carolina Divi-
sion of Land Resources, GM 214-SE and MRS 214- SE, 8 p.
__ , and S. E. LEWIS, 1984, Evolution of Grenville mas-
sifs in the Blue Ridge geologic province, southern and
central Appalachians; pp. 229-254 in M. J. Bartholomew
(ed.), The Grenville Event in the Appalachians and
Related Topics: Geological Society of America, Special
Paper 194, 287p.
__ , and S. E. LEWIS, 1988, Peregrination of Middle
Proterozoic massifs and terranes within the Appalachian
orogen, eastern U.S.A.: University of Oviedo (Spain), Tra-
bajos de Geologia, v. 17, p. 153-163.
__ , and S. E. LEWIS, 1992, Appalachian Grenvillian
massifs: Pre-Appalachian translational tectonics; pp. 363-
374 in R. Mason (ed.), Proceedings of the 7th Inter-
national Conference on Basement Tectonics (Kings-
ton, Ontario, July 1992): Kluwer Academic Publishers,
Dordrecht, The Netherlands, 480 p.
__ , T. M. GATHRIGHT II, and W. S. HENIKA, 1981, A
tectonic model for the Blue Ridge in central Virginia:
American Journal of Science, v. 281, no. 9, p. 1164-1183.
__ , A. P. SCHULTZ, W. S. HENIKA, and T. M.
GATHRIGHT 11,1982, Geology of the Blue Ridge and Valley
and Ridge at the junction of the central and southern
Appalachians; pp. 121-170 in P. T. Lyttle (ed.), Central
Appalachian Geology, NE-SE GSA '82 Field Trip
Guidebooks: American Geological Institute, Falls
Church, Virginia, USA, 266 p.
__ , S. E. LEWIS, J. R. WILSON, and J. J. GRYTA, 1983,
Defonnational history of the region between the Grand-
father Mountain and Mountain City windows, North
Carolina and Tennessee; Article 1, 30 p. in S. E. Lewis
(ed.), Geological Investigations in the Blue Ridge of
Northwestern North Carolina: 1983 Guidebook for the
Carolina Geological Society: North Carolina Division of
Land Resources, Raleigh, North Carolina, USA, 236 p.
__ , S. E. LEWIS, S. S. HUGHES, R. L. BADGER, andA. K.
SINHA, 1991, Tectonic history of the Blue Ridge basement
and its cover, central Virginia; pp. 57-90 in A. Schultz and
E. Compton-Gooding (eds.), Geologic Evolution of the
Eastern United States: Fieldguide for the NE-SE GSA
Meeting: Virginia Museum of Natural History, Fieldguide
Number 2, 304 p.
__ , S. E. LEWIS, S. S. HUGHES, and R. J. WALKER,
1992a, Late Proterozoic tectonics along the Laurentian
margin [abstract]: Geological Society of America Abstracts
with Programs, v. 24, no. 2, p. 3.
__ , A. P. SCHULTZ, S. E. LEWIS, and R. C. McDOWELL,
1992b, Bedrock Geologic Map of the Radford 0030'x
roo' Quadrangle, Virginia and West Virginia: U.S.
Late Proterozoic continental margin of the Laurentian craton 465
Geological Survey, Map 1- 2170B, 1:100,000 scale map with
text, in press.
BLOOMER, R. 0., and H. J. WERNER, 1955, Geology of the
Blue Ridge region in central Virginia: Geological Society
of America Bulletin, v. 66, no. 5, p. 579-606.
BRENT, W. B., 1960, Geology and Mineral Resources of
Rockingham County: Virginia Division of Mineral Re-
sources, Bulletin 76,174 p.
BROWN, W. R., 1958, Geology and Mineral Resources of
the Lynchburg Quadrangle, Virginia: Virginia Divi-
sion of Mineral Resources, Bulletin 74, 99 p.
BRYANT, B. H., and J. C. REED, Jr., 1970, Geology of the
Grandfather Mountain Window and Vicinity, North
Carolina and Tennessee: U.S. Geological Survey, Pro-
fessional Paper 615,190 p.
CONLEY, J. F., 1978, Geology of the Piedmont of Virginia
- Interpretations and problems; pp. 115-149 in Contri-
butions to Virginia Geology - III: Virginia Division of
Mineral Resources, Publication 7, 154 p.
ESPENSHADE, G. H., 1986, Geology of the Marshall
Quadrangle, Fauquier County, Virginia: U.S. Geo-
logical Survey, Bulletin 1560, 60 p.
EVANS, N. H., 1984, Late Precambrian to Ordovician
Metamorphism and Orogensis in the Blue Ridge and
Western Piedmont, Virgina Appalachians: PhD dis-
sertation, Virginia Polytechnic Institute and State Uni-
versity, Blacksburg, Virginia, USA, 313 p.
FAVRET, P. D., and R. T. WILLIAMS, 1988, Basement
beneath the Blue Ridge and Inner Piedmont in north-
eastern Georgia and the Carolinas: A preserved, Late
Proterozoic, rifted continental margin: Geological Society
of America Bulletin, v. 100, no. 12, p. 1999-2007.
FICHTER, L. S., and R. J. DIECCHIO, 1986, Stratigraphic
model for timing the opening of the Proto-Atlantic Ocean
in northern Virginia: Geology, v. 14, no. 4, p. 307-309.
FULLAGAR, P. D., and J. R. BUTLER, 1980, Radiometric
dating in the Sauratown Mountains area, North Carolina;
Article II, 10 p. in V. Price Jr., P. A. Thayer, and W. A.
Ranson (eds.), Geological Investigations of Piedmont
and Triassic Rocks, Central North Carolina and
Virginia: 1980 Guidebook for the Carolina Geological
Society: Savannah River Laboratory, E. I. du Pont de
Nemours & Co., Aiken, South Carolina, USA, 194 p.
GATHRIGHT, T. M., II, 1976, Geology of the Shenandoah
National Park, Virginia: Virginia Division of Mineral
Resources, Bulletin 86, 93 p.
GOLDBERG, S. A., J. R. BUTLER, and P. D. FULLAGAR,
1986 The Bakersville dike swarm: Geochronology and
petr;genesis of Late Proterozoic basaltic magmatism in the
southern Appalachian Blue Ridge: American Journal of
Science, v. 286, p. 403-430.
GRIFFIN, V. S., Jr., 1971, Fabric relations across the
Catoctin Mountain-Blue Ridge anticlinorium in central
Virginia: Geological Society of America Bulletin, v. 82, no.
2, p. 417-432.
GRYTA, J. J., and M. J. BARTHOLOMEW, 1989, Factors
influencing the distribution of debris avalanches asso-
ciated with the 1969 Hurricane Camille in Nelson County,
Virginia; pp. 15-28 in A. P. Schultz and R. W. Jibson (eds.),
Landslides of Eastern United States and Puerto Rico:
Geological Society of America, Special Paper 236,102 p.
HARRIS, L. D., A. G. HARRIS, W. DEWITT, Jr., and K. C.
BAYER, 1981, Evaluation of southern eastern overthrust
belt beneath Blue Ridge-Piedmont thrust: American
Association of Petroleum Geologists Bulletin, v. 65, p. 2497-
2505.
HENIKA, W. S., 1981, Geology of the Villamont and
Montvale Quadrangles, Virginia: Virginia Division of
Mineral Resources, Publication 35, 18 p.
HERZ, N., and E. R. FORCE, 1984, Rock suites in Gren-
villian terrane of the Roseland district, Virginia; pp. 187-
214 in M. J. Bartholomew (ed.), The Grenville Event in
the Appalachians and Related Topics: Geological
Society of America, Special Paper 194, 287 p.
__ , and E. R. FORCE, 1987, Geology and Mineral
Deposits of the Roseland District of Central Virginia:
U.S. Geological Survey, Professional Paper 1371,56 p.
__ , D. C. MOSE, and M. S. NAGEL, 1981, Mobley Mou-
ntain granite and the Irish Creek tin district, Virginia: A
genetic and temporal relationship [abstract): Geological
Society of America Abstracts with Programs, v. 13, no. 7, p.
472.
HUDSON, T. A., 1981, Geology of the Irish Creek Tin
District, Virginia Blue Ridge: MS thesis, University of
Georgia, Athens, Georgia, USA, 144 p.
__ , and R. D. DALLMEYER, 1981, Age and origin of
mineralized greisens from the Irish Creek tin district,
Virginia Blue Ridge [abstract): EOS, Transactions of the
American Geophysical Union, v. 62, p. 429.
KALAGHAN, T. A., 1987, Deformation in the Striped
Rock Pluton, Southwest Virginia: MS thesis, Virginia
Polytechnic Institute and State University, Blacksburg,
Virginia, USA, 68 p.
KAYGI, P. B., 1979, The Fries Fault Near Riner, Vir-
ginia: An Example of a Polydeformed, Ductile Defor-
mation Zone: MS thesis, Virginia Polytechnic Institute
and State University, Blacksburg, Virginia, USA, 165 p.
KAMO, S. L., C. F. GOWER, and T. E. KROGH, 1989,
Birthplace for the Iapetus Ocean? A precise U-Pb zircon
466 M. J. Bartholomew
and baddeleyite age for the Long Range dikes, southeast
Labrador: Geology, v. 17, p. 602- 605.
LUKERT, M. T., 1973, The Petrology and Geochrono-
logy of the Madison Area Virginia: PhD dissertation,
Case Western Reserve University, Cleveland, Ohio, USA,
218p.
__ , and P. O. BANKS, 1984, Geology and age of the
Robertson River Pluton; pp. 161-166 in M. J. Bartholomew
(ed.), The Grenville Event in the Appalachians and
Related Topics: Geological Society of America, Special
Paper 194, 287 p.
__ , and C. R. HALLADAY, 1980, Geology of the Mas-
sies Corner Quadrangle, Virginia: Virginia Division of
Mineral Resources, Publication 17 (Geologic Map 197 A).
__ , and E. B. NUCKOLS, 1976, Geology of the Linden
and FUnt Hill Quadrangles, Virginia: Virginia Division
of Mineral Resources, Report of Investigation 44, 83 p.
MOSE, D. G., and S. NAGEL, 1984, Rb-Sr age for the
Robertson River pluton in Virginia and its implication on
the age of the Catoctin Formation; p. 167-173 in M. J. Bar-
tholomew (ed.), The Grenville Event in the Appala-
chians and Related Topics: Geological Society of Ameri-
ca, Special Paper 194,287 p.
ODOM, A. L., and P. D. FULLAGAR, 1984, Rb-Sr whole-rock
and inherited zircon ages of the plutonic suite of the
Crossnore Complex, southern Appalachians, and their
implications regarding the time of opening the Iapetus
Ocean; pp. 255-261 in M. J. Bartholomew (ed.), The Gren-
ville Event in the Appalachians and Related Topics:
Geological Society of America, Special Paper 194,287 p.
RADER, E. K., T. H. BIGGS, and W. E. NUNAN, 1975,
Geologic Map of the Front Royal Quadrangle, Vir-
ginia: Virginia Division of Mineral Resources, Report of
Investigation 40, Plate 1.
RANKIN, D. W., 1970, Stratigraphy and structure of Pre-
cambrian rocks in northwestern North Carolina; pp. 227-
245 in G. W. Fisher, F. J. Pettijohn, J. C. Reed, Jr., and K.
N. Weaver (eds.), Studies of Appalachian Geology:
Central and Southern: Interscience, New York, New
York, USA, 460 p.
__ , 1975, The continental margin of eastern North
America in the southern Appalachians: The opening and
closing of the proto-Atlantic Ocean: American Journal of
Science, v. 275-A, p. 298- 336.
__ , 1976, Appalachian salients and recesses: Late Pre-
cambrian continental breakup and the opening of the
Iapetus Ocean: Journal of Geophysical Research, v. 81, p.
5605-5619.
__ , 1988, The Jefferson terrane of the Blue Ridge tec-
tonic province: An exotic accretionary prism [abstract):
Geological Society of America Abstracts with Programs, v.
20, no. 4, p. 310.
__ , and R. P. TOLLO, 1987, Late Proterozoic anorogenic
granitoids and Iapetan rifting of Laurentia as preserved in
the Appalachian orogen [abstract): Geological Society of
America Abstracts with Programs, v. 19, no. 7, p. 813.
__ , T. W. STERN, J. C. REED, Jr., and M. F. NEWELL,
1969, Zircon ages of felsic volcanic rocks in the upper Pre-
cambrian of the Blue Ridge, Appalachian Mountains:
Science, v. 166, p. 741-744.
__ , G. H. ESPENSHADE, and R. B. NEUMAN, 1972,
Geologic Map of the West Half of the Winston-Salem
Quadrangle, North Carolina, Virginia, and Tennes-
see: U.S. Geological Survey, Miscellaneous Investigations,
Map I-709-A.
RAST, N., and K. M. KOHLES, 1986, The origin of the
Ocoee Supergroup: American Journal of Science, v. 286, p.
593-616.
RATCLIFFE, N. M., 1987, High Ti0
2
metadiabase dikes of
the Hudson Highlands, New York and New Jersey: Pos-
sible Late Proterozoic rift rocks in the New York recess:
American Journal of Science, v. 287, p.817 -850.
REED, J. C., Jr., 1955, Catoctin Formation near Luray,
Virginia: Geological Society of America Bulletin, v. 66, no.
7, p. 871-896.
__ , and B. A. MORGAN, 1971, Chemical alteration and
spilitization of the Catoctin greenstones, Shenandoah
National Park, Virginia: Journal of Geology, v. 79, no. 5, p.
526-548.
RIECKEN, C. C., 1966, Petrology of the Striped Rock
Granite and Surrounding Rocks, Grayson County,
Virginia: MS thesis, Virginia Polytechnic Institute and
State UniverGity, Blacksburg, Virginia, USA, 161 p.
SCHWAB, F. L., 1974, Mechum River Formation: Late Pre-
cambrian(?) alluvium in the Blue Ridge Province of Vir-
ginia: Journal of Sedimentary Petrology, v. 44, p. 862-871.
__ , 1986, Latest Precambrian-earliest Paleozoic sedi-
mentation, Appalachian Blue Ridge and adjacent areas:
Review and speculation; pp. 115-137 in R. C. McDowell and
L. Glover III (eds.), Studies in Appalachian Geology
(Lowry Volume): Virginia Tech, Department of Geo-
logical Sciences, Memoir 3, 137 p.
SIMPSON, C., and T. KALAGHAN, 1989, Late Precambrian
crustal extension preserved in Fries fault zone mylonites,
southern Appalachians: Geology, v. 17, no. 2, p. 148-151.
SIMPSON, E. L., and K. A. ERIKSSON, 1989, Sedimentology
of the Unicoi Formation in southern and central Virginia:
Evidence for Late Proterozoic to Early Cambrian rift-to-
passive margin transition: Geological Society of America
Bulletin, v. 101, no. 1, p. 42-54.
Late Proterozoic continental margin of the Laurentian craton 467
__ , E. L., and F. A. SUNDBERG, 1987, Early Cambrian
age for synrift deposits of the Chilhowee Group of south-
western Virginia: Geology, v. 15, no. 2, p. 123-126.
SINHA, A. K., and M. J. BARTHOLOMEW, 1984, Evolution
of the Grenville terrane in the central Virginia Appala-
chians; pp. 175-186 in M. J. Bartholomew (ed.), The
Grenville Event in the Appalachians and Related
Topics: Geological Society of America, Special Paper 194,
287p.
STOSE, A. J., and G. W. STOSE, 1957, Geology and
Mineral Resources of the Gossan Lead District and
Adjacent Areas in Virginia: Virginia Division of Min-
eral Resources, Bulletin 72, 291 p.
THOMAS, W. A., 1977, Evolution of Appalachian-Ouachita
salients and recesses from reentrants and promontories in
the continental margin: American Journal of Science, v.
277, no. 10,p. 1233-1278.
__ , 1983, Continental margins, orogenic belts, and
intracratonic structures: Geology, v. 11, no. 5, p. 270-272.
__ , 1991, The Appalachian-Ouachita rifted margin of
southeastern North America: Geological Society of Ameri-
ca Bulletin, v. 103, no. 3, p. 415-431.
TOLLO, R. P., and J. N. ALEINIKOFF, 1992, The Robertson
River Igneous Suite, Virginia Blue Ridge: A case study in
mUltiple-stage magmatism associated with the early
stages of Iapetan rifting [abstract): Geological Society of
America Abstracts with Programs, v. 24, no. 2, p. 70.
__ ,J. N. ALEINIKOFF, and K. J. GRAY, 1991, New UlPb
zircon isotopic data from the Robertson River Igneous
Suite, Virginia Blue Ridge: Implications for the duration
of Late Proterozoic anorogenic magmatism [abstract):
Geological Society of America Abstracts with Programs, v.
23, no. 1, p. 139.
WEHR, F., 1985, Stratigraphy of the Lynchburg Group and
Swift Run Formation, Late Proterozoic (730-570 Ma),
central Virginia: Southeastern Geology, v. 25, no. 4, p. 225-
239.
__ , and L. GLOVER III, 1985, Stratigraphy and tectonics
of the Virginia-North Carolina Blue Ridge: Evolution of a
Late Proterozoic-early Paleozoic hinge zone: Geological
Society of America Bulletin, v. 96, p. 285-295.
WERNER, H. J., 1966, Geology of the Vesuvius Quad-
rangle, Virginia: Virginia Division of Mineral Resources,
Report of Investigations 7,53 p.
WILLIAMS, H., and R. D. HATCHER, Jr., 1983, Appala-
chian suspect terranes; pp. 33-53 in R. D. Hatcher, Jr., H.
Williams, and 1. Zietz (eds.), Contributions to the Tec-
tonics and Geophysics of Mountain Chains: Geological
Society of America, Memoir 158, 223 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Effect of the Transylvania fracture zone on evolution of the
western margin of the Central Appalachian basin
SAMUEL I. ROOT
Department of Geology, The College of Wooster, Wooster, Ohio 44691, USA
(received August 5, 1988; revision accepted January 3D, 1989)
ABSTRACT
A major crustal fracture zone in concealed Precambrian gneisses of the Grenville Province, transects - at
a large angle - the western flank of the Appalachian Paleozoic depositional basin and superimposed Mesozoic
rift basin. This zone, which has an extended deformational history, may be traced intermittently more than
400 km from the Mesozoic Gettysburg rift-basin westward toward Pittsburgh and Cleveland. Reactivation of
the zone during the Cambrian to Pennsylvanian is well established on the Appalachian Plateau. Thickness
variations of such economic units as Silurian salt, Pennsylvanian coals, and high-alumina clays are partially
a result of syndepositional fracture-zone movement. In the Valley and Ridge Province of central
Pennsylvania, reactivation of the zone is evident at least during Devonian sedimentation. In southeastern
Pennsylvania during the Ordovician Taconic orogeny, the basin north of the fracture zone was deep and was
the site of emplacement of Taconic allochthons. During the Permian Alleghanian orogeny, Blue Ridge
structures were rotated 25 across the zone. Where the zone intersects the major Jurassic-age normal fault
that bounds the Gettysburg half-graben basin, local wrenching interrupted the regional extensional fabric.
The structurally deepest part of the extensional basin is adjacent to the fracture zone.
INTRODUCTION
Many structural features and lineaments deve-
loped at large angles to the NE-trending regional
structural grain of the Central Appalachians
(Gwinn, 1964; Wagner and Lytle, 1976; Kowalick
and Gold, 1976; Root and Hoskins, 1977; Rodgers
and Anderson, 1984; Shumaker, 1986; Wheeler,
1986). The largest structural feature of this type, in
the Pennsylvania-Ohio part of the Central Appala-
chians, is the Transylvania fault zone (Root and
Hoskins, 1977) - herein termed the Transylvania
fracture zone. This paper is a review of published
studies that suggest reactivation of this fracture
zone during Paleozoic sedimentation and orogenesis
as well as Mesozoic rifting. This fracture zone pro-
foundly influenced evolution of the western margin
of the Appalachian basin. It extends >400 km
westward from the Mesozoic Gettysburg rift-basin,
toward Pittsburgh, then trends northwestward
toward Cleveland, Ohio (Figure 1). Its trace is
intermittent in Paleozoic cover rocks but may be
continuous in Precambrian basement. Locally com-
plex structures within the Mesozoic rift basin are
sited at the intersection of the fracture zone and
Mesozoic normal faults.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 469-480. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
469
470 S. I. Root
STRUCTURE OF THE TRANSYLVANIA
FRACTURE ZONE
A principal segment of this fracture zone, the
Carbaugh-Marsh Creek fault (Figure 1), is signifi-
cant because: (1) it is oriented E-W, nearly normal
to the NE-trending Appalachian structural grain;
(2) comparable amounts of shortening, represented
by faults and folds of similar geometries and
orientation, developed semi-independently on either
side of the fault during the Alleghanian orogeny;
and (3) the fault is a locus of Appalachian arcuation
because 20_30 regional rotation of the structural
grain, including cleavage, about a vertical axis
occurs across the fault (Root, 1970). The occurrence
of the Shippensburg fault, another EIW-trending
fault 15 km to the north, suggests that these faults
comprise a zone rather than a single discrete fea-
ture. Kowalik and Gold (1976) identified a linea-
ment extending intermittently from the Carbaugh-
Marsh Creek fault in the Blue Ridge, west across
much of the Valley and Ridge. Mapping there also
shows a discontinuous zone of E-W faulting with
LAKE ERIE
\:DDLEBURG F.
AKRON-SUFFIELD F.Z.
I"
o
I
I
o
semi-independent shortening across the zone (Root
and Hoskins, 1977).
The aeromagnetic map of western Pennsyl-
vania and eastern Ohio (Popenoe and others, 1964)
has contrasting patterns along an EIW-trending
zone extending from the eastern edge of the Appala-
chian Plateau, past Pittsburgh and into Ohio.
Because the sedimentary cover there is magneti-
cally transparent, contrasting magnetic patterns
reflect features in Precambrian basement; an in-
ferred E-W fault there could juxtapose rocks of
differing magnetic susceptibilities producing this
anomaly pattern. Sited above the zone of basement
magnetic contrasts in Pennsylvania is a narrow
zone of structural discontinuities (Figure 2) where
Appalachian folds, above the regional decollement,
terminate or are sharply deflected (Wagner and
Lytle, 1976). Based on these structural and aero-
magnetic discontinuities, the Transylvania zone is
projected westward across the Valley and Ridge and
then northwestward across the Plateau. Principally
on gravity and magnetic evidence, Lavin and others
(1982) had projected the NW-trending zone on the
MILES
I
KM
I
50
50
I
Figure 1. Geologic provinces of the Central Appalachians showing mapped surface and near-surface faults within the
Transylvania fracture zone. Mapped fault, heavy solid line; C.-MC.F, Carbaugh-Marsh Creek fault; SHP.F., Shippensburg
fault; BR, Blue Ridge Province; Y.B., Yellow Breeches thrust sheet; HBG, Harrisburg; PBG, Pittsburgh; CLV, Cleveland;
Pennsylvania faults after Root and Hoskins (1977), Ohio faults after Gray (1982). Light dashed line in Valley and Ridge
shows folded basal Silurian strata.
Transylvania fracture zone and evolution of the Central Appalachian basin 471
plateau southeastward to Washington, DC, as one
boundary of a major crustal block. However, I be-
lieve the disruption of folds is more compelling
evidence for linking the plateau structure eastward
to faults in the Valley and Ridge-Blue Ridge. An
unambiguous basement fracture zone is recognized
= 100 km northeast of and subparallel to the Tran-
sylvania zone. This structure, which extends south-
east from Erie, Pennsylvania, has a comparable
geologic history (Rodgers and Anderson, 1984) and
may be related to the process that formed the
Transylvania fracture zone.
To the west, in Ohio, Appalachian folds are very
low amplitude and difficult to map. However, Gray
(1982), in a subsurface study, prepared structure
contour maps on top of the Mississippian Berea
sandstone and the Devonian Onondaga limestone.
These maps defined zones of faulting, referred to
collectively as the Highlandtown fault system
(Figure 1), that Gray (1982) showed are not related
to Appalachian detachment tectonics on Silurian
salt horizons but involve basement faulting with the
southwest block downthrown by = 50 m. This work
extends the Transylvania fracture zone westward to
Cleveland. A ground-derived total magnetic inten-
sity map across the central part of the Akron-
Suffield fault zone suggests that the fault is a SW-
dipping normal fault (Root and MacWilliams, 1986).
Original fault displacement during the Early Cam-
brian (or earlier) may have been different, as
inferred from the en echelon geometry of the Akron-
Suffield faults, which suggests they formed as
synthetic right-lateral wrench faults.
Root and Hoskins (1977) considered these faults
and lineaments as part of a fundamental fracture
zone extending well into the continental plate. They
cited as support for this conclusion: (1) its great
extent (>400 km) transverse to the Appalachian
structural grain; (2) its extreme rectilinearity;
(3) its occurrence in both surface and subsurface
layers; (4) the large variety, and age, of rocks
faulted - from Precambrian greenstones in the
Figure 2. Position of the Transylvania fracture zone on the Appalachian Plateau shown by heavy line. Light lines show total
intensity magnetic field relative to an arbitrary datum (from Popenoe and others, 1964); C.I.=50 gammas. Fold axes shown
by intermediate lines; after Wagner and Lytle (1976). Fracture zone in Pennsylvania sited on zone of fold deflection,
terminations, and NW -trending aeromagnetic anomaly. In Ohio, fracture zone sited on Highlandtown fault of Gray (1982).
472 S. I. Root
Blue Ridge to Pennsylvanian coals along the Ohio
River; (5) its probable sub vertical attitude inferred
from its trace in the high-relief Blue Ridge Moun-
tains and magnetic studies in Ohio; (6) its 15-20 km
width in the Blue Ridge; and (7) indications from
aeromagnetic mapping that Precambrian basement
rocks were substantially offset.
FRACTURE ZONE REACTIVATION
Evidence of recurrent movement along the frac-
ture zone was presented by Root and Hoskins (1977).
The zone was an active discontinuity surface in the
Blue Ridge and Great Valley during Alleghanian
deformation, allowing rocks to be deformed semi-
independently on opposite sides of the Carbaugh-
Marsh Creek fault. Offset of the border fault of the
Mesozoic Gettysburg basin on the Shippensburg
fault indicates Jurassic reactivation. It was also
suggested (Root and Hoskins, 1977) that during the
Ordovician Taconic orogeny, the Appalachian basin
north of the fracture zone was structurally depres-
sed and thus became the site for emplacement of
allochthons in what is now the Great Valley.
The Highlandtown fault, mapped by Gray
(1982) at the Ohio River (Figure 1), differs in
position by several kilometers between its locations
on structure maps of the tops of the Onondaga and
Berea. These subparallel faults are suggestive of a
fracture zone in which reactivation occurred at
different times and in somewhat different positions.
The effect of such movement during deposition of
some Paleozoic strata adjacent to the Akron-Suffield
fault was studied by Root and MacWilliams (1986).
They observed that subsurface distribution patterns
define minor syndepositional fault movement
during accumulation of the Silurian Salina salts
and again during deposition of the Mississippian
Berea sands. Published regional stratigraphic
studies (see below) document recurrent syndeposi-
tional activity along the fracture zone throughout
the Paleozoic.
Appalachian Plateau and Valley and Ridge
Province
Because Paleozoic strata on the plateau are
involved in relatively little cover shortening and
have been extensively drilled for oil and gas, the
subsurface stratigraphy of this region is better
known than that to the east. Calvert (1974) ex-
amined Cambro-Ordovician strata in Ohio and
produced two isopach maps pertinent to this study.
An isopach map of the interval from the top of
Precambrian basement to the top of the Upper
Cambrian Maynardville dolomite shows a graben-
like area of thick sedimentary fill southwest of the
Highlandtown fault system (Figure 3). Syndeposi-
tional Cambrian faulting on the Highlandtown zone
could have resulted in the southwest side of the fault
subsiding more rapidly than the northeast side.
Comparable syndepositional effects are obser-
ved in Ordovician strata. Calvert's (1974) isopach
map of the Middle Ordovician Trenton limestone
shows that along the Ohio River, the Highlandtown
fault as it is now mapped (Figure 4) separates belts
of thin (38-42 m) and thicker (45-52 m) Trenton.
Salt horizons (Upper Silurian Salina Group)
show distribution patterns (Clifford, 1973) that may
be attributed to syndepositional movement along
the fracture zone. The Akron-Suffield fault zone
appears to form the southwest limit of a local F-4
salt embayment in this area (Figure 5), and the
Highlandtown fault separates areas of thin (18 m)
o
I
o
I
50
KM
LAKE ERIE
Figure 3. Isopach of interval from top of Precambrian
basement in Ohio, to top of Upper Cambrian Maynardville
dolomite from Calvert (1974); Highlandtown fault system
(Gray, 1982) shown by heavy line.
Transylvania fracture zone and evolution of the Central Appalachian basin 473
salt (SW) from thicker (21-27 m) salt (NE). It is dif-
ficult to determine displacement sense on the faults
from salt distribution.
A stratigraphic analysis of the Lower Devonian
Oriskany sandstone by Diecchio (1985) shows inter-
ruption of the NE-trending Appalachian basin along
the EIW-trending Transylvania fracture zone (Fig-
ure 6). Strata are thicker south of the fracture zone,
suggesting greater subsidence and accumulation
south of a hinge line or a fault. In Ohio, the Akron-
Suffield fault zone crosses an area where the Oris-
kany is absent, indicating it was a locally positive
area during, or shortly after, deposition. Lower
Devonian carbonates of the Lower Onesquethaw
also show an isopach configuration (Mesolella, 1978)
(Figure 7) suggestive of syndepositional influence
of the Transylvania zone. Strata are thicker south
of the zone but, as may also be observed in the
Oriskany (Figure 6), a small carbonate depocenter
occurs north of the Highlandtown fault (Figure 7).
However, Mesolella's (1978) map of Upper Onesque-
thaw limestones show no distribution patterns that
can be related to activity of the Transylvania zone.
Middle and Upper Devonian black shales show a
o
I
o
I
50
KM
LAKE ERIE
C.1. = 25'
Figure 4. Isopach of Middle Ordovician Trenton limestone
in Ohio (from Calvert, 1974); faults shown by heavy line.
thickness pattern (Harris, 1975) consistent with a
positive area south of the basement faults, where
accumulation was less than half that north of the
Transylvania zone.
Distribution of Lower Mississippian Berea
sandstones and other associated sandstones in Ohio,
examined by Pepper and others (1954), are consis-
tent with syndepositional movement on the zone. A
major lobe of thick Berea sandstone, with a NNE
source, extends south from Cleveland but is inter-
rupted where it crosses the Akron-Suffield fault
zone. However, a more compelling syndepositional
relation is exhibited by the slightly older Cussewago
sandstone, which has a depocenter axis that paral-
lels the trend of the Highlandtown fault and the
eastern portion of the Akron-Suffield fault zone
(Figure 8). This distribution suggests down-to-the-
north fault movement. Although these units are
part of a complex depositional system, syndeposi-
tional activity on the Transylvania fracture zone
may account for some of the distribution patterns.
Several Pennsylvanian units demonstrate syn-
depositional reactivation of the Transylvania zone.
Hansen (1984) identified a persistent Pottsville
LAKE ERIE
\
I
MILES 20
I
I
KM 20
"
o z
i:-Z

I


<'0 .
I
20_i
4
0
--+

Figure 5. Isopach of Upper Silurian F4 salt from Clifford
(1973) relative to the faults mapped by Gray (1982); faults
shown by heavy lines; C.l. =20'.
474
S. I. Root
Figure 6. Isopach of Lower
Devonian Oriskany sandstone
(after Diecchio, 1985); faults
shown by heavy lines;
C.l.=50'.
Figure 7. Isopach of Lower
Devonian carbonates (after
Mesolella, 1978); faults shown
by heavy lines; C.l. = 50'.
LAKE ERIE
LAKE ERIE
marine embayment north of the Transylvania zone
and suggested that the southern basin margin was
structurally controlled by the zone. Along the Ohio
River, Hook and Ferm (1988) mapped faults that are
part of the Highlandtown fault zone. Here, the
Allegheny Freeport coal shows syndepositional
movement on the fault; thick coals form on the
upthrown (NE) side of the fault, whereas limestones
accumulated in somewhat deeper water on the
downthrown (SW) side. The Conemaugh Ames
-'
..... -... 1
._._._._._._._._.J
o MILES 50
I I I I
o KM 50
o MILES 50
I I
I I
o KM 50
I
I
I
'-.
t.. .
'"
-'
....
)
Limestone was studied regionally by Brezinski
(1983). In southwestern Pennsylvania, a reentrant
of deeper water nodular limestone and calcareous
shale occurs south of the zone (Figure 9). The
northeast margin of the reentrant parallels the
Transylvania zone, consistent with syndepositional
south-side-down movement on the zone.
Some Devonian and Pennsylvanian stratigra-
phic thicknesses or lithofacies patterns in extreme
southwestern Pennsylvania may be due, in part, to
f
0
1
1
0
'\'
. .
.. .. . ... .... :
. .
. .
MILES 20
I
1
KM 20
Transylvania fracture zone and evolution of the Central Appalachian basin 475
Erosi ono 1
Edge .
\\ (\.1
L

,-.J
(,'"' ;!
' .;:::
->
J
activity of the Rome Trough, a major NE-trending
graben initiated in the Cambrian, that showed syn-
depositional movements into at least the Ordovician
(Wagner, 1976).
In the Broad Top basin of the central Valley and
Ridge an EIW-trending complex of Alleghanian
orogeny-related surface faults was mapped by Root
and Hoskins (1977) (Figure 1) and considered as
evidence of a pre-existing basement fracture zone.
They are a series of subvertical faults across which
semi-independent layer-parallel shortening occurs.
Wilson and Shumaker (1988) calculated 30% short-
ening in a profile in the Broad Top coal basin north
vf the zone, and 40% shortening in a profile south of
the zone. Depth to basement there is estimated to be
9 km (Kulander and Dean, 1986).
Figure 8. Distribution of two Lower Mississippian
sandstones (after Pepper and others, 1954). Lobes of Berea
sandstone >60' shown by stippled pattern; Cussewago
sandstone shown by 40' contour interval; faults shown by
heavy lines.
In the Valley and Ridge, a major detachment
horizon is present in Cambrian shales. The seq-
uence above is allochthonous, so basement fractures
may not propagate across an actively overriding
thrust sheet. Wheeler (1986) observed that linea-
ments in Devonian strata formed over basement
faults in West Virginia may not overlie those faults
because the thrust sheets on which they are located
were transported to the northwest. Because the
Transylvania zone is as much as 15 km wide, and
the angles between regional thrust transport and
the trend of the zone diverge by only 20, some EIW-
trending surface faults may still be located above a
portion of the basement fracture zone.
LAKE ERIE

!'l'l .

1
oj
I'
0 1
I
I
N.Y.
_._._._. _ ._ ._._._. _._._. _ ._ ._ ._._._. _.
PENNA.
m LIMESTONE
NODULAR
LI MESTONE
CALCAREOUS
SHALE
SHALE
---
"1
.
\
i
')
i
i
')
(
'-' )
I
.-'
, ' -.J
o MILE S 50
I I
1 )
o KM
I
'.
,)
Figure 9. Sedimentary facies
of Pennsylvanian Ames Lime-
stone (after Brezinski, 1983);
faults shown by heavy lines .
476 S. I. Root
Blue Ridge-Great Valley
Identification of structural activity on the Tran-
sylvania fracture zone in this region is based en-
tirely on analysis of the EfW-trending Carbaugh-
Marsh Creek and Shippensburg faults. This area is
an allochthonous structurally complex, cleavage-
dominated, W-verging anticlinorium exposing large
areas of Precambrian metavolcanic rocks in the
Blue Ridge fold core. A 46 km NW-transport of the
Blue Ridge is suggested from comparison of
predeformed and deformed lengths of a profile in
this region (Wilson and Shumaker, 1988). Several
kilometers of transport should be added to account
for shortening on the plateau, and this figure could
be somewhat larger if plastic flow and solution
effects are considered. Basement below the Blue
Ridge thrust sheet has been estimated to be at a
depth of 11 km south along strike in Virginia
(Kulander and Dean, 1986).
NE-trending Alleghanian age folds and faults
cannot be matched across the EfW-trending Car-
baugh-Marsh Creek fault, nor is the sense of appar-
ent displacement consistent, but structural geome-
try and kinematics are identical on both sides, in-
dicating a process of semi-independent shortening.
Furthermore, abrupt regional rotation of 20-30
about a vertical axis occurs at the fault; south of the
fault the structural grain is NI5E, whereas to the
north it is N45E (Root, 1970). These features are
consistent with a pre-existing zone of weakness in
these units at the time of Alleghanian deformation.
Across the Shippensburg fault, structures in the
Great Valley show a combination of simple offset as
well as semi-independent shortening. In the Blue
Ridge similar relations occur but, in addition, the
northern block appears to be uplifted 1300 m rela-
tive to the southern block. Uplift may be associated
with Jurassic rift-related faulting, not Alleghanian
shortening.
In a region such as this, where autochthonous
basement may be at a depth of 11 km, the problems
associated with the upward propagation of a base-
ment fracture are large because they involve pene-
tration of materials with high ductility contrasts.
Considering the estimated = 50 km of displacement
of the Blue Ridge thrust sheet, faults at surface
cannot overlie the basement fracture zone. If tec-
tonic transport is N700-75W, normal to the struc-
tural grain of NI5-200E, then the EfW-trending
fracture zone in basement is located 12-17 km south
of the faults at surface. Thus, the fracture zone
necessarily was present in the thick Precambrian-
Cambrian-Ordovician sequence prior to Alleghan-
ian thrusting and folding. Such reasoning is also
applicable in the Valley and Ridge.
Inferences from regional relationships indicate
other periods of activity on the Transylvania frac-
ture zone. In the Great Valley, 20 km southeast of
Harrisburg (Figure 1), are exotic lithologies in the
Ordovician Martinsburg Formation. These repre-
sent the southern limit of the extensive Taconic
nappes sequence in the Martinsburg shale basin.
Root and Hoskins (1977) suggested that the fracture
zone may have been active during the Ordovician
Taconic orogeny, forming a hinge line to a deep
basin on the north into which Taconic nappes were
emplaced. In the late Paleozoic, shortly after
regional folding and thrusting formed the Central
Appalachians, the plunging nose of the Blue Ridge
(Figure 1) was overridden on a sub horizontal thrust
by the Yellow Breeches thrust sheet of multiply
deformed lower Paleozoic rocks. This latest Alle-
ghanian deformational event may be related to
some form of gravity emplacement if the Tran-
sylvania fracture zone formed a hinge line with
structurally lower area on the north into which the
Yellow Breeches thrust sheet slid - a process com-
parable to that suggested for Taconic nappe em-
placement.
Gettysburg Basin and Piedmont
Jurassic extensional faulting formed the
Gettysburg half-graben, preserving Carnian-Norian
rift-related red beds that now dip 20-30 north-
westward toward a SE-dipping, basin-bounding
listric normal fault. Stratal dip results from post-
Pliensbachian monoclinal rotation and concomitant
major border faulting accompanied by igneous and
volcanic activity (principally as extensive saucer-
shaped sills of diabase). Locally adjacent to the
Shippensburg and Carbaugh-Marsh Creek faults,
wrench structures and basement, horst-like blocks
are present (Figure 10).
The basin border has a sharp 3.5 km right-
lateral offset where it intersects the Shippensburg
fault. In the Mesozoic basin, at this offset, is a
narrow, fault-bounded sliver of Cambro-Ordovician
(basement) limestones. En echelon with the fault
sliver is an EfW-trending fold produced by right-
lateral wrenching (Root, 1988). These structures
are considered transtensional and transpressional
Transylvania fracture zone and evolution of the Central Appalachian basin 477
features produced as the pre-existing Shippensburg
fault was locally reactivated as a right-lateral
wrench during the extension that formed the border
fault. The sense of wrenching derives from offset of
the border; however, the 3.5 km of border fault offset
should not be construed as the amount of wrench
displacement. With that amount of displacement,
an array of en echelon folds should be developed.
Instead, the offset is attributed to a combination of
right-lateral wrenching concomitant with normal
faulting that is oriented at a large angle to the
wrench (Root, 1988). The right-lateral motion may
be related to eastward regional arcuation of
Appalachians.
Within the Gettysburg basin, two large, NW-
trending, high-angle faults occur south of the
Carbaugh-Marsh Creek fault, offsetting the simple
pattern of a NE-trending border fault (Figure 10).
Transpressional wrenching occurred on these two
faults, producing several en echelon folds whose
geometry indicates a sense of left-lateral motion
(Root,1988). Complex interaction of wrenching and
normal faulting formed two small horst-like fault
blocks exposing Cambro-Ordovician limestone
(basement) adjacent to the border fault. The orien-
tation and sense of motion of these two faults
relative to the Carbaugh-Marsh Creek fault sug-
gests they may be antithetic.
The lack of offset of the border fault here
(unlike that at the Shippensburg fault) is possibly
related to the abrupt 25-30
0
change of Appalachian
structural grain across the Carbaugh-Marsh Creek
fault. The principal anisotropy of this grain is the
Blue Ridge cleavage parallel by the border fault.
Offset along the Carbaugh-Marsh Creek fault may
have been inhibited by cleavage convergence at the
fault and movement transferred to antithetic faults.
The structurally deepest part of the Gettysburg
basin is where the border fault is adjacent to the
Carbaugh-Marsh Creek and Shippensburg faults.
Vertical displacement there approaches 10 km.
Displacement rapidly decreases southward and the
basin terminates in Maryland. Basement rocks
northwest of the basin (including the EIW-trending
faults) behaved as a passive shoulder during forma-
tion of the half-graben. Nevertheless, within the
graben maximum fault displacement occurs along
the extension of these EIW-trending faults into the
basin. At this time in the geologic evolution of the
Central Appalachians, the original basement frac-
ture zone is located some 12-17 km south of the
Carbaugh-Marsh Creek fault so they do not reflect

0 MI L ES 10
I
I
I
0 KM 10
4000'
"

<Q
c,

-\.
-<..
c,
w
-<..
'Y-
<v
<Q
(!)
"
0
a:::
PENNA
MO
W
:::>
-J
al
'\


<v<V -0

0
0
l"-
I"-
Figure 10. Geometric relationships of Jurassic rift-related
normal faults (shown by hachured line) in the Gettysburg
basin. The reactivated Shippensburg fault offsets the
border fault and forms a fault sliver of basement lime-
stones (black area). The reactivated Carbaugh-Marsh
Creek fault (C-MC.F) does not offset the border fault
directly, but two associated left-lateral wrench faults do
offset it. The wrench faults generate a series of en echelon
folds and horst blocks of basement limestone (black areas).
Jurassic reactivation of this original structure.
Reactivation of the Carbaugh-Marsh Creek and
Shippensburg faults appears to be a local conse-
quence of Jurassic normal faulting and not a
causative component as in earlier structural events.
The fracture zone has not been recognized east
of the Mesozoic rift basin in the Piedmont area,
which is characterized by Precambrian granitic
rocks, belts of serpentinite, and metamorphosed
lower Paleozoic sedimentary and volcanic strata -
all of which were multiply deformed and involved in
large-scale nappe and thrust tectonics.
478 S. I. Root
Summary of Reactivation
On the Appalachian Plateau, recurrent move-
ment on the Transylvania zone affected sedimen-
tary distribution patterns in Cambrian to Pennsyl-
vanian strata. In most instances, fracture zone
activity is reflected by a positive area to the
northeast on which a thinner or shallower water
sequence was deposited relative to strata to the
southwest. However, in a number of instances the
opposite occurs.
In the Valley and Ridge Province, two regional
stratigraphic studies show intermittent reactivation
of the fracture zone during the Devonian. Surface
mapping defines a series of EIW-trending faults
caused by further reactivation during the terminal
Paleozoic deformation that formed this fold belt. In
the Blue Ridge-Great Valley, surface mapping and
regional relations suggest fracture zone activity
during the Taconic deformation and during the
main and late phases of Alleghanian deformation.
During the Jurassic continental breakup, the sur-
face faults of the basement fracture zone were zones
of mechanical anisotropy that interrupted develop-
ment of normal faults. Where normal faults in-
tersect the pre-existing faults, complex structures
developed as these earlier faults are "jostled" in the
extensional stress field. Post-Jurassic activity on
the fracture zone was probably minimal in this
highly stable region after formation of the Atlantic
passive margin. The area is presently aseismic.
REGIONAL RELATIONS
A number of deep wells in Ohio penetrated
Precambrian basement, predominantly granite
gneiss, deformed during the Grenville orogeny
(Lucius and Von Frese, 1988). Similar rocks occur
unconformably below late Precambrian metavol-
canic rocks in the core of the Blue Ridge anti-
clinorium in Maryland. Thus the Appalachian
basement is all part of the Grenville Province; hence
the fracture zone must originate from intraplate
processes (and not as a cryptic Grenville suture).
Because no wells penetrate to basement in the
Valley and Ridge, this assumption cannot be con-
firmed. Furthermore, on geophysical evidence King
and Zietz (1978) suggested a major NE-trending
crustal break in the Appalachian basin near the
western limit of the Valley and Ridge.
The origin of the fracture is still undetermined;
however, several plausible origins exist. Root and
Hoskins (1977) suggested that the fracture zone is
the EIW-trending failed arm of the late Precam-
brian triple junction proposed by Rankin (1976) to
account for the Appalachian salient centered on the
Blue Ridge in Pennsylvania. Shumaker (1986) con-
sidered the fracture zone to be the northern margin
of a rift that links, to the south, with the Rome
trough in northeastern West Virginia. Alterna-
tively, the fracture zone may be associated with a
late Precambrian transform fault that Thomas
(1977) located in the Pennsylvania Blue Ridge,
separating NE-trending spreading ridges that
formed the eastern margin of the Appalachians. In
this case, because the fracture zone is located on
continental crust, the position of the transform is
localized on oceanic crust, and therefore the fracture
zone pre-dates the transform.
Supporting a transform-related origin is the
lateral displacement of geophysical anomalies in the
basement by many kilometers across the fracture
zone (Lavin and others, 1982) and the geometry of
the Akron-Suffield faults, which suggest an original
right-lateral wrenching (Root and MacWilliams,
1986). Post-Precambrian movement along this fault
would be mainly vertical. Supporting a triple junc-
tion and failed arm origin is the petrologic and other
evidence that Rankin (1976) cited in the Blue Ridge.
Limited knowledge of Precambrian tectonics be-
neath the Appalachian basin does not permit specu-
lation beyond this.
ACKNOWLEDGMENTS
Critical review of this manuscript by B. Kulan-
der and E. W. Spencer is greatly appreciated. Ear-
lier versions were reviewed by C. Summerson and
M. Wilson. This study was supported by a Faculty
Development Grant from The College of Wooster.
REFERENCES
BREZINSKI, D. K., 1983, Development model for an Appa-
lachian Pennsylvanian marine incursion: Northeastern
Geology, v. 5, p. 92-99.
CALVERT, W. L., 1974, Sub-Trenton structure of Ohio with
views on isopach maps and stratigraphic sections as basis
for structural myths in Ohio, Illinois, New York, Penn-
Transylvania fracture zone and evolution of the Central Appalachian basin 479
sylvania, West Virginia, and Michigan: American Asso-
ciation of Petroleum Geologists Bulletin, v. 56, p. 957-972.
CLIFFORD, M. J., 1973, Silurian Rock Salt in Ohio: Ohio
Geological Survey, Report of Investigations 90, 42 p.
DIECCHIO, R. J., 1985, Regional controls on gas accumula-
tion in Oriskany sandstones, Central Appalachian Basin:
American Association of Petroleum Geologists Bulletin, v.
69, p. 722-732.
GRAY, J. D., 1982, Subsurface structure mapping in
eastern Ohio: U.S. Department of Energy, DOEIETI12131-
1399, p. 3.1-3.13.
GWINN, V. E., 1964, Thin-skinned tectonics in the Plateau
and northwestern Valley and Ridge Provinces of the Cen-
tral Appalachians: Geological Society of America Bulletin,
v. 75, p. 863-900.
HANSEN, M. C., 1984, Middle Carboniferous depositional
events: The Sharon Sandstone, a basal Pennsylvanian
quartzarenite in the North Appalachian Basin: Appala-
chian Basin Industrial Associates, Proceedings of the
Spring Meeting, no. 6, p. 148-173.
HARRIS, L. D., 1975, Oil and GfUJ Data from the Lower
Ordovician and Cambrian Rocks in the Appalachian
Basin: U.S. Geological Survey, Map I-917D.
HOOK, R. W., and J. C. FERM, 1988, Paleoenvironmental
controls on vertebrate-bearing abandoned channels in the
Upper Carboniferous: Palaeogeography, Palaeoclimato-
logy, Palaeoecology, v. 63, p. 159-181.
KING, E. R., and 1. ZIETZ, 1978, The New York-Alabama
lineament: Geophysical evidence for a major crustal break
in basement beneath the Appalachian Basin: Geology, v. 6,
p.312-318.
KOWALICK, W. S., and D. P. GOLD, 1976, The use of
LANDSAT-1 imagery in mapping lineaments in Penn-
sylvania; pp. 236-249. in R. A. Hodgson, P. S. Gay, and J.
Y. Benjamin (eds.), Proceedings of the First Inter-
national Conference on BfUJement Tectonics: Utah
Geological Association, Publication 5, 636 p.
KULANDER, B. R., and S. L. DEAN, 1986, Structure and
tectonics of Centraland Southern Appalachian Valley and
Ridge and Plateau Provinces, West Virginia and Virginia:
American Association of Petroleum Geologists Bulletin, v.
70, p. 1674-1684.
LAVIN, P. M., D. L. CHAFFIN, and W. F. DAVIS, 1982,
Major lineaments and the Lake Erie-Maryland crustal
block: Tectonics, v. 1, p. 431-440.
LUCIUS, J. E., and R. R. B. VON FRESE, 1988, Aero-
magnetic and gravity constraints on the crustal geology of
Ohio: Geological Society of America Bulletin, v. 100, p.
104-116.
MESOLELLA, K. J., 1978, Paleogeography of some Silurian
and Devonian reef trends: American Association of Petro-
leum Geologists Bulletin, v. 62, p. 1607-1644.
PEPPER, J. F., W. DEWITT, and D. F. DEMAREST, 1954,
Geology of the Bedford Shale and Berea Sandstone in
the Appalachian BfUJin: U.S. Geological Survey, Profes-
sional Paper 259, 119 p.
POPENOE, P., A. J. Petty, and N. S. Tyson, 1964, Aero-
magnetic Map of Western Pennsylvania and Parts of
EfUJtern Ohio, Northern West Virginia and Western
Maryland: U.S. Geological Survey, Geophysical Investi-
gations Map. GP445.
RANKIN, D. W., 1976, Appalachian salients and recesses:
Late Precambrian breakup and opening of the Iapetus
Ocean: Journal of Geophysical Research, v. 81, p. 5605-
5619.
RODGERS, M. R., and T. H. ANDERSON, 1984, Tyrone-
Mount Union cross-strike lineament of Pennsylvania: A
major Paleozoic basement fracture and uplift boundary:
American Association of Petroleum Geologists Bulletin, v
68, p. 92-105.
ROOT, S. I., 1970, Structure of the northern terminus of the
Blue Ridge in Pennsylvania: Geological Society of America
Bulletin, v. 81, p. 815-830.
__ , 1988, Structure and hydrocarbon potential of the
Gettysburg Basin, Pennsylvania and Maryland; pp. 353-
367 in W. Manspeizer (ed.l, TrifUJsic-JurfUJsic Rifting:
Continental Breakup and the Origin of the Atlantic
Ocean and PfUJsive Margins: Developments in Geotec-
tonics 22 (A and B), Elsevier, Amsterdam, The Nether-
lands, 998 p.
__ , and D. M. HOSKINS, 1977, Lat. 400N fault zone,
Pennsylvania: A new interpretation: Geology, v. 5, p. 719-
723.
__ , and R. H. MacWILLIAMS, 1986, The Suffield Fault,
Stark County, Ohio: Ohio Journal of Science, v. 86, p. 161-
163.
SHUMAKER, R. C., 1986, The effect of basement structure
on sedimentation and detached structural trends within
the Appalachian basin; pp. 67-81 in R. C. McDowell and L
Glover III (eds.l, Studies in Appalachian Geology
(Lowry Volume): Virginia Polytechnic Institute and State
University, Department of Geological Sciences, Memoir 3,
137p.
THOMAS, W. A., 1977, Evolution of Appalachian-Ouachita
salients and recesses from reentrants and promontories in
the continental margin: American Journal of Science, v.
277, p. 1233-1278.
WAGNER, W. R., 1976, Growth faults in Cambrian and
Lower Ordovician rocks of western Pennsylvania: Ameri-
can Association of Petroleum Geologists Bulletin, v. 60, p.
414-427.
____ , and W. S. LYTLE, 1976, Greater Pittsburgh
Region Revised Surface Structure and Its Relation to
Oil and Gas Fields: Pennsylvania Geological Survey, 4th
Series, Information Circular 80,20 p.
480 S. I. Root
WHEELER, R. L., 1986, Stratigraphic evidence for
Devonian tectonism on lineaments at Allegheny Front,
West Virginia; pp 47-66 in R. C. McDowell and L. Glover,
III (eds.), Studies in Appalachian Geology (Lowry
Volume): Virginia Polytechnic Institute and State
University, Department of Geological Sciences, Memoir 3,
137p.
WILSON, T. H., and R. C. SHUMAKER, 1988, Three-
dimensional structural interrelations within Cambro-
Ordovician lithotectonic unit of Central Appalachians:
American Association of Petroleum Geologists Bulletin, v.
72, p. 600-614.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Late Paleozoic destruction of the western proto-Atlantic margin
in the southern Appalachians
STEPHEN A. KISH
Department of Geology, B-160, Florida State University, Tallahassee, Florida 32306, USA
(received November 3,1988; revision accepted September 15,1989)
ABSTRACT
Basement rocks (Suwannee terrane) of Florida and adjacent Georgia are composed of undeformed,
unmetamorphosed lower Paleozoic shallow-water sedimentary rocks that rest upon late Precambrian-
Cambrian volcanic and plutonic rocks. Paleontological and stratigraphic evidence indicates a probable
affinity of this terrane with the West African rather than the North American craton. Several plate tectonic
models propose that the Suwannee terrane was sutured to North America during a late Paleozoic continental
collision. Delineation of this suture, based on both bore hole and geophysical data, provides no unique
location. Seismic reflection profiles in southern Georgia reveal a 50-km-wide zone of SE-dipping reflectors
interpreted to root at the suture zone; however, these reflectors are probably equivalent to other SE-dipping
reflectors (beneath the Piedmont and adjacent Coastal Plain of south-central Georgia and South Carolina)
probably associated with earlier Paleozoic structures. If basement reflectors observed in southern Georgia are
products of late Paleozoic suturing, their geometry indicates a subduction zone extended beneath the
Suwannee terrane before collision. Paleozoic subduction-related igneous or tectonic features are not known in
the Suwannee terrane; however, extensive late Paleozoic calcalkaline plutonic activity and more restricted
Barrovian-style regional metamorphism is known in the southern Appalachian Piedmont. A more consistent
model for late Paleozoic collision may involve an initial phase associated with subduction beneath the
Piedmont followed by collision of Africa and North America. The presence of a major embayment along the
southern margin of the Appalachian-Ouachita orogen allowed juxtaposition of the Suwannee terrane next to
North America by transform faulting rather than by direct collision.
INTRODUCTION
J. Tuzo Wilson (1966) proposed that Paleozoic
orogenic belts bordering the Atlantic Ocean were
products of multiple cycles of ocean basin creation
and destruction - now known as a Wilson cycle.
One of Wilson's main lines of evidence for a closing
cycle was juxtaposition of different faunal realms
within orogenic belts of eastern North America and
Europe. He considered this indicative of two widely
separated shelf regions being joined by the closing of
a proto-Atlantic Ocean. He also noted that a region
of undeformed Paleozoic rocks, covered by Coastal
Plain sediments of southwestern Georgia and north-
ern Florida, was one location where a shelf sequence
was present on the "eastern side" of the crystalline
Appalachians. He speculated that this region may
represent a fragment of African crust sutured to
North America. Subsequent paleontological studies
support African affinities for this region.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 481-489. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
481
482 S.A. Kish
A prominent magnetic anomaly in southwest-
ern Georgia (Brunswick magnetic anomaly, Figure
1) has been interpreted as an indicator of the actual
suture between Appalachian and African terranes
(Williams and Hatcher, 1983; Horton and others,
1984). Nelson and others (1985a,b) proposed a SE-
dipping suture associated with this anomaly.
The purpose of this paper is to present evidence
that constrains both the geometry of subduction
associated with the late Paleozoic collision between
North America and Africa, and the nature of the
boundary between Florida basement (African rocks)
and basement of the Appalachian orogen.
GEOLOGIC SETTING
Late Paleozoic Circum-Atlantic Orogenic Belts
Four major late Paleozoic orogenic belts border
the present Atlantic Ocean (Figure 1). The Appala-
chians of North America and the Hercynides of
Europe both underwent tectonic events during the
early and mid-Paleozoic. The Mauranitides of
Africa (Villeneuve and Dallmeyer, 1987) and the
Ouachitas of the south-central United States had a
relative quiet Paleozoic tectonic history until late
Paleozoic times. The Appalachians and Hercynides
also had intense igneous activity during late Paleo-
zoic orogenic events; the Mauranitides and Oua-
Mognel1c onomolle-s
@ Bruns'."c.
chita orogens had limited late Paleozoic plutonic
and metamorphic activity.
Southern Appalachian Orogen
The southern Appalachian orogen (Figure 2) is
divided into three major subdivisions: (1) a western
fold-and-thrust belt (Valley and Ridge), which con-
tains unmetamorphosed Cambrian- to Pennsyl-
vanian-age shallow-water marine and fluvial sedi-
mentary rocks; (2) the central Blue Ridge Province,
which contains Grenville-age basement and an
upper Precambrian-lower Paleozoic cover sequence
metamorphosed and multiply deformed during early
to mid-Paleozoic orogenic events, and which was
thrust westward over the Valley and Ridge during
the Alleghanian orogeny; and (3) an eastern terrane
of upper Precambrian to lower Paleozoic meta-
volcanic and plutonic rocks (Piedmont), which was
accreted to the North America craton during the
early to mid-Paleozoic. An extensive number of late
Paleozoic calcalkaline granite plutons occur in the
Piedmont (Figure 2) (Fullagar andButler, 1979).
The most evident aspect of the late Paleozoic
orogeny in the southern Appalachians is the major
thrust-and-fold belt of the Valley and Ridge. Blue
Ridge and Piedmont crystalline rocks are present in
higher level sheets of the thrust stack. Seismic
reflection studies (COCORP and U.S. Geological
Figure 1. Late Permian reconstruction
of circum-Atlantic orogens (modified
from Pindell, 1985): FS, Florida Stratis
block. Hercynides modified from
Windley (1984) and Laurent (1972),
Mojave-Sonoya megashear from
Anderson and Schmidt (1983).
E0'51 - \ es Af r' lcon COO'SI
--- - post-PaleOZOIC
'5 flke - SliP fClJlI zone'S
Foreland bouodOty
of late Paleozol4;
fold- t tl"lSI belts
OUACHI TAS


LOle Poleozolc
inlflJSIOn'S
Undeformed eorly TO
mid Poleozoic bO'S1ns
5B - Suwonnee basin
BB - bosin
Late Paleozoic destruction of the western proto-Atlantic margin 483
o
Q FOfeklnd .oId 8 ItltuSI
D Lore Poleozoic meiamorphic b@lIs
Lote Paleozoic plutons
D
Und efol med early 10 mid Poleozoic
sed imenlol y rOCkS
Unmelomorphosed late Precambrian
volcanic and pl ufonic rocks
L """';.,....:1 MetamorphiC rocks with
540- 560 Ma ag
-::::
Figure 2. Major geologic features of the southeastern
United States; volcanic rocks in southern Florida (shown
with question marks) are of uncertain age - either J uras-
sic or Eocambrian. Widely spaced horizontal lines repre-
sent subsurface Mesozoic basins; BMA, Brunswick
magnetic anomaly; ECMA, East Coast magnetic anomaly;
WU, Wiggins uplift.
NW
Survey) indicate that all of these thrusts may be
associated with a master decollement that extends
beneath the Piedmont (Cook and others, 1979; 1981;
Harris and others, 1981; Ando and others, 1983).
Areas of late Paleozoic prograde metamorphism are
restricted to narrow, fault-bounded anticlinorial
zones in the eastern portion of the Piedmont (Fig-
ure 2). During the late Paleozoic, much of the
Piedmont behaved as a rigid block deformed by both
thrust and strike-slip faults (Rast, 1984).
Thomas (1985) noted that the westward sub-
surface extension of the shallow-water stratigraphic
sequence of the Valley and Ridge grades into black
shales and sandstones that are similar to deep-
water sedimentary facies of the Ouachita fold belt.
The two orogens may be linked, but a major change
in structural trend must occur at the westernmost
portion of the Appalachian fold belt (Figure 1).
The southward extension of the Ouachita fold
belt is uncertain (King, 1973), but Ouachita facies
rocks can be traced into northern Mexico (Figure 1).
Anderson and Schmidt (1983) proposed that pre-
Mesozoic portions of the southwestern North Ameri-
can margin have several hundred kilometers ofleft-
lateral displacement along the Mojave-Sonora
megashear (Figure 1); hence, Paleozoic structural
features in this region may have been displaced by
Cordilleran tectonic movement.
Suwannee Terrane, South Florida and Georgia
Wells drilled into pre-Mesozoic rocks of north-
ern Florida, southwestern Georgia, and adjacent
W21 Mesozoic basin
Suwannee basin
Devonian
[::: .... ' .J Si l urian - Ordovician
Ordovician
Late Precambr ian
volcanic and p lutonic
rocks
Figure 3. Configuration of pre-Coastal
Plain rocks in subsurface of Florida and
adjacent parts of Georgia and Alabama
(modified from Barnett, 1975). Units with-
in Coastal Plain: T, Tertiary, U.C., Upper
Cretaceous, L.C., Lower Cretaceous. Ig-
neous rocks of southern Florida (circled
question marks) of uncertain age - either
Jurassic or Eocambrian.
484 S. A. Kish
Alabama encountered a sequence of rocks quite
different from metamorphic and plutonic rocks of
the adjacent Appalachian orogen. Barnett (1975)
and Chowns and Williams (1983) summarized data
obtained from these wells (Figure 3). The oldest
rocks are unmetamorphosed Early Cambrian to
Eocambrian (560-540 Ma) volcanic and plutonic
rocks that are nonconformably overlain by Early
Ordovician sandstones and Silurian to Middle
Devonian shales. No Paleozoic intrusive or volcanic
rocks are documented within the Paleozoic sedimen-
tary rocks, that are essentially undeformed and
unmetamorphosed and locally are cut by faults that
may be associated with Mesozoic rifting. Paleonto-
logical affinities (Cramer, 1971; Projeta and others,
1976) link these sedimentary units with rocks of
similar age in west Africa. Metamorphic rocks that
yield Pan-African radiometric dates occur in base-
ment of central Florida (Dallmeyer, 1989a).
Because of distinct stratigraphic and structural
differences between Paleozoic rocks of the Florida
subsurface and adjacent Appalachians, this region
was interpreted as a separate geologic province un-
related to the Appalachians. Using tectonic models
developed for the North America Cordillera,
Williams and Hatcher (1983) classified this Paleo-
zoic-Precambrian province as a suspect or exotic
terrane - the Tallahassee-Suwannee Terrane, and
the term Suwannee terrane is used here. Recent
studies (Williams and Hatcher, 1983; Nelson and
others, 1985a; Pindell, 1985; Venkatakrishnan, and
Culver, 1988; Dallmeyer, 1989a,b) consider the
boundary between the Suwannee terrane and the
Appalachians to be a late Paleozoic suture between
North America and Africa. The location of the
suture cannot be precisely constrained with avail-
able well data because of extensive, thick, subcrop
cover of Mesozioc rocks (Figure 2). For this reason,
geophysical data were utilized in several studies to
locate the possible suture.
SUTURE LOCATION - GEOPHYSICAL MODELS
A major gravity gradient within the southern
Appalachian Piedmont (Figure 4) has been inter-
preted to represent the boundary between Gren-
ville-age cratonic rocks of Laurentia and accreted,
allochthonous Piedmont rocks (Thomas, 1983).
COCORP profiles across the gradient (Figure 4)
reveal a 50-km-wide zone of SE-dipping reflectors.
On the northern COCORP line (line A in Figure 4),
reflectors were interpreted to be a sedimentary rock
sequence associated with the proto-Atlantic con-
tinental margin of North America overridden by
accreted terranes of the Piedmont (Cook and others,
1979; Ando and others, 1983; Thomas, 1983). In the
southern Georgia profile (Figure 4), the zone was
rooted at the late Paleozioc suture associated with
the Brunswick anomaly (Nelson and others, 1985a).
A prominent offshore magnetic anomaly (Fig-
ure 2) parallels the eastern Northern American
coast (East Coast magnetic anomaly) and is
interpreted to be produced by an edge effect between
continental and oceanic crust (Taylor and others,
1968). A similar magnetic anomaly is present off
the West African coast. The two anomalies are
nearly identical in shape (Figure 1) and may repre-
sent the original continental margins formed during
Mesozoic rifting and formation of the present
Atlantic Ocean (Roussel and Liger, 1983). The
Brunswick magnetic anomaly is a series of offshore
magnetic highs that parallel the East Coast magne-
tic anomaly (Figure 2). A WINW-trending magne-
tic high and low, present in southeastern Georgia, is
considered to be an onshore extension of the Bruns-
wick anomaly. Higgins and Zietz (1983) interpreted
the magnetic low as an independent magnetic
anomaly extending from offshore Georgia westward
to eastern Alabama (Altamaha anomaly). Horton
and others (1984) demonstrated truncation of NE-
trending magnetic anomalies of the Pedmont
against the Altamaha anomaly. In western Geor-
gia, the Piedmont structural grain appears to swing
subparallel to the anomaly.
Several tectonic reconstructions (Williams and
Hatcher, 1983; Horton and others, 1984; Nelson and
others, 1985a,b) place the Brunswick anomaly over
the late Paleozioc suture between Africa and North
America.
Tauvers and Muehlberger (1987) noted that the
southern Georgia COCORP profile crossed a struc-
turally complex area at the convergence of many
geologic features of different ages. In particular,
they noted that rocks of the Suwannee terrane are
present north of the Brunswick anomaly in south-
eastern Georgia. They offered an alternative to the
COCORP model that agrees with subsurface base-
ment data of Chowns and Williams (1983). In this
model, the Brunswick anomaly is the outboard limit
of the Appalachian terrane, but the actual suture
with the African terrane is north of the anomaly due
to overthrusting of African crust over North
American crust. Tauvers and Muehlberger (1987)
Late Paleozoic destruction of the western proto-Atlantic margin 485
also proposed that subsequent transform fault
movement offset the original suture. Recent studies
by Dallmeyer (1989b) show that Appalachian-age
metamorphic rocks occur south of the anomaly,
within the Wiggins uplift (Figure 2). This region
may have been offset by Mesozoic-Cenozoic faulting
associated with the Bahamas fracture zone
(Dallmeyer, 1989b).
Modeling of the Brunswick anomaly (McBride
and Nelson, 1988) demonstrated that the magnetic
high portion of the anomaly could result from
several geometric arrangements, including an
ancient continental margin, tectonic slabs of oceanic
crust, or sheet-like intrusions within Mesozoic
basins. Thus, no unique position for a suture is
established at this time.
Zones of inclined reflectors found in both
COCORP profiles (Figure 4) occupy a common
along-strike position. Reflectors in the northern
profile are attributed to early Paleozoic structural
features, those in the southern profile to late
Paleozoic suturing. Without independent geo-
chronologic constraints, the age and the tectonic
significance of the reflectors in the southern
COCORP profile are uncertain, and they may be the
products of an early or mid-Paleozoic tectonic event.
o
mGol
- 50
B z.
LATE PALEOZOIC SUTURING AND
SUBDUCTION POLARITY
One particularly difficulty in reconciling geo-
metric models oflate Paleozoic suturing (based upon
COCORP profiling and observed geologic features)
is that NW-transported thrust sheets of the Pied-
mont, Blue Ridge, and Valley and Ridge imply a
synthetic (parallel) relationship to the orientation of
the subducting plate. The presence of a late Paleo-
zoic, SE-dipping subduction zone beneath the
Suwannee terrane should have produced features
there associated with an active plate margin -
continental arc, igneous activity, accretionary
sedimentary wedges, mafic-ultramafic melanges.
These features have not been reported in the
Suwannee terrane or in basement near the Bruns-
wick anomaly.
The Piedmont of the southeastern United
States was the site of intense igneous activity
during the late Paleozoic (Fullagar and Butler,
1979; Sinha and others, 1989). Over forty late
Paleozoic granite to granodiorite plutons, several
reaching batholith size, form two arcuate belts
extending from Georgia to Virginia (Figure 2).
Most studies (Fullagar and Butler, 1979; Sinha and
-
Figure 4. A) Gravity and COCORP
profile through Georgia and South Caro-
lina (profile A on Figure 2) (modified
from Thomas, 1983); shaded pattern de-
notes major accreted terranes of the Pied-
mont. B) Gravity and COCORP profile
through southwestern Georgia and Flori-
da (profile B on Figure 2) (modified from
Nelson and others, 1985a). In both pro-
files the Brevard zone (B.Z.) is used as a
common reference point.
o 100 KM
L _____ --',
mGal 0
NO DATA
KM
-------------------
40

B
Coostal Plain
...
___ - --,- ?
486 S. A. Kish
Zietz, 1982; Sinha and others, 1989) considered
these plutons to have a nonnal, calcalkaline charac-
ter. Available geochronologic data (Figure 5a)
indicate post-Acadian magmatic activity through-
out the late Paleozoic, but it was most intense
"" 320-290 Ma. Most of these plutonic rocks are
felsic (Si0
2
>68%), although syenitic and gabbroic
plutons are also present (Figure 5b). Initial
87Sr/86Sr data (Figure 5c) indicate derivation from a
lower crustal or subcrustal source. The combined
chemical and isotopic characteristics of these
plutons are similar to subduction-related plutons of
the Sierra Nevada batholith (Bateman and Dodge,
1970; Kistler and Peterman, 1978). Absence of
extensive late Paleozoic volcanic rocks in the Pied-
mont is probably due to significant late Paleozoic
and early Mesozoic erosion. Within the southern
Appalachians, "" 10-15 km of cover was eroded
during this period of time (Jamieson and Beaumont,
1988). Few late Paleozoic Piedmont plutons are
similar to two-mica granites interpreted as products
of post-collisional anatexis in the Himalayas.
Voluminous calcalkaline magmatism in the
Piedmont and the absence of such magmatism in the
Suwannee terrane during the late Paleozoic is a
strong argument for a NW-dipping subduction zone
between these terranes, as also proposed by Sinha
and Zeitz (1982).
SOUT ER APPALACHIA S
LATE PALEOZOIC I TRUSI ES
350 300 250 Mo
INTRUSIVE AGES
b
DISCUSSION
Both the Appalachian and Hercynides orogens
are characterized by extensive mid- to late Paleozoic
calcalkaline plutonism. In contrast, the Maurita-
nides orogen was the site of a stable continental
shelf until a terminal defonnational episode that
produced large-scale, E-transported thrusts of cry-
stalline basement and the overlying cover sequence.
The Ouachita orogen was the site of deep-water
sedimentation until a single late Paleozoic orogenic
event .
Pre-drift geometry of the Appalachian-Maurita-
nides orogens is difficult to reconcile in tenns of a
single-stage late Paleozoic collisional event. The
Appalachians and the Mauritanides have large-
scale thrusts systems with symmetrically opposed
directions of transport (Figure 1). This contrasts
with collisional fold belts such as the Alps and the
Himalayas in that thrust and nappe development is
largely restricted to the subducting plate. An
alternative model may involve more than one
collisional event in the late Paleozoic. In the Cana-
dian Cordillera, at least two collisional events were
associated with the Mesozoic orogeny (Price and
Hatcher, 1983): (1) an initial orogenic phase asso-
ciated with an E-dipping subduction zone beneath a
large batholith belt; and (2) subsequent collision
with a narrow outboard continental terrace wedge
c
, 0
0700 0.705 0.710
(8
7
Sr / 86Sr )
0.715
Figure 5. Radiometric ages (a), silica content
(b), and 87Srf!
6
Sr intial ratios (c) for late
Paleozoic plutons of southern Appalachians;
from Butler and Ragland (1969), Fullagar
and Butler (1979), Merz and Sinha (1977),
Sando (1979), Kish (1983).
Late Paleozoic destruction of the western proto-Atlantic margin 487
Lourentio
ERCY IDEs
c=7

"
Gondwana
Late PaleOZOIC 'gneous terranes
Illlllllll Eorly to late PaleozoIC contlnen 01
margin sed men ory roo<S
c:::::J Poleozolc sedlmen or rocks 0
the S '.',onnee on r:rve bas ns
.J... . .L La e PaleOZOIC Benioff zone
Figure 6. Schematic reconstruction of late Paleozoic
collisional geometry between Laurentia and Gondowana;
heavy lines denote continental margins.
that produced a reversal of subduction polarity.
This was followed by an oblique collisional event
with small, accreted terranes.
Applying a Cordilleran model to late Paleozoic
events in the southern Appalachians would include
an initial stage (340-300 Ma) with widespread plu-
tonism and volcanism associated with a W-dipping
subduction zone (Figure 6). Collision with a small
microplate may have reversed subduction polarity
shortly before collision with the African continent.
Lack of plutonism within the Mauritanides may be
due to the brief time period between initiation of
subduction and the terminal collision that produced
extensive thrust development in the Appalachians
(=300-km-wide zone) and a much narrower zone
(= 150 km wide) in the Mauritanides.
The absence of orogenic effects in the Suwannee
terrane and the Bove Basin south of the Maurita-
nides (Figure 1) probably resulted from the pre-
collision geometry of the North American plate.
Thomas (1985) noted that a major southern Appala-
chian promontory-embayment pair, produced by
late Precambrian rifting, had a profound influence
on Paleozoic orogenic events. The Alabama pro-
montory, located near the southern terminus of
Appalachian structural trends (Figure 2), may have
acted as an indenter with respect to W-transported
arcs and continental plates. Much of the eastern
Ouachita orogen may have been the site of escape
tectonic movement associated with oblique colli-
sional events. Likewise, the EfW-trending boun-
dary between the Piedmont and Suwannee terranes
is more likely the site of a major transform fault
system, as proposed by Higgins and Zietz (1983) ,
rather than a direct collisional suture. This would
explain the lack of Paleozoic deformation within the
Suwannee terrane (Figure 6). Such a model is not
new, as it has been used in several previous plate
tectonic models describing the closing of the proto-
Atlantic (Helwig, 1973; Pindell, 1985).
ACKNOWLEDGMENTS
R. D. Hatcher, Jr. and A. 1. Odom provided
helpful comments on the manuscript. R. Raymond
and G. Lee assisted with the illustrations.
REFERENCES
ANDERSON, T. H., and V. A. SCHMIDT, 1983, The
evolution of Middle America and the Gulf of Mexico-
Caribbean Sea region during Mesozoic time: Geological
Society of America Bulletin, v. 94, p. 941-966.
ANDO, C. J ., F. A. COOK, J . E. OLIVER, L. D. BROWN, and
S. KAUFMAN, 1983, Crustal geometry of the Appalachian
orogen from seismic reflection studies; pp. 83-101 in R D.
Hatcher, Jr., H. Williams, and I. Zietz (eds.), Contribu-
tions to the Tectonics and Geophysics of Mountain
Chains: Geological Society of America, Memoir 158, 223 p.
BARNETT, R, 1975, Basement structure of Florida and its
tectonic implications: Gulf Coast Association of Geological
Societies Transactions, v. 25, p. 122-142.
BATEMAN, P. C., and F. W. DODGE, 1970, Variations of
major chemical constituents across the central Sierra
Nevada batholith: Geological Society of America Bulletin,
v. 81, p. 409-420.
BUTLER, J . R , and P. C. RAGLAND, 1969, A petrochemical
survey of plutonic intrusions in the Piedmont, south-
eastern Appalachians, U.S.A.: Contributions to Minera-
logy and Petrology, v. 24, p. 164-190.
CHOWNS, T. M., and C. T. WILLIAMS. 1983, Pre-
Cretaceous rocks beneath the Georgia Coastal Plain -
Regional implications; pp. L1-L42 in G. Gohn (ed.),
Studies Related to the Charleston, South Carolina
Earthquake of 1886 - Tectonics and Seismicity: U.S.
Geological Survey, Professional Paper 1313, 375 p.
COOK, F., D. ALBAUGH, L. BROWN, S. KAUFMAN, J . E.
OLIVER, and R D. HATCHER, Jr., 1979, Thin-skinned
tectonics in the crystalline southern Appalachians:
COCORP seismic-reflection profiling of the Blue Ridge and
Piedmont: Geology, v. 7. , p. 563-567.
488 S. A. Kish
COOK, F., L. BROWN, S. KAUFMAN, J. OLIVER, and T.
PETERSEN, 1981, COCORP seismic profiling of the
Appalachian orogen beneath the Coastal Plain of Georgia:
Geological Society of America Bulletin, v. 92, p. 738-748.
CRAMER, F., 1971, Position of the north Florida lower
Paleozoic block in Silurian time - Phytoplankton evi-
dence: Journal of Geophysical Research, v. 76, p. 4754-
4757.
DALLMEYER, RD., 1989a, Petrologic and geochronologic
linkages between the Rodelide orogen and the St. Lucie
metamorphic complex in the Florida subsurface: Im-
plications for terrane accretion: Journal of Geology, v. 89,
p.183-195.
__ , 1989b, 40 Ar/39 Ar ages from subsurface crystalline
basement of the Wiggins uplift and southwesternmost
Appalachian Piedmont: Implications for late Paleozoic
terrane accretion during assembly of Pangea: American
Journal of Science, v. 289, p. 812-838.
FULLAGAR, P. D., and J R BUTLER, 1979, 325-265 myoId
granitic plutons in the Piedmont of the southeastern
Appalachians: American Journal of Science, v. 279, p. 161-
185.
HARRIS, L. D., A. G. HARRIS, and W. DEWITT, Jr., 1981,
Evaluation of the southeastern overthrust belt beneath the
Blue Ridge-Piedmont thrust: American Association of
Petroleum Geologists Bulletin, v 65, p. 2497-2505.
HELWIG, J., 1975, Tectonic evolution of the southern
continental margin of North America from a Paleozoic
perspective; pp. 243-255 in A. E. M. Nairn and F. G. Stehli
(eds.), The Ocean Basins and Margins, Volume 3: The
Gulf of Mexico and the Caribbean: Plenum Press, New
Y ork, New York, USA, 706 p.
HIGGINS, M. E., and I. ZIETZ, 1983, Geologic inter-
pretation of geophysical maps of the pre-Cretaceous "base-
ment" beneath the coastal plain of the southeastern United
States; pp. 125-130 in R D. Hatcher, Jr., H. Williams, and
I. Zietz (eds.), Contributions to the Tectonics and Geo-
physics of Mountain Chains: Geological Society of Am-
erica, Memoir 158,223 p.
HORTON, W., 1. ZIETZ, and T. NEATHERY, 1984,
Truncation of the Appalachian Piedmont beneath the
Coastal Plain of Alabama: Evidence from new magnetic
data: Geology, v. 12, p. 51-55.
JAMIESON, R A., and C. BEAUMONT, 1988, Orogeny and
metamorphism: A model for deformation and pressure-
temperature-time paths with applications to the central
and southern Appalachians: Tectonics, v. 7, p. 417 -445.
KING, P. B., 1975, The Ouachita and Appalachian orogenic
belts; pp. 210-241 in A. E. M. Nairn and F. G. Stehli (eds.),
The Ocean Basins and Margins, Volume 3: The Gulf
of Mexico and the Caribbean: Plenum Press, New York,
New York, USA, 706p.
KISH, S. A., 1983, A Geochronological Study of Defor-
mation and Metamorphism in the Blue Ridge and
Piedmont of the Carolinas: PhD dissertation, University
of North Carolina, Chapel Hill, North Carolina, USA, 220
p.
KISTLER, R W., and Z. E. PETERMAN, 1978, Recon-
struction of Crustal Blocks of California on the Basis
of Initial Strontium Isotopic Compositions of Mesozoic
Granitic Rocks: U.S. Geological Survey, Professional
Paper 1071, 17p.
LAURENT, R, 1972, The Hercynides of south Europe, a
model: Proceedings of the 24th International Geological
Congress, Section 3, p. 363-370.
McBRIDE, J. H., and K. D. NELSON, 1988, Integration of
COCORP deep reflection and magnetic anomaly analysis
in the southeastern United States: Implications for origin
of the Brunswick and East Coast magnetic anomalies:
Geological Society of America Bulletin, v. 100, p. 436-445.
MERZ, B. A. and A. K. SINHA 1977, Rolesville batholith;
pp. B1-B7 in J. K. Costain, L. Glover III, and A. K. Sinha
(eds.), Evaluation and Targeting of Geothermal En-
ergy Resources in the Southeastern United States:
Virginia Polytechnic Institute and State University,
Department of Geological Sciences, Report of Investigations
VPI-SU-51 034-4 [ERDA], 101 p.
NELSON, K. D., J. A. ARNOW, J. H. McBRIDE, J. H.
WILLEMIN, J. HAUNG, L. ZHENG, J. E. OLIVER, L. D.
BROWN, and S. KAUFMAN, 1985a, New COCORP profiling
in the southeastern United States, Part I: Late Paleozoic
suture and Mesozoic rift basin: Geology, v. 13, p. 714-717.
__ , J. H. McBRIDE, J. A. ARNOW, J. E. OLIVER, L. D.
BROWN, and S. KAUFMAN, 1985b, New COCORP profiling
in the southeastern United States, Part II: Brunswick and
East Coast magnetic anomalies, opening of the north-
central Atlantic Ocean: Geology, v. 13, p. 718-721.
POJETA, J., Jr., J. KRIZ, and J. M. BERDAN, 1976,
Silurian-Devonian Pelecypods and Paleozoic Strati-
graphy of Subsurface Rocks in Florida and Georgia
and Related Silurian Pelecypods from Bolivia and
Turkey: U.S. Geological Survey, Professional Paper 879,
32p.
PRICE, R A., and R. D. HATCHER, Jr., 1983, Tectonic
significance of similarities in the evolution of the Ala-
bama-Pennsylvania Appalachians and the Alberta-British
Columbia Cordillera; pp. 149-160 in R. D. Hatcher, Jr., H.
Williams, and I. Zietz (eds.), Contributions to the Tec-
tonics and Geophysics of Mountain Chains: Geological
Society of America, Memoir 158,223 p.
Late Paleozoic destruction of the western proto-Atlantic margin 489
PINDELL, J. L., 1985, Alleghenian reconstruction and
subsequent evolution of the Gulf of Mexico, Bahamas, and
proto-Caribbean: Tectonics, v. 4, p. 1-39.
RAST, N., 1984, The Alleghenian orogeny in eastern North
America; pp. 197-217 in D. H. W. Hutton and D. J. San-
derson (eds.), Variscan Tectonics of the North Atlantic
Region: Geological Society of London, Special Publication
14,270p.
ROUSSEL, J., and J. L. LIGER, 1983, A review of deep
structure and ocean-continent transition in the Senegal
basin (West Africa): Tectonophysics, v. 9, p. 183-214.
SANDO, T. W., 1979, Trace Elements in Hercynian
Granitic Rocks of the Southeastern Piedmont, U.S.A.:
MS thesis, University of North Carolina, Chapel Hill,
North Carolina, USA, 91 p.
SINHA, A. K., and I. ZIETZ, 1982, Geophysical and geo-
chemical evidence of a Hercynian magmatic arc, Maryland
to Georgia: Geology, v. 10, p. 593-596.
__ , E. A. HUND, and J. P. HOGAN, 1989, Paleozoic
accretionary history of the North American plate margin
(central and southern Appalachians): Constraints from
the age, origin, and distribution of granitic rocks; pp. 219-
238 in J. W. Hillhouse (ed.), Deep Structure and Past
Kinematics of Accreted Terranes: IUGGIAGU Mono-
graph 50, 283 p.
TAYLOR, P. I., I. ZIETZ, and L. S. DENIS, 1968, Geological
implications of aeromagnetic data from the eastern con-
tinental margin of the United States: Geophysica, v. 33, p.
755-780.
TAUVERS, P. R., and W. R. MUEHLBERGER, 1987, Is the
Brunswick magnetic anomaly really the Alleghanian
suture?: Tectonics, v. 6, p. 331-342.
THOMAS, M. D., 1983, Tectonic significance of paired gra-
vity anomalies in the southern and central Appalachians;
pp. 113-124 in R. D. Hatcher, Jr., H. Williams, and I. Zietz
(eds.), Contributions to the Tectonics and Geophysics
of Mountain Chains: Geological Society of America,
Memoir 158, 223 p.
THOMAS, W. A., 1985, The Appalachian-Ouachita con-
nection: Paleozoic orogenic belt at the southern margin of
North America: Annual Reviews of Earth and Planetary
Sciences, v 13, p. 175-199.
VENKATAKRISHNAN, R., and S. J. CULVER, 1988, Plate
boundaries in West Africa and their implications for Pan-
gean continental fit: Geology, v. 16, p. 322-325.
VILLENEUVE, M., and R. D. DALLMEYER, 1987, Geo-
dynamic evolution of the Mauritanide, Bassaride, and
Rokelide orogens (West Africa): Precambrian Research, v.
37, p. 19-28.
WILLIAMS, H., and R. D. HATCHER, Jr., 1983, Appala-
chian suspect terranes; pp. 33-53 in R. D.Hatcher, Jr., H.
Williams, and I. Zietz (eds.), Contributions to the Tec-
tonics and Geophysics of Mountain Chains: Geological
Society of America, Memoir 158, 223 p.
WILSON, J. T., 1966, Did the Atlantic close and then
reopen?: Nature, v. 211, p. 676-68l.
WINDLEY, B. F., 1984, The Evolving Continents (2nd
Edition): John Wiley & Sons, New York, New York, USA,
399p.
INTERNATION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
The Pine Mountain window of Alabama: Basement-cover
evolution in the southernmost exposed Appalachians
MARK G. STELTENPOHL
Department of Geology, Auburn University, Auburn, Alabama 36849-5305, USA
(received December 27,1988; revision accepted April 20, 1989)
ABSTRACT
The Pine Mountain window is the most southern and internal Grenville basement massif in the
Appalachian Mountain chain that retains its primary miogeoclinal stratigraphic cover. Previously
metamorphosed billion-year-old granite gneisses have a modified stratigraphic contact with a younger
metasedimentary sequence interpreted to correspond with either the upper Precambrian Ocoee Supergroup or
Cambrian Weisner-Shady sequence of the Blue Ridge. Middle to upper amphibolite-facies assemblages in
cover rocks sharply contrast with granulite-facies assemblages reported for basement units, which were
retrograded to amphibolite-facies assemblages possibly as a result of thrust emplacement of the Inner
Piedmont terrane during the middle to late Paleozoic. Late Paleozoic right-slip faulting, characterized by
midcrustal-level mylonites with subvertical foliations and shallow NE-plunging elongation lineations, was
pervasive throughout the basement/cover complex in Alabama and apparently overprinted earlier thrust
zones. Right-slip faulting, possibly aided by ramping along lower, as yet unroofed thrusts, elevated the
basement/cover contact to its present-day high structural level. It is unknown at present how far displaced the
Pine Mountain window is from its original paleogeographic position along the ancient Laurentian margin.
INTRODUCTION
Billion-year-old basement gneisses (Odom and
others, 1973) and the cover sequence of the Pine
Mountain basement massif record the middle
Proterozoic to early Paleozoic evolution of the south-
eastern margin of the ancestral North American
craton and its subsequent orogenic overprinting.
The area containing the Pine Mountain window and
adjacent terranes in Alabama is of critical impor-
tance to understanding the Appalachian mountain
system for four principal reasons (see Figures 1
and 2).
It contains the southernmost exposed crystal-
line rocks in the orogen, which are concealed by Gulf
Coastal Plain sediments to the south.
It contains the orogen's most southern and in-
ternal Grenville basement massif that retains its
primary miogeoclinal cover sequence in the Pine
Mountain window.
Some of the orogen's widest mylonite zones (the
Brevard, Towaliga" Bartletts Ferry, and Goat Rock
fault zones) appear to merge into this area forming
the southern terminus of the eastern Piedmont fault
system (Hatcher and others, 1977; Steltenpohl,
1988a).
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 491-501. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
491
492
M. G. Steltenpohl
Much heated debate has arisen over COCORP's
suggestion that the Towaliga, Bartletts Ferry, and
Goat Rock fault zones represent rare exposures of a
portion of the proposed southern Appalachian
master decollement.
This report outlines the geologic setting of the
Alabama portion of the Pine Mountain window.
Emphasis is placed on the complex polyphase
structural evolution, particularly the kinematics of
recently recognized late Paleozoic right-slip ductile
deformation zones that lace the basement/cover
complex.
LITHOSTRATIGRAPHIC CONTEXT
Billion-year-old basement gneisses (Odom and
others, 1973, 1985) and distinct cover units of the
Pine Mountain window in Georgia are structurally
continuous with units in Alabama. The basement
complex comprises a group of variably mylonitized
granitic to tonalitic gneisses (Wacoochee Gneiss of
Bentley and Neathery, 1970). In Alabama, the Pine
Mountain Group cover sequence (Crickmay, 1933)
><
39 96
34
\
"
,
TENN
'\ KY
"
- --"
' I H' IIIITN..
BELT
"-._ ' -""1-'-
comprises, in ascending order, basal quartzite
(Hollis Quartzite), graphitic schist (Manchester
Schist) and dolomitic marble (Chewacla Marble),
which were interpreted by Odom and others (1973)
and Sears and others (1981) to correspond with the
late Precambrian to early Paleozoic miogeoclinal
cover sequence observed in the foreland. The con-
tact with underlying basement units is everywhere
tectonic in Alabama but is considered to be a modi-
fied nonconformity on the basis of analogy with
relationships described in Georgia (Clarke, 1952).
No reported fossils or isotopic data constrain the
depositional age of any Pine Mountain rock unit in
Alabama.
Some of the widest zones of mylonite (:s 10 km
wide) in the Appalachian Mountain chain mark the
northwest and southeast flanks of the Pine Moun-
tain window (Towaliga and Bartletts Ferry/Goat
Rock fault zones, respectively; Figures 1 and 2) and
juxtapose it with the Inner Piedmont and Uchee
terranes, respectively. The Inner Piedmont terrane
comprises a sillimanite-grade metavolcanic/meta-
sedimentary complex (Bentley and Neathery, 1970)
intruded by 369 Ma synmetamorphic Farmville
42 92
X
44 79
><
,
/
"-
"-


X
SC --... /
r------_- -'
92 32 Ne
100
X KILOMETERS
78 34
Figure 1. The Pine Mountain window is the most southern and internal Grenville basement massif in the Appalachian
orogen that retains its primary miogeoclinal stratigraphic cover sequence (from Hatcher, 1984). Black, Grenville basement.
Pine Mountain window of Alabama: Basement-cover evolution 493

l') '-------
Figure 2. Generalized tectonic map of
portions of the Alabama and south-
western Georgia Piedmonts (after
Bentley and Neathery, 1970; Schamel
and others, 1980; Sears and others,
1981). The Pine Mountain window is
situated between the Towaliga fault
and the Goat Rock fault zone. Inter-
pretive map of aeromagnetic linea-
ments projects contacts beneath the
Coastal Plain (see Horton and others,
1984).
0 Schists
C
Hollis Quart zi te

;;11& liD Mylonite zones
D Basement complex
GRFZ Uchee belt
Co .. "".IIt o""'. 0....n
,L."
Scale
o 10 20 30 40 ml
o 10 20 30 40 50 km
" Alleghanian $ulUre"
granites (Goldberg and Burnell, 1987). These litho-
logies probably represent an island arc with uncer-
tain pre accretionary relationships to the ancestral
North American craton (Neilson and Stow, 1986).
To the southeast, granitic and amphibolitic
gneisses, sillimanite-bearing schists, and migma-
tites of the Uchee terrane (Uchee belt of Bentley and
Neathery, 1970) constitute the remainder of the
crystalline rocks exposed in Alabama. Most
workers correlate these rocks with those of the Inner
Piedmont (Clarke, 1952; Bentley and Neathery,
1970; Schamel and others, 1980; Sears and others,
1981).
METAMORPHISM
Metapelitic cover rocks of the Pine Mountain
massif in eastern Alabama commonly contain
sillimanite- and/or kyanite-grade assemblages -
i.e., AI
2
Si0
5
-garnet-white mica-biotite-plagioclase-
quartz (Figure 3). However, aluminosilicates have
not been reported from cover units of the south-
western portion of the massif although these are
interpreted to be structurally continuous with the
aluminosilicate-bearing rocks to the north (Sears
and Cook, 1984). These observations may indicate
that rocks to the southwest are of lower metamor-
phic grade. This is consistent with calcite-dolomite
thermometric results on the Chewacla Marble
(Steltenpohl and Moore, 1988) that range from
""417C to 502C; eight of the ten specimens lie
between 440C and 489C. These temperatures are
consistent with assemblages observed in calcareous
10
II
W
cz:
,
'"
w
g:
300 400 600
TEMPERATURE (clog"" C)
Figure 3. Metamorphic conditions suggested for the Pine
Mountain window, after Bentley and Neathery (1970),
Schamel and others (1980), McDaniel (1984), Sears and
Cook (1984), and Steltenpohl and Moore (1988) . Ruled
lines indicate the range of temperatures estimated from
the Chewacla marble specimens (Steltenpohl and Moore,
1988).
494 M. G. Steltenpohl
rocks (i.e., calcite-dolomite-tremolite-quartz) and
metapelites (garnet-white mica-biotite-plagioclase-
quartz) and suggest that a wide range of P-T con-
ditions are recorded in Pine Mountain Group
metasedimentary rocks.
Lower to upper amphibolite-facies metamor-
phism in rocks of the Pine Mountain Group cover
sequence sharply contrasts with deeper crustal
granulite-facies conditions reported for some base-
ment units in neighboring Georgia (Clarke, 1952;
Schamel and Bauer, 1980). The granulite-facies
assemblages were retrograded to amphibolite-facies
assemblages apparently during prograde metamor-
phism of the Pine Mountain Group sedimentary
rocks. As Grenville basement massifs throughout
North America commonly contain relicts of
granulite-facies metamorphism (Wynne-Edwards,
1972), this event in the billion-year-old Pine
Mountain basement rocks is interpreted to have
occurred as a result of the Precambrian Grenville
orogeny (Odom and others, 1973, 1985). However,
neither granulite-facies mineral assemblages nor
Grenville ages are documented in basement units
from the Alabama portion of the Pine Mountain
window.
On the basis offield observations, most workers
suggested that amphibolite-facies metamorphism of
Pine Mountain Group rocks resulted from synmeta-
morphic over-thrusting of the Inner Piedmont
terrane (Schamel and others, 1980; Sears and
others, 1981). No isotopic age determinations for
metamorphism of the Pine Mountain Group are
reported, however, and consequently this event is
poorly constrained to be post-late Precambrian to
pre-Mesozoic. If, however, the Pine Mountain
Group rocks were metamorphosed as a result of syn-
metamorphic emplacement of the Inner Piedmont
terrane, then the 369 Ma Rb-Sr isochron on the
synmetamorphic Farmville granite may date this
metamorphic event (Goldberg and Burnell, 1987).
This implies that metamorphism accompanied
deformation of the Farmville granite during the
latest Devonian, which is consistent with isotopic
data reported for Inner Piedmont rocks in neigh-
boring Georgia (Dallmeyer, 1978; Odom and others,
1982; see Steltenpohl and Moore, 1988, for a
discussion). Limited conventional K-Ar hornblende
and biotite cooling ages from rocks of the Inner
Piedmont and Uchee terranes in Alabama range in
age from 323 to 303 Ma and 311 to 274 Ma, respec-
tively. This implies that temperatures remained
> 500C as late as 303 Ma and >300C until 274 Ma
(assuming temperatures for Ar retention in horn-
blende and biotite are 500C and 300C, respec-
tively; see Cliff, 1985), compatible with slow cooling
from the Late Devonian metamorphic peak. Tull
(1982, 1984), Tull and others (1985), Ferrill and
Thomas (1988), and Tull and Telle (1989) have
discussed the possible sedimentalogical responses of
rocks in the foreland to Devonian plutonism,
metamorphism and deformation in the Alabama
Piedmont.
The midcrustal-Ievel, middle to upper green-
schist-facies (Lawrence and La Tour, 1988; Stel-
tenpohl, 1988a) mylonite zones that bound the Pine
Mountain window have overprinted the dominant
schistosity in this and adjacent terranes (Figure 4)
and thus formed after the Late Devonian. Aero-
magnetic evidence indicates that Mesozoic diabase
dikes crosscut the shear zones (see Georgia Geo-
logical Survey, 1976; Bentley and Neathery, 1970),
thus providing a minimum age for this event.
Goldberg and Steltenpohl (1988) suggested a late
Paleozoic age for mylonitization based on a 2954
Ma Rb-Sr mineral isochron for a sample of tec-
tonized Farmville granite. The isochron was de-
fined by separates of plagioclase, potassium feld-
spar, and biotite, plus an analysis of the whole-rock
specimen. The whole rock sample plotted on the
previously reported 369 5 Ma Rb-Sr whole rock
isochron for the granite (Goldberg and Burnell,
1987). Because Faure (1986) reported that stron-
tium isotopic homogenization in minerals may occur
as a result of dynamothermal metamorphism while
the whole-rock system remained closed, the Farm-
ville mineral isochron could record the age of
penetrative deformation and associated greenschist-
facies metamorphism.
STRUCTURE
Several deformational phases are distinguished
in the study area using microscopic to macroscopic
relationships. These deformations can be broadly
divided into "early-phase" pre(?)- to syn-
amphibolite-facies metamorphic and "late-phase"
postmetamorphic deformations. Early-phase struc-
tures are interpreted to record the peak of meta-
morphism and emplacement of the Inner Piedmont
allochthon. These were succeeded by late phase
structures that record part of a complex, progressive
development of the right-slip Towaliga fault zone.
The same structural sequence is present in the
Pine Mountain window of Alabama: Basement-cover evolution 495
basement gneisses but the present database does not
allow for distinction of earlier structures in these
that are not present in the cover rocks.
Below, structural nomenclature is based on the
convention that structural elements produced
during a deformational event are numbered the
same. For example: 8
1
and Ll formed during D
1
; 8
2
and L2 formed during D
2
; and so on.
Early-Phase Deformations (D
I
)
The earliest recognizable deformational event,
Db is poorly understood. It apparently was a pro-
tracted event that initiated before and continued
sometime after the Devonian (?) amphibolite-facies
metamorphic peak. Microscopic evidence for early,
pre-D
1
(?) deformation is in the form of a weak align-
ment of fine-grained inequant mineral inclusions in
poikiloblasts, which are truncated by the external
foliation, 8
1
, defining the dominant schistosity.
This internal foliation is most common in Dl garnet
poikiloblasts where it is defined by inclusion trails
of quartz, white mica, biotite, plagioclase, dusty
graphite, and opaques. The foliation is interpreted
to represent an early-D
1
fabric but because it is
overprinted by kyanite/sillimanite-grade metamor-
phism and intensely developed D1 structures, the
significance of its occurrence is uncertain.
F 1 folds are rare but are shallowly NE- and 8W-
plunging and NW-verging. They are isoclinal,
subvertical to inclined, similar folds. F 1 folds con-
tain the dominant schistosity, 8
b
as an axial planar
foliation passing through their hinges. The folded
surface is always a compositional layering (8
0
), The
isoclinal form of F1 folds results in coplanar 8
1
and
8
0
foliations everywhere except in F 1 hinges and
may account for the oblique internal/external
foliations observed in D1 porphyroblasts.
Two prominent fabric elements were produced
during D
1
, the 8
1
schistosity and L1 lineation. 8
b
the primary lepidoblastic schistosity, is defined by
quartz, muscovite, biotite, plagioclase, garnet, horn-
blende, kyanite, and/or sillimanite. The most pro-
minant type of linear fabric, L
l
, is a dimensional
alignment of elongate minerals, primarily quartz
and hornblende. Little was ascertained about the
D
1
-transport direction, which is presumed to have
been top-to-the-NW on the basis of common NW-
verging F
1
-folds. This, however, remains to be
proven on independent grounds.
Figure 4. Photomicrograph of a billion-year-old micaceous
basement gneiss, perhaps showing vestiges of each
orogenic event - Middle Proterozoic (Grenville), middle
Paleozoic (Acadian?), and late Paleozoic (Alleghanian).
The large biotite grains forming the dominant sillimanite-
grade gneissosity are interpreted to have formed during
the Paleozoic. This gneissosity has been truncated by a
fine-scale, late Paleozoic sinistral shear band that dips
steeply to the left in the photograph; quartz grains were
dynamically recrystallized, producing an oblique grain
shape preferred orientation (GSPO). Gypsum plate is
inserted. Long dimension of photomicrograph is 5 mm.
Late-Phase Deformations
Late-phase structures are easily distinguished
because they deform structures formed during the
amphibolite-facies metamorphic peak (i.e., 8018
1
and L
1
). The late-phase deformations represent a
protracted event characterized by ductile deforma-
tion zones defined by phyllonites and mylonites,
496
M. G. Steltenpohl
which are described below. Various types ofF2 folds
were produced concurrent with this retrogressive
greenschist-facies metamorphic event. F2 folds
have variable styles, ranging from recumbent to
inclined, tight to isoclinal folds and inclined to
upright, tight to rather broad, chevron and kink-
like folds. F2 folds fold the dominant schistosity and
compositional layering (SO/Sl) and contain a shear-
band foliation or intense mylonitic foliation that
mayor may not be strictly parallel to hinge sur-
faces. D2 structures formed in response to nonco-
axial laminar flow as indicated by the common
occurrence of variably developed phyllonites (button
schists) in metapelites and strongly lineated and
foliated mylonites in quartzofeldspathic rocks.
F3 cross folds deform the mylonitic foliation and
have variable styles depending on lithology. The
mesoscopic folds range from broad, gentle warps in
gneisses and quartzites to tighter, often straight-
limbed, chevron- and kink-like styles in metape-
lites. The folds have shallow-plunging (0_20), NE-
and SW-trending hinges and are upright.
Shear Zone Structures
Pine Mountain window rocks of variable compo-
sition have been mylonitized, preserving numerous,
sub horizontal , dextral shear-sense indicators (Simp-
son, 1986). These are, in order of decreasing fre-
quency of occurrence, sigma and delta porphyro-
clasts, extensional shear bands, mica fish, lattice
(LPO) and grain shape preferred orientation (GSPO)
in quartz, asymmetric folds, false foliation, strain-
related growth of myrmekite, and S-C composite
planar fabrics (Figure 5; Steltenpohl, I988a). Mul-
tiple shear-sense indicators are commonly observed
in a single rock specimen, providing unequivocal
evidence for dextral shear. Quartz and feldspar
elongation lineations are generally sub horizontal
(Figures 6 and 7, respectively). Both quartz and
feldspar were dynamically recrystallized suggesting
middle to upper greenschist-facies conditions for de-
formation (see Simpson, 1985).
Figure 7 illustrates fabric elements measured
in mylonites from the investigated shear zones. The
schistosity symbol incorporates measurements of a
GSPO formed before mylonitization, S-planes in S-C
mylonites, the "backs" of mica fish (generally paral-
lel to [001]), and the "backs" of buttons in phylloni-
tic schists. The contoured plots represent both the
dominant mylonitic foliation in rocks where com-
posite planar fabrics were not observed, which was
generally the case, and C-planes in S-C mylonites.
Dextral shear also affected rocks of the Inner
Piedmont terrane and Pine Mountain window for
some undetermined distance along either side of the
Towaliga fault zone. To the south, ductile deforma-
tion zones, ranging in width from centimeters to
tens of meters, in basement gneisses contain dextral
S-C composite planar fabrics, unidirectional exten-
sional shear bands, asymmetric folds, and sigma
and delta porphyroclasts (see Steltenpohl, 1988a).
Deformational structures in quartz and feldspar are
identical to those observed in Towaliga fault zone
mylonites. The shear-zone boundaries are grada-
tional and parallel to the dominant steep NW-
dipping mylonitic foliation (Figure 7). Orienta-
tions of fabrics within the basement shear zones are
very similar to those observed in mylonites of the
Towaliga fault zone (cr, Figures 7a, 7b, and 7c).
Dextral shear affected rocks of the Inner
Piedmont terrane as far as 5 km northwest of the
Towaliga fault zone. Inner Piedmont button schists
contain pervasive unidirectional extensional shear
bands and sigmoidal quartz-feldspar composites.
Along the shear bands, minerals forming the kya-
nite/sillimanite-grade schistosity were retrograded,
resulting in mats of fine-grained biotite-quartz-
plagioclase-white mica and rare chlorite consistent
with greenschist-facies conditions for deformation.
Farmville granite intrusions into these schists are
mylonitized along their margins. Mylonitic fabrics
are asymmetric and reflect similar deformational
conditions. Shear fabrics in both rock types indicate
slightly oblique, dextral, down-to-the-NINW shear
along moderate- to steep-dipping NE-trending shear
planes (see Figure 7c).
Several map-scale movement-sense indicators
corroborate dextral shearing along the Towaliga
fault zone. First, the vergence, right-stepping en
echelon arrangement, and single wave train occur-
rence of late-phase folds affecting the basement-
cover contact in Alabama and Georgia are features
characteristic of a dextral wrench zone (Figure 8;
Wilcox and others, 1973). There appears to be a
decrease in the intensity of development of these
folds with distance away from the Towaliga fault
zone (Figure 9). Sears and others (1981) recognized
right-slip displacements along faults in the Inner
Piedmont, one of which appears to have offset the
kyanite/sillimanite isograd (see Sears and others,
1981; Steltenpohl and Moore, 1988). A frequent and
systematic repetition of slices containing the NW-
Pine Mountain window of Alabama: Basement-cover evolution 497
TOWALIGA FAULT ZONE
6 eIongallon Inntlons
fold axel
o .chl'foalty
eonlouted pbt, .,.
mytonfUc 'oIItJOn
BASEMENT SHEAR ZONES
N
c.
OPELIKA COMPLEX SHEAR ZONES
Figure 5. Extensional shear bands
(shallow dipping to the right) and
sigma-type potassium feldspar por-
phyroclasts in basement gneiss.
View is parallel to elongation linea-
tion and perpendicular to the sub-
vertical mylonitic foliation; north-
east is to the right.
Figure 6. L>S mylonite from the
Towaliga fault zone. Left-hand face
is cut perpendicular to the mylonitic
foliation and parallel to the elonga-
tion lineation, right-hand face is
perpendicular to both features, and
the upward-facing surface is parallel
to the mylonitic foliation. Scale bar
is 1cm.
Figure 7. Fabric observations (see
text); poles illustrated for all planar
fabrics: a) Towaliga fault zone, con-
tours 0%, 6%, 11% per 1% area, n =
122; b) shear zones in granitic base-
ment rocks of the Pine Mountain
window, contours 0%, 3%, 8% per 1 %
area, n = 24; c) shear zones in rocks
of the Inner Piedmont, contours 0%,
3%,5% per 1% area, n = 24.
498
M. G. Steltenpohl
A
N
,..... ,.....
"
20Km
1
----"
Figure 8. Comparision of megascopic structures associated with: (A) the Towaliga fault zone (see Figure 2 for location); and
(B) those of the wrench zone model of Wilcox and others (1973).
facing sequence, from bottom to top, basement-
Hollis Quartzite-cover schist has been documented
within the Towaliga fault zone (i.e., Pine Mountain
imbricate zone in Figure 2; Schamel and others,
1970; Sears and others, 1981). Steltenpohl (I988b),
on the basis of the geometry of the slices and shear-
sense studies within the sheared rocks, has inter-
preted these slices to reflect a right-slip contrac-
tional duplex (see Woodcock and Fischer, 1986, for a
discussion of strike-slip duplexes).
CONCLUSIONS
Early tectonic models for the Pine Mountain
window hinged upon the interpretation that the
Towaliga, Bartletts Ferry, and Goat Rock fault
zones represent the faulted and folded remains of a
NW-directed thrust zone (Clarke, 1952; Bentley and
Neathery, 1970; Cook and others, 1979; Schamel
and others, 1980; Sears and others, 1981; Nelson
and others, 1985, 1987). Subsequent detailed ob-
servations described for the northeast terminus of
the window (Hooper, 1986; Hooper and Hatcher,
1988), however, render this interpretation unten-
able. Hooper (1986) and Hooper and Hatcher (1988)
reported that fault zones rimming the northeastern
terminus of the window have polyphase structural
histories reflecting early, pre- or synthermal peak
thrusting followed by right-slip faulting. Results of
the present investigation document a similar struc-
tural history for the southernmost exposures of the
window, that is, WNW-directed thrust emplacement
of the Inner Piedmont allochthon followed by right-
slip faulting along fault zones flanking the window.
Consequently, earlier models must be modified to
incorporate late-stage dextral shearing, which ap-
pears to be a kinematic pattern common to internal
windows throughout the southern Appalachians
(see Secor and others, 1986).
REFERENCES
BENTLEY, R. D., and T. L. NEATHERY, 1970, Geology of
the Brevard Zone and Related Rocks of the Inner
Piedmont of Alabama: Alabama Geological Society, 8th
Annual Field Trip Guidebook, 119 p.
CLARKE, J. W., 1952, Geology and Mineral Resources
of the Thomaston Quadrangle, Georgia: Georgia Geo-
logical Survey, Bulletin 59, 99 p.
CLIFF, R. A., 1985, Isotopic dating in metamorphic belts:
Journal of the Geological Society of London, v. 142, p. 97-
110.
COOK, F. A., D. S. ALBAUGH, L. D. BROWN, S. KAUFMAN,
J. E. OLIVER, and R. D. HATCHER, Jr., 1979, Thin-skinned
tectonics in the crystalline southern Appalachians:
COCORP seismic-reflection profiling of the Blue Ridge and
Piedmont: Geology, v. 7, p. 563-568.
Pine Mountain window of Alabama: Basement-cover evolution 499
O. n=56 E. n 84


F. n- 87
EXPLANATI ON
m
---------------
t:
. iii
C
::J
a

OJ
Schists
Hollis Quart zite

o
Basement complex
Contours represent :
A. 1%,2'%4% per 4)/ 0 area
B. 1%2%-4% per % area
C. 1% 2/Qo _40/0, per '!o1Q. area
D. 2%-8%t4% per % area
E. 1
%
.4"/0 -7% per % area
F. 10/04'%70/0 per "% area
Figure 9. Stereoplots of poles to SoIS1 along the basement-cover contact in Alabama and Georgia. Contours: A, B, and Care
1 %,2%,4% per % area; D, 2%, 8%, 14% per % are; and E and F, 1%,4%,7% per % area. Plots A, B, and C incorporate Sears
(unpublished data) and the author's observations. Plot C incorporates data from McDaniel (1984). Plot D is after Higgins and
others (1986) and the author's observations. Plots E and F are after Higgins and others (1986).
CRICKMA Y, G. W., 1933, The occurrence of mylonites in
the crystalline rocks of Georgia: American Journal of
Science, v. 26, p. 161-177.
DALLMEYER, R. D., 1978, 4oAr/39Ar incremental release
ages of hornblende and biotite across the Georgia Inner
Piedmont: Their bearing on late Paleozoic-early Mesozoic
tectonothermal history: American Journal of Science, v.
286, p. 403-430.
FAURE, G., 1986, Principles of Isotope Geology: John
Wiley & Sons, New York, New York, USA, 589 p.
FERRIL, B. A., and W. A. THOMAS, 1988, Acadian dextral
transpression and synorogenic sedimentary successions in
the Appalachians: Geology, v. 16, p. 604-608.
GEORGIA GEOLOGICAL SURVEY, 1976, Geologic Map of
Georgia, 1:500,000 scale.
GOLDBERG, S. A., and J. R. BURNELL, 1987, Rubidium-
strontium geochronology of the Farmville granite, Ala-
bama Inner Piedmont; pp. 251-257 in M. S. Drummond and
N. L. Green (eds.), Granites of Alabama: Alabama
Geological Survey, Tuscaloosa, Alabama, USA, 343 p.
__ , and M. G. STELTENPOHL, 1988, Evidence for Alle-
ghanian penetrative deformation in the Inner Piedmont of
Alabama [abstract): Geological Society of America Ab-
stracts with Programs, v. 20, no. 4, p. 267.
HATCHER, R. D., Jr., 1984, Southern and Central Appala-
chian basement massifs; pp. 149-160 in M. J. Bartholomew
(ed.), The Grenville Event in the Appalachians and
Related Topics: Geological Society of America, Special
Paper 194,294 p.
__ , D. E. HOWELL, and P. TALWANI, 1977, Eastern
Piedmont fault system: Speculations on its extent: Geo-
logy, v. 5, p. 636-640.
HIGGINS, M. W., R. L. ATKINS, T. J. CRAWFORD, R. F.
CRAWFORD III, R. BROOKS, and R. B. COOK, 1986, The
Structure, Stratigraphy, Tectonostratigraphy and
Evolution of the Southernmost Part of the Appala-
500
M. G. Steltenpohl
chian Orogen. Georgia and Alabama: U.S. Geological
Survey, Open File Report 86-0372, 282 p.
HOOPER, R J., 1986, Geologic Studies at the East End
of the Pine Mountain Window and Adjacent Piedmont,
Central Georgia: PhD dissertation, University of South
Carolina, Columbia, South Carolina, USA, 374 p.
HOOPER, R J., and R D. HATCHER, Jr., 1988, Pine
Mountain terrane, a complex window in the Georgia and
Alabama Piedmont: Evidence from the eastern termina-
tion: Geology, v. 16, p. 307-310.
HORTON, W., 1. ZIETZ, and T. L. NEATHERY, 1984,
Truncation of the Appalachian Piedmont beneath the coas-
tal plain of Alabama: Evidence from new magnetic data:
Geology, v. 12, p. 51-55.
LAWRENCE, T. S., and T. E. LA TOUR, 1988, A comparison
of the Goat Rock and Bartletts Ferry faults, Georgia, using
thermobarometry [abstract]: Geological Society of America
Abstracts with Programs, v. 20, no. 4, p. 276.
McDANIEL, C. R, 1984, Geology of the Pine Mountain
Window and Adjacent Uchee Belt: MS thesis, Auburn
University, Auburn, Alabama, USA, 105 p.
NEILSON, M. J., and S. H. STOW, 1986, Geology and geo-
chemistry of the mafic and ultramafic intrusive rocks,
Dadeville belt, Alabama: Geological Society of America
Bulletin, v. 97, p. 296-304.
NELSON, K. D., J. A. ARNOW, M. GIGUERE, and S.
SCHAMEL, 1987, Normal-fault boundary of an Appala-
chian basement massif? Results of COCORP profiling
across the Pine Mountain belt in western Georgia: Geo-
logy, v. 15, p. 832-836.
__ , J. A. ARNOW, J. H. McBRIDE, J. H. WILLEMIN, J.
HUANG, L. ZHENG, J. E. OLIVER, L. D. BROWN, and S.
KAUFMAN, 1985, New COCORP profiling in the south-
eastern United States, Part I: Late Paleozoic suture and
Mesozoic rift basin: Geology, v. 13, p. 714-718.
ODOM, A. L., R D. HATCHER, Jr., and R J. HOOPER, 1982,
A premetamorphic tectonic boundary between contrasting
Appalachian basements, southern Georgia Piedmont [ab-
stract]: Geological Society of America Abstracts with
Programs, v. 14, p. 579.
__ , S. A. KISH, and P. J. LEGGO, 1973, Extension of
"Grenville basement" to the southern extremity of the
Appalachians: U -Pb ages of zircons [abstract]: Geological
Society of America Abstracts with Programs, v. 10, p. 196.
__ , S. A. KISH, and C. W. RUSSELL, 1985, U-Pb and Rb-
Sr geochronology of the basement rocks in the Pine
Mountain belt of southwest Georgia; pp. 12-14 in S. Kish,
T. A. Hanley, and Steve Schamel (eds.), Geology of the
Southwestern Piedmont of Georgia: Geological Society
of America, Field Trip Guidebook, 47 p.
SCHAMEL, S., and D. T. BAUER, 1980, Remobilized
Grenville basement in the Pine Mountain window; pp. 313-
316 in D. R. Wones (ed.), The Caledonides in the USA:
Proceedings of the Symposium: Virginia Polytechnic
Institute and State University, Department of Geosciences,
Memoir 2, 329 p.
__ , T. B. HANLEY, and J. W. SEARS, 1980, Geology of
the Pine Mountain Window and Adjacent Terranes in
the Piedmont Province of Alabama and Georgia: Geo-
logical Society of America, Southeastern Section Meeting
Guidebook, 69 p.
SEARS, J. W., and R B. COOK, 1984, An overview of the
Grenville basement complex of the Pine Mountain window,
Alabama and Georgia; pp. 281-287 in M. J. Bartholomew
(ed.), The GrenvUle Event in the Appalachians and
Related Topics: Geological Society of America, Special
Paper 194, 294 p.
__ , R B. COOK, and D. E. BROWN, 1981, Tectonic
evolution ofthe western part of the Pine Mountain window
and adjacent Inner Piedmont province; pp. 1-13 in J. W.
Sears (ed.), Contrasts in Tectonic Style Between the
Inner Piedmont Terrane and the Pine Mountain
Window: Alabama Geological Society, 18th Annual Field
Trip Guidebook, 61 p.
SECOR, D. T., Jr., A. W. SNOKE, and RD. DALLMEYER,
1986, Character of the Alleghanian orogeny in the south-
ern Appalachians, Part III: Regional tectonic relations:
Geological Society of America Bulletin, v. 97, p. 1345-1353.
SIMPSON, C., 1985, Deformation of granitic rocks across
the brittle-ductile transition: Journal of Structural Geo-
logy, v. 7, p. 503-511.
__ ,1986, Determination of movement sense in mylo-
nites: Journal of Geological Education, v. 34, p. 246-261.
STELTENPOHL, M. G., 1988a, Kinematics of the Towaliga,
Bartletts Ferry, and Goat Rock fault zones, Alabama: The
late Paleozoic dextral shear system in the southernmost
Appalachians: Geology, v.16, p. 852-855.
__ , 1988b, Geology of the Pine Mountain Imbricate
Zone, Lee County, Alabama: Alabama Geological Sur-
vey, Circular 136,15 p.
__ , and W. B. MOORE, 1988, Metamorphism in the
Alabama Piedmont: Alabama Geological Survey, Circu-
lar 138,29 p.
TULL, J. F., 1982, Stratigraphic framework of the
Talladega Slate belt, Alabama Appalachians; pp. 3-18 in D.
N. Bearce, W. W. Black, S. A. Kish, and J. F. TuB (eds.),
Tectonic Studies in the Talladega and CaroUna Slate
Belts, Southern Appalachian Orogen: Geological So-
ciety of America, Special Paper 191,164 p.
__ ,1984, Polyphase late Paleozoic deformation in the
southeastern foreland and northwestern Piedmont of the
Pine Mountain window of Alabama: Basement-cover evolution 501
Alabama Appalachians: Journal of Structural Geology, v.
6, p. 223-234.
__ , and W. R TELLE, 1989, Tectonic setting of
olistostromal units and associated rocks in the Talladega
slate belt, Alabama Appalachians; pp. 247-271 in J. W.
Horton, Jr., and N. Rast (eds.), Melanges and Olisto-
stromes of the U.S. Appalachians: Geological Society of
America, Special Paper 228, 276 p.
__ , J. F., D. N. BEARCE, and G. M. GUTHRIE (eds.),
1985, Early Evolution of the Appalachian MiogeocUne:
Upper Precambrian-Lower Paleozoic Stratigraphy of
the Talladega Slate Belt: Alabama Geological Society,
22d Annual Field Trip Guidebook, 92 p.
WILCOX, R E., T. P. HARDING, and D. R SEELY, 1973,
Basic wrench tectonics: American Association of Petro-
leum Geologists Bulletin, v. 57, no. 1, p. 74-96.
WOODCOCK, N. H., and M. FISCHER, 1986, Strike-slip
duplexes: Journal of Structural Geology, v. 8, no. 7, p. 725-
735.
WYNNE-EDWARDS, H. R, 1972, The Grenville Province;
pp. 265-334 in R A. Price and R J. W. Douglas (eds.),
Variations in Tectonic Styles in Canada: Geological
Association of Canada, Special Paper 11, 688 p.
INTERNATION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Tectonic implications of the brittle fracture history of the
Permian Narragansett Pier Granite, Rhode Island
MICHAEL J. HOZIK
Geology Program, Stockton State College, Pomona, New Jersey 08240, USA
(received August 12, 1988; revision accepted May 24, 1989)
ABSTRACT
The Permian Narragansett Pier Granite pluton underlies the southern coast of Rhode Island west of
Narragansett Bay and extends into eastern Connecticut. Detailed study of dikes, joints, faults, microjoints,
zones of microjoints, and quartz lenses permits (1) definition of domains, (2) determination of patterns of
brittle structures and their age relationships, and (3) inference of a stress history that records the transition
from Alleghanian ocean closure to Mesozoic rifting. Structural features arranged in the order in which they
formed are: (1) early aplite and pegmatite dikes; (2) early normal faults; (3) youngest set of pegmatite dikes;
(4) quartz lenses; (5) oblique-slip faults; (6) strike-slip faults; (7) late dip-slip faults; and (8) common joints,
microjoints, and zones of micro joints.
Early dikes appear to be controlled by local stress orientations apparently associated with intrusion of the
pluton. Early normal faults, the youngest set of pegmatite dikes, and the quartz lenses require stress systems
compatible with the last phase of Alleghanian deformation documented in the Narragansett Basin. The
motion sense for nearly all faults indicates a stress system with a N-S to NW-SE axis of minimum compressive
stress. These orientations are compatible with stress orientations hypothesized for this part of North America
during opening of the Atlantic. Common joints, microjoints, and zones of microjoints appear to be related to a
younger stress system with an E-W to ESE-WNW axis of minimum compressive stress. These features could
have developed along pre-existing weaknesses developed during the Jurassic or in response to stresses
associated with post-rifting uplift and tilting.
INTRODUCTION
The Narragansett Pier Granite (Nichols, 1956)
is an = 5 X 40 km batholith inferred to underlie the
southern coast of Rhode Island (Figure 1) (Quinn,
1971). The granite's Permian age (Zartman and
Hermes, 1987), rapid cooling (Dallmeyer, 1982;
Murray, 1988), and field relations indicate intrusion
during the latter part of Alleghanian deformation of
the Narragansett Basin (Mosher, 1983; Reck and
Mosher, 1988) and suggest that the granite recorded
the transition from Alleghanian compression to
Mesozoic rifting.
This study was undertaken to (1) determine the
fracture patterns in the Narragansett Pier Granite,
(2) infer the stress systems which produced these
patterns, and (3) correlate the inferred stress history
with models for the evolution of southern New Eng-
land. Throughout this paper, compressive stresses
are considered positive.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 503-525. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
503
504 M.J. Hozik
REGIONAL SETTING
The Narragansett Pier Granite occupies a
unique geographic location paralleling the coast of
North America along one of the largest trend
changes of the continental margin (Figure 2).
Figure 1. Geologic setting of the Narra-
gansett Pier Granit e and surrounding rocks.
Probable Carboniferous rocks include Narra-
gansett Basin, Scituate Basin (SB), Woon-
socket Basin (WB), Norfolk Basin (NB),
Worcester Basin (WoB), and Pin Hi1llocality
(PH) . Honey Hill Fault (HHF), Lake Char
Fault (LCF), Clinton-Newbury Fault (C-NF),
Bloody Bluff Fault (BBF), Hope Valley Shear
Zone (HVS), Beaverhead Shear Zone (BHS).
Modified from Snoke and Mosher (1989),
O'Hara and Gromet (1985), Reck and Mosher
(1988).
D

0
[2J
fa
EJ
Alleghenlan delOfmallon
InvolVing Perml .. nOun ard
Group of Western Pet.
Alleghanian Deformation
Hercynian Deformation
Fold Trends
Faults
Thrust Faults
Permian Inlruslons
Cobequid Fault
\\
,\:
,
,
,
,
Coincident with the coast line bend is a bend in the
trend of Appalachian structures. Drake and
Woodward (1963) postulated the existence of the
Cornwall-Kelvin fault connecting the Kelvin Sea-
mounts with a proposed offset of Appalachian trends
in New Jersey and Pennsylvania. They suggested
WoF
Q
0510 Graben
Permian
ilnd IfltrtJ.5IVeS
Urahan ThtuSIS
Figure 2. Narragansett Pier Granite in relation to major late Paleozoic tectonic elements on a pre-drift reconstruction by
LePichon and others (1977).
Brittle fracture history of the Permian Narragansett Pier Granite 505
that this fault had right-lateral offset of "'90 km.
Wise (personal communication, 1980) examined the
thickness of Eocambrian and Cambrian sedimen-
tary rocks in Pennsylvania and concluded that
major stike-slip motion along the Cornwall trend is
incompatible with the known distribution of these
strata. Rankin (1976) suggested that the bend was
the location of an aulacogen which formed during
the initial rifting ofIapetus. Thomas (1977) argued
that the area was the site of a transform fault/ocean
ridge junction. In either case, the implication is that
this bend in the Appalachians is a very old feature
(Precambrian), mantled by younger sediments, and
modified by multiple collisions with other land
masses.
This ancient continental margin bend could
well have influenced the location of successful
rifting and the development of a transform fault in
this area. The possibility of a Mesozoic transform
fault along the Kelvin Seamounts has been in-
timated by several workers (Drake and Woodward,
1963; Diment and others, 1972; Sbar and Sykes,
1973; Wise and others, 1975; Sykes, 1978; and
Fletcher, Sbar, and Sykes, 1978). Motion on the
transform fault would have been left-lateral, in
accord with Wilson's (1965) model for transform
faults. The seamounts formed later when a change
in spreading direction caused the transform to leak.
Age Relationships
The age of the Narragansett Pier Granite is
well constrained by the following:
A Pennsylvanian fossil (Brown and others,
1978) is contained in an inclusion in the pluton.
Radiometric age determinations are consistent
(Early Permian) - 276 Ma (Kocis and others, 1978),
2724 Ma (Hermes and others, 1981) and 2732
Ma (Zartman and Hermes, 1987).
Structural studies (Mosher, 1983; and Reck and
Mosher, 1988) indicate that intrusion began during
their third period of deformation (D
3
) and that at
least some granite dikes were deformed by their
fourth period of deformation and metamorphism
(D
4
), tightly placing the granite in the deforma-
tional history of the Narragansett Basin.
Argon release patterns (Dallmeyer, 1982) and
studies of metamorphism (Murray, 1988) indicate
that the granite and sediments it intruded had
cooled below argon retention temperatures of horn-
blende by 250-255 Ma (Late Permian) and of biotite
by 239-250 Ma (Early Triassic).
The age of the Narragansett Pier Granite
makes it one of a small group oflate Paleozoic, New
England granites that intruded Avalonian base-
ment. The Westerly Granite occurs as dikes in both
the Narragansett Pier Granite and older gneissic
rocks; it was radiometrically dated at 276 7 Ma.
The similarities in age, mineralogy, and geochemis-
try suggest the Narragansett Pier and Westerly
Granites are genetically related (Zartman and
Hermes, 1987). Aleinikoff and others (1979) re-
ported an age of ",275 Ma for the granite near
Milford, New Hampshire. Wintsch and Aleinikoff
(1987) also reported an age of 268 Ma for a granite
dike that intrudes Avalonian basement gneisses
(620 3 Ma) near Centerbrook, Connecticut. Other
Carboniferous granitic rocks intrude Paleozoic rocks
in New Hampshire and southern Maine (Hayward
and Gaudette, 1984; Aleinikoff and others, 1985).
Carboniferous and Permian magmatism is much
more widely distributed in the central and southern
Appalachians (Snoke and Mosher, 1989). Other
granites of similar age include Permian granites in
the Oslo Graben (Rutten, 1969) and granites on
Corsica (Maluski, 1976) (Figure 2).
Detailed studies of deformed metasediments in
the Narragansett Basin (Mosher, 1983; Reck and
Mosher, 1988; Murray, 1988) yielded a detailed
kinematic and thermal history. Cazier's (1987)
work on the Norfolk Basin and studies outside of the
basin in Rhode Island and Connecticut (O'Hara and
Gromet, 1983, 1985; Goldstein, 1982; Wintsch and
Aleinikoff, 1987) provided an earlier Alleghanian
framework. Radiometric dating (Zartman and
others, 1970; Zartman and others, 1988) docu-
mented the extent of Alleghanian thermal effects in
southern New England.
Pre-Carboniferous History
Many workers (Fairbairn and others, 1967;
Zarman and Naylor, 1972; Cameron and Naylor,
1976; Osberg and Skehan, 1979; Skehan and
Murray, 1979; Rast, 1980) have suggested that most
of southern New England is built on a late Precam-
brian Avalonian platform extending south and east
of the Bloody Bluff-Lake Char-Honey Hill fault
system. This platform is characterized by late Pre-
cambrian rocks intruded by '" 600 Ma granitic
bodies and overlain by Cambrian sedimentary rocks
506 M. J. Hozik
containing fossils with Acado-Baltic affinities
(Skehan and others, 1976; Skehan and Murray,
1978; Lyons, 1978). These rocks are typically less
deformed than rocks farther west in the orogen, they
lack Acadian metamorphism, and they are inter-
preted to have docked with the North American
continent some time in the Paleozoic.
O'Hara and Gromet (1985) subdivided the
southern New England Avalonian platform into two
distinct Avalonian terranes (Figure 1). The west-
ernmost (Hope Valley) terrane occupies most out-
crop areas of the Hope Valley alaskite, the Ten Rod
granite, and the Potter Hill granite gneiss of Quinn
(1971). It is characterized by its strong tectonite
fabric, totally recrystallized nature, and monoton-
ously leucocratic composition. The Esmond-
Dedham terrane to the southeast, separated from
the Hope Valley terrane by the Hope Valley shear
zone, contains rocks of widely varying compositions
which preserve primary igneous textures. The
Esmond-Dedham terrane includes most outcrop
areas of the Dedham, Milford, Esmond, and Pona-
ganset groups in northern and western Rhode Island
and Massachusetts. This terrane also includes
granites southeast of the Narragansett Basin. In-
trusive into the Esmond-Dedham Terrane is a group
of granitic rocks ranging in age from Ordovician to
Devonian (the Scituate, Quincy, Cape Ann, Rattle-
snake, and Peabody granites - Chute, 1969; Gates,
1969; Zartman and Marvin, 1971; Naylor and
Sayer, 1976; Lyons and Krueger, 1976; Hermes and
Zartman, 1985). O'Hara and Gromet (1985) inter-
preted their data to indicate that the Hope Valley
terrane collided with North America in the Acadian
but that the Esmond-Dedham terrane did not dock
until the Alleghanian.
Carboniferous Basins
The Carboniferous age of rocks in the Narra-
gansett Basin is based on floral evidence (Shaler
and others, 1899; Knox, 1944; Lyons, 1971, 1978;
Lyons and Chase, 1976; Oleksyshyn, 1976; Lyons
and Darrah, 1977). Plant fossils also establish the
Carboniferous age of both the Norfolk Basin and
rocks near Worcester, Massachusetts (Lyons and
others, 1976; Grew, 1976). Similar rocks of the
Scituate Basin, the Woonsocket Basin, and the Pin
Hill Locality near Harvard, Massachusetts, have
long been considered Carboniferous. All six Car-
boniferous basins in southern New England are
filled with terrestrial sediments with the exception
of some volcanic rocks in the Wamsutta Formation
in the Narragansett Basin (Quinn and Oliver, 1962;
Mutch, 1968; Cameron and Naylor, 1976; Snoke and
Mosher, 1989). In the Narragansett Basin, these
rocks were multiply deformed and metamorphosed
to sillimanite grade (Murray and Skehan, 1979;
Mosher, 1983; Murray, 1988; Reck and Mosher,
1988).
The deformational history, delineated by Mos-
her (1976, 1983) and Reck and Mosher (1988), is
summarized as follows:
Dl N- to NNE-trending, W-verging folds; shallowly
to steeply E-dipping axial surfaces;
D2 N- to NNE-trending, E-verging folds;
D3 sinistral shearing on NNE- to NE-trending
faults;
D4 dextral shearing on NE- to ENE-trending
faults.
Dl and D2 are attributed to E-W compression accom-
panied by prograde metamorphism (Mosher, 1983;
Reck and Mosher, 1988; Murray, 1988). Precise
timing of metamorphism and expression of various
fold subsets vary with location in the basin. D3 is
attributed to a sinistral shear couple (Mosher,
1983), clearly displayed along the Beaverhead shear
zone. Minor structures are also consistent with N-S
compression. D4 is related to N-S extension.
Post-Carboniferous Events
Post-Carboniferous events in the region sur-
rounding the Narragansett Pier Granite include
formation of large Mesozoic basins in the Hartford-
Deerfield and Newark Basins, a few small basins,
and graben structures on the continental shelf. The
large basins on land are bounded on one side by a
zone of en echelon normal faults, with the opposite
side acting as a hinge. DeBoer and others (1988)
suggest a complex stress history beginning with
Late Triassic-Early Jurassic rifting associated with
NW(WNW)-SE(ESE) extension, followed by rota-
tion of maximum compression from the vertical to a
horizontal N(NE)-S(SW) trend. Using Swanson's
(1982) suggestion, they attribute this stress to
Middle Jurassic clockwise rotation of Africa. They
point out, however, that evidence for compression is
not convincing and that similar features could be
produced by shearing due to a change in extension
direction relative to the grain of the Appalachians.
Their final phase is dominated by NW-SE to E-W
Brittle fracture history of the Permian Narragansett Pier Granite 507
compression, with much variation in stress orienta-
tion. Manspeizer and Cousminer (1988) suggest a
simpler history caused by an E-W sinistral shear
couple acting on a basement cut by old thrust faults.
The strain ellipse, elongated NW-SE, consistent
with their model is very similar to the strain ellipse
suggested by Ballard and Uchupi (1975) derived
from data in the Gulf of Maine. Both models require
NW -SE extension.
In addition to both lava flows and intrusives
within the Mesozoic basins, an extensive series of
Mesozoic mafic dikes occurs both inside and outside
the basins (May, 1971; McHone, 1978; McHone and
Butler, 1984). Two representatives of this series
may be present in Rhode Island: (1) in the Watch
Hill Quadrangle at the eastern edge of the town of
Westerly, where a mafic dike cuts a large dike of
Westerly Granite (Moore, 1967); and (2) in the Ash-
away Quadrangle == 1 km northwest of Chapman
Pond, where mafic dikes intrude the Narragansett
Pier Granite (Feininger, 1965; Hermes, 1987).
The White Mountain Magma Series of New
Hampshire is a NW -trending belt of young intru-
sions. Although the tendency has been to assign all
Mesozoic igneous rocks in southern New England to
this series, McHone and Butler (1984) and deBoer
and others (1988) argue for a more coherent group
(restricted to Jurassic plutons in or near New
Hampshire) with characteristic petrology. Foland
and Faul (1977) dated many intrusions of the White
Mountain Magma Series at 220-90 Ma (K-Ar). More
recent data (Doherty and Lyons, 1980; McHone and
Butler, 1984) allow separation of the Jurassic White
Mountain Magma Series rocks (200-155 Ma) from
Early Cretaceous intrusives related to the Montere-
gian Hills and New England Seamounts (125-100
Ma) (deBoer and others, 1988).
Regional Fracture Data
The faults critical to this study include the
Clinton-Newbury, Bloody Bluff, Lake Char, Honey
Hill, Hope Valley shear zone, and Beaverhead shear
zone, and have been extensively studied (Billings,
1929; Billings and Rahm, 1966; Dixon and Lund-
gren, 1968; Skehan, 1968, 1969; Castle and others,
1976; Barosh and others, 1977; Goldstein, 1980,
1982; Mosher, 1983; Eberley, 1985; O'Hara and
Gromet, 1985). With the exception of faults asso-
ciated with the Mesozoic basins, all these faults
were active prior to intrusion of the Narragansett
Pier Granite and therefore were pre-existing sur-
faces. Many of the faults have a pronounced NE
strike. Mosher (1983) and Reck and Mosher (1988)
documented motion on the Beaverhead shear zone
during and after intrusion of the Narragansett Pier
Granite. Reck and Mosher (1988) argued for motion
on the Hope Valley shear zone after intrusion of the
Narragansett Pier Granite but noted a conflict with
data of Getty and Gromet (1986). The current study
shows no evidence in the Narragansett Pier Granite
fracture data to support motion on the Hope Valley
shear zone after intrusion of the granite.
FRACTURE DATA
Summary of Age and Orientation Data (Table 1)
For analysis, the Narragansett Pier batholith
was divided into three domains (Figure 3): Narra-
gansett, Westerly, and Cormorant Point. Each area
is geographically distinct and seems to have a frac-
ture pattern different from the others.
Dikes
The oldest fracture elements are dikes of felsic
composition (Figure 4) - dikes are defined as any
tabular igneous intrusive cutting the granite. Dikes
intruded as a more or less continuous series and
hence have overlapping age relationships. The
chronological order for most dike orientations has
been established (Hozik, 1981). Given only six age
relations in the Cormorant Point domain, the
Cormorant Point and Narragansett domains were
merged for examination of age relations. Critical
relationships in the Narragansett/Cormorant Point
domain are as follows:
Most dike sets exhibit overlapping age relation-
ships and are grouped together as older dikes.
The youngest, and most abundant, NW-strik-
ing, SW -dipping set of dikes (labeled 3, Figure 4) is
the only set younger than the earliest set of faults.
The Westerly domain has four sets of dikes.
The early dikes probably formed in response to local
stresses as the granite cooled. After formation of the
early normal faults, the youngest and most abun-
dant set of dikes (EIW-striking, S-dipping) intruded.
These dikes are subparallel to the major Westerly
dikes (Figure 5) and suggest that 03 is steeply
NINE-plunging.
508
Figure 3. Map of the Narragansett,
Westerly, and Cormorant Point do-
mains, Narragansett Pier Granite.
M.J. Hozik

Table 1. Sequence of brittle fracturing
events in the Narragansett Pier Granite,
from oldest to youngest.
1)
2)
Intrusion of the batholith
Intrusionof several sets of early
dikes (mostly pegmatitic and
aplitic)
Formation of early normal faults
(Set l)
"".' Cormorant Point
. <-.:'."Narragansett Domain
3)
4)
5)
6)
Intrusion of youngest set of dikes
(mostly pegmatitic and aplitic)
Injection of quartz lenses
Formation of oblique-slip faults
(Set II with several subsets)
Formation of strike-slip faults
(SetIIl)
r.., ...... ;
Quartz Lenses
Quartz lenses are planar, quartz-filled features
closely resembling very small quartz veins except
that the ends are usually visible and invariably
pointed, with large aspect ratios of length to thick-
ness (100:1). Their lengths parallel to strike are
commonly several meters, with thicknesses of a few
centimeters. In the Narragansett Pier Granite, they
tend to be steeply dipping, with vertical extent
unknown.
Quartz lenses cut the youngest pegmatite dikes
(next in the age sequence), with possible overlap
between these two groups. The Narragansett
domain (Figure 4) is characterized by E-striking, S-
dipping quartz lenses. Twelve quartz lenses in the
Cormorant Point domain are indistinguishable in
orientation from quartz lenses in the Narragansett
domain. In the Westerly domain (Figure 4), two
sets of quartz lenses are nearly at right angles to
each other. The more prominent is a NE-striking,
NW-dipping set. Relevant age relations between
these two sets are unknown. Because the quartz
lenses appear to have formed in extension, they
imply a N-S axis of minimum compressive stress in
the Narragansett and Cormorant Point domains.
The more widely distributed orientation in the
Westerly domain indicates a NE-SW orientation of
Domain
7)
8)
9)
10)
Formation oflate dip-slip faults
(Set IV)
Opening of common joints
Opening of micro joints and zones
of microjoints
03, which leaves the NE-striking set unaccounted
for. N-S-extension in the Narragansett domain may
well be correlative with late-stage boudin-related
quartz veins and other features indicative of late N-
S extension in the Narragansett Basin (Mosher,
1983; Mosher, personal communication, 1988).
Faults
The term fault was restricted to features on
which motion parallel to the surface could be
demonstrated either by offset of dikes cut by the
fault or by the presence of slickensides. Where
slickensides were observed, their attitudes were
measured and the attitude of the line lying in the
fault plane perpendicular to the slickensides was
measured and recorded as the attitude of the
rotation axis. Wherever possible, relative motion
sense was recorded along with its method of deter-
mination.
Only quartz mineralization was detected along
faults and no deformation or drag adjacent to them
was mesoscopically visible. Most fault surfaces
developed an iron oxide coating (goethite and clay)
on weathering. There are two exceptions: (1) a
group of left-lateral and normal oblique-slip faults
(Set IIa, below) in the Narragansett domain occur
Brittle fracture history of the Permian Narragansett Pier Granite 509
along pegmatitie dikes, and dike minerals usually
preserve excellent slickensides; and (2) a few faults
in the Westerly domain are mineralized with fluo-
rite, chlorite, epidote, and sulfide, but these were
not recognizable as a group with parallel orienta-
tions or motions.
The origin offaults in the Narragansett domain
is the most clearly understood (Figure 6) . These
faults are classified into four groups based on
motion sense, orientation, and relative age. The
oldest faults (Set I) are NE-striking and SE-dipping
Figure 4
Top: Poles to all dikes in
Westerly, Narragansett,
and Cormorant Point do-
mains. Orientation
labeled 3 refers to orienta-
tion of youngest set of
dikes in Narragansett
domain.
Middle: Poles to quartz
WESTERLY
nl11
and they generally have normal dip-slip motion,
locally combined with a small right-lateral compon-
ent. These faults are more irregular than other
faults, and quartz mineralization is more extensive.
They are offset by all other groups and are the only
group of faults cut by dikes. A member of the young-
est group of dikes cuts these faults at two stations.
Set II, the next younger group of faults, has
oblique-slip motion and is divided into four sub-
groups with unknown age relations (lIa, lIb, lIe, and
lId: Figure 6). [For more information, see Hozik
rOPllORP.lIT PO \lIT
n ')96 n
"
lenses.
contours S\ ("on' ours ..; ,8.10
Bottom: Poles to faults.
Roman numerals place
faults in chronological
order
I, early, normal faults
II, various groups of
oblique-slip faults
III, strike-slip faults
n 2l
IV, late dip-slip faults
n 91
n
"
('onlOUI!!l 2 .....
IVa Ite
n 1"12
contour$
12
n .. 83
COntours 2. S\
510
M. J. Hozik
-r--r 25 7
12 37
,...- 357

4222,5 17
- __ __ =c+--:-;--c28cc=:c;-- 65
( -,---),.... _ 15-25 ,,-,22'/2
/ 45 --....... i ..
(
_. __ ' IMILE
I
Figure 5. Bodies of Westerly Granite in the Westerly-
Bradford area, southwestern Rhode Island (from Quinn
and Moore, 1968).
(1981), his figures 56-63 in Appendix II.] Subgroup
IIa (NW-striking, SW-dipping) is characterized by
left-lateral normal motion and is coplanar with the
youngest dikes. Subgroup IIb (WINW-striking, SW-
dipping) is characterized by left-lateral and normal
motion. Subgroups IIc and IId are ENE-striking
and dip steeply. Those which dip northward are left-
lateral and normal (Set IIc); those that dip south-
ward are left-lateral and reverse (Set IId).
Set III faults (WNW- to W-striking) exhibit
strike-slip motion (dominantly left-lateral) and
have moderate to steep dips. In the field, oblique-
slip faults along pegmatite dikes show an oblique-
slip set of slickensides overprinted by a more
shallowly plunging stike-slip set. suggesting that
oblique-slip motion graded into strike-slip motion.
The youngest set of faults (Set IV) are EIW - to
ENE-striking with steep dips. The motion is dom-
inantly north-side-down with a small left-lateral
component of motion as well, but some show a right-
lateral component.
The same age sequence determined in the
Narragansett domain is inferred in the Cormorant
Point domain where 82 faults were examined, but
age relations among various fault groups were not
independently established because of difficulty in
finding intersections. Elements of the NE-striking
set (Set I) occur at three stations. This set is asso-
ciated with faults of similar orientation but with
strike-slip motion. A few EIW-striking strike-slip
faults (Set III) are present as are a few of the late
Early Normal Faults
Set ;
Left Lateral a Normal Faults
with pegmatite
Set IIa
\sl

\\\"
"
Lett Lateral a Normal Faults Left Lateral a Normal Faults
(South dip) (North dip)
Set lIb Set I Ie
,
-,-
Left Lateral a Reverse Faults Left Lateral Strike-Slip Faults
Set lId Set III
Late Dip-Slip Faults
(Normal)
Set IVa

Late Dip-Slip Faults
(Reverse)
Set I Vb
Figure 6. Stress orientations inferred from faults in the
Narragansett and Cormorant Point domains. Solid great
circles = fault plane; dashed great circles = 01-03 plane; sl
= slickensides.
Brittle fracture history of the Permian Narragansett Pier Granite 511
dip-slip set (Set IV). The Cormorant Point domain
also has a group ofN-striking faults.
Three distinct periods offaulting are inferred in
the combined Narragansett and Cormorant Point
domains. Assuming that the youngest set of dikes
cuts the oldest set of faults and that the dikes are
related to the last stages of magma crystallization,
then the first period of faulting is associated with
the last stages of magma crystallization. The oldest
faults are early normal faults (Set I), indicative of a
minimum compressive stress with a moderate
plunge. This implies NW-SE extension (Figure 6),
which is roughly consistent in both age and orienta-
tion with NE-striking normal faults with a right-
lateral component (D
4
of Mosher, 1983; Mosher,
personal communication, 1988). Because D3 of the
Narragansett Basin involved sinistral motion on
NE-trending faults and E-W extension (Mosher,
1983), early normal faults may record stresses tran-
sitional between the third and fourth deformational
phases.
The second period of faulting, which post-dated
the last dike injections, started with oblique-slip
motion on Set IIa-d but gradually became domin-
ated by strike-slip motion (Set III). Left-lateral and
normal faults may well be the oldest in the second
period of faulting. Inspection of plots of inferred
stress orientation for left-lateral and normal faults
localized along pegmatites (Set IIa) and for a similar
group (Set lIb) not along pegmatites indicates
nearly identical orientations of the intermediate
principal stress (see Figure 6). The orientations of
01 and 03 are based on the assumption of a 30 angle
between the fault and 01 which may not have been
true if the pegmatites were still soft enough to act as
a planar weakness.
If Subgroups IIc and lId are older than the
strike-slip faults (Set III), then 01 became tilted
down to the southwest as inferred from steeply
dipping oblique-slip faults (see Figure 6). The
maximum compressive stress(01) would have rota-
ted back to nearly horizontal to produce the strike-
slip faults. This implies that while the orientation
of 03 remained fairly stable, similar magnitudes of
01 and 02 could permit orientation changes easily.
The youngest faults were E-striking, steeply
dipping dip-slip faults (Set IV) lacking a conjugate
relationship but consistently north-side-down. This
implies an EINE-trending intermediate principal
stress ( Figure 6).
In the Westerly domain, the fault pattern is
much simpler; virtually all faults are NE-striking
and moderately to steeply E- or W-dipping (Figure
4). Although they are divided into classes based on
motion sense (Table 2), age relationships were
rarely observed because the faults were coplanar.
Of 76 faults on which both the type and sense of
motion are known, the majority (54%) are oblique-
slip. Faults with a reverse component of motion are
minor. The faulting is dominantly oblique-slip with
a normal component and a slight preference for
right-lateral versus left-lateral.
Normal faults in the Westerly domain indicate
that 01 was vertical while 02 and 03 were horizontal
and SW-trending and NW-trending, respectively.
These orientations are not very different from those
deduced in the Narragnsett domain for early normal
faults (Set I) (compare Figures 6 and 7).
In the Narragansett domain, faults with left-
lateral and normal components of motion (Set II)
seem to have formed second. The only faults in the
Westerly domain that might correlate with them
are those with left-lateral and reverse components
of motion (Figure 7). The basis of this correlation is
that both require a SW-trending 02. If the correla-
tion is correct, curving stress trajectories or differen-
tial tilting between the two ends of the intrusion are
required to enable 01 and 03 to coincide in the two
areas.
In the Westerly domain, the right-lateral
strike-slip faults and faults with right-lateral and
normal components require the same orientations of
principal stresses. A similar stress orientation is
required for left-lateral strike-slip faults in the
Narragansett domain, suggesting a counterclock-
wise rotation of 01 and 03 about 02. Curving tra-
jectories may be responsible, and this may indicate
that right-lateral strike-slip faults in the Westerly
and left-lateral srike-slip faults in Narragansett
domains are conjugate (Figures 6 and 7).
The remaining faults (three sets) in the Wester-
ly domain (Figure 7) are not correlated with faults
in the Narragansett domain.
If orientations of principal stresses (Figures 6
and 7) are plotted separately (Figure 8), then all
faults in the Narragansett domain and most in the
Westerly domain require a NW/SE-oriented 03
(Figure 8). Both 01 and 02 were rotated about 03 to
produce the various types of fault motion. This
implies a stable orientation of 03, perhaps because it
was very different in magnitude from 01 and 02.
512 M. J. Hozik
Table 2. Statistical distribution of fault motions.
Fault Type
Left-lateral strike-slip
Left-lateral and normal
Left-lateral and reverse
Right-lateral strike-slip
Right-lateral and normal
Right-lateral and reverse
Normal
Reverse
Strike-slip
Oblique-slip
Dip-slip
Motion unknown
Total
Number offaults with completely known motions
Sum of all pure strike-slip
Sum of all pure oblique-slip
Sum of all pure dip-slip
Sum of all with left-lateral component
Sum of all with right-lateral component
Sum of all with normal component
Sum of all with reverse component
Westerly
No.
9
9
11
12
19
2
12
2
6
13
4
4
103
76
21
41
14
29
33
40
15
%
9
9
11
12
28
2
12
2
6
13
4
4
102
28
54
18
38
43
53
20
;-:-
!
Narragansett
No. %
71 20
103 29
22 6
4 1
29 8
8 2
69 20
13 4
4 1
16 5
11 3
3 1
353 100
319
75 24
162 51
82 26
196 61
41 13
201 63
43 13
51
Cormorant
Point
No. %
12 15
24 29
4 5
0 0
19 23
3 4
7 9
2 2
0 0
3 4
2 2
6 7
82 100
71
12 17
50 70
9 13
40 56
22 31
50 70
9 13
Normal Faults Left Lateral 6 Reverse
Faults
Left Lateral Strike-Slip Faults Left Lateral 8 Normal Faults
Set I
Set IIa
Right Lateral 8. Normal Faults Right Lateral Strike-Slip
Feu Its
Set I I b Set I I I
Left Lateral 8. Reverse Faults
Figure 7. Stress
orientations inferred
from faults in the
Westerly domain.
Solid great circles =
fault plane; dashed
great circles =
01-(}3 plane; sl =
slickensides.
Brittle fracture history of the Permian Narragansett Pier Granite 513
Common Joints
Shear fractures and faults of very small dis-
placement are not included here as joints; thus,
joints are interpreted as forming perpendicular to
minimum compressive stress, assuming an isotropic
homogeneous rock.
In the Narragansett domain are three promin-
ent sets of joints: (1) ""N15E-striking, steeply W-
dipping; (2) "" N800W -striking, steeply N - or S-dip-
ping; and (3) N45E-striking, steeply NW-dipping.
The first and second sets are equally well developed
throughout the whole domain (Figure 9), although
each is better developed in the southern and north-
ern parts, respectively. Members of the second set
abut members of the first set. The third set, found
mainly in the southern part of the area is less well
developed,. Age relationships between the third set
and the first two joint sets are uncertain.
Westerly Area ,\rea



OrientatlllnS of minimum compressive stress
...

0
0
o I I
0" 0
j-

00

0


Orientatiuns l)f maximum cl)mpressive stress (.)
OrientatillnS of intermedialp l'\)mpressivf::' stress (0)
Figure 8. Orientations of principal stresses derived from
faults in the Narragansett and Westerly domains. Points
are derived from Figures 6 and 7.
Common joints appear to have opened later
than all other fracture elements except microjoints
and zones of microjoints. Microjoints and zones of
microjoints terminate against common joints or are
developed at different places on opposite sides of the
joint they intersect. Common joints exhibit the
same relationship when they intersect faults. They
are not offset because, in places where the net slip
on the fault can be determined, joints on opposite
sides of the fault cannot be matched. Common
joints cut all dikes and quartz lenses, suggesting
that the latter are older.
In the Cormorant Point domain, there are two
orthogonal sets of joints, a third well developed set,
and a significant amount of scatter. Age relations
among the three major set are less clear here than in
Narragansett. The orthogonal sets strike N800E
and N200W and dip steeply. The former exhibits a
second maximum (N800W) and may correlate in age
with the N800W set in the Narragansett domain.
The N200W set may be the same age as the N15E
set. The N45E set (SE-dipping) may have formed
at the same time as the N45E set in the Narragan-
sett domain.
In the Westerly domain (Figure 9), the most
prominent set (NE-striking) may be correlative in
age with the NE-striking set in the Narragansett
and Cormorant Point domains. The orientation of
this set varies somewhat across the Westerly do-
main but not in a simple manner.
The next most prominent set (NW -striking) in
the Westerly domain is nearly at right angles to the
first set. A third, weakly developed set (EIW-
striking), present at nine stations, may be the same
as the EIW-trending set in the Narragansett do-
main. Finally, a fourth set (N-striking, E-dipping)
occurs at five stations. Age relationships among
these sets are unknown, as consistent abutting rela-
tionships were not observed.
The age of the joints in the Westerly domain
relative to other fracture elements seems to be the
same as determined in the Narragansett domain .
Microjoints and Zones of Microjoints
Microjoints are defined as two or more macro-
scopic, subparallel fractures with spacing < 2 cm.
Zones of microjoints are long, narrow regions, a
meter to a few meters wide, in which microjointing
is intensely developed. They are the youngest and
most consistently oriented of all fracture elements.
514
M.J. Hozik
Figure 9
Top: Poles to common
joints.
Middle: Poles to zones of
microjoints.
Bottom: Poles to micro-
joints.
liESTERLY
"
'"
con t- ours 2 . '5
n .. 6S
cCJntQul'S J. . S. I
n 4. 17
contOu r s 2 . '5 .10\
In the Narragansett domain, zones of microjoints
strike northerly. In the Westerly domain, consis-
tently NE-trending zones of microjoints approxi-
mately parallel the most prominent strike of faults
and common-joints (Figures 4 and 9).
Individual microjoints show more scatter in
orientation than zones of micro joints. In the Narra-
gansett domain, the dominant (NINE-trending) set
shows some scatter, including the suggestion of a
NARRAGA Ism
"
'50
oC:ont.ollrs
n .. 106
1't.Ir'lUIII"$ 2. 50 . 10 . 10
n .. 8 l <l
contou r li 2. 50 . 10'
POINT
n
'"
o:..."OntOu:I"$ 2 . ... (;
" 2
n .. 110
cQntQl,lr'$ :2. S
NW -striking set, but this set does not show on the
summary plot. In the Westerly domain, NE-trend-
ing microjoints generally parallel zones of micro-
joints.
In summary, microjoints and, more important,
zones of microjoints, indicate E-W to WNW exten-
sion. The extension direction appears to become
more northwesterly toward the Westerly domain.
Brittle fracture history of the Permian Narragansett Pier Granite 515
TECTONICS
This section summarizes the stress history ,
derived from fracture data, ofthe Narragansett Pier
Granite and the inferred tectonic framework. In
these speculations, four major tectonic episodes are
discussed: (1) pre-intrusive Alleghanian events re-
corded in metasediments of the Narragansett Basin;
(2) intrusion of the pluton and post-intrusive Alle-
ghanian events recorded in both the Narragansett
Pier Granite and metasediments of the Narragan-
sett Basin; (3) rifting of the continent and change in
drift direction; and (4) post-rifting fracturing.
Stress History of the Narragansett Pier Granite
Based on Fracture Data
For discussion of the tectonics, a selection was
made of classes of fracture elements for which age
relations are reasonably certain, examples occur in
both the Narragansett and Westerly areas, and
relationships to the principal stresses are reason-
ably known. The youngest dike set orientation in
the Narragansett area and the orientation of the
maximum for all dikes in the Westerly area were
also included. The most nearly horizontal of the
principal stresses was selected in each case, and a
set of presumed stress trajectories was plotted.
Perpendicular to these stress trajectories are lines
representing the trace of planes containing the
other two principal stresses (Table 3 and Figure
10). These chronological plots are highly specula-
tive.
The patterns described by the stress trajectories
can be divided into four groups based on stress
orientation and age:
For early normal faults, 03 has a NW -SE orien-
tation.
For quartz lenses and dikes, 03 has a more N-S
to NE-SW orientation.
For strike-slip and oblique-slip faults, 03 trajec-
tories return to a more NW -SE direction and are
concave toward the south.
For zones of micro joints and microjoints, the E-
W trajectories of 03 are concave to the north.
Each of these patterns relates to a separate
phase of the stress history recorded by the granite.
The initial NW-SE orientation of 03 is related in
time to emplacement of the granite and probably
correlates with the transition from E-W extension to
N -S to NE-SW extension suggested by dikes and
quartz lenses that record the latest stage of Alle-
ghanian events. It is consistent with the post-
intrusive D4 in the Narragansett Basin (Mosher,
1983; Reck and Mosher, 1988). The strike-slip and
oblique-slip faults record left-lateral motion pro-
duced by stresses involved in the beginning of
rifting, consistent with Mesozoic stresses proposed
by Manspeizer and Cousminer (1988) and by deBoer
and others (1988). E-W extension, indicated by
microjoints and zones of microjoints, represents the
youngest part of the record and marks a fundamen-
tal change in stress orientation, which may cor-
relate with N/S-trending rift structures suggested
by McHone and Butler (1984) as a means of loca-
lizing the White Mountain Magma Series.
Pre-Intrusive Alleghanian Events
The Narragansett Basin is inferred to have
formed as a complex ofrhomb-grabens or pull-apart
basins in response to sinistral shear on NE-trending
strike-slip faults with concomitant normal motion
on N/S-trending dip-slip faults driven by E-W
extension in the Pennsylvanian (Mosher, 1983).
Pennsylvanian sinistral motion on NE-trending
faults following Mississippian dextral motion was
also recognized in maritime Canada (Webb, 1969).
Subseqently, stress orientations changed to E-W
compression and the basin began to close. Two
periods of deformation producing NNW- to NNE-
trending folds are documented (Mosher, 1983; Reck
and Mosher, 1988) in the southern Narragansett
Basin. High-grade metamorphism was associated
with these deformations (Mosher, 1983; Murray,
1988).
Intrusion and Cooling of the Narragansett Pier
Granite Batholith
D
3
, coincident with initial intrusion of the N ar-
ragansett Pier Granite, was associated with sinis-
tral shear on NE-trending zones such as the Beaver-
head shear zone (Mosher, 1983; Reck and Mosher,
1988).
As the granite solidified, pegmatite dikes were
intruded. Some dikes show multiple injections,
perhaps due to periodic small deformational pulses
as a result of changing magmatic pressure. Such
pulsing action could enhance formation of fractures
in the cooler periphery of the intrusion. Eventually,
516
M. J. Hozik
Table 3. Stress history of the Narragansett Pier Granite sequence from oldest (top) to youngest (bottom).
Westerly Domain
Event
Feature
1 2
Intrusion Pluton EW
Early
dikes Variable Local Stress
Normal St.NE vert. SSW
faults dipNW
Set I
Late St.EW
dikes shallow
S-dip
Quartz St.NW
pods steep dip
Oblique- St.NE PI. PI.
slip steep dip NW SW
faults (LLR)
Set II
Strike- ST. NNE NE Vert.
slip NW-dip
faults (RLSS)
Set III (RLN)
Late dip-
slip
Faults
Set IV
Micro- St.NE
joints steep dip
and zones
of micro-
joints
01 became vertical, with a NW/SE-trending 03
allowing development of early normal faults (some
of which have a right-lateral component of motion).
These faults may be correlative with dextral shear
on NE-trending zones within the basin, especially
the Beaverhead shear zone. When late stage dikes
intruded, 02 diminished until it was less than 03 and
these two orientations interchanged. With con-
tinued cooling and crystallization, 03 became more
nearly horizontal again but remained aligned in a
N-NE direction as suggested by quartz lens deve-
lopment. Boundinage in the Narragansett Basin,
apparently produced by N-S extension, cuts all other
03
ESE
PI.
stp.
N
NE
PI.
ESE
SE
SE
Narragansett Domain
Feature
1 2
03
Pluton EW
Variable Local Stress
St.NE Vert. NE SE
SE-dip
St.NW PI.
moderate mod.
SW-dip toNE
St.EW NS
steep
SW-dip
St. Wto E WSW NNW
NW-dipSW
(with peg.)
(LLN S-dip)
St. WNW NE Vert. SE
S-dip
(LLSS)
St.EW SSE ENE NNW
steep
N- or S-dip
St. NNE ESE
steep dip
structures and probably correlates with the quartz
lenses.
The Close of the Alleghanian
By the end of the Alleghanian, pre-existing
surfaces had developed, setting the stage for frac-
ture zones in the Atlantic. Sheridan (1974) recog-
nized some of the essential elements involved, such
as NW/SE-trending elements (Figure 11). NINE-
trending elements, which are not shown, include
Webb's (1969) network offaults in eastern Canada:
Brittle fracture history of the Permian Narragansett Pier Granite 517
A
B
I

c
o
\
':i
"i
I
L. I <?;
" --1-___ / ..... '
I I
-.,.-- (",,'" ....
I
J ,.-- I
.... I
t
E
F
Figure 10. Orientations of stress trajectories
inferred from various fracture elements, Trajec-
tories dashed in central portion of pluton because
data restricted to ends; trajectory maps arranged
in chronological order from youngest to oldest:
A, zones of microjoints; B, microjoints; C, strike-
slip faults; D, oblique-slip faults; E, quartz lenses;
F, youngest dikes; G, early normal faults.
518 M. J. Hozik
the Bloody Bluff, Clinton-Newbury, Lake Char, and
Honey Hill faults and the Hope Valley and
Beaverhead shear zones.
The timing of the latest motion on the Hope
Valley shear zone is not clear. O'Hara and Gromet
(1985) correlated right-lateral motion with the last
stages of closure of the Narragansett Basin, but they
also indicated that the Narragansett Pier Granite is
not cut by the shear zone. Reck and Mosher (1988)
explicitly address the problem: "the shear zone can-
00
.,.
".
' 0
\
':'0 -
>0. -
1. NORTHERN GULF OF MEXICO BASIN
2 SOUTH FLORIDA- BAHAMAS BASIN
3. BI.AKE PLATEAU BASIN
4. BALTIMORE CANYON TROUGH
5. GEORGES BANK BASIN
6. NOVA SCOTIAN SHELF BASIN
7 GRANO BANKS TROUGHS
e. NORTHEAST NEWFOUNDLAND SHELF
BASIN
9. LABRADOR SHELF BASIN
.,.
" .
n
not correlate with the last period of deformation in
the Narragansett Basin unless it is younger than
the granite, because the granite intruded prior to
that deformation."
The present study supports the contention that
the latest motion on the Hope Valley shear zone pre-
dates intrusion of the Narragansett Pier Granite
because of the absence of fracture elements with
moderate to shallow N-NE dips and because of the
absence of ductile deformation of the granite.
.0
DEPTH TO lI1I:[oJuA4SSIC usrlllE"'"
___ .. .....
.,.
".
,...tdP-lOOCM I$O!IA.n_
.....____ COPtTtN(HTAI. ... UCIH fA\JtT <:. :
'0 ". "'.
Figure 11. Generalized basement map on pre-Jurassic (from Sheridan, 1974).
Brittle fracture history of the Permian Narragansett Pier Granite 519
Furthermore, the latest stages of intrusion, the
injection of the youngest dikes and quartz lenses,
and the earliest faults all require N-S to NW
extension, which is not compatible with right-
lateral motion and S-directed thrusting described
for the shear zone. In short, the fracture data
provide no evidence that the shear zone cuts the
granite or that deformation of the granite could be
correlated with motion on the shear zone if it curved
into the Ten Rod Granite north of the Narragansett
Pier pluton. There is, however, evidence (in the
orientation of dikes, quartz lenses, and earliest
faults) that the pluton records the latest period of
deformation in Narragansett Basin.
Figure 11 shows that the continental margin is
offset with an apparent right-lateral sense on the
Kelvin and Orpheus (Cobequid) trends. According
to Wilson's (1965) model for transform faults, the
faults themselves originate at the time of rifting,
presumably along pre-existingweaknesses. The
motion on transform faults should be opposite that
of the apparent offset of the continental margin.
Thus, the Kelvin and Orpheus transform faults
should have left-lateral motion.
Continental Rifting
Within the Narragansett Pier Granite, stresses
associated with initial rifting are apparent from
orientation of oblique-slip faults, which grade into
strike-slip faults (Figures 6 and 7). The oblique-
slip faults imply that 02 was inclined whereas the
strike-slip faults imply a vertical 02. Price (1966)
discussed the special conditions necessary to pro-
duce strike-slip faulting, including concomitant
extension parallel to 03, which is consistent with the
suggestion of a stress history controlled by the
orientation of 03. Data from this study suggest 01
was essentially vertical until after development of
the quartz lenses. McHone and Butler (1984)
argued for a broad, domal uplift of southern New
England, greatest in areas of Alleghanian orogeny,
beginning soon after granite emplacement. During
this phase of development, 01 would have been
decreasing due to reduction of overburden and, by
the time strike-slip faults formed, the pluton had
been elevated sufficiently so that 02 was vertical.
Major rifting occurred at 170-180 Ma (LePichon
and Fox, 1971; Klitgord and Schouten, 1986),
probably initiating significant motion on transform
faults, consistent with data from this study that
indicate similar timing for motion on a transform
fault along the Kelvin Seamount trend curving into
the New York Bight. Assuming that such a left-
lateral transform fault (striking E-W) developed in
an environment of pure shear, with NE-trending
horizontal 01, vertical 02' and NW-trending hori-
zontal 03, then not only the transform but left-
lateral EIW-trending faults in the Narragansett
domain and NINE-trending faults with right-lateral
motion in the Westerly domain could have formed.
Oblique-slip faults with right-lateral and normal
components of motion have very similar stress
orientations and could have developed about the
same time in the Westerly domain if 01 and 02 had
rotated slightly or if an earlier NE-trending struc-
tural trend such as the Watch Hill Fault was acting
as a significant anisotropy.
McHone's (1978) NW-SE orientation of mini-
mum compression from his 160-220 Ma dikes and
May's (1971) "Triassic" stress orientations agree in
general with orientations inferred from left-lateral
faults in the Narragansett domain. Similar stress
orientations were reported for the Newark Basin
(Ratcliffe and Burton, 1985) using NW -SE extension
to reactivate old faults. Manspeizer and Cousminer
(1988) suggest an EIW-trending left-lateral shear-
couple and show the appropriate strain ellipsoid and
suggest left-lateral motion on the Kelvin seamount
transform. deBoer and others (1988) similarly re-
quire NW-SE extension in their first phase of rift-
ing.
Change in Spreading Direction
A change in spreading direction in the Atlantic
Ocean could have produced N -S extension and could
also explain the origin of late steeply dipping dip-
slip faults at Narragansett (Set IV). Their motion
sense consistently indicates north-side-down. Pos-
sible changes in spreading direction were suggested
by LePichon and Fox (1971) at 80 Ma and by Klit-
gord and Schouten (1986). However, Wise and
others (1975) suggested that these faults may have
formed during coastal subsidence and rotation.
Post-Rifting Fracturing
Since the close of the Paleozoic, extensive uplift
and erosion have occurred in New England (Thomp-
son and Norton, 1968). Studies of a Late Cretaceous
520 M.J. Hozik
planation surface (unpublished data of Fair-
bridge,1977, cited in Sykes, 1978) suggest most
uplift occurred prior to the Late Cretaceous. In the
Pleistocene, the entire New England region was
glaciated, which involved erosion and removal of
overburden on the Narragansett Pier Granite.
Reduction of compressive stresses necessary to
fonn joints could have occurred as overburden was
removed by erosion, particularly if joints developed
along pre-existing surfaces. Based on relative age
relationships, the nearly N/S-trendingjoints fonned
first, and EIW-trending joints opened slightly later.
Some movement parallel to N-S-joints could accom-
modate N-S expansion necessary for development of
E-Wjoints.
Microjoints and zones of microjoints are
believed to develop perpendicular to 03 (Wise, 1964)
and are among the youngest fracture elements
examined in the Narragansett Pier Granite.
Because of their very unifonn orientation, they pro-
bably represent an event restricted in time, pre-
dating some common joints and post-dating others;
perhaps their orientation is controlled by pre-
existing weaknesses. Stress trajectories, based on
attitudes of these features (see Figure 8), are differ-
ent from those for other fracture elements (with the
possible exception of common joints), suggesting a
significant separation in time of origin.
CONCLUSIONS
1) The Narragansett Pier Granite intruded sub-
parallel to the regional foliation in the Pennian,
after the peak of thennal metamorphism during
shearing-dominated phases of the Alleghanian oro-
geny recorded in the adjacent Narragansett Basin.
The granite truncates Alleghanian isograds indica-
tive of temperatures of 600C and pressure of 5-6 kb
(Murray, 1988). The granite apparently cooled
rather quickly and recorded some effects of the
latest Alleghanian stresses (N -S extension recorded
by early nonnal faults, youngest dikes, and quartz
lenses). The depth of intrusion was probably moder-
ate, as indicated by open-space fillings in some veins
and the near absence of any indication of ductile
defonnation or mylonites in dikes and faults that
overlapped, in part, the late stages of cooling.
The Narragansett Pier Granite retains evi-
dence of three episodes of extensional faulting
but no evidence of a major compressive defor-
mation.
Early nonnal faults fonned in response to N-S
extension at the end ofthe Alleghanian orogeny
- compatible with the last stages of Alleghan-
ian deformation recorded in the Narragansett
Basin.
2) Oblique-slip and strike-slip faults appear rela-
ted to initial Late Triassic or Early Jurassic rifting.
In detail, the fault pattern indicates a relatively
constant NW-SE orientation of 03 while 01 and 02
interchanged and rotated. This orientation is con-
sistent with that proposed by May (1971), McHone
(1978), McHone and Butler (1984), Ratcliffe and
Burton (1985), Manspeizer and Cousminer (1988)
and de Boer and others (1988) and is also compatible
with a transform fault along the Kelvin Seamount
line, as suggested by May (1971).
The youngest faults, which are evident only in
the Narrgansett domain, are dip-slip faults that
may have developed during a change in spread-
ing direction "" 80 Ma or in response to more
recent coastal zone tilting.
3) The development of microjoints and common
joints seems to be part of younger separate stress
systems involving dominantly E-W extension.
Common joints may have developed under a
similar stress system with one major joint set
parallel to microjoints. A second (orthogonal)
common-joint set could reflect additional near-
surface expansion at right angles to the major
release direction.
Microjoints post-date at least some common
joints. Consistent microjoint orientation sug-
gests development along pre-existing weak-
nesses, perhaps in response to stresses sugges-
ted by McHone and Butler (1984) for localizing
the Jurassic White Mountain Magma Series or
in a single, discrete event of unknown origin.
4) Tectonic models of the Rhode Island and Con-
necticut coastline which rely on simple left-lateral
or right-lateral shearing are not supported by data
from the Narragansett Pier Granite. Models in-
volving N-S to NW-SE extension in the transfonn
should be given more serious consideration.
Brittle fracture history of the Permian Narragansett Pier Granite 521
5) A consistent (NE-trending orientation) of frac-
ture elements in the Westerly domain suggests
some underlying anisotropy other than rift and
grain.
ACKNOWLEDGMENTS
I would like to thank my field assistants, Gerald
Williams and Jay Silverman for their help in the
field, as well as Donald U. Wise, George E. McGill,
Stephen W. Field, and Sharon Mosher for their
helpful reviews of this paper. I am grateful to the
Geological Society of America for financial support
in the form of Penrose Bequest Research Grants
1893-74 and 2027-75.
REFERENCES
ALEINIKOFF, J. H., R. H. MOENCH, and J. B. LYONS,
1985, Carboniferous U-Pb age of the Sebago batholith,
southwestern Maine: Geological Society of America Bulle-
tin, v. 96, p. 990-996.
__ , R. E. ZARTMAN, and J. B. LYONS, 1979, Geo-
chronology of the Massabesic Gneisss and the granite near
Milford, south-central New Hampshire: New evidence for
Avalonian basement and Taconic and Alleghenian in east-
ern New England: Contributions to Mineralology and
Petrology, v. 71, p. 1-11.
BALLARD, R. D., and E. UCHUPI, 1975, Triassic rift
structures in Gulf of Maine: American Association of
Petroleum Geologists Bulletin, v. 59, p. 1041-1072.
BAROSH, P. J., M. H. PEASE, Jr., R. W. SCHNABEL, K. G.
BELL, and J. D. PEPPER, 1977, Aeromagnetic Linea-
ment Map of Southern New England Showing Rela-
tions of Lineaments to Bedrock Geology: U.S. Geologi-
cal Survey, Miscellaneous Field Studies Map, MF- 885.
BILLINGS, M. P., 1929, Structural geology of the eastern
part of the Boston Basin: American Journal of Science, v.
218, p. 97-137.
__ , and D. A. RAHM, 1966, Geology of the Malden Tun-
nel, Massachusetts: Journal of the Boston Society of Civil
Engineers, v. 53, p. 116-141.
BROWN, A., D. P. MURRAY, and E. S. BARGHOORN, 1978,
Pennsylvanian fossils from metasediments within the Nar-
ragansett Pier Granite, Rhode Island [abstract): Geologi-
cal Society of America Abstracts with Programs, v. 10, no.
2, p. 34-35.
CAMERON, B., and R. S. NAYLOR, 1976, General geology of
southeastern New England; pp. 117-134 in B. Cameron
(ed.), Geology of Southeastern New England: A Guide-
book for Field Trips to the Boston Area and Vicinity
(68th Annual Meeting of the New England Intercolle-
giate Geological Conference): Science Press, Princeton,
New Jersey, USA, 513 p.
CASTLE, R. 0., H. R. DIXON, S. GREW, A. GRISCOM, and 1.
ZIETZ, 1976, Structural Dislocations in Eastern Massa-
chlUletts: U.S. Geological Survey, Bulletin 1410, 39 p.
CAZIER, E. C., 1987. Late Paleozoic tectonic evolution of
the Norfolk Basin, southeastern Massachusetts: Journal
of Geology, v. 95, p. 55-73.
CHUTE, N. E.,1969, Bedrock Geologic Map of the Blue
Hills Quandrangle, Norfolk and Suffolk Counties,
MassachlUletts: U.S. Geological Survey, Geologic Quad-
rangle Map, GQ-796.
DALLMEYER, R. D., 1982, 4oArp9 Ar ages from the Narra-
gansett Basin and southern Rhode Island basement ter-
rane: Their bearing on the extent and timing of Alleghan-
ian tectonothermal events in New England: Geological
Society of America Bulletin, v. 93, p. 1118-1130.
DEBOER, J. Z., J. G. McHONE, J. H. PUFFER, P. C.
RAGLAND, and D. WHITTINGTON, 1988, Mesozoic and
Cenozoic magmatism; pp. 217-241 in R. E. Sheridan and J.
A. Grow (eds.), The Geology of North America, Volume
12: The Atlantic Continental Margin: U.S.: Geological
Society of America, Boulder, Colorado, USA, 610 p.
DIMENT, W. H., T. C. URBAN, and F. A. REVETTA, 1972,
Some geophysical anomalies in the eastern United States;
pp. 544-574 in E. Robertson (ed.), The Nature of the Solid
Earth: Interscience, New York, New York, USA, 677 p.
DIXON, H. R., and L. W. LINDGREN, Jr., 1968, Structures
of eastern Connecticut; pp. 219-229 in E. Zen, W. S. White,
J. B. Hadley, and J. B. Thompson, Jr. (eds.), Studies of
Appalachian Geology: Northern and Maritime: Inter-
science, New York, New York, USA, 475 p.
DOHERTY J. T., and J. B. LYONS, 1980, Mesozoic erosion
rates in northern New England: Geological Society of
America Bulletin, Part 1, v. 91, p. 16-20.
DRAKE, C. L., and H. P. WOODWARD, 1963, Appalachian
curvature, wrench faulting, and offshore structures:
Transactions of the New York Academy of Science, v. 26, p.
48-63.
EBERLY, P.O., 1985, Brittle Fracture Petrofabric
Along a West-East Traverse From the Connecticut
VaUey to the Narragansett Basin: MS Thesis, Univer-
sity of Massachusetts, Amherst, Massachusett, USA, 137
p.
FAIRBAIRN, H. W., S. MOORBATH, A. O. RAMO, W. H.
PINSON, and P. M. HURLEY, 1976, Rb-Sr age of granitic
rocks in southeastern Massachusetts and the age of the
Lower Cambrian at Hoppin Hill: Earth and Planetary
Science Letters, v. 2, p. 321-328 ..
522
M.J.Hozik
FEININGER, T., 1965, Bedrock Geologic Map of the Ash-
away Quadrangle, Connecticut-Rhode Island: U.S.
Geological Survey, Geological Quandrangle Map GQ-403.
FLETCHER, J. B., M. L. SBAR, and L. R. SYKES, 1978, Seis-
mic trends and travel-time residuals in eastern North
America and their tectonic implications: Geological So-
ciety of America Bulletin, v. 89, p. 1656-1676.
FOLAND, K. A., and H. FAUL, 1977, Ages of the White
Mountain intrusives, New Hampshire, Vermont, and
Maine: American Journal of Science, v. 227, p. 888-904.
GATES, 0., 1969, Lower Silurian-Lower Devonian rocks of
the New England coast and southern New Brunswick; pp.
484-503 in M. Kay (ed.), North Atlantic Geology and
Continental Drift: American Association of Petroleum
Geologists, Memoir 12,1082 p.
GETTY, S.R. and L.P. GROMET, 1986, The southern ter-
minus of the Hope Valley shear zone, Rhode Island
[abstract]: Geological Society of America Abstracts with
Programs, v. 18, p. 18.
GOLDSTEIN, A.G., 1980, Tectonics of Ductile Faulting
in a Portion of the Lake Char Mylonite Zone, Massa-
chusetts and Connecticut: PhD Dissertation, University
of Massachusetts, Amherst, Massachusetts, USA, 177 p.
__ , 1982, Geometry and kinematics of ductile faulting
in a portion of the Lake Char mylonite zone, Massachu-
setts and Connecticut: American Journal of Science, v.
282, p. 1378-1405.
GREW, E. S, 1976, Pennsylvanian rocks of east-central
Massachusetts; pp. 151-167 in B. Cameron (ed.), Geology
of Southeastern New England: A Guidebook for Field
Trips to the Boston Area and Vicinity (68th Annual
Meeting of the New England Intercollegiate Geologi-
cal Conference): Science Press, Princeton, New Jersey,
USA,513p.
HAYWARD, J. A., and H. E. GUADETTE, 1984, Carboni-
ferous age of the Sebago and Effingham plutons, Maine
and New Hampshire [abstract]: Geological Society of
America, Abstracts with Programs, v. 16, p. 23.
HERMES, O. D., 1987, Geologic relationships of Permian
Narragansett Pier and Westerly granites and Jurassic
lamprophyric dike rocks, Westerly, Rhode Island; pp. 181-
186 in D. C. Roy (ed.), Northeastern Section, Geological
Society of America: Centennial Field Guide, Volume
5: Geological Society of America, Boulder, Colorado, USA,
481p.
HERMES, O. D., and R. E. ZARTMAN, 1985, Late Pro-
terozoic and Devonian plutonic terrane within the Avalon
zone of Rhode Island: Geological Society of America
Bulletin, v. 96, p. 272-282.
__ , P. J. BAROSH, and P. V. SMITH, 1981, Contact rela-
tionships of the late Paleozoic Narragansett Pier Granite
and country rock; pp. 125-152 in J. C. Boothroyd and O. D.
Hermes (eds.), Guidebook to Geologic Field Studies in
Rhode Island and Adjacent Areas (73rd Annual Meet-
ing of the New England Intercollegiate Geological
Conference): University of Rhode Island, Kingston,
Rhode Island, USA, 383 p.
HOZIK, M. J., 1981, Brittle Fracture History of the
Narragnansett Pier Granite, Rhode, Island: PhD
Dissertation, University of Massachusetts, Amherst, Mas-
sachusetts, USA, 309 p.
KLITGORD, K. D., and H. SHOUTEN, 1986, Plate kine-
matics of the central Atlantic; pp. 351-378 in P. R. Vogt
and B. E. Tucholke (eds.), The Geology of North Ameri-
ca, VolumeM: The Western North Atlantic: Geological
Society of America, Boulder, Colorado, USA, 696 p.
KNOX, A. S., 1944, A Carboniferous flora from the Wam-
sutta Formation of southeastern Massachusetts: American
Journal of Science, v. 242, p. 130-138.
KOCIS, D. E., O. D. HERMES, and J. A. CAIN, 1978, Petro-
logical comparison of the pink and white facies of the
Narragansett Pier Granite, Rhode Island [abstract]: Geo-
logical Society of America Abstracts with Programs, v. 10,
no. 2,p. 71.
LEPICHON, X., and P.H. FOX, 1971, Marginal offsets,
fracture zones and the early opening of the North Atlantic:
Journal of Geophysical Research, v. 76, p. 6294-6308.
LYONS, P. C., 1971, Correlation of the Pennsylvanian of
New England and the Carboniferous of New Brunswick
and Nova Scotia [abstract]: Geological Society of America,
Abstracts with Programs, v. 3, p. 43-44.
__ , 1978, A late Middle Pennsylvanian flora of the
Narragansett Basin, Massachusetts: Geological Society of
America Bulletin, v. 89, p. 433-438.
__ , and H. B. CHASE, 1976, Coal stratigraphy and flora
of the northwestern Narragansett Basin; pp. 405-427 in B.
Cameron (ed.), Geology of Southeastern New England:
A Guidebook for Field Trips to the Boston Area and
Vicinity (68th Annual Meeting of the New England
Intercollegiate Geological Conference): Science Press,
Princeton, New Jersey, USA, 513 p.
__ , and W. C. DARRAH, 1977, Floral evidence for Upper
Pennsylvanian in the Narragansett Basin, southeastern
New England [abstract]: Geological Society of America,
Abstracts with Programs, v. 9, no. 3, p. 297.
__ , and H. W. KRUEGER, 1976, Petrology, chemistry,
and age of the Rattlesnake Pluton and implications for
other alkalic granite plutons of southern New England; pp.
71-102. in P. C. Lyons and A. H. Brownlow (eds.), Studies
in New England Geology: Geological Society of America,
Memoir 146, 389 p.
Brittle fracture history of the Permian Narragansett Pier Granite 523
__ , B. TIFFNAY, and B. CARMENT, 1976, Early
Pennsylvanian age of the Norfolk Basin, southeastern
Massachusetts, based on plant megafossils; pp. 181-197 in
P. C. Lyons and A. H. Brownlow (eds.), Studies in New
England Geology: Geological Society of America, Memoir
147,389p.
MALUSKI, H., 1976, K-Ar ages ofbiotites from Corsica and
arguments for Permian age alkaline granitic intrusion:
Contribution to Mineralogy and Petrology, v. 58, p. 305-317.
MANSPEIZER, W., and H. L. COUSMINER, 1988, Late Tri-
assic-Early Jurassic synrift basins of the U.S. Atlantic
margin; pp. 197-216 in R E. Sheridan and J. A. Grow
(eds.), The Geology of North America, Volume 12: The
Atlantic Continental Margin: U.S.: Geological Society
of America, Boulder, Colorado, USA, 610 p.
MAY, P. R, 1971, Pattern of Triassic-Jurassic diabase
dikes around the North Atlantic in the context of predrift
position of the continents: Geological Society of America
Bulletin, v. 82, p. 1285-1292.
McHONE, J. G., 1978, Distribution, orientations, and ages
of mafic dikes in central New England: Geological Society
of America Bulletin, v. 89, p. 1645-1655.
__ , and J. R BUTLER, 1984, Mesozoic igneous provinces
of New England and the opening of the North Atlantic
Ocean: Geological Society of America Bulletin, v. 95, p.
757-765.
MOORE, G. E., 1967, Bedrock Geologic Map of the
Watch Hill Quadrangle, Washington County, Rhode
Island, and New London County, Connecticut: U.S.
Geological Survey, Geological Quadrangle Map GQ-655.
MOSHER, S., 1976, Pressure solution as a deformation
mechanism in Pennsylvanian conglomerate from Rhode
Island: Journal of Geology, v. 84, p. 355-364.
__ ,1983, Kinematic history of the Narragansett Basin,
Massachusettsand Rhode Island: Constraints on late
Paleozoic plate reconstruction: Tectonics, v. 2, p. 327-344.
MURRAY, D. P., 1988, Post-Acadian metamorphism in the
Appalachians; pp. 597-609 in A. 1. Harris and D. J. Fettes
(eds.), The Caledonian-Appalachian Orogen: Geologi-
cal Society of America, Special Publication 38, 643 p.
__ , and J. W. SKEHAN, SJ, 1979, A traverse across the
eastern margin of the Appalachian-Caledonide Orogen,
southeastern New England; pp. 1-35 in P. S. Osberg and J.
W. Skehan, SJ (eds.), The Caledonides in the United
States of America: Geological Excursions in the
Northeast Appalachians (Contributions to IGCP Pro-
ject 27 - Caledonide Orogen): Weston Observatory,
Department of Geology and Geophysics, Boston College,
Weston, Massachusetts, USA, 250 p.
MUTCH, T. A., 1968, Pennsylvanian non-marine sediments
of the Narragansett Basin of Massachusetts and Rhode
Island; pp. 177-209 in G. de v. Klein (ed.), Late Paleozoic
and Mesozoic Continental Sedimentation, Northeast-
ern North America: Geological Society of America, Spe-
cial Paper 106, 309 p.
NAYLOR, R S., and S. SAYER, 1976, The Blue Hills
Igneous Complex, Boston area, Massachusetts; pp. 135-157
in B. Cameron (ed.), Geology of Southeastern New Eng-
land: A Guidebook for Field Trips to the Boston Area
and Vicinity (68th Annual Meeting of the New Eng-
land Intercollegiate Geological Conference): Science
Press, Princeton, New Jersey, USA, 513 p.
NICHOLS, D. R, 1956, Bedrock Geology of the Narra-
gansett Pier Quadrangle, Rhode Island: U. S. Geo-
logical Survey, Geological Quadrangle Map GQ-91.
O'HARA, K.D., and L. P. GROMET, 1983, Textural and Rb-
Sr isotopic evidence for late Paleozoic mylonitization
within the Honey Hill fault zone, southeastern Connec-
ticut: American Journal of Science, v. 283, p. 762-779.
__ , and L. P. GROMET, 1985, Two distinct late Pre-
cambrain (Avalonian) terranes in southeastern New
England and their late Paleozoic juxtapositon: American
Journal of Science, v. 285, p. 673-709.
OLEKSYSHYN, J., 1976, Fossil plants of Pennsylvanian
age from northwestern Narragansett Basin; pp. 143-180 in
P. C. Lyons and A. H. Brownlow (eds.), Studies in New
England Geology: Geological Society of America, Memoir
146,389p.
OSBERG, P. S., and J. W. SKEHAN, SJ (eds.), 1979, The
Caledonides in the United States of America: Geo-
logical Excursions in the Northeast Appalachians
(Contributions to IGCP Project 27 - Caledonide Oro-
gen): Weston Observatory, Department of Geology and
Geophysics, Boston College, Weston, Massachusetts, USA,
250p.
PIEPUL, R G, 1975, Analysis of Jointing and Faulting
at the Southern End of the Eastern Border Fault,
Connecticut: MS Thesis, University of Massachusetts,
Amherst, Massachusetts, USA, 109 p.
PRICE, J. 1966, Fault and Joint Development in Brittle
and Semi-Brittle Rock: Pergamon Press, New York,
New York, USA, 176 p.
QUINN A. W., 1971, Bedrock Geology of Rhode Island:
U.S. Geological Survey, Bulletin, 1295,64 p.
__ , and G. E. MOORE, Jr., 1968, Sedimentation, tec-
tonism and plutonism of the Narragansett Basin region;
pp. 269-280 in E. Zen, W. S. White, J. B. Hadley, and J. B.
Thompson, Jr. (eds.), Studies in Appalachian Geology
(Billings Volume): Northern and Maritime: Intersci-
ence, New York, New York, USA, 475 p.
__ , and W. A. OLIVER, 1962, Pennsylvanian rocks in
New England; pp. 60-73 in C. C. Branson (ed.), Pennsyl-
524 M.J. Hozik
vanian System in the United States: American Asso-
ciation of Petroleum Geologists, Tulsa, Oklahoma, USA,
505p.
RANKIN, D. W., 1976, Appalachian salients and recesses:
Late Precambrian continental breakup and the opening of
the Iapetus Ocean: Journal of Geophysical Research, v. 81,
p. 5605-5619.
RAST, N., 1980, The Avalonian Plate in the northern
Appalachians and Caledonides; pp. 63-66 in D. R. Wones
(ed.), The Caledonides in the United States of Ameri-
ca: Proceedings of IGCP Project 27 - Caledonide
Orogen: Virginia Polytechnic Institute and State Univer-
sity, Department of Geological Sciences, Memoir 2, 329 p.
RATCLIFFE, N. M., and W. C. BURTON, 1985, Fault
reactivation models for origin of the Newark Basin and
studies related to eastern U.S. seismicity; pp. 36-45 in G. R.
Robinson and A. J. Froelich (eds.), Proceedings of the
2nd U.S. Geological Survey Workshop on the Early
Mesozoic Basins of the Eastern United States: U.S.
Geological Survey, Circular 946, 147 p.
RECK, B. H., and MOSHER, S., 1988, Timing of intrusion of
the Narragansett Pier Granite relative to deformation in
the southwestern Narragansett Basin, Rhode Island:
Journal of Geology, v. 96, p. 677-692.
RUTTEN, M. G., 1969, The Geology of Western Europe:
Elsevier Publishing Company, Amsterdam, The Nether-
lands, 476 p.
SBAR, M. L., and L. R. SYKES, 1973, Contemporary com-
pressive stress and seismicity in eastern North America:
An example of intra plate tectonics: Geological Society of
America Bulletin, v. 84, p. 1861-1882.
SHALER, N. S. S., J. B. WOODWORTH, and A. F. ROERSTE,
1899, Geology of the Narragansett Basin: U.S. Geologi-
cal Survey, Monograph 33, 402 p.
SHERDIAN, R. E., 1974, Atlantic continental margin of
North America; pp. 391-407 in C. A. Burke and C. L. Drake
(eds.), The Geology of Continental Margins: Springer-
Verlag, New York, New York, USA, 1009p.
SKEHAN, J. W., SJ, 1968, Fracture tectonics of south-
eastern New England as illustrated by Wachusett-
Marlborough tunnel, east-central Massachusetts; pp. 281-
290 in E. Zen, J. B. White, J. B. Hadley, and J. B. Thomp-
son, Jr. (eds.), Studies of Appalachian Geology (Billings
Volume): Northern and Maritime: Interscience
Publishers, New York, New York, USA, 475 p.
__ ,1969, Tectonic framework of southern New England
and eastern New York; pp. 793-814 in M. Kay (ed.), North
Atlantic Geology and Continental Drift: American
Association of Petroleum Geologists, Memoir 12, 1082 p.
__ , and D. P. MURRAY (eds.), 1978, The Coal-Bearing
Narragansett Basin of Massachusetts and Rhode Is-
land: Geology, Volume I: Weston Observatory, Depart-
ment of Geology and Geophysics, Boston College, Weston,
Massachusetts, USA, 99 p.
__ , and D. P. MURRAY, 1979, Woonsocket and North
Scituate Basins; pp. A14-A15 in J. W. Skehan, SJ, D. P.
Murray, J. C. Hepburn, M. P. Billings, P. C. Lyons, and R.
G. Doylee (eds.), The Mississippian and Pennsylvanian
(Carboniferous) Systems in the United States - Mas-
sachusetts, Rhode Island, and Maine: U.S. Geological
Survey, Professional Paper 111 O-A, 30 p.
__ , D. P. MURRAY, E. S. BELT, O. D. HERMES, N. RAST,
and J. DEWEY, 1976, Alleghanian deformation, sedimen-
tation, and metamorphism in southeastern Massachusetts
and Rhode Island; pp. 447-471 in B. Cameron (ed.), Geo-
logy of Southeastern New England: A Guidebook for
Field Trips to the Boston Area and Vicinity (68th
Annual Meeting of the New England Intercollegiate
Geological Conference): Science Press, Princeton, New
Jersey, USA, 513 p.
SNOKE, A. W., and S. MOSHER, 1989, The Alleghanian
orogeny as manifested in the Appalachian internides; pp.
288-318 in R. D. Hatcher, Jr., W. A. Thomas, and G. W.
Viele (eds.), The Geology of North America, Volume F2:
The Appalachian-Ouachita Orogen in the United
States: Geological Society of America, Boulder, Colorado,
USA,791p.
SWANSON, M. T., 1982, Preliminary model for an early
transform history in central Atlantic rifting: Geology, v.
10, p. 317-320.
SYKES, L. R., 1978, Intraplate seismicity, reactivation of
preexisting zones of weakness, alkaline magmatism and
other tectonism post-dating continental fragmentation:
Reviews of Geophysics and Space Physics, v. 16, p. 621-688.
THOMAS, W. A., 1977, Evolution of Appalachian salients
and recesses form reentrants and promontories in the con-
tinental margin: American Journal of Science, v. 277, p.
1233-1278.
THOMPSON, J. B., Jr., and S. A. NORTON, 1968, Paleozoic
regional metamorphism in New England and adjacent
areas; pp. 319-327 in E. Zen, J. B. White, J. B. Hadley, and
J. B. Thompson, Jr. (eds.), Studies of Appalachian Geo-
logy (Billings Volume): Northern and Maritime:
Interscience Publishers, New York, New York, USA, 475
p.
WEBB, G. W., 1969, Paleozoic wrench faults in Canadian
Appalachians; pp. 754-786 in M. Kay (ed.), North Atlantic
Geology and Continental Drift: American Association of
Petroleum Geologists, Memoir 12,1082 p.
WISLON, J. T., 1965, A new class of faults and their
bearing on continental drift: Nature, v. 195, p. 135-138.
WINTSCH, R. P., and J. N. ALEINIKOFF, 1987, U-Pb
isotopic and geological evidence for late Paleozoic anatexis,
Brittle fracture history of the Permian Narragansett Pier Granite 525
deformation, and accretion of the Late Proterozoic Avalon
terrane, south central Conneticut: American Journal of
Science. v. 287, p. 107-126.
WISE, D. U., 1964, Microjointing in basement, Middle
Rocky Mountains of Montana and Wyoming: Geological
Society of America Bulletin, v. 75, p. 287-306.
__ , M. J. HOZIK, A. G. GOLDSTEIN, and R.G. PIEPUL,
1975, Minor fault motions in relation to Mesozoic tectonics
of southern New England [abstract): EOS, Transactions of
the American Geophysical Union, v. 56, p. 451.
ZARTMAN, R. E., and O. D. HERMES, 1987, Archean
inheritance in zircon from late Paleozoic granites from the
Avalon zone of southeastern New England: An African
connection: Earth and Planetary Science Letters, v. 82, p.
305-315.
__ , and R. F. MARVIN, 1971, Radiometric age (Late Or-
dovician) of the Quincy, Peabody, and Cape Ann Granites
from eastern Massachusetts: Geological Society of America
Bulletin, v. 82, p. 932-958.
__ , and R. S. NAYLOR, 1972, Structural implications of
some U-Th-Ph zircon isotopic ages of igneous rocks in
eastern Massachusetts [abstract): Geological Society of
America Abstracts with Programs, v. 4, no. 1, p. 54.
__ , O. D. HERMES, and M. H. PEASE, Jr., 1988, Zircon
crystallization ages, and subsequent isotopic disturbance
events, in gneissic rocks of eastern Connecticut and west-
ern Rhode Island: American Journal of Science, v. 288, p.
376-402.
__ , P. M. HURLEY, H. W. KRUEGER, and B. J. GILETTI,
1970, A Permian disturbance of K-Ar radiometric ages in
New England: Its occurrence and cause: Geological
Society of America Bulletin, v. 82, p. 937-958.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Comparison of sedimentary petrologic sequences produced in
Mesozoic extensional and convergent tectonic settings
in North America
PAUL T. RYBERG
Department of Geology and Earth Sciences, Clarion University, Clarion, Pennsylvania 16214, USA
(received October 17,1988; revision accepted June 16, 1989)
ABSTRACT
In eastern North America, two Mesozoic basins related to the early rifting phases of the breakup of the
supercontinent Pangaea are the Newark-Gettysburg Basin of the mid-Atlantic states and the Hartford-
Deerfield Basin of New England. These basins show classic listric normal-fault, half-graben structure, with
older back-tilted strata generally showing slightly more rotation with depth. The Newark and Hartford
Basins were probably never depositionally connected, but they do show similar cyclic patterns of basin filling.
Arkosic and lithic basin fill were dominated by alluvial fan deposition adjacent to border faults and braided
stream deposition in proximal parts of basins. Sediment influx was dominantly from south to north in the
Newark Basin during the Late Triassic. Distal and low lying areas of the basins were dominated by lacustrine
deposition exhibiting Milankovitch-type lake-level cyclicity.
The western margin of North America was dominated by sedimentation in convergent tectonic settings
during the Mesozoic. The Galice Formation ofthe Western Jurassic Belt (Klamath Mountains) is a classic
example of sedimentary fill accumulated in a marginal interarc basin (which later contracted) adjacent to
North America. Original sedimentary patterns of the Galice are obscured by folding and slaty cleavage
developed during collision and accretion of the Rogue arc terrane with the continent during the Nevadan
Orogeny. The Galice sediments have detrital modes indicating recycled orogen clast input from the continent
to the east, and volcanic clast input from the Rogue island arc to the west. Post-Nevadan Dothan (Oregon) and
Franciscan (California) strata suggest deposition in a subduction complex setting of an Andean-type tectonic
regime along the western edge of North America. Detrital modes of Dothan-Franciscan (subduction complex)
and Myrtle Group-Great Valley Sequence (forearc basin) sandstones show considerable compositional overlap,
indicating a common provenance - the Sierran-Klamath volcanoplutonic arc and collisional orogen complex.
INTRODUCTION
Interpretations of ancient continental margins,
especially those involving Precambrian rocks, must
rely on much younger analogs for which the sedi-
mentary record is more complete and less deformed.
The purpose of this paper is to present a comparison
of sedimentary styles in well-studied Mesozoic ex-
tensional and convergent margin basins of North
America. Geologists working on Precambrian meta-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 527-538. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
527
528
P. T. Ryberg
sedimentary sequences may find useful analogies
from the concise summaries presented here, al-
though no direct comparison of Mesozoic and Pre-
cambrian basins is attempted herein. Attention is
focused on rift-related sedimentation in the
Newark-Gettysburg Basin (which extends across
New York, New Jersey, Pennsylvania, Maryland,
and Virginia), and the Hartford Basin (which
extends from Connecticut to Massachusetts). Meso-
zoic convergent-margin basins of western North
America are also examined, including the marginal
basin between the Rogue island-arc system and the
earlier accreted margin of the continent along the
Klamath Mountains. Basins associated with the
change to an Andean-type continental margin with
a subduction complex along the inner trench slope
(Dothan-Franciscan), and a forearc basin complex
(Myrtle-Great Valley) between the trench and the
N
Sierran-Klamath volcanic plutonic arc are also dis-
cussed.
MESOZOIC EXTENSIONAL BASINS OF EASTERN
NORTH AMERICA
The supercontinent of Pangaea began rifting in
Late Triassic time, eventually separating North
America from Africa during the Jurassic. Numer-
ous rift basins closely follow the trend of the Appala-
chian orogen all along the eastern margin of North
America (Figure 1). Westernmost basins along this
trend are exposed from Canada to the southeastern
United States, and many more are now buried
farther east beneath younger passive margin strata
of the coastal plain and continental shelf. Early
phases of rifting allowed thousands of meters of con-
\
Chedabucfo Basln-
Chedabuclo Fault
200 400 11m
Minas Basln-
Cobequid Foul'
Fundy Basln-
Fundy Fault
Gulf of Maine Foull Zone
Pomperaug Bosm - '
Pomperaug Faull
Newark BaSin _ -------...... ,4
Ramapo Faull
Getlysburg Basln-----..
"- Hartford Basln-
Minerai Hill Foull
Culpeper Basln-
Stafford Fault
Brevard Zone -.......
-""rlf--- Tay.lorsville H las Zone
--"",","--- Richmond Basln""""""- Y
Farmville Basm - LakeSide Faull
Roanoke Creek. BaSin
Scottsburg BaSI n
Ou rhom BaSI n
'---+---- Sanford BaSin
Wadesboro BaSin
Crow burg Basin
--1---- Dunbarton BaSin
Rlddleville Basin - Augusta Fault
"--+----- Main South Georgia Rift Basm-
Paleozoic Suture (?)
Figure 1. Generalized map of Late Triassic
to Early Jurassic rift basins and major fault
systems along the eastern margin of North
America (after Swanson, 1986).
Sedimentary sequences in Mesozoic extensional and convergent tectonic settings
Figure 2. Generalized geologic map of
the Newark.Gettysburg Basin and the
HartfordDeerfield Basin (Tr.Jr rift
basins), northeastern United States
(modified from Glaeser, 1966; Schlisch,
1988; Huber and others, 1978).
Unit abbreviations: Jrp, Portland
Formation; Tr sl, Sugarloaf Formation;
Tr nh, New Haven Formation; TrJ r b,
Brunswick Formation; Tr 1, Lockatong
Formation; Tr s, Stockton Formation;
Tr no, New Oxford Formation. For pat-
terns and more detailed stratigraphy,
see Figure 3.
GETTYS BURG
BAS I N __
Figure 3. Generalized thickness of
stratigraphic sequences of the Newark
and Hartford Basins - note that this is
not a time stratigraphic chart (modi -
fied from Hubert and others, 1978).
TI IICK
'ESS
( Km)
y
7
I>
o
!
I
;
"- ,
,
,.. __ . .
.I
!
DEERFIELD
BASI N
I
/Tr
/ 1
/ l
J r p
HARTFORD
BASIN
BASIN
o 40
Km
' EI\' ARK B,\ SI'>
( ,y . 'J . P'\ )
rahs,ildes Sill M'Id
Oile!!!o
St<x tOn I Ne\", Oxford
f"n,:Triassl(' Rocks
N
II,., ___ ----
HARTF RD BASI'
( T )
rk)rdil::!r ftlngloml!' r.lLt'
r-orm.ltion
H,1mpdl,.'n Flow
East Derlin
Hm
I &::.all
... H.wen I Sugar 10M
Form;;1tions
fOre-TriaS!.il ROC"ks
_ Basalt Flow I Diaba,e Intrusion
Siltstone and Argillite
m:l Lacustrine Shale
Arkosic Sandstone and
Conglomerate
529
530 P. T. Ryberg
tinental sediments to accumulate in these basins,
along with subordinate mafic volcanic flows, dikes,
and sills. Basalt flows and diabase intrusions (in-
cluding the Palisades sill) associated with extension
were not emplaced until early Jurassic time (Fig-
ures 2 and 3).
The largest of the exposed basins in the United
States have received the most study, including the
Newark-Gettysburg Basin and the Hartford-Deer-
field Basin (Figures 2 and 3). Both are half-graben
structures bounded on one flank by complex listric
normal fault systems (Schlische, 1988), and strata
within these basins are tilted back toward the
border faults, with older strata exposed farther from
the fault. The Newark Basin system is divided by
accommodation faults (Schlische, 1988) into sub-
basins with their own distinctive patterns of filling.
The principal extension direction, perpendicular to
dike orientation, is a controlling factor of basin
geometry and depth (Schlische, 1988). Where the
border fault system is perpendicular to the exten-
sion direction, the faults are steep and the basin is
deepest, with the longest continuous sedimentation
history. Where the border fault system is oblique to
the extension direction, the faults dip more shal-
lowly, and the sedimentary record in the adjacent
part of the basin is not as thick and of less duration.
Some of the Newark border fault structures have
been identified as reactivated Paleozoic thrust
faults, based on seismic reflection profiles and drill
cores (Ratcliffe and Burton, 1985; Ratcliffe and
others 1986; Swanson, 1986).
Stratigraphic Sequence and Sediment Com-
position of the Newark-Gettysburg Basin
Sedimentation begins with the Stockton Forma-
tion (Figure 3) composed of arkosic conglomerate
and sandstone grading upward to interbedded sand-
stone, siltstone, and shale. Fining upward fluvial
cycles are common in this unit. Figure 4 shows the
quartzose and quartzofeldspathic compositions of
sandstones suites collected from this unit (Glaeser,
1966). Compositions of the Stockton arkose in New
Jersey (Van Houten, 1969, 1988) have a slightly
higher feldspar content (15-40% modal), with pla-
gioclase to potassium feldspar ratios of .., 2:1.
The gray-black argillites of the Lockatong For-
mation represent a widespread lacustrine sequence
(Glaeser, 1966; Van Houten, 1969) in the Delaware
sub-basin (eastern Pennsylvania and western New
NEWARK RIFT lIAS I"!
CLASTICS
F
I'ROVENA CE
Ci\ TEGORIES
o Cunlilu ..'nt,1t Blv(k
o Rt.'cyclt.d Orogt.'n
F
Q
Qm
L
Lt
Figure 4. Sandstone compositions for the Stockton
Formation, Newark Basin, Pennsylvania (data from
Glaeser, 1966). Key: fine dot pattern, cratonibasement
uplift source; coarse dot pattern, recycled orogen source;
diagonal line pattern, mixed source.
Jersey), which was deposited in the widest part of
the Newark Basin. A remarkable pattern of cycli-
city on several orders is present in this formation.
Cycles of .., 5-6 m thickness are the basic units and
have recently been referred to as Van Houten cycles
(in honor of Franklin Van Houten). Each basic cycle
(Figure 5) is generally asymmetric, consisting of
(1) a thin transgressive phase, (2) a highstand in
lake level, and (3) a thick regressive phase of sedi-
mentation dominated by desiccation features.
These basic cycles are nested within higher order
cycles (Van Houten, 1988). Very fine laminations
(interpreted as annual varves) within the deeper-
water black muds show the least amount of dis-
ruption or desiccation. These varved sequences are
Sedimentary sequences in Mesozoic extensional and convergent tectonic settings 531
Rtl('k
C\'de SI.."c tion
J)j , 'i =--inns
3
2
-
BI.lCk laminated siltst one
D
Gray mJssi n ' mudstont.>
with dcss1cdtion cr,l ck:,
D
RNl IThlssivl' n\\ldston>
with cr""cl s
D
Gray line sandstone-
RcI.lt ivc
1..,Kc Depth
Total
Organic
C. ... rbClIl
o
t t l lill
<@
""
..-.<

(,
111111 1
Clam shrimp

!'ish
Rl!ptill' skdl'l un::-
lo,l,ltrrllll:-
Figure 5. Generalized Van Houten cycle of lake sedi-
mentation; left showing: 1, transgressive phase; 2, deepest
lake level phase; 3, regressive phase. Diagram modified
from Olsen and Van Houten (1988).
Figure 6. Generalized sediment dispersion patterns of
the Newark-Gettysburg and Hartford-Deerfield Basins;
basin outlines show present extent of outcrops but were
probably more extensive at time of deposition. (Triassic-
Jurassic Sediment Dispersal Patterns)
source
used to time-calibrate sedimentation rates for the
entire basin (Van Houten, 1988).
The thick Brunswick Formation in Pennsyl-
vania is composed of red shale and siltstone, with
minor amounts of interbedded sandstone. The
Triassic-Jurassic boundary is within this unit. In
New Jersey, the Late Triassic Passaic Formation is
equivalent to the lower part of the Brunswick (Fig-
ure 3). The appearance of basaltic lava flows and
interbedded sediments characterizes the Early Jur-
assic sequence ofthe long-lived Delaware sub-basin.
Recent test corings for a Corp of Engineers flood
diversion project in central New Jersey provide a
complete sedimentary record of the Early Jurassic
sedimentary record (Fedosh and Smoot, 1988). The
sequence is locally capped by quartz-rich fanglo-
merate and limestone fanglomerate deposited by
debris flows in alluvial fans from point sources near
border faults (Glaeser, 1966).
Sediment Dispersal PaUerns in the Newark-
Gettysburg Basin
Basin fill began with quarzofeldspathic sedi-
ment carried northwestward by streams from a
granitic-metamorphic Piedmont source area to the
south (Figure 6). Sedimentation continued with
,---.---,
o 40
Km
r _.- R;;-ELD
/ BASIN
/
i

i
i
I -
I I
I t
N
1
532
P. T. Ryberg
input of Paleozoic clastic debris derived from the
fold-thrust belt to the north. Most of this sediment
entered the basin in the narrow neck area between
the two sub-basins and dispersed laterally into both
the Newark and Gettysburg Basins. Large lakes
developed in the widest and lowest parts of the
Newark Basin. Milankovitch orbital precession
cycles, which control the amount of solar energy
received at a particular latitude, are used to explain
the cyclic sedimentation patterns preserved in the
Lockatong Formation (Olsen and Fedosh, 1988).
The sedimentation rate diminished (exponentially)
as the basin widened and the depositional surface
area increased (Schlische and Olsen, 1988). The size
of the lakes depended on the amount of sediment
influx, as well as on precipitation and temperature
fluctuation. The last phase of deposition occurred as
alluvial fan aprons developed along the border fault.
Stratigraphic Sequence and Sediment Compo-
sition of the Hartford and Deerfield Basins
Deposition in the Hartford Basin began with
the New Haven Arkose (Figures 2 and 3), with
basal conglomerate grading upward to arkosic sand-
stone. The Deerfield Basin shows a similar pattern
in the correlative Sugarloaf Arkose. The Early Jur-
assic Talcott basalt flow overlies the New Haven
Arkose. The Shuttle Meadow Formation, which is
composed of arkosic conglomerate, grades upward to
sandstone and siltstone. The Holyoke basalt flow
overlies the Shuttle Meadow Formation, which is in
turn overlain by the East Berlin Formation, com-
posed of arkosic conglomerate and sandstone grad-
ing upward to lacustrine sandstone, siltstone, and
shale. The Hampden basalt flow separates the East
Berlin Formation from the overlying Portland For-
mation, composed of arkosic conglomerate grading
upward to lacustrine sandstone, siltstone, and shale.
Sediment Dispersal Patterns in the Hartford
and Deerfield Basins
Broad alluvial fans developed adjacent to the
eastern border fault (Figure 6). Braided stream
systems carried sediment southwestward into the
lower parts of the basin. Lakes developed locally in
the western part of the basin, resulting in deposition
of symmetrical lacustrine cycles, such as those in
the East Berlin Formation (Hubert and others,
Km Jg
Km
N
Km 20
I
124W
Figure 7. Generalized geologic map of the Western Jur-
assic Belt in southwestern Oregon and northern California
(map modified from Harper, 1984; Irwin, 1966; Wells and
Peck, 1961). Unit abbreviations: Tu, Tertiary Umpqua
Formation; Kr, Upper Cretaceous sediments; Km, Lower
Cretaceous Myrtle Group; Jcs, Colebrooke schist; Jsf,
South Fork Mountain schist; Jd, Dothan Group; Jf, Fran-
ciscan Group; Ji, Diorite plutons; Js, serpentinite; Jjo, Jo-
sephine ophiolite; Jcc, Chetco complex; Jr, Rogue Forma-
tion; Jg, Galice Formation; WKB, Western Klamath Belt.
1978). Milankovitch cycles also control lacustrine
sediment cyclicity here as well (Schlische and Olsen,
1988). Some Early Jurassic lake cycles can be cor-
related with incredible accuracy with those in the
Newark Basin.
Sedimentary sequences in Mesozoic extensional and convergent tectonic settings 533
MESOZOIC CONVERGENT BASINS OF WESTERN
NORTH AMERICA
The western margin of North America was
changing rapidly during Mesozoic time, as numer-
ous terranes accreted to the continent (Coney and
others, 1980). Accretion of the Western Klamath
Belt (Figure 7; formerly the "Western Paleozoic and
Triassic Belt" ofIrwin, 1966) to North America most
likely occurred during Early to Middle Jurassic time
(Gray, 1983; Hamilton, 1978; Davis and others,
1978; Irwin and others, 1978). Davis and others
(1978) suggested that the entire belt was imbricated
and disrupted by faulting since Middle Jurassic
time.
The following sections describe examples of
sedimentation in two general types of convergent
margin settings: marginal (back-arc or inter-arc)
basin, and forearc basin and adjacent subduction
complex.
Western Jurassic Belt, Klamath Mountains
The Western Jurassic Belt includes the Jose-
phine ophiolite, the Rogue metavolcanic rocks, and
the Galice metasedimentary rocks (Figure 7). The
Middle to Late Jurassic Josephine ophiolite (con-
cordant U-Pb age of 1502 Ma for plagiogranite;
Saleeby and others, 1982) is apparently marginal
basin ophiolitic crust, based on interpretation of
geochemical studies (Saleeby and others, 1982;
Harper,1983). The coeval Rogue Formation (K-Ar
ages of 150-157 Ma; Hotz, 1971) consists of basaltic
andesites and rhyodacites, interpreted to represent
a calcalkaline island-arc sequence (Garcia, 1979).
The Chetco complex in southwestern Oregon
(Figure 7) is interpreted as the plutonic roots of the
island arc (Harper, 1983). The Galice Formation,
although metamorphosed between prehnite-
pumpellyite and lower greenschist facies, does
contain a few fossils of the pelecypod Buehia eoneen-
triea, which indicate a late Oxfordian to early Kim-
meridgian age (Imlay and others, 1959). Galice
strata lie east of the Josephine and Rogue units and
are interpreted as sediments that accumulated in a
marginal basin between the Rogue island arc to the
west and the North American continent to the east
(Harper, 1980, 1983). Harper postulated that the
Galice graywackes, with their large component of
chert and lithic fragments, were in large part
derived from erosion of older accreted terranes of the
MARCI AL I TERAR BASI N
Galiet., (.\1 CaliforJ u(1)
Type At<. .Ii ,. ( IV
PROVENA CE
CA TEGORIES
o
o Rl'Cyckd Orohl'n
F
Qm
Lt
Figure 8. Composition of Galice Formation sandstones,
southwestern Oregon and northern California (data from
Harper, 1984).
Klamath Mountains (Figure 8). Submarine fans
prograded into the basin from the east. Basal Galice
argillites and green chert lie in depositional contact
with the Josephine ophiolite in southernmost
Oregon and northern California (Harper, 1983).
These strata apparently represent pelagic deposits
accumulating near the spreading center of the mar-
ginal basin. However, in the type area of the Galice
Formation (Diller, 1907; Wells and Walker, 1953),
volcanic clastics strata of the Rogue Formation
interfinger with Galice argillites along the western
margin of the basin. Figure 9 shows how the basin
may have looked prior to the Late Jurassic Nevadan
Orogeny.
Nevadan Orogeny - Tectonic Assembly and
Plutonism
The Late Jurassic Nevadan Orogeny ("'" 150
Ma) is well defined in California and Oregon by
intense deformation, presumably generated by the
final island-arc collision and accretion in the Sierra
Nevada and Klamath Mountains (Cashman, 1988;
Davis, 1966; Irwin, 1964, 1966, 1981). Hamilton
(1969) envisioned a major change at this time for the
western margin of North America - from similar
to the present-day western Pacific (e.g., Asia and
Japan) to similar to the present-day Andean-type
western margin of South America. The imbricated
patterns within the Klamath Mountains as we see
534 P. T. Ryberg
Figure 9. Sedimentary-tec-
tonic model for Galice strata
deposited in a marginal inter-
arc basin, prior to closure
during the Nevadan Orogeny
(modified from Harper, 1984).
ISLAND ARC REMNANT INTERARC BASIN WESTERN K LAMATH BELT
Rogue volcanics
Km
Interfingering
volcaniclastic-
epiclastic
sediments
Type Galice
pelagic
sediments
Galice
turbidites
-
remnant arC
-.-

50
? ?
collapse of basin
during Nevadan Orogeny
them today were largely complete by the end of the
Nevadan Orogeny (Davis, and others, 1978). The
inter-arc or marginal basin containing Galice strata
closed at that time, with coincident metamorphism
of the sediments to slates, schists, and phyllites
(Saleeby and others, 1982). Telescoping along the
major thrust boundary between the Western Kla-
math and Western Jurassic Belts was at least 30
km, if the Condrey Mountain schist (exposed in a
fenster beneath the Western Klamath Belt just
south of the Oregon-California border) is stratigra-
phically equivalent to the Galice strata. Helper
(1985) challenged this correlation, believing the
Condrey Mountain schist to be an older, isolated
thrust plate because of dissimilar structural fabrics
and geochemistry.
A new E-dipping subduction zone formed along
the western margin of North America,just outboard
of the newly accreted Western Jurassic Belt (Hamil-
ton, 1969, 1978). As a result of continued subduc-
tion, magmas were intruded into the accreted Kla-
math terranes in Late Jurassic to Early Cretaceous
time (Hamilton, 1969). Composition of magmas
ranges from gabbro to true granite, but by far the
greatest volume are of intermediate compositions
(Hotz, 1971; Hamilton, 1969). Calcalkaline pluton-
ism was widespread at this time in the Klamath
Mountains, but not in a well-defined continuous belt
as in the Sierra Nevada (Hamilton, 1969). Many of
the smaller, older plutons in the Klamath Moun-
tains were offset by imbricate faulting, whereas
younger plutons are not offset (Davis and others,
1978).
FOREARC BASI,
Grt'oll \' dlll'y tC At
o ,OR)
sUlloucnO:>l COMPLEX
SLOPE IlASll'
" ,\;;,.. !\,-
torU
I'ROVEN,\1'\:CF
CAHGORIES
CJ
CJ
CJ
F
F
Q
L
Qm
Lt
Figure 10. Compositions of Late Jurassic to Early Cre-
taceous sandstones of southwestern Oregon and California.
California data (Franciscan and Great Valley) are sum-
marized and tabulated in Dickinson and others (1982),
Oregon data from Ryberg (1984).
Sedimentary sequences in Mesozoic extensional and convergent tectonic settings 535
km t
DOTHAN-FRANCISCAN
50
SUBDUCTION COMPLEX
longitudinal
MYRTLE GROUP KLAMATH-SIERRAN
GREAT VALLEY
MAGMATIC ARC
active volcanic arc
trench fill slope baaln forearc basin fill eroded plutons
\
1",,"\' I" "" =ri'\I<-ct"
1/11/;1"" II \\ \\.\.-j .... ""
III( /(; I// j \
Figure 11. Sedimentary-tectonic model for Oregon and iii 11/ jill,. fl a
previously accreted terran..
California continental margin during Early Cretaceous //1 ////
time (modified from Dickinson, 1982). ////
Late Mesozoic Westward Growth
Post-Nevadan Mesozoic rocks along the fringe
of the Klamath and Sierra Nevada Mountains in-
clude the Franciscan and Great Valley sequences in
Northern California and Dothan and Myrtle Group
rocks in southwestern Oregon. Structural trends
along the Klamath Mountain fringe essentially
parallel those of the Klamath Mountains (Hamilton,
1969), except in the coastal areas near Cape Mendo-
cino (Diller, 1902) and near Port Oxford, Oregon
(Dott, 1965). Franciscan and equivalent rocks are
overthrust along their eastern margin (Figure 7) by
Western Jurassic Belt rocks along the South Fork
Mountain Fault (Blake and others, 1967; Lanphere
and others, 1968, 1978). Distinct, mappable belts of
Fransican rocks recognized by Bailey and others
(1964) and by Suppe (1969) include the Late
Jurassic-Early Cretaceous Central Belt and the
Late Cretaceous-early Tertiary Coastal Belt.
Similar belts continue into southern Oregon,
although they have not been mapped in great detail
(Ryberg,1984). The Dothan Group in Oregon is the
stratigraphic equivalent of the Franciscan Group in
California (Figure 7). The composition of sand-
stones from both groups is quite similar (Figure 10).
Both units are interpreted to be forearc basin fill,
although the Great Valley is much more widespread
and continuous. Myrtle Group strata now occupy
fault-bounded depressions along the Klamath
Mountain fringe (Figure 7) and are especially
common along the boundary fault with the Western
Jurassic Belt. The Myrtle Group grades upward
from basal conglomerates and pebbly sandstones to
rhythmically bedded sandstone, siltstone, and shale.
Folding of these Late Jurassic to Early Cretaceous
units was presumably accomplished during the
//
South Fork Mountain thrusting event in Late
Cretaceous time (Jones and Irwin, 1971).
Hamilton (1969) recognized the significance of
subduction of a large amount of Pacific sea floor
under the western margin of North America during
late Mesozoic time, and he was first to describe the
relationships between the coeval Franciscan me-
lange (along the trench wall), the Great Valley
"outerarc" basin (or forearc basin ofIngersol, 1978),
and the Klamath-Sierran plutonic/volcanic arc. A
later study by Dickinson and others (1982) sum-
marized the sedimentary-tectonic settings (Figure
11) and sandstone provenance relationships of Cali-
fornia. The margin of North America at that time
probably looked similar to the present-day Andes of
South America. As subduction continued, more ma-
terial scraped off the subducting plate was even-
tually accreted and uplifted to the continental
margin, forcing westward growth of North America
(Ernst, 1970). The trench slope was also the site of
locally ponded "slope basins" (Figure 11), which
trap sediments behind parallel thrust ridges.
REFERENCES
BAILEY, E. H., W. P. IRWIN, and D. L. JONES, 1964, Fran-
ciscan and Related Rocks, and their Significance in
the Geology of Western California: California Division
of Mines and Geology, Bulletin 183, 177 p.
BEHRMAN P. G., and G. A. PARKINSON, 1978, Paleogeo-
graphic significance of the Callovian to Kimmeridgian
strata, central Sierra Nevada foothills, California; pp. 349-
360 in D. G. Howell and K. A. McDougall (eds.), Mesozoic
Paleogeography of the Western United States: Society
of Economic Paleontologists and Mineralogists, Pacific
Coast Paleogeography Symposium 2, 573 p.
536 P. T. Ryberg
BLAKE, M. C., Jr., W. P. IRWIN, and R G. COLEMAN, 1967,
Upside-down metamorphic zonation, blueschist facies,
along a regional thrust in California and Oregon: U. S.
Geological Survey, Professional Paper 575-C, p. C1-C9.
CASHMAN, S. M., 1988, Finite-strain patterns of Nevadan
deformation, western Klamath Mountains, California:
Geology, v. 16, p. 839-843.
CONEY, P. J., D. L. JONES, and J. W. H. MONGER, 1980,
Cordilleran suspect terranes: Nature, v. 288, p. 329-333.
DAVIS, G. A., 1966, Metamorphic and granitic history of
the Klamath Mountains; pp. 39-50 in E. H. Bailey (ed.),
Geology of Northern California: California Division of
Mines and Geology, Bulletin 190, 508 p.
__ , J. W. H. MONGER, and B. C. BURCHFIEL, 1978,
Mesozoic construction of the Cordilleran "collage," central
British Columbia to central California; pp. 1-32 in D. G.
Howell and K. A. McDougall (eds.), Mesozoic Paleogeo-
graphy of the Western United States: Society of Econo-
mic Paleontologists and Mineralogists, Pacific Coast Paleo-
geography Symposium 2,573 p.
DICKINSON, W. R, 1982, Compositions of sandstones in
circum-Pacific subduction complexes and forearc basins:
American Association of Petroleum Geologists Bulletin, v.
66, no. 2, p.121-137.
__ , 1985, Interpreting provenance relations from
detrital modes of sandstone; p. 333-361 in G. G. Zuffa (ed.),
Provenance of Arenites: Reidel Publishing Company,
Dordrecht, The Netherlands, 478 p.
__ , and C. A. SUCZEK, 1979, Plate tectonics and
sandstone compositions: American Association of Petro-
leum Geologists Bulletin, v. 63, no. 12, p. 2164-2182.
__ , L. S. BEARD, G. R BRACKENRIDGE, J. L. ERJAVEK,
R C. FERGUSON, K. F. INMAN, R A. KNEPP, F. A.
LINDBERG, and P. T. RYBERG, 1983, Provenance of North
American Phanerozic sandstones in relation to tectonic
setting: Geological Society of America Bulletin, v. 94, p.
222-235.
__ , R W. INGERSOLL, D. S. COWAN, K. P. HELMHOOD,
and C. A. SUCZEK, 1982, Provenance of Franciscan gray-
wackes in coastal California: Geological Society of America
Bulletin, v. 93, p. 95-107.
DILLER, J. S., 1902, Topographic development of the
Klamath Mountains: U.S. Geological Survey, Bulletin 196,
69p.
__ , 1907, The Mesozoic sediments of southwestern
Oregon: American Journal of Science, 4th Series, v. 23, no.
138, p. 401-421.
__ , and G. F. KAY, 1924, Description of the Riddle quad-
rangle, Oregon: U.S. Geological Survey, Geological Atlas
of the United States, Riddle Folio 218.
DOTI, R. H., JR, 1965, Mesozoic-Cenozoic tectonic history
of the southwestern Oregon coast in relation to Cordilleran
orogenesis: Journal of Geophysical Research, v. 70, p.
4687-4707.
ERNST, W. G., 1970, Tectonic contact between the Francis-
can melange and the Great Valley sequence, crustal
expression of a late Mesozoic Benioff zone: Journal of
Geophysical Research, v. 75, p. 886-901.
FAILL, R T., 1988, Mesozoic tectonics of the Newark
Basin, as viewed from the Delaware river; pp. 19-41 in J.
M. Husch and M. J. Hozik (eds.), Geology of the Central
Newark Basin: Fifth Annual Meeting of the Geo-
logical Association of New Jersey - Field Guide and
Proceedings: Rider College, Lawrenceville, New Jersey,
USA,158p.
FEDOSH, M. S., and J. P. SMOOT, 1988, A cored strati-
graphic section through the northern Newark Basin, New
Jersey; pp. 67-81 in J. M. Husch and M. J. Hozik (eds.),
Geology of the Central Newark Basin: Fifth Annual
Meeting of the Geological Association of New Jersey
- Field Guide and Proceedings: Rider College, Law-
renceville, New Jersey, USA, 158 p.
GARCIA, M. 0., 1979, Petrology of the Rogue and Galice
Formations, Klamath Mountains, Oregon: Identification
of a Jurassic island arc sequence: Journal of Geology, v. 86,
p.29-41.
GLAESER, J. D., 1966, Provenance, Dispersal, and
Depositional Environments of the Triassic Sediments
in the Newark-Gettysburg Basin: Pennsylvania Geo-
logical Survey, General Geology, Fourth Series, Report 43,
168p.
GRAY, G. G., 1983, A kinematic analysis ofthrust faulting,
west-central Klamath Mountains, California [abstract):
Geological Society of America Abstracts with Programs, v.
15, p. 585.
__ , 1985, Structural. Geochronologic and Deposi-
tional History of the Western Klamath Mountains,
California and Oregon: Implications for the Early to
Middle Mesozoic Tectonic Evolution of the Western
North American Cordillera: PhD thesis, University of
Texas, Austin, Texas, USA, 224 p.
HAMILTON, W., 1969, Mesozoic California and the under-
flow of Pacific mantle: Geological Society of America
Bulletin, v. 80, p. 2409-2430.
__ , 1978, Mesozoic tectonics of the western United
States; pp. 33-70 in D. G. Howell and K. A. McDougall
(eds.), Meosozic Paleogeography of the Western United
States: Society of Economic Paleontologists and Minera-
logists, Pacific Coast Paleogeography Symposium 2, 573 p.
Sedimentary sequences in Mesozoic extensional and convergent tectonic settings 537
HARPER, G. D.,1980, The Josephine Ophiolite - Remains
of a Late Jurassic marginal basin in northwestern Cali-
fornia: Geology, v. 8, p. 333-337.
__ , 1983, A depositional contact between the Galice for-
mation and a Late Jurassic ophilite in northwestern
California and southwestern Oregon: Oregon Geology, v.
45, no. 1, p. 3-9.
__ , and J. E. WRIGHT, 1984, Middle to Late Jurassic
tectonic evolution of the Klamath Mountains, California-
Oregon: Tectonics, v. 3, p. 759-772.
HELPER, M. A., 1985, Structural, Metamorphic and
Geochronologic Constraints on the Origin of the Con-
drey Mountain Schist, North-Central Klamath Moun-
tains, Northern California: PhD dissertation, University
of Texas, Austin, Texas, USA, 264 p.
HOTZ, P. E., 1973, Blueschist metamorphism in the Yreka
Fort Jones area, Klamath Mountains, California: U. S.
Geological Survey Journal of Research, v. 1, no. 1, p. 53-61.
HUBERT, J. F., A. A. REED, W. L. DOWDALL, and J. M.
GILCHRIST, 1978, Guide to the Red Beds of Central
Connecticut: SEPM Eastern Section, 1978 Field Trip:
University of Massachusetts-Amherst, Department of
Geology and Geography, Contribution No. 32, 129 p.
IMLAY, R. W., H. M. DOLE, F. G. WELLS, and D. L. PECK,
1959, Relations of certain Jurassic and Lower Retaceous
formations in southwestern Oregon: American Association
of Petroleum geologists Bulletin, v. 43, no. 12, p. 2770-2785.
INGERSOLL, R. V., 1978, Paleogeography and paleo-
tectonics of the late Mesozoic forearc basin of northern and
central California; pp. 471-482 in D. G. Howell and K. A.
McDougall (eds.), Mesozoic Paleogeography of the
Western United States: Society of Economic Paleonto-
logists and Mineralogists, Pacific Coast Paleogeography
Symposium 2,573 p.
IRWIN, W. P., 1964, Late Mesozoic orogenies in the ultr-
amafic belts of northwestern California and southwestern
Oregon: U.S. Geological Survey, Professional Paper 501-C,
p. C1-C9.
__ , 1966, Geology of the Klamath Mountains Province;
pp. 17-38 in E. H. Bailey (ed.), Geology of Northern
CaUfornia: California Division of Mines and Geology,
Bulletin 190, 508 p.
__ ,1981, Tectonic accretion of the Klamath Mountains;
pp. 29-49 in W. G. Ernst (ed.), The Geotectonic Develop-
ment of CaUfornia (Rubey Volume 1): Prentice-Hall,
Englewood Cliffs, New Jersey, USA, 706 p.
__ , D. L. JONES, and T. A. KAPLAN, 1978, Radiolarians
from pre-Nevadan rocks of the Klamath Mountains,
California and Oregon; pp. 303-310 in D. G. Howell and K.
A. McDougall (eds.), Mesozoic Paleogeography of the
Western United States: Society of Economic Paleonto-
logists and Mineralogists, Pacific Coast Paleogeography
Symposium 2, 573 p.
JONES, D. L., and W. P. IRWIN, 1971, Structural impli-
cations of an offset early Cretaceous shoreline in northern
California: Geological Society of America Bulletin, v. 82, p.
815-822.
LANPHERE, M. A., M. C. BLAKE, and W. P. IRWIN, 1978,
Early Cretaceous metamorphic age of the South Fork
Mountain schist in the northern Coast Ranges of Cali-
fornia: American Journal of Science, v. 278, p. 798-815.
__ , W. P. IRWIN, and P. E. HOTZ, 1968, Isotopic age of
the Nevadan orogeny and older plutonic and metamorphic
events in the Klamath Mountains, California: Geological
Society of America Bulletin, v. 79, p. 1027-1052.
MOORE, G. F., and D. E. KARIG, 1976, Development of
sedimentary basins on the lower trench slope: Geology, v.
4, p. 693-697.
NORRELL, G. T., and G. D. HARPER, 1988, Detachment
faulting and magmatic extension at mid-ocean ridges: The
Josephine ophiolite as an example: Geology, v. 16, p. 827-
830.
OLSEN, P. E., 1980, Fossil great lakes of the Newark
Supergroup in New Jersey; pp. 352-398 in W. Manspeiser
(ed.), Field Studies in New Jersey Geology and Guide
to Field Trips: 52nd Annual Meeting of the New York
State Geological Association: Rutgers University, Ne-
wark, New Jersey, USA, 492 p.
__ , 1986, A 40-million year lake record of early Meso-
zoic orbital climate forcing: Science, v. 234, p. 842-848.
__ , and M. S. FEDOSH, 1988, Duration of the early
Mesozoic extrusive igneous episode in eastern North Am-
erica determined by use of Milankovitch-type lake cycles
[abstract): Geological Society of America Abstracts with
Programs,v.20,no.1,p.59.
RATCLIFFE, N. M., and W. C. BURTON, 1985, Fault
reactivation models for the origin of the Newark Basin and
studies related to U.S. eastern seismicity: U.S. Geological
Survey, Circular 946, p. 36-45.
__ , W. C. BURTON, R. M. D'ANGELO, and J. K.
COSTAIN, 1986, Low-angle extensional faulting, reacti-
vated mylonites, and seismic reflection geometry of the
Newark Basin margin, eastern Pennsylvania: Geology, v.
14, p. 766-770.
RYBERG, P. T., 1984, Sedimentation, Structure and
Tectonics of the Umpqua Group (Paleocene to Early
Eocene), Southwestern Oregon: PhD dissertation, Uni-
versity of Arizona, Tucson, Arizona, USA, 280 p.
SALEEBY, J. B., G. D. HARPER, A. W. SNOKE, and W. D.
SHARP, 1982, Time relations and structural-stratigraphic
patterns in opholite accretions, west-central Klamath
538 P. T. Ryberg
Mountains, California: Journal of Geophysical Research,
v. 87, no. B5, p. 3831-3848.
SCHLISCHE, R. W., and P. E. OLSEN, 1988, Structural
evolution of the Newark Basin; pp. 43-65 in J. M. Husch
and M. J. Hozik (eds.), Geology of the Central Newark
Basin: Fifth Annual Meeting of the Geological Asso-
ciation of New Jersey -Field Guide and Proceedings:
Rider College, Lawrenceville, New Jersey, USA, 158 p.
SNEIDERS, V. M., 1988, Origin of conglomerate strati-
graphy in the Fransciscan assemblage and Great Valley
Sequence, northern California: Geology, v. 16, p. 783-787.
SUPPE, J., 1969, Times of metamorphism in the Fran-
ciscan terrain of the northern Coast Ranges, California:
Geological Society of America Bulletin, v. 80, p. 135-142.
SWANSON, M. T., 1986, Preexisting fault control for Meso-
zoic basin formation in eastern North America: Geology, v.
14, p. 419-422.
VAN HOUTEN, F. B., 1988, Late Triassic-Early Jurassic
deposits, Newark Basin, New Jersey; pp.1-17 inJ. M.
Husch and M. J. Hozik (eds.), Geology of the Central
Newark Basin: Fifth Annual Meeting of the Geo-
logical Association of New Jersey - Field Guide and
Proceedings: Rider College, Lawrenceville, New Jersey,
USA,158p.
__ , 1969, Late Triassic Newark Group, north-central
New Jersey and adjacent Pennsylvania and New York; pp.
314-347 in S. Subitzky (ed.), Geology of Selected Areas
in New Jersey and Eastern Pennsylvania and Guide-
book of Excursions: Geological Society of America,
Field Trip No.4: Rutgers University Press, New Bruns-
wick, New Jersey, USA, 440 p.
__ , 1977, Triassic-Liassic deposits of Morocco and
eastern North America: A comparison: American Asso-
ciation of Petroleum GeolOgists Bulletin, v. 61, p. 79-99.
WELLS, F. G., and G. W. WALKER, 1953, Geologic Map of
the Galice Quadrangle, Oregon: Oregon Department of
Geology and Mineral Industries, scale 1:62,500.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Structure of the Mesozoic Marietta-Tryon graben,
South Carolina and adjacent North Carolina
JOHN M. GARIHAN and WILLIAM A. RANSON
Department of Geology, Furman University, Greenville, South Carolina 29613, USA
(received August 12, 1988; revision accepted April 15, 1989)
ABSTRACT
The Marietta-Tyron graben is a system of 21 brittle faults, cataclastic zones, and lineaments in the
Piedmont of South Carolina and adjacent North Carolina, developed as part of a broad zone of crustal
deformation during Mesozoic continental rifting. Sinistral normal-oblique movements dominate the N500-
700E fault system; dextral normal-oblique movements, possibly younger, also occur along three ENE-trending
faults. Recurrent, incongruous movements are suggested by scatter in regional slickenline data; nonetheless,
the mean vector orientations for slickenlines from each fault zone are remarkably consistent, produced by a
strong strike-slip component of oblique slip. A complex Mesozoic brittle history superimposed on Paleozoic
folding, metamorphism, magmatism, and jointing is recorded also in the multiple shearing and quartz vein-
filling textures offault-related, siliceous cataclastic rocks.
Locally, Late Triassic-Early Jurassic(?) diabase (dolerite) dikes are sinistrally strike-separated several
hundred meters across brittle faults. Positioning and steep attitudes of the faults likely are structurally
inherited from a northeasterly regional joint set.
Conjugate zeolite/epidote-coated brittle shears in bedrock gneiss and drusy quartz extension fractures in
cataclastic rocks indicate that dilational opening, oblique 90) to the N600E fault trend, operated during at
least the later stages of development of the graben. Regional extension oriented S35-45E may be responsible
for the sinistral normal-oblique longitudinal shearing experienced by all faults and many nearby joints of the
system.
REGIONAL SETTING
Steep brittle faults, discontinuously lined with
siliceous cataclastic rocks, trend eastward and
northeastward and attain widespread development
across the Blue Ridge, Inner Piedmont, Lowndes-
ville belt, and Charlotte belt provinces of the
Carolinas and northeastern Georgia. Some con-
spicuous examples include: (1) a N700E-trending
zone of "siliceous mylonite" skirting the northern
margin of the Tallulah Falls dome in the Blue
Ridge, called the Warwoman lineament (Hatcher,
1974); (2) in Laurens County, South Carolina, "not-
able" microbreccia exposures at the regional bend in
the Lowndesville belt (Griffin, 1977); (3) numerous
"flinty crush rock" exposures scattered across the
South Carolina Piedmont (Birkhead, 1973); and
(4) between northeastern Georgia and North Caro-
lina, a set of sixty prominent regional topographic
lineaments, including regional systematic joint
trends and the more conspicuous Mesozoic brittle
faults, detected on ERTS imagery by Penley and
Pickering (1973).
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 539-555. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
539
540 J. M. Garihan and W. A. Ranson
In the Knoxville 1 X 2 quadrangle southeast of
the Brevard fault zone, Hadley and Nelson (1971)
mapped zones of silicified breccia in the Piedmont
(20-60 km long and 4-8 km apart), corresponding to
parts of the Pax Mountain, Gap Creek, Poinsett, and
Cross Plains faults (Garihan and others, 1988).
Mapping by Griffin (1978) and a compilation by
Snipes (1981) show data for over 40 linear cata-
clastic zones 1-17 km long; 5 m-O.4 km wide;
trending N65-75E) in Oconee, Pickens, and Ander-
ROSE DIAGRAMS OF JOINT STRIKES
WHETSTONE OUACRANGlE HOLLY SPRINGS OOAllRANGLE TUGALOOLAKEOUACRANGLE

\ 7fY'
-
W E
12 10 8'"46 .<; Z:
ALL TABLE ROCK
AND CLEVELA D QUADRANGLES
( Modified ofter Hoselton, 1974)

2 0 8 o!.
...
PICKE S,OCO EE, AND A OERSON
COUNTIES
14410ints
(s...nmorlzed from map of Snipes
cod others, 1981 . F'19 II)
I88joints 164 joints
(MoClifie(1 ofler Acker and Hatcher,I970)
son Counties, South Carolina, south and west of the
study area (Figure 1). Cataclastic rocks (trending
N500-700E) along the Lowndesville belt and four
additional shear zones in Abbeville County, South
Carolina, are described by Griffin (1979). A com-
plicating factor is that brittle deformation is super-
imposed on mylonite and phyllonite of earlier Paleo-
zoic ductile deformation. In fact, several major
ductile fault zones such as the Modoc and Lowndes-
ville shear zones, not associated with Mesozoic
Gll!I!nIi(leCo.
__ .1S;!lTU!QO!.-' ..i.,..oo
Figure 1. Regional compilation of cataclastic zones and joint data, northwestern counties of South Carolina. Sources of
information for "silicified breccia zones" in Pickens, Oconee, and Anderson Counties: Birkhead (1973); Griffin (1978); Snipes
and others (1979); for faults and cataclastic zones in Pickens and northern Greenville Counties: Snipes and others (1979);
Garihan and Ranson (1986, 1987).
Structure of the Mesozoic Marietta-Tryon graben in the Carolinas 541
sedimentary basins in South Carolina and Georgia,
have significant amounts of vertical, post-Alle-
ghanian uplift; very young metamorphic biotite
mineral ages 260 Ma) are from areas along the
Mesozoic upthrown fault block (Kish, 1988).
Even a cursory review of published literature
indicates brittle crustal deformation, post-dating
Paleozoic orogenesis, produced probably hundreds of
normal faults in the western counties of South
Carolina. Many likely remain undetected or their
significance has gone unrecognized.
The purpose of this paper is to summarize meso-
scopic structural data which indicate that a signifi-
cant component of left normal-oblique movement
and S35-45E oblique extension operated along the
Marietta-Tryon fault system - a relatively narrow,
trough-like graben structure within a part of a
broad zone of Mesozoic crustal extension (Figure 1).
The study area in extreme northwestern South
Carolina and adjacent North Carolina encompasses
all or parts of thirteen 7 1I2-minute quadrangles.
Towns along the periphery of the area are Brevard,
Figure 2. Mesozoic brittle faults and lineaments in northwestern South Carolina. Line of cross-
section for Figure 4. Inner graben is dotted. GR, Green River fault; BC, Bobs Creek fault;
WS, West Lake Summit fault; ES, East Lake Summit fault; BP, Blakes Peak fault; CE, Cove
Creek fault; JC, Joels Creek fault; G, Gap Creek fault; RF, River Falls fault; CB, Coldspring
Branch fault; OCC, Oil Camp Creek lineament; CT, Colt Creek fault; P, Poinsett fault; M-
G, Melrose-Gosnell fault; CC-SB, Cox Creek-Short Branch fault; RS, Rocky Spur lineament;
HM, Hogback Mountain fault; CR, Chestnut Ridge lineament; CP, Cross Plains fault; PM, Pax
Mountain fault; OC, Oolenoy Church fault; DF, Devils Fork fault; MS, Mill Spring fault;
T, Tryon fault; BM, Bullard Mountain fault; d, diabase dike.
8245' 8230'
3515'-,F=====;;:===========y=:..t::=======;;::r=r:;;::::::::;r=7'=':==3==7=:..L::::::=J
8245
3500'
OBLIQUE SLIP FAULTS
dominantly strike slip
dominantly dip Slip
ex:



CP
,

: P
i i
i i
o 5KM 5M1 IOKM I5KMIOMI
;i:
DlP DIRECTION } LINEAMENT OR
75 FAULT ZONE
DIP DIRECTION, AMOUNT
GENERAL FOLIATION TREND
542 J. M. Garihan and W. A. Ranson
Tryon, Columbus, Tigerville, and Marietta (Figure
2). A recent detailed geologic roadlog through part
of the region is given in Garihan and others (1988).
CHARACTER OF BRITTLE FAULTS
Fourteen major Mesozoic faults and associated
splays and cross-faults of the Marietta-Tryon fault
system trend N50o-70E. At least 21 major and
minor faults are present, spaced 1-6 km apart. Dis-
continuous siliceous cataclastic zones less than 1
meter and locally up to several 100 m wide lie along
each fault zone. At one extreme, the New Liberty
Church exposure through the Pax Mountain fault,
ledges of cataclasite and quartz microbreccia (11 % of
the width) alternate with saprolitic gneiss in a fault
zone over 500 m wide (Garihan and others, 1985). A
spectrum of deformational textures in the cataclas-
tic rocks indicates a complex brittle history invol-
ving brecciation, re-brecciation, and multiple dila-
tional quartz vein-filling episodes. At places direct-
ly along the trace and adjacent to individual faults
of the system where cataclastic rock is absent,
intensely shattered and sheared Paleozoic gneiss
displays slickenlines and quartz-zeolite-epidote
mineralization in veins and along small sympa-
thetic faults and anastomosing fractures.
Topographically, the brittle faults are expres-
sed as: (1) prominent linear or arcuate valleys (500
m relieD, the faults perfectly coincident with por-
tions of major drainages; (2) aligned linear segments
of lower order stream tributaries; (3) very promin-
ent to very subtle NE-trending ridges; (4) aligned
headwater tributaries, saddles, and ridges along
strike; (5) prominent linear NE-trending slope
breaks; (6) linear shorelines (Lake Summit, North
Saluda reservoir); and (7) aligned springs. Un-
usually good exposures of fault-related rocks (by
Piedmont standards) are the result of modern
stream incision along joints and faults accelerated
by active Holocene uplift (2.6 mm per year relative
velocity, according to precise leveling surveys). The
uplift in this section of the southern Appalachians
centers on the Gulf-Atlantic drainage divide just
west of the Brevard zone (references in Gable and
Hatton, 1983). SLAR images and low resolution
1:20,000-scale USDA county photo-mosaics are
invaluable in tracing lineaments on the ground,
particularly in rugged gneiss terrane northwest of
the base of the Blue Ridge Front (elevations above,
roughly, the 450 m topographic contour of Hack,
1982).
PETROGRAPHY OF CATACLASTIC ROCKS
Excellent exposures of cataclastic rocks (see, for
example, Stops 2, 8, 10, and 13 in Garihan and
others, 1988) permit detailed study of brittle
deformational features in hand sample and thin
section. Mesoscopically, cataclastic rocks have tex-
tures ranging from very fine grained and struc-
tureless to coarse clasts of country rock in a finer
grained, crushed-rock matrix. In thin section, a
broader spectrum of deformational textures is
observed, which can be correlated with progressive
destruction of original rock fabric (Hallman and
others, 1987) by microscopic fracturing and break-
ing of grains and frictional sliding along intercon-
nected fractures, usually involving reduction of
grain size to varying degrees, but without visible
thermal effects. Episodes of recurrent motion along
brittle faults, well-demonstrated in structural data
(discussed below), led to disruption of earlier formed
cataclastic textures by means of renewed fracturing,
sliding, and reduction of grain or clast size.
Fault-related rock terminology remains proble-
matic save that rocks encountered in this study
result from brittle deformation and are nonfoliated.
According to the simplified classification of Wise
and others (1984), these rocks are "cataclasite."
Higgins (1971) distinguished other types of cata-
clastic rocks without fluxion structure but with
primary cohesion based on predominant grain size
of fragments in the rock. He used the term "micro-
breccia" for rocks containing> 30% fragments;;:: 0.2
mm or larger in size, and reserved the term "cata-
clasite" for rocks in which fragments are <0.2 mm
and make up < 30% of the rock. Hence, all fault-
related rocks discussed herein are classified simply
as microbreccia or cataclasite using Higgins' ter-
minology. These rocks may be further described
after the manner of Hallman and others (1987),
based on the degree to which their original texture
was disrupted.
A fabric believed to be precursory to the onset of
significant cataclasis produced veining of the other-
wise undisturbed country rock. Fractures of vari-
able width are commonly filled with stilbite, heu-
landite, chabazite, prehnite, epidote, fluorite, and
quartz (Miner, 1985). Associated with these veins or
Structure ofthe Mesozoic Marietta-Tryon graben in the Carolinas 543
Figure 3. Cataclastic textures in thin section and hand specimen; arranged in order of increasing degree of disruption of
original rock fabric. A) Photomicrograph (crossed polars) of narrow zone of cataclasis (between arrows) in sheared and veined
country rock from margin of Joels Creek fault. Note micro-faulted plagioclase along margin of cataclastic zone. Bar scale is
0.6 mm. B) Photomicrograph (crossed polars) of part of a large broken plagioclase with fine-grained opaque material and
small pieces of crushed plagioclase filling cracks. Bar scale is 0.6 mm. C) Photomicrograph (plane light) of pervasively
brecciated country rock from Joels Creek fault zone. Rounded clast of country rock consisting primarily of quartz (Q) and
feldspar (F) in matrix of crushed quartz and feldspar plus epidote and fine-grained opaque material. Bar scale is 0.6 mm.
D) Photomicrograph (crossed polars) of microbreccia from Joels Creek fault zone. Country rock clasts disaggregated to
component minerals at this stage. Bar scale is 0.6 mm. E) Photomicrograph (plane light) of cataclasite from Pax Mountain
fault zone. Note light-colored quartz veins criss-crossing matrix stained with iron oxide. Opaque squares are magnetite
crystals. Bar scale is 0.3 mm. F) Photograph of brecciated microbreccia from Gap Creek fault . Note dilational quartz veins
within microbreccia clasts (C) now truncated by matrix (M) and presence of younger generation of quartz veins within matrix.
544
J. M. Garihan and W. A. Ranson
fracture fillings are minor, very narrow zones of
cataclasis along vein walls (Figure 3A).
Disruption beyond simple veining results in
sheared country rock in which shear planes are
delineated by parallel micas and opaque phases.
Large grains are fractured, and fractures are filled
with fragmented minerals, epidote, and opaques
(Figure 3B). Regional Mesozoic crustal thinning,
elevation of isotherms, and attendant hydrothermal
activity are responsible for this striking mineral
assemblage. Such a complex system of zeolite/epi-
dote-coated brittle fracture surfaces (faults) and
veins is exposed in hanging wall rocks 0.65 km
southeast of the trace of the Gap Creek fault (Stop 4
in Garihan and others, 1988).
With continued fracturing the country rock be-
comes pervasively brecciated. Brecciation is readily
visible in hand specimens, and the rocks are appro-
priately referred to as microbreccia. They contain
abundant recognizable clasts of country rock. These
country rock fragments, angular to rounded clasts
(0.2-3 cm) usually of granitic gneiss, are rotated and
separated by large fractures filled with fragmented
quartz and feldspar plus epidote and opaques (Fig-
ure3C).
Concentration of strain into narrower zones or
intensification of brittle deformational stresses pro-
duces microbreccia with no or few remaining frag-
ments of country rock. Most fragments at this stage
are disaggregated into angular, broken feldspar and
quartz grains displaying undulatory extinction and
seriate texture (Figure 3D). Quartz veins are com-
mon and may be abundant.
Additional mechanical deformation further re-
duces grain size of mineral constituents until a fine-
grained, structureless rock results (Figure 3E).
Because grain size is mostly <0.2 mm, the rock is
referred to as cataclasite. Mineralogy is typically
difficult to identify, even in thin section, but consists
primarily of quartz with lesser amounts of feldspar
and opaque minerals. Multiple generations of criss-
crossing, dilational quartz veins are the most distin-
guishing feature of this rock type in hand speci-
mens.
Such textural progression described above may
be partially or fully developed along anyone fault.
For example, fault rock discontinuously developed
along the extensive Pax Mountain fault zone (500 m
wide) is predominantly cataclasite, whose resistance
to erosion is responsible for the knife-edge topo-
graphy of Pax Ridge south of Tigerville. In contrast,
along the Joels Creek fault (Stop 8 in Garihan and
others, 1988), a variety of cataclastic textures is
beautifully developed within the fault zone (2 m
wide), including sheared country rock, microbreccia,
and - near the center of the fault zone - catacla-
site and brecciated microbreccia. The latter consists
of rounded and angular fragments of cataclasite and
microbreccia S; 1cm set in a microbreccia. Breccia-
ted microbreccia is spectacularly developed locally
along the Gap Creek fault (Stop 10 in Garihan and
others, 1988), where larger clasts (s;5 cm) of micro-
breccia containing dilational veins are truncated at
clast margins by the microbreccia matrix (Figure
3F). These textures indicate multiple movement
along faults of the Marietta-Tryon graben.
STRUCTURE IN PALEOZOIC ROCKS
Metamorphosed Paleozoic rocks in the study
area include various biotite-microcline granitoid
gneisses (orthogneiss and paragneiss), augen gneiss,
Caesars Head "quartz monzonite" orthogneiss (on
the North Carolina state map, Devonian to Silurian
age; Fullagar, 1983), and subordinate amphibolite
(both interlayered units and discordant mafic
dikes), sillimanite-biotite schist, calc-silicate rock,
and migmatite. Metamorphic foliation and composi-
tionallayering, generally parallel except in noses of
early isoclinal folds, are well developed. Major met-
amorphism occurred during the Taconic orogeny
"" 450-480 Ma (references in Nelson and others,
1987).
Detailed geologic mapping of bedrock structure
and splitting out of major units at 7 1!2-minute
quadrangle scale generally has not been undertaken
in the study area, except in the region between
Marietta, Cleveland, and Pax Ridge in the southern
half of the Tigerville quadrangle (Snipes and others,
1984; Warner and others, 1985). There, a NW-
trending, Late Triassic-Early Jurassic(?) diabase
(dolerite) dike shows left-strike separation of 125 m
across the Cross Plains fault, based on float (Snipes
and others, 1979). Three kilometers northwest-
ward, on the same basis, Garihan and others (1988)
mapped 30-100 m of left-strike separation for the
same dike across the Cox Creek-Short Branch fault.
The maximum age of at least the later stages of
brittle fault movement is thereby established by
offset of diabase.
At broader perspective, in the adjacent Green-
ville lOX 2
0
quadrangle, regional tectonic relation-
ships of these Paleozoic sillimanite-grade rocks of
Structure of the Mesozoic Marietta-Tryon graben in the Carolinas 545
the Inner Piedmont were mapped by Clarke (1981)
and Nelson and others (1985, 1987), who recognized
five stacked crystalline thrust sheets in the Pied-
mont. An unsolved question (related to lack of
detailed mapping) is how steep Mesozoic faults
relate to these shallower dipping Paleozoic faults. A
currently fashionable concept, for example, in the
Virginia Piedmont (Glover and others, 1980) or in
the Pennsylvania Newark Basin (Ratcliffe and
others, 1986), is structural heredity of Mesozoic
fault positioning and attitude from earlier Paleozoic
thrusts. However, at present no such controlling
relationship has been established for faults of the
Marietta-Tryon graben. Our data tentatively sug-
gest inheritance from steep regional systematic
joints.
We measured regional metamorphic foliation
attitudes in all quadrangles, and at several of the
better exposures; fold element attitudes also were
recorded (see road-log of Garihan and others, 1988,
Stops 4 and 6), to determine the meso scopic struc-
tural chronology. At least three phases of folding
are present in the metamorphic units: (1) (oldest)
F 1 tight to isoclinal, steeply inclined to overturned
fold hinges plunge 100-25/N5OW-N35E; (2) F2
gentle to isoclinal, upright to overturned to recum-
bent fold hinges plunge 00-25/N300-500W; and
(3) (youngest) F3 open to gentle, upright to steeply
inclined fold hinges of gentle NE-plunge.
The latter folds are represented on the macro-
scopic scale by broad regional warpings of moder-
ately to gently dipping metamorphic foliation
surfaces across the study area; stereoplot analysis
indicates {3= 15/N62E. The orientation of the
maximum density of contoured poles is N18W,
15NE - the best-fit foliation attitude in the 13-
quadrangle area. Gentle dips to foliation dominate
over extensive regions. A broad antiformal F3 warp
or arch of metamorphic foliation larger than the
study area and plunging N62E is present. The Blue
Ridge synform, a conspicuous, smaller fold covering
the SE 114 ofthe Tigerville 7 1I2-minute quadrangle
(map of Snipes and others, 1984), plunges 22/N48E
and is an F3 fold (Figure 2). It lies on the flank of
the antiform. Nelson (1985) attributed all four
major regional fold phases, in the Greenville lOx 2
sheet, to a single metamorphic-deformational con-
tinuum - the Taconic orogeny. Nonetheless, little
modern geochronology has been applied to the study
area. In the light of recent detailed geologic and
geochronologic studies in the eastern Piedmont
demonstrating significant episodes of Alleghaniam
deformation and metamorphism (Secor and others,
1986), it is not known whether the folds are early,
mid-, or late Paleozoic in age.
MARIETTA-TRYON FAULT SYSTEM
Where they can be measured, cataclastic zones
typically dip 75; rarely are they vertical, as con-
sistently described in the literature. Moreover,
reversal in attitude of cataclastic zones along a line
of cross-section (Figure 2) and slickenline data
indicate the presence of a graben-within-a-graben
structure: an outer graben at least 725 km
2
in aer-
ial extent between the arcuate Green River and the
Pax Mountain faults and an inner graben of 235
km
2
between the Gap Creek and Cross Plains faults
(Figure 4). Southwestward, each of these four im-
\. . .... 0 ute
Ii r u hen ........ ,
no vertical exaggeration
Figure 4. Cross-section showing brittle fault system of Marietta-Tryon graben. Wavy lines are attitudes of metamorphic
foliation.
546 J. M. Garihan and W. A. Ranson
portant brittle faults subdivides into joint-
controlled, W-trending splays. Examples of splays
are the River Falls and Coldspring Branch faults.
In the vicinity of Cleveland and River Falls,
South Carolina, at the juncture of northeasterly and
westerly trends, the inner graben tapers markedly
and terminates as its bounding faults die out. The
outer graben continues westward beyond the study
area. Southeast of Brevard, North Carolina, the
NE-trending Green River fault merges with a pro-
minent N600W joint-controlled lineament along
Carson Creek and assumes its arcurate trace. To
the south, the W-trending splay of the Pax Moun-
tain fault continues from Marietta, South Carolina,
across Pickens County, terminating just east of
Lake Jocassee (Stafford and others, 1983).
The floor of the inner graben is highly faulted
(Figures 2 and 4, dotted areas). Beneath the North
Saluda reservoir, the NE-trending Cox Creek-Short
Branch fault terminates against the ENE-trending
Melrose-Gosnell fault. The latter, possibly younger,
fault crosses the Poinsett fault beneath the reservoir
without obvious separation. The northeastern por-
tion of the brittle fault system beyond the Mill
Spring fault (Davenport, 1987) near Columbus,
North Carolina, is outside the study area. However,
scattered occurrences of linear cataclastic zones
mapped by Conley and Drummond (1965) between
Forest City and Morganton, North Carolina, sug-
gest an arcuate swing of the graben-related struct-
ure - from an ENE to a more NNE trend.
Relevant specific information concerning char-
acteristics of the important faults of the Marietta-
Tryon system is presented in Garihan and others
(1988).
REGIONAL SYSTEMATIC JOINTS
In northwestern Oconee County, South Caro-
lina, Acker and Hatcher (1970) first recognized pro-
minent rectangular stream drainage patterns
produced by strong N400-500W joint control on head-
ward erosion and resulting stream piracy (Figure
1). Several cataclastic zones in the Tugaloo Lake
and Holly Springs quadrangles also have that trend,
suggesting that the brittle faults reactivated an
important joint weakness direction. In addition to
the dominant N300-80OW and N300-700E systematic
joint trends established in the Piedmont (Figure 1),
an easterly direction is common but less well deve-
loped. An analysis of systematic jointing confined to
the immediate study area confirms that, excluding
sheeting joints, well-developed N500-700E, N800E-
N800W, and N500-700W oblique sets superimposed
on Paleozoic fold patterns are present (Preddy and
others, 1987).
A summary rose diagram (N=1,683 joint
planes) is given in Figure 5. We interpret the N500-
700E trend of brittle faults as having developed by
shearing along the most widespread, steep joint set
at N600E (Garihan and Ranson, 1986). In order to
substantiate this sweeping conclusion, subdomains
of joint data were chosen from areas immediately
adjacent to three faults (Joe Is Creek, Colt Creek,
Cross Plains) and two aerial photographic linea-
ments (Rocky Spur, Oil Camp Creek). In each case,
the map trace of the fault or lineament can be de-
monstrated to coincide exactly or within 10 of the
statistical strike of a significant joint concentration
(contoured on the diagram). Commonly the trace
coincides with the NE-trending joint set. The ano-
malous N15E Joels Creek fault trend is identical to
a subordinate N15E regional joint trend found
north of the state line; the N700E joint set, however,
dominates in that area. Similarly, the Oil Camp
Creek lineament is parallel to the easterly regional
joint set, although the NW -trending set is more
frequently encountered in that particular region.
N65E and N85E segments or jogs in the trace of
the Cross Plains fault are likely structurally con-
trolled by shearing along, respectively, N600E and
N85W joint sets (Figure 5). As confirmation, reac-
tivated zeolite/epidote-coated joint surfaces, which
are slickensided, are notably common adjacent to
the Cross Plains fault. In general, the gross geome-
try of the broadly arcuate outer graben, bound by
the Green River and Pax Mountain faults, indicates
utilization of all three pre-existing regional syste-
matic joint trends during longitudinal shearing that
accompanied oblique Mesozoic crustal extension.
KINEMATICS
Longitudinal Shearing
Other than Mesozoic diabase dikes, apparently
sinistrally strike-separated on the order of several
100 m across the Cross Plains (Snipes and others,
1979) and Cox Creek-Short Branch faults (Figure
2), no markers within Paleozoic gneisses demon-
strate separations along the Marietta-Tryon fault
system. Displacement along many individual
Structure of the Mesozoic Marietta-Tryon graben in the Carolinas 547
et2t 45'
35'ls'- i; '
ALL JOINTS fRClM STUOV AREA fII I'6SJ

CONTOUH'S-l5%.,37%,67"4,96"'"
(MJlX DENSITy '10:1 1%3

t
CAMP
CREEl(
UNf,llMCNf
-S;:: IS'
- "",.'

SPUR
UNE:AMENr
Tryon
-!!II


\
COLT SC
I r'
FAUlT ......... ' .........-
'-, ,/
CONTOlF!S '1 2'4.2 4%.
2%, 14 1% ( oII A
16:;"4)
CCtHOUR$
09%,26%,52%.
78%,10 4% (MAX
Oe:NSlTY'122%I
Figure 5, Comparison of contoured joint data with map traces of selected faults and lineaments.
brittle faults probably is not great - a factor dif-
ficult to demonstrate unequivocally due to generally
poor exposure of sa pro Ii tic gneiss and schist,
In a few unusually good bedrock exposures
(shown in Figure 2), for example: (1) a road cut
along Route 25 (Zirconia quadrangle), (2) a quarry
near Mill Spring, and (3) a railroad cut near Tryon,
detailed examination indicates that systematic
brittle fracturing and slickenline development
(Figure 6) are indeed widespread and pervasive in
areas between and away from major faults. In
newer manmade exposures, en echelon fractures -
the result of a normal shear sense - are plainly
visible, with incipient or slight offsets (a few centi-
Til 'YON RAILROAD CUT
Sl.ICKE.. ...l.I ... S$ ON
FR ....CTURE sURFACes
., ",O-<IOE
fkACTIJR E SlIT
_=
FRACTURe SET
EASTER.l.Y
FRACTUlH'! SET
MIU. QUARRY
SPRISO, NC
Sl..iCt.:"LU"E
ON
fRACTURE
SUR-PACE
Figure 6. Slickenline and fracture surface data from
localities near Tryon and Mill Spring, North Carolina; see
Figure 2 for locations.
548
J. M. Garihan and W. A. Ranson
meters) of meso scopic gneissic compositional layers
(for example at Stops 4 and 5, Garihan and others,
1988). Outcrop-scale faults and associated slicken-
lines substantiate the predominance of normal-
oblique displacement and relatively small net slip,
which we believe also characterizes the larger
brittle fault zones. Only near a few major faults
with mostly strike-slip movement, such as the
Green River fault, are minor mesoscopic faults
themselves essentially strike slip, based on the
slickenlines they display.
In both stereoplots (Figure 6), three sets of
bedrock fracture surfaces are present: a NNE-, a
NE-, and an E-trending set. Fracture surfaces here
clearly are not joints, yet their trends are very
similar to regional joint directions. We note essen-
tially dip-slip or slightly sinistral normal-oblique
slip dominates on the E-trending set of fractures
(faults) at the Tryon railroad cut and the Mill
Spring quarry localities (Figure 6). Sinistral
normal-oblique slip with a dominant dip-slip
component exists on NNE-trending fractures at the
Tryon railroad cut. The cut lies within footwall
rocks of the Tryon fault, suggesting that the frac-
tures formed sympathetic to sinistral normal-
oblique movements along the major fault.
Slickenline data were collected regionally from
those fracture surfaces - developed within each
fault zone or in adjacent high-grade, sheared
gneisses - which shared the same approximate
map trend as the fault. Data were collected from
shears in bedrock as close to each fault as exposures
would permit. . Although all three fracture sets
described above typically are present in the data
collected for each brittle fault zone, only one or two
sets tend to dominate. For each fault, the plunges of
slickenline lineations lying on great circle traces
representing, for example, the NE-trending fracture
or joint set may be distributed: (1) as a relatively
tight grouping (Figure 7, data for the Bullard
Mountain and Poinsett faults); or (2) as a fan of
slickenline plunges consistent with a wide range of
strike-slip to dip-slip, sinistral to dextral normal-
oblique movements. Examples of the latter are data
for the Hogback Mountain fault, the Mill Spring
fault, or, especially, the Cross Plains fault. Recur-
rent, incongruous motions along faults and fract-
ures, presumably over a prolonged period, could
cause the scatter.
We used stereoplots (Figure 7) to document the
complex kinematic plan for faults of the Marietta-
Tryon graben. Structural analysis indicates that
the following faults are dominated by
left normal-oblique movements (net slip unknown):
(1) Bobs Creek; (2) Bullard Mountain; (3) Cox
Creek-Short Branch; (4) Gap Creek; (5) Hogback
Mountain; (6) Mill Spring; (7) Poinsett; and
(8) Green River faults. The somewhat more ENE-
trending faults, on the other hand, are dominated by
right normal-oblique movements (net slip un-
known): (1) Cross Plains; (2) Melrose-Gosnell; and
(3) Pax Mountains faults. Based on structural ana-
lysis of regional slickenlines (Figure 2), oblique-slip
faults dominated by a strike-slip component are
distinguished from those dominated by a dip-slip
component. The Green River, Bobs Creek, and Gap
Creek faults along the northwestern margin of the
outer and inner graben, for example, are members of
the Marietta-Tryon fault system which display
principally strike-slip movement.
A contoured stereoplot of slickenlines from the
Pax Mountain fault zone (Figure 7) shows two
concentrations. The concentration at 75JN53W
indicates that dip-slip movement along the fault
dominated in the vicinity of the New Liberty
Church cut (Figure 2) (Garihan and others, 1985);
elsewhere, particularly in the Pax Ridge area, the
concentration of slickenline plunges at 15JN55E
indicates that significant dextral normal-oblique
slip also occurred along this NE-trending fault.
Moreover, "" 2% of the fractures investigated for the
Pax Mountain fault display multiple, non-over-
printed slickenline lineations on the same surface.
Commonly, one set of lineations roughly parallels
the strike and the other the dip of the fracture
surface. We interpret this observation as consistent
with recurrent, dramatically dissimilar movements
along the fault during its kinematic history. Un-
fortunately, and perhaps surprisingly, convincing
risers on fault surfaces are rare.
The sinistral strike-separation of a diabase dike
across the Cross Plains fault (Snipes and others,
1979) is seemingly inconsistent with the predomin-
ance of right normal-oblique fault-slip, based on our
slickenline data for that fault. If, however, un-
obliterated slickenlines are accepted to represent on
Figure 7. Synoptic southern hemisphere, equal-area
stereoplots of fracture surface traces and associated
slickenlines for faults of the Marietta-Tryon system.
Last diagram is a stereoplot of vector means of slicken-
lines for each of 10 major faults.
Structure of the Mesozoic Marietta-Tryon graben in the Carolinas
OJtE'RI'lltfAl.1.T
0'"
"'-ORTIiEAST so Of
f'RACTl.-it S,-,Il'A,("ts
"'OJOI!l.TS
808SCltE(K fAULT
OAP CREEl( FAut..T
SU:::JtEJool.lSESON
f1lACTll'R1! SUltP"CS
_1!A.$TfStLY
FRACTURE SeT
81.'1..lAllD MOU:o.'TAIN fAIJLT
SLICr.f.''LIN"ES OS
NORTHEAST
FRACTURe
SURFACES
\lfl.11.0sE OOSND..L PAIJt.T
. 1I. $l..K;:I(SlINESO.
''Oil rJtA,ST SET Of!
FRACTURE SVl.fACa
!l,IO;t!NLI"'ES OX
EASTDLY SET Of!
FRACTURE SURfACD


ON
FRACT\,:1tE!
SIJRPACe
OOX CREE':: PAl,;'l.T
SUCKE''t.IS"ES os
"'l<'lRTIIWST

NORTH
NORTHEAST
I'kACTUI<E
SURPACf
NORTtII!An
AtAC1'L"I:[
SuaFACE
POINSETT' P .... IJt..T
", Sl.K:K[SLIMi
JOII"o'" :!fURPACe
SLtct:l!XLI'E
FRACTURE SURfACE
HOOBACK MOUNTAIN PAULT
.st.ICKIlSLlNES ON
_NOII.THl!An

f1tACT\I1t
S'LlllPACES
eASTERLY ..101"'-.,.
(:R.OSS ft..\I\;S f'Al.I.T
SLICt:.EI'OLlI'roIES Os
FRACTIJa.E SURFACES
NCmTtI NORTHEAST
F1t.ACTlJ'RE SET
NOi-TlleAn
Fl!.ACTU'ft1! SET
EASTt'llLY
FRACTURE: StT
VECTOR OF
AU. NOftTItIlAST r.
flt.ACTuae
SURPACe
SUCK!NLINES. .."",
FOR EACH :4.5
PAU\.T

ODl..)QUE!
. .. RlOlrr NOJ.MAL
OflUQUE
549
550
J. M. Garihan and W. A. Ranson
a statistical basis the younger movements along a
fault, then sinistral normal-oblique slip post-dating
and displacing Late Triassic-Early Jurassic(?) dia-
base may have been important at an earlier stage of
the kinematic history of the fault - that is, at a
stage earlier than represented by the slickenlines
we now observe. By implication, the younger right-
oblique movements were insufficient to reverse the
sense of sinistral strike separation on the dike
produced earlier. The same diabase is sinistrally
strike separated across the Cox Creek-Short Branch
fault to the northwest; the slickenline array on the
appropriate fault stereoplot in this case indicates no
important component of dextral normal-oblique slip.
Rather it shows sinistral normal-oblique slip con-
sistent with left-strike separation of the dike.
An additional, somewhat equivocal argument
for a difference in character - and possibly a youn-
ger age - for the dextral normal-oblique move-
ments is the slight variation in trend displayed by
the ENE-trending Pax Mountain, Cross Plains, and
Melrose-Gosnell faults relative to the general north-
east trend of most members of the fault system. The
Melrose-Gosnell fault (dextral normal-oblique)
truncates the Cox Creek-Short Branch fault
(sinistral normal-oblique) and crosses the Poinsett
fault (sinistral normal-oblique), thereby indicating
it may be younger.
A summary of kinematic data is presented (Fig-
ure 7) in terms of ten vector means calculated for all
slickenlines found in each of ten major fault zones.
Considering the apparent scatter that is present in
individual slickenline plots of Figure 7, the result of
incongruous motions, the consistency in orientation
is quite remarkable when one compares the calcu-
lated mean slip-vector for border faults of the inner
and outer graben (on the diagram, respectively: GC,
CP and GR, PM; plunges are mostly <45). We con-
clude that contemporaneous with or subsequent to
initial crustal rifting that produced the graben,
significant longitudinal shearing, dominated locally
by a strong component of strike-slip, occurred along
major normal faults of the system (Ranson and
Garihan, 1988).
Crustal Extension
Two general categories of meso scopic structural
features are used to determine crustal extension
direction affecting these rocks in Mesozoic time:
(1) brittle conjugate shear faults in gneiss, marginal
to the Gap Creek fault zone; and (2) regional ex-
tension fractures in cataclastic rocks of four major
fault zones. Data from both types of features are
mutually corroborative, and their widespread deve-
lopment lends credence to the conclusion that
extension oriented S35-45E operated across the
graben system in at least the later stages of opening.
Two conspicuous conjugate sets of shear sur-
faces (N100-200E and N45E) in the Route 25 cut
(Figure 2) display well-developed slickenlines
superimposed on zeolite-epidote fracture coatings
(Figure 8). Where detected, displacement of com-
positional layering across these minor faults is
small, generally a few centimeters or less, and of a
normal-oblique slip sense. A great circle drawn
through the girdle of points produced by plunging
slickenlines represents the plane of the maximum
and minimum principal stresses, sigma one and
sigma three. From these data, orientation of the
axis of extension - sigma three - operating during
this episode of minor fa ul ting was "., 1 0 IN 4 7W. It
is similar to the regional direction indicated by
extension fractures.
Planar to lenticular, open fractures (generally
< 1-2 cm wide and lined with small drusy quartz
crystals) are encountered as late-stage extensional
features in a small proportion of exposures of
cataclastic rocks. These brittle features normally
B
REOK>SAL I:XTE.XSIO'" Dla. OCTIOSS
FQR EAC1t FAl.'l.T, N
Figure 8. A) Conjugate zeolite-epidote
fracture surfaces and associated slicken-
lines used to determine extension direction
in hanging-wall rocks of the Gap Creek
fault; Route 25 cut. B) Stereoplot of mean
extension directions determined from poles
to drusy quartz extension fractures, for
various faults; 12/S38E vector mean is the
regional extension direction for the graben.

F'IlACTIJRE
:n..'"RPAce
OYA MEA}r, POLl! TO
()Run QUARTZ
E)(TE.'StO:o. FRACT\laES
"'-

OC OAP OtUK IN 15;
8\1 8UL.l.A1l1) MTN (:.. 'I
/IIOC NOkTHF.AST PAlT
16)
PM SOUTtn.' EST PART
OF PAX MTN' (N
Structure of the Mesozoic Marietta-Tryon graben in the Carolinas 551
are undeformed; delicate quartz needles grow from
each margin inward toward the center of the
extension fracture. Rhombohedral faces of opposing
quartz crystals locally at the center are mutually
interlocking, where the fractures are completely
healed. The drusy quartz extension fractures pro-
bably represent the latest stages of dilational vein
filling, certainly developed subsequent to the epi-
sodes of intense shearing that brecciated the coun-
try rock gneisses and in turn rebrecciated the
cataclastic rocks themselves.
Along four faults of the Marietta-Tryon system,
a sufficient number of attitudes of planar exten-
sional fractures determine a mean pole to extension
fractures (Figure 8B). The mean pole attitude is
essentially the orientation of sigma three operating
across each fault zone at a time late in its kinematic
history. A vector mean for all six data sets is calcu-
lated as 12/S38E. We regard this determination of
the regional extension direction as one significant
contribution of our regional studies, agreeing well
with the S40o-50oE extension direction postulated
by Ratcliffe and others (1986) for the Newark Basin
in Pennsylvania. A S35-40oE direction is confirmed
by an expanded study covering the entire upstate
region (Garihan and others, 1990).
Transtension in the Joels Creek Area
In terms of kinematics, an analysis of faulting
in the Joels Creek area is informative (Figure
9A,B). It is plausible that transfer of stress between
the NE-terminating Gap Creek fault and the Cove
A
Creek fault produced in the transtension zone of left
en echelon overlap a small, SE-dipping fault at Joels
Creek. Development of that fault may have utilized
a pre-existing regional joint weakness direction
(N15E). The great circle distribution of slicken-
lines on faults from the Joels Creek area (Figure
9C) suggests extension oriented N45W-S45E. If
the compressional axis was oriented N45E, con-
sistent with sinistral normal-oblique shear-sense on
the two master faults, then dextral normal-oblique
slip suggested by the slickenlines of the Joels Creek
fault zone might be expected to be produced by
transtension.
DISCUSSION
Widespread development of N50o-70oE cata-
clastic lineaments in northwestern South Carolina
indicates a broad zone of brittle crustal deformation
post-dating Paleozoic tectonism. Crustal extension
affects a region far larger than the study area,
wherein a relatively narrow, trough-like structure
is present. We have termed this feature the
Marietta-Tryon graben, including an inner and
outer graben. Trends and textures of fault-related
cataclastic rocks are so similar to features along the
margins of Triassic-Jurassic sedimentary basins of
the Appalachians - for example, similarity to
microbreccia of the Chatham fault zone that flanks
the northwest border of the Danville basin - that
we have tacitly assumed they are also of Mesozoic
age. All previous workers reached the same
conclusion.
c
JOELS CREEK
SLICKENLINES ON
FRACTURE SURFACES
N
Figure 9. A,B) Development of Joels Creek fault in transtensional zone of overlap between Gap Creek and Cove Creek faults.
C) Stereoplot of slickenlines on fracture surface traces from Joels Creek area used to determine N45"W-S45E extension
direction.
552
J. M. Garihan and W. A. Ranson
Circumstantial lines of evidence led Garihan
and Ranson (1987) to postulate that the major fault
zones comprise the basement substructure of a
former Mesozoic sedimentary basin: (l)the position
of the Marietta-Tryon graben directly along the
trend of the Dan River and Davie County basins
(190 km northeastward); (2) the similarity in aerial
dimensions (Danville and Dan River basin, 878
km2); (3) the presence of local cross faults offsetting
the major graben fault, as well as minor sympa-
thetic and antithetic faults within the graben; and
(4) the relationship of faults at the margin of the
graben to offset NW-trending Triassic-Jurassic(?)
diabase dikes. The inner graben may be interpreted
as part of the extensive step faulting shown, for
example, in the residual gravity data inward from
the margin of the Newark-Gettysburg Basin
(Sumner,1977). According to our hypothesis, uplift
and erosion in Mesozoic and Cenozoic time along
this portion of the Blue Ridge Front presumably
entirely removed the sedimentary fill.
Arguments such as these we used to interpret
the Marietta-Tryon graben as the deeply eroded
substructure of a Mesozoic structural basin may be
weakened by the fact that our border faults are
steep, averaging 75. Contradictory evidence in one
example comes from a recent study of the border
fault of the Newark Basin in Pennsylvania. On the
basis of a Vibroseis profile and continuously cored
drill holes, Ratcliffe and others (1986) demonstrated
low-angle (25_35) dips for the border fault. Meso-
zoic extensional faulting in a S400-500E direction
reactivated a mylonitic Paleozoic imbricate thrust
system originally with a shallow dip. If it is valid to
generalize: from a typical ramp-thrust geometry, it
would seem unlikely that a low-angle border fault at
high structural levels would steepen at greater
depths, to coincide with the attitudes we observe for
our cataclastic zones. The structural heredity of
Mesozoic fault positioning and attitude from Paleo-
zoic thrusts is probably widespread throughout the
basins of the Virginia Piedmont (Glover and others,
1980), as well as in the Virginia Blue Ridge west of
the Triassic-Jurassic basins (Bartholomew, 1988).
However, it has not been demonstrated to this point
that brittle faults of the Marietta-Tryon graben bear
any relationship to the low-angle Paleozoic thrusts
mapped in the adjacent IX2 Greenville quad-
rangle (Nelson and others, 1985; 1987). It is more
likely that, in the study area, they are structurally
inherited from regional systematic joints oriented
N500-700E and therefore mimic their steep atti-
tudes. This being the case, it is plausible that Meso-
zoic sedimentary basin fill was present in the
Marietta-Tryon graben, with underlying and mar-
ginal steep faults (Figure 5). Lack of definitive
evidence - namely, the sedimentary rocks them-
selves - ultimately renders the hypothesis moot.
The argument above suggests that the major
brittle faults are not at positions of earlier Paleozoic
thrusts; however, it certainly accommodates - and
even predicts - the possibility that the north-
eastern extensions of thrust faults within and
bounding the Chauga-Walhalla thrust slice and the
Six-Mile thrust slice of the adjacent Greenville
1 X 2 sheet (Nelson and others, 1987) have had
recurrent Mesozoic tectonic activity along them.
SUMMARY
Some generalizations concerning the brittle
fault system can be made. A significant conclusion
is that sinstral normal-oblique movements, based on
slickenlines, were common throughout the system
- along inner and outer graben border faults, as
well as the highly faulted inner graben. Approxi-
mately 125 m of left strike-separation of a Late
Triassic-Early Jurassic(?) diabase dike across the
Cox Creek-Short Branch and the Cross Plains
(Snipes and others, 1979) faults is consistent with
this sinistral normal-oblique kinematic plan. Net
slip on faults may be of the order of a few hundred
meters. Adding our slickenline data for the Cross
Plains fault to the 125 m of separation mentioned
above allows an estimate, via graphic construction
techniques, of 180 m of net slip on the Cross Plains
fault. Total slip, of course, may be substantially
larger. A more accurate and reliable determination
of displacement awaits a detailed ground magneto-
meter survey, which will define more precisely the
offset of diabase dikes across brittle faults. More-
over, future detailed bedrock mapping may define
markers whose separations, if unlike those for the
dikes, could be used to determine whether offsets
occurred prior to intrusion of diabase.
Transtension produced the anomalous NNE-
trending Joels Creek fault by following shear(?)
joint weakness directions. The strike-slip compon-
ent of the typical sinistral normal-oblique kinematic
plan dominated along the northwestern faults of the
system (Green River, Bobs Creek, Gap Creek).
Along the southeastern part of the system, where
the more ENE- than NE-trending faults occur
Structure of the Mesozoic Marietta-Tryon graben in the Carolinas 553
(Melrose-Gosnell, Cross Plains, Pax Mountain),
dextral normal-oblique movements dominated (with
roughly equal components of strike- and dip-slip).
Somewhat equivocal evidence suggests these may
be younger shear movements.
Some additional geologic evidence for the pre-
Mesozoic antiquity of ENE-trending lineaments is
the general outline of the Caesars Head "quartz
monzonite" batholith in the southern Tigerville
quadrangle (map of Snipes and others, 1984). Al-
though obviously polydeformed during the Paleo-
zoic, two prominent extensions of the intrusion (apo-
physes?) in map view strike ENE and coincide in
trend and position precisely with the Cross Plains
and Pax Mountain faults. ENE-weakness directions
(joints or faults) may have controlled emplacement
of the Caesars Head batholith in mid-Paleozoic time
and, in turn, these directions and positions were
reactivated by brittle shearing and dilation once
again in Mesozoic time.
Crustal extension direction in the study area,
based on conjugate shears and drusy quartz exten-
sion fractures, is S35-45E in at least the later
stages of graben rifting. With a typical strike of
N600E for faults of the system (inherited from steep
regional NE-trending systematic joints), the re-
gional extension direction lies at an angle somewhat
< 90 away. As the Mesozoic graben developed and
the crust was extended in a roughly S38E direction,
longitudinal shearing developed along the normal
faults. Moreover, it probably caused the observed
sinistral oblique-sense that dominates regionally.
A tectonic framework for the recurrent history
outlined in this paper is described by DeBoer and
Snider (1979), involving: (1) Late Triassic crustal
extension and development of NE-trending rift
valleys (grabens) in the Carolinas; followed by (2) a
Jurassic shift in the position of the hot spot and the
triple junction to the vicinity of the Blake Platform.
Important for our study is their hypothesis that,
with continued spreading after the shift, new exten-
sion fractures would develop, Jurassic diabases
would be intruded along sigma one stress trajec-
tories, and longitudinal shearing on pre-existing
NE-trending fractures and faults would play a
significant role in the kinematic scheme, causing
considerable strike-slip movement.
ACKNOWLEDGMENTS
We are pleased to acknowledge financial sup-
port for our field studies: in 1985, the Furman Uni-
versity R&PG Fund, and, in the summer of 1986,
the Dana Foundation, which supported the very
capable field work of Mark Preddy and Treg
Hallman by means of two Dana Research Fellow-
ships. The Greenville Water System granted per-
mission to do limited field work in the vicinity of the
North Saluda Reservoir. Since 1982, a number of
senior research projects at Furman University have
been carried out by geology undergraduates in
northern Greenville County; we are pleased to in-
clude and acknowledge the results of their studies.
REFERENCES
ACKER, L. L., and R. D. HATCHER, Jr., 1970, Relationships
between structure and topography in northwest South
Carolina: South Carolina Geologic Survey/South Carolina
State Development Board, Geologic Notes, v. 14, no. 2, p. 35-
48.
BARTHOLOMEW, M. J., 1988, Northwest extent of Meso-
zoic extensional faulting in the Virginia Blue Ridge [ab-
stract]: Geological Society of America Abstracts with Pro-
grams, v. 20, no. 4, p. 253.
BmKHEAD, P. K., 1973, Some flinty crush rock exposures
in northwest South Carolina and adjacent areas of North
Carolina: South Carolina Geologic Survey/South Carolina
State Development Board, Geologic Notes, v. 17, no. 1, p. 18-
25.
CLARKE, J. W., 1981, Fold events in the Inner Piedmont of
South Carolina [abstract]: Geological Society of America
Abstracts with Programs, v. 13, no. 1, p. 4-5.
CONEY, J. F., and K. M. DRUMMOND, 1965, Ultramylonite
zones in the western Carolinas: Southeastern Geology, v. 6,
no. 4, p. 201-211.
DAVENPORT, S., 1987, The Geology and Mesozoic Fault
History of Portions of the Landrum, South Carolina-
North Carolina Quadrangle and the Mill Spring,
North Carolina Quadrangle: BS thesis, Furman U niver-
sity, Greenville, South Carolina, USA, 11 p.
DEBOER, J., and F. G. SNIDER, 1979, Magnetic and chemi-
cal variations of Mesozoic diabase dikes from eastern
North America: Evidence for a hot spot in the Carolinas?:
Geological Society of America Bulletin, v. 90, p. 1185-1198.
554 J. M. Garihan and W. A. Ranson
FULLAGAR, P. D., 1983, Geochronological Analyses for
North Carolina Geological Survey Mapping Projects
- Summary of Results 1983-1984: North Carolina Geo-
logical Survey Section, unpublished open-file data.
GABLE, D. J., and T. HATION, 1983, Maps of Vertical
Crustal Movements in the Conterminous United
States over the Last 10 Million Years: U.S. Geological
Survey, Miscellaneous Geological Investigations Map 1-
1315, scale 1:5,000,000-1:10,000,000.
GARIHAN, J. M., 1987, Brittle fault zones and lineaments
in Piedmont of northern Greenville County, South Caro-
lina - Substructure of an eroded Triassic-Jurassic sedi-
mentary basin? [abstract]: Geological Society of America
Abstracts with Programs, v. 19, no. 2, p. 84.
__ , and W. A. RANSON, 1986, Brittle fault development
in the Piedmont of extreme northwestern South Carolina
and adjacent North Carolina [abstract]: Geological Society
of America Abstracts with Programs, v. 18, no. 3, p. 222.
__ , W. A. RANSON, K. A. ORLANDO, and M. S. PREDDY,
1990, Kinematic history of Mesozoic faults in northwestern
South Carolina and adjacent North Carolina: South Caro-
lina Geology, v. 33, no. 1, p. 19-32.
__ , W. A. RANSON, M. PREDDY, and T. D. HALLMAN,
1988, Brittle faults, lineaments, and cataclastic rocks in
the Slater, Zirconia and part of the Saluda 7 112-minute
quadrangles, northern Greenville County, SC, and adja-
cent Henderson and Polk Counties, NC; pp. 266-350 in D.
T. Secor (ed.), Southeastern Geological Excursions:
Guidebook of the Southeastern Section of the Ge-
ological Society of America: South Carolina Geologic
Survey, Columbia, South Carolina, USA, 350 p.
__ , K. A. SARGENT, and W. A. RANSON, 1985, Geology
of Green River, Bobs Creek, Gap Creek, and Pax Mountain
lineaments, northwestern South Carolina and adjacent
North Carolina [abstract]: Geological Society of America
Abstracts with Programs, v. 17, no. 2, p. 91.
GLOVER, L., F. B. POLAND, R. D. TUCKER, and W. C.
BOURLAND, 1980, Diachronous Paleozoic mylonites and
structural heredity of Triassic-J urassic basins in Virginia
[abstract]: Geological Society of America Abstracts with
Programs, v.12, no. 4, p. 178.
GRIFFIN, V .S., Jr., 1977, Preliminary geologic map of the
South Carolina Inner Piedmont belt: South Carolina Geo-
logic Survey, South Carolina State Development Board,
Geologic Notes, v. 21, no. 4, p. 198-204.
__ , 1978, Geology of the Calhoun Creek, Calhoun
Falls, and Portions of the Chennault and Verdery
Quadrangles, South Carolina: South Carolina Geologic
Survey/South Carolina State Development Board MS-23,
46p.
__ , 1979, Geology of the Abbeville East and Abbe-
ville West, Larimer, and Lowndesville Quadrangles,
South Carolina: South Carolina Geologic Survey MS-24,
58p.
HACK, J. T., 1982, Physiographic Divisions and Dif-
ferential Uplift in the Piedmont and Blue Ridge: U.S.
Geological Survey, Professional Paper 1265, 49 p.
HADLEY, J. B., and A. E. NELSON, 1971, Geologic Map of
the Knoxville Quadrangle, North Carolina, Tennes-
see, and South Carolina: U.S. Geological Survey, Mis-
cellaneous Geological Investigations Map 1-654, 1:250,000-
scale.
HALLMAN, T. D., M. S. PREDDY, and W. A. RANSON, 1987,
Brittle faulting and the production of cataclastic rocks
along lineaments in Piedmont of northwestern South
Carolina and adjacent North Carolina [abstract]: Geologi-
cal Society of America Abstracts with Programs, v. 19. no.
2,p.88.
HASELTON, G. M., 1974, Some reconnaissance geomor-
phological observations in northwestern South Carolina
and adjacent North Carolina: South Carolina Geologic
Survey/South Carolina State Development Board, Geologic
Notes, v. 18, no. 4, p. 60-74.
HATCHER, R. D., Jr., 1974, An Introduction to the Blue
Ridge Tectonic History of Northeast Georgia: Georgia
Geological Society, Field Trip Guidebook 13-A, 6 p.
HIGGENS, M. W., 1971, Cataclastic Rocks: U.S. Geo-
logical Survey, Professional Paper 687, 97 p.
KISH, S. A., 1988, Post-Alleghanian uplift adjacent to
major ductile fault.zones in the southern Appalachian
Piedmont [abstract]: Geological Society of America Ab-
stracts with Programs, v. 20, no. 4, p. 274.
MINER, R. S., 1985, Hydrothermal Zeolite Assemblages
Associated with the Gap Creek Joint System: BS the-
sis, Furman University, Greenville, South Carolina, USA,
39p.
NELSON, A. E., 1985, Major Tectonic Features and
Structural Elements in the Northwest Part of the
Greenville Quadrangle, Georgia: U.S. Geological Sur-
vey, Bulletin 1643, 22 p.
__ ,1987, Generalized Tectonic Map of the Green-
ville ]Ox 2 Quadrangle, Georgia, South Carolina, and
North Carolina: U.S. Geological Survey, Map MF-1898,
scale 1:250,000.
__ , J. W. HORTON, Jr., and J. W. CLARKE, 1985,
Stacked thrust sheets in the Blue Ridge and Piedmont
Provinces northeast Georgia and western South Carolina
[abstract]: Geological Society of America Abstracts with
Programs, v. 17, no. 7, p. 674.
Structure of the Mesozoic Marietta-Tryon graben in the Carolinas 555
PENLEY, H. M., and S. M. PICKERING, 1973, Topo-
graphic Lineaments Which Transect the Northeast
Georgia Blue Ridge as Seen from Landsat Imagery:
Georgia Geological Survey Open File Report, 7 p.
PREDDY, M. S., T. D. HALLMAN, and J. M. GARIHAN,
1987, Regional joint patterns and their relationship to
drainage and brittle fault structure in northern Greenville
County, South Carolina and adjacent North Carolina
[abstract): Geological Society of America Abstracts with
Programs, v. 19, no. 2, p. 125.
RANSON, W. A., and J. M. GARIHAN, 1988, Mesoscopic
structural evidence for longitudinal shearing, oblique crus-
tal extension, and fault block rotation during development
of Mesozoic Marietta-Tryon graben, Piedmont of northwest
South Carolina [abstract): Geological Society of America
Abstracts with Programs, v. 20, no. 4, p. 310.
RATCLIFFE, N. M., W. C. BURTON, R. M. D'ANGELO, and
J. K. COSTAIN, 1986, Low-angle extensional faulting, re-
activated mylonites, and seismic reflection geometry of the
Newark basin margin in eastern Pennsylvania: Geology,
v.14, p. 766-770.
SECOR, D. T., Jr., A. W. SNOKE, K. W. BRAMLETT, O. P.
COSTELLO, and O. P. KIMBRELL, 1986, Character of the
Alleghanian Orogeny in the Southern Appalachians, Part
I: Alleghanian deformation in the eastern Piedmont of
South Carolina: Geological Society of America Bulletin, v.
97, no. 11, p. 1319-1328.
SNIPES, D. S., 1981, Groundwater Quality and Quantity
in Fracture Zones in the Piedmont of Northwestern
South Carolina: Clemson University, Water Resources
Research Institute Report 93, 87 p.
__ , L. L. BURNETT, J. A. WYLIE, L. A. SACKS, S. B.
HEATON, G. A. DALTON, B. A. ISRAEL, and G. G.
PADGETT, 1984, Indicators of Groundwater Quality
and Yield for a Public Water Supply in Rock Fracture
Zones of the Piedmont: Clemson University, Water
Resources Research Institute Report 115, 80 p.
__ , M. W. DAVIS, and P. R. MANOOGIAN, 1979, Cross
Plains fault in Piedmont of northwest South Carolina
[abstract): Geological Society of America Abstracts with
Programs, v. 11, no. 4, p. 213.
STAFFORD, D. B., J. T. LIGON, and D. S. SNIPES, Fracture
Trace Mapping and Well Water Yield in the Piedmont
Region of South Carolina: Clemson University, Water
Resources Research Institute Report 112, 66 p.
SUMNER, J. R., 1977, Geophysical investigation of the
structural framework of the Newark-Gettysburg Triassic
basin, Pennsylvania: Geological Society of America
Bulletin, v. 88, p. 935-942.
WARNER, R. D., D. S. SNIPES, S. S. HUGHES, J. C.
STEINER, M. W. DAVIS, P. R. MANOOGIAN, and R. A.
SCHMITT, 1985, Olivine normative dolerite dikes from
western South Carolina: Mineralogy, chemical composi-
tion, and petrogenesis: Contributions to Mineralogy and
Petrology, v. 90, no. 4, p. 386-400.
CHAPTER 5
PROBLEMS AT MARGINS IN
EUROPE AND ASIA
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Strain pattern in the Purbeck-Isle of Wight Monocline: A case
study of folding due to dip-slip fault in the basement
MOHAMMED S. AMEEN
Structural Geology Section, Department of Geology, Royal School of Mines,
Imperial College, Prince Consort Road, London, SW7 2BP, England, UK
Current address: GeoScience Ltd, Silwood Park, Buckhurst Road, Ascot, Berks, SL5 7QW, England, UK
(received August 12, 1988; revision accepted January 2, 1989)
ABSTRACT
The Purbeck-Isle of Wight Monocline, southern England, is a forced fold developed in the Upper
Cretaceous and Tertiary rocks as a result of an inversion of motion on a pre-existing normal fault in the
basement. Mesofractures are used to determine the strain pattern associated with folding in the cover rocks.
The results show that on a mesoscopic scale the strain pattern was controlled by layering. Directions of
maximum extension and contraction were predominantly symmetrical relative to bedding. Furthermore,
layering helped in releasing layer-parallel shear in the synclinal part of the monocline by layer-parallel slip
(bedding faults). On a macroscopic scale, the strain pattern in the monocline varies with both time and space.
The results show similarities to strain patterns obtained in experimental studies.
INTRODUCTION
The Purbeck-Isle of Wight Monocline is a major
structure exposed in southern England (Figure 1).
It has a regional E-W trend and extends from the
eastern end of the Isle of Wight to the Dorset coast.
The monocline faces northward with a steep or
overturned northern limb. It flattens out rapidly to
both the north and the south (Figure 2). The
amplitude of the flexure varies from place to place
with estimated maximum values of 1200 m on the
Dorset coast and 1500 m on the Isle of Wight (Selley
and Stoneley, 1983).
The exposed rocks in the flexure are of Jurassic,
Cretaceous and Tertiary age, although older (i.e.,
Devonian, Carboniferous, Permian, and Triassic)
rocks were encountered in drill cores obtained
during hydrocarbon exploration in the region
(Hinde, 1980; Colter and Howard, 1981). These
older rocks are exposed in Devon (Chatwin, 1960;
Webby, 1965a,b; Laming, 1966; Edmonds and
others, 1975; Selley and Stoneley, 1983). The
stratigraphy of the area was described in detail by
earlier writers (Bristow, 1889; White, 1921; Arkell,
1947; Edmonds and others, 1975; Hinde, 1980;
Colter and Howard, 1981; Melville and Freshney,
1982), and a brief description of the rock sequence
from which the data for the present study were
collected is given in Table 1.
The tectonic history of the monocline has long
been a subject of controversy - that is, whether the
monocline has developed as a buckle fold or forced
fold (Bristow, 1889; White, 1921; Arkell, 1936a,b,
1947; Lees and Cox, 1937; Phillips, 1964; Stoneley,
1982; Bevan, 1985a,b,c). Recently, seismic data
(Simpson and others, 1989) showed that the mono-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp 559-578. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
559
560
M.S. Ameen
Figure 1. Location map (A) and a
general geologic map (B) of the
Purbeck-Isle of Wight Monocline,
southern England.
.+.
-- 1--
<Jllhc:hn.
,.,nch".
N
...
I
;
--'--'-
I."", O--1
0Km
0
'."'iU'(

Chalk
A
0
Upp., Gr nung and GAult
OJ
L(lwe' G, nsand
0
W ,den S.,l
.. .
Purlbec:-': eed, - OafOrd CI .W'
N

COrn!)' ..... . Inf.rlor Oolile
...
50"50'
N
Figure 2. Profile section of
the Purbeck-Isle of Wight
Monocline along A-A' in
Figure 1 (after Stoneley, 1982,
Fig. 3D). The part of the se-
quence discussed in the pre-
sent study is ornameted with
circles; the dotted area is
Paleozoic-Precambrian base-
ment. Perm, Permian; TR,
Triassic; L, Liassic; BS, Brid-
port Sands; MJ, Middle Juras-
sic; 0, Oxford Clay; C, Coral-
lian; Kim, Kimmeridge Clay;
P, Portland and Purbeck; W,
Wealden; UK, Upper Creta-
ceous; T, Tertiary. Arrows 1
and 2 indicate earlier and
later sense of movement on
the basement fault; vertical
equals horizontal scale.
A
1000m
-- ..........
cline in the Upper Cretaceous and Tertiary rocks
developed as a result of reverse movement on a pre-
existing Variscan fault in Palaeozoic and older rocks
(Figure 2). Three phases are recognized: (1) Late
Carboniferous Variscan N-S compression, which
caused thrusting to the north (2) Permian to mid-
Cretaceous (Aptian) extension, during which
Variscan thrusts were reactivated as normal faults
that produced half-graben basins; and (3) post-
"
A-
=t-N
? Perm
Aptian N-S compression, which inverted the basin
and formed the Purbeck-Isle of Wight Monocline.
Early studies of mesostructures in the mono-
cline (Bristow, 1889; Strahan, 1895, 1898; White,
1921) were general, brief, and descriptive. Later,
more detailed studies (Arkell, 1936a,b, 1947;
Phillips 1964; Bevan, 1985a,b,c) dealt mostly with
the Dorset coast part of the monocline. In the
present study, mesostructures were studied in detail
Strain pattern in the Purbeck-Isle of Wight Monocline 561
Table 1. Summarized stratigraphy of the studied rocks on the Isle of Wight (lOW) and the Dorset coast (0).
Thickness (m)
Period Epoch Formation Lithology lOW D
Quaternary- Plio-Pleistocene clays, gravels
Neogene and sands
Hampstead Beds clays
Bembridge Marls laminated muds
E. Oligocene and silts, thin
limestone
Bembridge shell limestone
Limestones and thin clays
Osborne Beds clays, marls and
limestone-lentils
Palaeogene Headon Beds Fm clays, sands and 575 215
limestones
I I
(poorly lithified Eocene
I I
rocks except Barton Beds Fm sands and clays 724 182
limestone) Brackleshaml sands, clays and
BagshotFm pebble bands
London Clay Fm sandy and silty clays
L. Palaeocene Oldhaven, clays, sands
Blackheath, and gravel beds
Woolwich,
Reading and
ThanetBeds
Upper Chalk mostly smooth 311 152
Senonian (including massively bedded
I I
Chalk Rock) chalk with flints 404 305
Middle Chalk homogenous 30 27
Turonian (including and nodular chalk
I I
Cretaceous
Melbourn Rock) 60 34
(lithified rocks) Cenomanian Lower Chalk marly and 'flaser' 25 15
I I
chalk 60 46
Methods and Techniques in an attempt to determine the strain pattern in the
monocline. The results are integrated with other
field and experimental studies of forced folding to
outline possible criteria for prediction of the type of
dip-slip basement faults from strain patterns in the
cover rocks.
Mesoscopic kinematic indicators (e.g., faults,
joints, veins, stylolites) were used to determine
strain distribution within Upper Cretaceous to
Tertiary sedimentary rocks of the Purbeck-Isle of
562 M.S.Ameen
Wight Monocline. Measurements were taken in
exposed parts of the monocline, mainly along the
Dorset and Isle of Wight coasts and inland - in
quarries and farm pits. Although the variations in
bedding attitude and rock type were considered
when selecting sites for data collection, the choice of
localities is limited by the distribution and accessi-
bility of exposures. Measurements mostly were
carried out on approximately equal areas of exposed
rocks (100 m
2
) to avoid biased data collection. The
number of measurements in individual localities
varies according to the degree of exposure and level
of deformation. A total of 2,607 measurements were
taken (including 2,077 planar structures and 530
linear structures) from 30 localities (Figure 3).
The collected data at each locality were plotted
on stereographic projections in order to determine
the symmetry of the structures relative to each
other and their relationship to bedding, to the
horizontal, and to the major structure. In addition,
the stereographic projections were used to deter-
mine the axes of maximum contraction and exten-
sion related to the structures. Where more than one
system of structures are present their relative
chronologies are established and hence the varia-
tion in strain pattern with time was determined.
The results for different localities were then inte-
grated in order to recognize subdomains or zones in
the monocline which have a similar strain pattern
at a particular stage of the folding.
The relative ages of the mesostructures were
determined by their abutting, crosscutting, and
offsetting relationships. Such relationships and
their implications have been thoroughly discussed
in the literature (Harris and others, 1960; Price
1966; Kulander and others, 1979). The kinematic
implications of the mesostructures were determined
using standard criteria based on experimental rock
deformation and theoretical analysis and were
reviewed by Hancock (1985).
A recently developed computerized technique
was used for determination of stress and strain axes
from displacement vectors (lineations) on faults in
which the displacement may occur as a result of a
stress field unrelated to that which initiated the
fracture (Angelier, 1984; Aleksandrowski, 1985). It
is reliable for areas with a large population (2: 100)
of lineation and fracture attitudes in individual
localities. In the present study, the number of slip
vector measurements in individual localities is con-
siderably less than 100 and therefore this technique
was not used.
Kinematic/Geometric Classification of
Mesostructures
The marjority of mesostructures observed in
this study are classified with respect to the bedding
rather than to the horizontal. Such a classification,
which is a modification of Norris's (1958) subdivi-
sion of faults into contractional and extensional
faults is justified in this area because most
mesostructures and their kinematic implication
(direction of maximum contraction and extension)
are systematically oriented with respect to the
bedding rather than to the horizontal.
The kinematic implications of individual planar
structures (stylolite, extension vein) or systems of
conjugate sets of planar structures (faults or joints)
show that the directions of maximum extension and
contraction are approximately parallel to one of
three mutually perpendicular axes: parallel to bed-
ding dip 'd'; parallel to bedding strike's'; and normal
to bedding 'n' - as shown in Figure 4A. Two nota-
tions, 'c' and 'E,' are used for the kinematic descrip-
tion of a planar structure or a system of conjugate
planar structures where C stands for contraction
and E for extension, respectively. Subscripts d, s,
and n are used to indicate the directions of extension
and contraction axes relative to bedding. According
to this classification, faults and shear joints that
occur as conjugate sets in the area fall into one of
five geometric/kinematic groups, irrespective of
their attitude relative to the horizontal: CdEn, CdE
s
,
CnEd, CnEs" and CsEd (Figures 4B-F). As the pre-
sent study is concerned with strain patterns rather
than stresses, the variation of the dihedral angle
between conjugate faults or joints sets is not con-
sidered in this classification. However, if the direc-
tion of contraction associated with a conjugate sys-
tem offaults or joints is not along the acute bisector
of the dihedral angle, it is indicated.
Individual planar extensional structures (e.g.,
extension joints and veins) fall into one of three
groups: EdC
sn
, EsC
dn
, and EnC
ds
(Figures 4G-I).
Individual contractional planar structures, (stylo-
lites) also fall into one of three groups: CdE
sn
'
CsEdn, and CnEdB (Figure 4J-I).
The traditional description of faults according
to their attitude and sense of slip relative to the
horizontal - normal, reverse, and strike slip - is
applied to two sets of mesofaults which have a
strike-slip sense of movement and which are
symmetrical about the horizontal rather than the
bedding.
Strain pattern in the Purbeck-Isle of Wight Monocline
Figure 3. Distribution of the
study localities in the Purbeck-
Isle of Wight Monocline on the
Dorset coast (A) and the Isle of
Wight (B).
UJ, Upper Jurassic,
C, Lower Cretaceous,
UCC, Upper Cretaceous Chalk;
PAL, Palaeocene;
EOC, Eocene;
T, Tertiary;
UC, Upper Chalk; MC, Middle
Chalk; LC, Lower Chalk.
\
A
B
\
@
...... .. @
lJI T
D uC
D MC
LC
Iilll C&U J
563
Figure 4. Schematic block diagrams illustrating the geometric/kinematic classification of mesofractures which are sym-
metrically oriented around the bedding. A) The three principal coordinates (n, d, s) used as reference to directions of extension
and contraction related to each system of fractures. B-F) Conjugate shear fracture systems CdEn' CdE
s
' CnEd' CnEs, and CsEd'
respectively. G-!) Extensional fractures EdCsn' EsCdn, and EnCds' respectively. J-L) Stylolitic seams CdEsn' CsEdn' and CnE
ds
'
respectively.
564 M.S.Ameen
MESOSTRUCTURES AND RELATED STRAIN
PATTERNS
General
Eleven sets of shear fractures (faults and shear
joints) were observed (Figures 5, 6, 7). The degree
of their development and their chronological se-
quence (Table 2) vary in different parts of the mono-
cline: steep limb part (Figures 5 and 6), semi-
horizontal upthrown and downthrown parts (Figure
7). These faults include bedding-parallel faults, four
conjugate systems (eight planes) symmetrically
oriented about the bedding (CdEn, CdEn, CnEd, CsEd)
and one conjugate system (two planes) symmetri-
cally oriented about the horizontal (strike-slip
faults). In vertical or nearly vertical beds, the latter
faults are classified as CnEs, whereas in horizontal
or nearly horizontal beds, like the gently dipping
downthrown limb, they can be classified as CdEs'
Extension joints or veins include three mutual-
ly perpendicular sets: EdC
sn
' EsC
dn
, EnC
ds
(bedding
joints). Stylolitic seams occur in the Upper Chalk
parallel to these joint sets and are CdE
sn
, CsE
dn
, and
Figure 5. Lower hemisphere
equal-area projection of mesofrac-
tures in the Upper Chalk of the
steep limb of the Purbeck-Isle of
Wight Monocline; see Figure 3 for
localities.
A) Localities 1, 17, 21, 24, and 30.
B) Locality 14.
C) Localities 2, 5, 6, and 13.
D) Locality 20.
Notice the symmetry of most struc-
tures around the bedding. Slip
vectors on shear fractures are
shown by solid symbols; strike-slip
faults (SS), which can be con-
sidered CnE. shear fractures in
vertical beds, are dextral in the
NE-SW quarters and sinistral in
the NW-SE quarters.
N
CnEds (bedding stylolite). Stylolitic seams also
occur on some of the ten faults or shear joints and
help in establishing chronology of different fault
sets.
In the Upper Cretaceous Chalk, faults may
show flint filling or gouge zones (Figure 8A,B) and
occur in well defined sets. No such flint filling or
gouge zone occur in the Middle and Lower Chalk.
Striations on faults are either fibrous crystal
growth, frictional wear, or solution hollows - i.e.,
slickolite columns (Figure 8C,D).
Faults may show evidence of reactivation either
by faulting (the occurrence of two sets of striations)
or by pressure solution (stylolitization) (Figure
9A,B). Shear zones, defined by an array of en
echelon extension fractures (some of which show
evidence of later stylolitization) are present in the
Upper Chalk, parallel to some faults (Figure 9C).
Joints also developed parallel to faults (Figure 9D).
Joints parallel to faults are assumed to be shear
fractures, and this assumption is supported by the
occurrence of striations on some joint surfaces
although no displacement was detected. In addition,
extension veins and joints are developed in chalk,
DO
o

.
o
"

B 0
N
o ..
00
0
0
00
0

D
LEGEND
B
CnEd
CdEs
CdEn
CsEd
CnEs
(55)
o
*
o
EdCsn *
EsCdn *
B
Strain pattern in the Purbeck-Isle of Wight Monocline 565
N N
N
N
t.

. .1:
A
N

o 00
Figure 6. Lower hemisphere equal-area projec-
00 tion of meso fractures; see Figure 5 for symbols.
B
N
LEGEND
Bedding 0
CnEd
CnES
* "
CdEn
.. t.
CdEs
} 0
(SS)
EdC
sn
EsC
dn
} .
CsEd
A) Middle Chalk, Localities 3 and 4.
B) Middle Chalk, Localities 7, 10, 12, 19, 25, 29.
C) Lower Chalk, Localities 8, 9, 11, 26, 28.
D) Tertiary rocks, Localities 15 and 23.
Notice the consistent symmetry of most struc-
tures around the bedding.
Figure 7. Lower hemisphere equal-area projec-
tion of meso fractures; see Figure 3 for localities .
A) The gently dipping downthrown part of the
Purbeck-Isle of Wight Monocline (Tertiary
rocks in Localities 16,18,22).
B) The gently dipping upthrown part of the
Purbeck-Isle of Wight Monocline (Lower Chalk,
Locality 27). Notice the symmetrical attitude of
most fractures around the bedding (cf. Figure
6C and 6D, respectively). Slip vectors on shear
fractures are shown by solid symbols; strike-slip
faults (SS), which can be considered CdEs
fractures in horizontal beds, are dextral in the
NE-SW quarters and sinistral in the NW -SE
quarters.
566
M.S.Ameen
Table 2. Chronological sequence of the mesofractures observed in the different tectonic and
lithostratigraphic parts, in the Purbeck-Isle of Wight Monocline according to which the
monocline is divided into six distinctive zones shown in Figure 15.
4
1 2 3 A B C D 5
SL
UC 1,2,5
CdEn
CnEd CdEn
D,S
6,13,14,
CdE.
CnEds CdEsn Z1
17,21,24
CdEsn
EdCsn EnCds
30 EsCdn
SL
UC
20 CsEd CnEd
CsEdn CnEds Z2
EdCsn EdCsn
SL
MC 3,4,
CdEs
CnEd CdEn
7,10,12, EsCdn
EdCsn EnCds Z3
19,25,29
SL
LC 8,9,
11,26,
Z3
28
SL
T 15,23
CdEn
CnEd EsCdn Z4
CsEd
DT
T
16,18
CdEn
CnEd
22
CdEs
CnEs Z5
CdEsn
EdCsn
EnCds
EsCdn
UT
LC
27
CnEd
CnEs
CsEd
EsCdn Z6
EdCsn
Column 1 - tectonic setting: SL, steep limb, DT, downthrown gently dipping flank, UT, upthrown gently
dipping flank. Column 2 - lithostratigraphic units: T, Tertiary rocks; UC, Upper Chalk; MC, Middle
Chalk; LC, Lower Chalk. Column 3 - location of the analysed data (from Figure 3). Column 4 - classi-
fication of the mesofractures according to their chronological sequence: A to D from oldest to youngest
(D, dextral fault; S, sinistral fault). Column 5 - zone numbers given to the parts that show the same
mesostructure patterns and chronological sequence.
and some show fibrous growth perpendicular to
their walls (Figure 10).
In poorly lithified Tertiary sands, faults occur
as swarms of bifurcating granulation seams with
displacements up to a few centimeters (Figure
lIA,B). Some ductile shear zones are developed in
the shale horizons (Figure lIe), although joints are
very rare (Figure lID) and no stylolitic seams were
observed. In the few Tertiary limestone horizons,
joints are better developed and faults are discrete
(Figure 12).
Mesostructures in Different Zones of the
Monocline
From the relative chronology and degree of
development of different mesostructures in the ex-
posed parts of the monocline (Table 2) it is possible
to recognize at least six different zones, each of
which experienced a different strain history (Figure
13). A stereographic plot of the directions of
maximum extension and contraction in each zone is
given in Figure 14. The boundaries between the
different zones are probably transitional; however,
Strain pattern in the Purbeck-Isle of Wight Monocline 567
c.
Figure 8. General features of mesofractures in the
Upper Chalk in the steep limb of the Purbeck-Isle of
Wight Monocline; see Figure 3B for localities.
A) Conjugate CdEn faults (1 and 2) are filled with later
diagenetic flint and are cut and displaced by younger
CnEd faults (3) and bedding faults (B) at Culver Cliff, Isle
of Wight (Locality 24); notice the symmetry of the
fractures around the bedding. B) CdEn fault with a
gouge zone (1) is cut and displaced by younger CnEd
faults at Bat's Head, Dorset coast (Locality 1); notice the
black flint fragments in the gouge. C) Fibrous calcite
striations on a bedding fault at St. Oswald's Bay, Dorset
coast, (Locality 6); the arrow indicates the sense of
movement of the eroded part. D) Frictional wear
striations and solution hollows on a CnEd fault at Man-
O-War Cove (Locality 5).
2
c
m

N

f
f

.
.

B

.
"
'
;
.
.
.
.
.
.
.
.
.
.

"
"
'
"
'
"

/

.

.
.
.
.
.
.

1

=
=
:
>
N

.
D

F
i
g
u
r
e

9
.

G
e
n
e
r
a
l

f
e
a
t
u
r
e
s

o
f

m
e
s
o
f
r
a
c
t
u
r
e
s

i
n

t
h
e

U
p
p
e
r

C
h
a
l
k

i
n

t
h
e

s
t
e
e
p

l
i
m
b

o
f

t
h
e

P
u
r
b
e
c
k
-
I
s
l
e

o
f

W
i
g
h
t

M
o
n
o
c
l
i
n
e
.

A
)

T
w
o

s
e
t
s

o
f

f
r
i
c
t
i
o
n
a
l

w
e
a
r

s
t
r
i
a
t
i
o
n
s

o
n

a

s
t
r
i
k
e
-
s
l
i
p

f
a
u
l
t

a
t

A
l
u
m

B
a
y
,

I
s
l
e

o
f

W
i
g
h
t

(
L
o
c
a
l
i
t
y

1
4
,

F
i
g
u
r
e

3
B
)
;

s
e
t

2

i
s

o
l
d
e
r

a
n
d

r
e
l
a
t
e
d

t
o

t
h
e

m
a
i
n

s
t
r
i
k
e
-
s
l
i
p

m
o
v
e
m
e
n
t
,

s
e
t

1

i
s

y
o
u
n
g
e
r

a
n
d

r
e
l
a
t
e
d

t
o

l
a
t
e
r

e
x
t
e
n
s
i
o
n
a
l

m
o
v
e
m
e
n
t

o
n

t
h
e

f
a
u
l
t
.

B
)

C
d
E
.

f
a
u
l
t

(
1
)

s
t
y
l
o
l
i
t
i
z
e
d

b
y

l
a
t
e
r

m
o
v
e
m
e
n
t

t
h
a
t

c
a
u
s
e
d

c
o
n
t
r
a
c
t
i
o
n

p
e
r
p
e
n
d
i
c
u
l
a
r

t
o

t
h
e

b
e
d
d
i
n
g

(
B
)
,

a
s

e
v
i
d
e
n
t

f
r
o
m

t
h
e

s
t
y
l
o
l
i
t
i
c

c
o
l
u
m
n
s
;

C
u
l
v
e
r

C
l
i
f
f
,

I
s
l
e

o
f
W
i
g
h
t

(
L
o
c
a
l
i
t
y

2
4
,

F
i
g
u
r
e

3
b
)
.

C
)

C
o
n
j
u
g
a
t
e

C
n
E
d

s
h
e
a
r

z
o
n
e
s

d
e
f
i
n
e
d

b
y

f
a
u
l
t

(
1
)

a
n
d

a
n

a
r
r
a
y

o
f

e
n

e
c
h
e
l
o
n

e
x
t
e
n
s
i
o
n

f
r
a
c
t
u
r
e
s

(
2
)
;

b
o
t
h

z
o
n
e
s

h
a
v
e

b
e
e
n

r
e
j
u
v
e
n
a
t
e
d

b
y

s
t
y
l
o
l
i
t
i
z
a
t
i
o
n

d
u
e

t
o

c
o
m
p
r
e
s
s
i
o
n

a
t

l
o
w

a
n
g
l
e

t
o

t
h
e

b
e
d
d
i
n
g

(
B
)
;

B
a
t
'
s

H
e
a
d
,

D
o
r
s
e
t

c
o
a
s
t

(
L
o
c
a
l
i
t
y

1
,

F
i
g
u
r
e

3
A
.
)
.

D
)

H
i
g
h
l
y

f
r
a
c
t
u
r
e
d

v
e
r
t
i
c
a
l

U
p
p
e
r

C
h
a
l
k

b
e
d
s

a
t

A
l
u
j
m

B
a
y
,

I
s
l
e

o
f

W
i
g
h
t

(
L
o
c
a
l
i
t
y

1
4
,

F
i
g
u
r
e

3
B
)
,

s
h
o
w
i
n
g

o
l
d
e
r

C
d
E
n

f
a
u
l
t
s

(
3
)

a
n
d

y
o
u
n
g
e
r

C
n
E
d

f
a
u
l
t
s

(
1

a
n
d

2
)

a
n
d

j
o
i
n
t
s

p
a
r
a
l
l
e
l

t
o

t
h
e

f
a
u
l
t
s
.

0
1

0
)

<
X
l


r
n

>

s


B,
Strain pattern in the Purbeck-Isle of Wight Monocline 569
Figure 10. Extensional veins
showing fibrous calcite at a high
angle to the vein wall.
A) EnCw, vein (1) is cut and dis-
placed by younger CnEd fault in
the Upper Chalk at Freshwater
Bay, Isle of Wight (Locality 30,
Figure 3B).
B) EdC", vein (1) in the Lower
Chalk at Mupe Bay, Dorset coast
<Locality 11, Figure 3A).
the lack of accessible continuous exposure makes it
difficult to outline accurately the position of these
boundaries. This can only be inferred by comparing
relative chronological sequences of structures in
different zones and suggesting that defonnation
started simultaneously in different zones. A des-
cription of structures and strain patterns in the
different zones is given below.
Zone 1. This zone includes the steepest exposed
parts of the Upper Chalk, which lie in the internal
synclinal area of the monocline (Figure 13). The
attitude of beds varies between 65N and 60
0
S. The
modes of failure are dominately fracturing (faults
and joints) and pressure solution. The patterns of
mesostructures and related extensional strain in
this zone are shown by stereographic projections
(Figure 5A,B,C and Figure 14A), and their
chronological development is illustrated in Table 2
and Figure 15.
The oldest fracture systems are the CdEn and
CdEs faults and joints, which show flint filling
(Figure SA). Some of these faults developed gouge
zones a few millimeters to several centimeters in
thickness (Figure SB) . The presence of flint filling
indicates a relatively early diagenetic origin of the
fractures. In faults of the CdEn system where
gouges developed. flint fragments indicate that
faults have a relatively complex displacement
history. They developed as either faults or joints
which were then opened and filled with diagenetic
570
M. S. Ameen
Figure 11. General features of mesofrac-
t ures in the poorly lithified Tertiary sand-
stones. A) Conjugate CdEn faults (1 and 2)
developed as granulation seams in a ver-
tical bed (B) at White Cliff Bay, Isle of
Wight (Locality 23, Figure 3B). B) Con-
jugate CnEd faults (1 and 2) developed as
granulation seams in a vertical sandstone
bed (B) at the same location as (A). C) Duc-
tile shear zone of the CnEd system in a
vertical shale horizon at Alum Bay, Isle of
Wight (Locality 15, Figure 15B); compare
with fault (1) in (B). D) Bottom view of a
vertical sandstone bed showing conjugate
C.Ed faults (1 and 2) and parallel joints at
the same location as (A).
flint, then a rejuvenation of shear movement on
these filled faults or joints resulted in crushing of
the flint filling and development of gouge zones.
The CdEn and CdEs faults and joints were accom-
panied by extension veins (Figure lOA), and joints
EnCds and EsCdn and stylolitization seams (CdE
sn
)
which cut and displace diagenetic bedding stylolite.
This system of mesostructures indicates a
strain pattern dominated by approximately bed-
ding-parallel contraction (along the dip direction)
accompanied by extension approximately perpen-
dicular to bedding for the CdEn system, and parallel
to bedding strike for the CdEs system (Figure 15A).
Younger systems of faults and related struc-
tures cut and displace the older system and consist
of CnEd fault and joint systems (Figure SA),
stylolitic seams CnEds, - i.e., bedding stylolites and
extension joints and veins EdCsn. Some CnEd frac-
tures show flint filling (indicating an early dia-
genetic origin), fault gouges (Figure 8A), and
frictional wear striations and fibrous growth fillings
(Figure 8C,D). Furthermore, some CnEd zones
developed as shear zones defined by en echelon
arrays of extension cracks (Figure 9C). All these
features of the younger fractures indicate continua-
tion of the same strain system over a considerable
period of time and range of deformation conditions.
The CnEd fractures and related structures indicate a
strain pattern dominated by contraction at a high
angle to bedding and extension in the dip direction
(Figure 15A). The older CdEn and CdEs systems
were stylolitized to various extents by later defor-
mation (Figure 9B).
A second generation of CdEn fractures and
related structures developed in the innermost
synclinal part of the zone where they cut and dis-
placed structures related to earlier deformations.
The early CnEd fractures were rejuvenated by this
Strain pattern in the Purbeck-Isle of Wight Monocline 571
Figure 12. Clear-cut mesofractures in Tertiary limestone
horizons in the gently dipping downthrown part of the
Purbeck-Isle of Wight Monocline. A) Conjugate CdEn
faults (1 and 2) at White Cliff Bay, Isle of Wight (Locality
22, Figure 3B). B) Top view showing conjugate CdEsjoints
at the same location as (A). C) EsC
dn
extensional joints (1)
at the same location as (A).
later movement - one set being stylolitized and the
other experiencing an inversion of movement and/or
stylolitization as shown in Figure 9C. The strain
pattern associated with this movement is similar to
that associated with the first generation of edEn
fractures (Figure 15C).
The youngest fracture system includes conju-
gate strike-slip faults, symmetrically oriented
around the fold axis in a manner that indicates
Figure 12 (continued). D) CnEd fault (1) cut and
displaced EdCnsjoint (2) at the same locality as (A). E) Top
view of a bedding fault surface in a limestone at the same
locality as (A). F) Bedding fault in a quarry on the Isle of
Wight (Locality 18, Figure 3B).
contraction approximately perpendicular to the fold
axis and extension approximately parallel to it
(Figure 15D). This system of fractures cuts and
displaces the older systems, and faults of this young-
est system are generally larger in their apparent
length and displacement. They show frictional wear
striations. Some strike slip faults were later re-
juvenated with extension parallel to the fold axis
and vertical or nearly vertical contraction (Figure
9A).
572 M.S.Ameen
Figure 13. Schematic block diagram of the Purbeck-Isle of Wight Monocline showing the locations of the different zones (1 to
6) according to mesostructural and extensional strain patterns. The thick solid lines outline the estimated boundaries
between the different zones.
Bedding faults also developed in this zone, in
addition to extensional, contractional, and strike-
slip faults. They are particularly well developed on
the Dorset coast (Figure BC). Their displacements
were directly measured from the offset of structures
- e.g., joints or faults which cross the bedding, or
indirectly estimated from the length of tool pit
striations (tool pit length :5 displacement) on the
fault surface. Cumulative (total) displacement on
successive bedding planes in a 10-m-thick unit of
rocks vary from 2 m to 4 m on the Dorset coast and
0.1 m to 2 m on the Isle of Wight. Bedding faults
occurred penecontemporaneously with CnEd frac-
tures. The two sets of faults cut across each other:
bedding faults displace CnEd faults in one place in
the zone, whereas CnEd faults displace bedding
faults at another location in the zone. Apart from
these CnEd fractures, the relative age of bedding
faults and the other fractures cannot be distin-
guished. Bedding faults likely continued to develop
right up until the latest stages of folding.
Zone 2. This covers the exposed Upper Chalk
in the higher positions of the steep limb of the
monocline, closer to the anticlinal hinge zone (Fig-
ure 13). The mesofractures and strain patterns in
this zone (Figures 5D and 14A) differ from Zone I
by the lack of significant bedding faulting and by
the occurrence of conjugate sets of joints and faults
of class CsEd (FigureI6A) associated with stylolites
CsEdn (Figure 16B), both of which indicate layer-
parallel contraction along the strike of the beds and
extension along the dip direction. These are youn-
ger than the CdEn and CdEs systems and older than
the CnEd system (Table 2 and Figure IS).
Zone 3. This covers the exposed Middle and
Lower Chalk in the lower part of the steep limb of
the monocline (Figure 13). Bedding is less tilted
than in Zone 1. Mesofractural patterns and the rela-
ted extensional strain pattern in this zone (Figures
6A,B,C and 14A) differ from Zone 1 by the absence
of the early CdEn system and relatively weak deve-
lopment of the CdEs system (Table 2 and Figure
IS). Zone 3 also differs from Zone I in that rejuvena-
tion of early fractures by stylolitization is rare com-
pared to rejuvenation by dilation, as evident from
the abundance of extensional fibrous calcite veins
(Figure lOB) compared with rare and weakly deve-
loped tectonic pressure-solution seams.
Bedding faults in this zone show less displace-
ment than in Zone 1. The cumulative (total) dis-
placement on bedding in a 10-m-thick unit is 0.2m
to 3 m on the Dorset coast and O.lm to 1 m on the
Isle of Wight.
Zone 4. This zone covers the upper part of the
steep limb where Tertiary rocks are exposed
(Figure 13). As can be seen in Figures 6D and 14A
and in Table 2, this zone is different from Zone 1.
Firstly, during the early stage of deformation only
the CdEn fracture system developed (Figure lIA).
Secondly, the CsEd system develped penecontem-
poraneously with the younger CnEd system (Figure
lIB,D). Some CnEd fractures occur as ductile zones
in zones in the shale or clay horizons (Figure lIC).
Strain pattern in the Purbeck-Isle of Wight Monocline
Figure 14. Lower hemisphere
equal-area projection of directions of
maximum contractions (solid sym-
bols) and extension (open symbols) in
the different zones of the monocline
obtained from mesofracture patterns
shown in Figures 5, 6, and 7 (num-
bers on lower left sides indicate zone
number as shown in Figure 13):
A) Steep limb of the monocline,
Zones 1-4.
B) Gently dipping flanks of the
monocline, Zones 5 and 6.
LEGEND
CdES * >:-
EdC
sn
E C } *
s dn
SS
0
Layer-parallel slip (faulting) is less obvious than in
Zone 1 but does occur, as indicated by striations on
bedding surfaces; however, no clear evidence of the
amount of bedding slip was found.
Zone 5. This zone covers the exposed down-
thrown parts of the monocline and consists of Upper
Chalk and Tertiary rocks. The exposures of Ter-
tiary rocks are largely limited to the eastern and
western coastal parts of the Isle of Wight and the
eastern end of the Dorset coast. Upper Chalk is also
exposed in the latter area. Mesostructures in these
rocks on the Dorset coast were dealt with separately
(Ameen and Cosgrove, 1990a,b), and only those in
the Tertiary rocks, Isle of Wight, are described here.
The mesofractures and related extensional strain
pattern are shown in Figures 7A and 14B. Ten
sets of joints and faults as well as bedding faults are
found in these rocks. The CdEn, CdEs systems
include both faults and joints (Figure 12A,B),
whereas the CnEd and CnEs systems are mostly
joints (Figure 12D). The cutting relationship indi-
cates that the CdEn' CdEs' EsCdn, and bedding paral-
lel faults are older than the CnEd, CnEs, and EdCns
systems (Table 2). This implies an early layer-
parallel contraction in a direction perpendicular to
the monoclinal axis, followed by later stages of
layer-parallel extension parallel and perpendicular
to the monoclinal axis (Figure 15A,B).
A
0
1
3
B
. Jo

r
0*
*
0
N
N
*
*
5
N
6
2
4
N
* ..

c:P.
N
*
.0 8
LEGEND
Bedding 0
C
n
Ed
573
Cd En ... l>
Cn Es * *
Cd Es 0
Cs Ed * "*
EdCsn } *
EsC
dn
574
M. S. Ameen
Figure 15. Schematic block
diagrams illustrating the
mesofractural patterns and
related extensional strain
patterns in different parts of
the Purbeck-Isle of Wight
Monocline and their varia-
tion during development of
the monocline. The actual
location of each fracture
system is indicated by num-
bers corresponding to Zones
1-6 shown in Figure 13; dot
pattern indicates basement.
Figure 16. A) Bottom view of the Upper Chalk
beds at Locality 20. (Figure 3B), showing
conjugate C.Ed faults (1 and 2) . B) C.Edn
stylolite at about 90 to bedding (B) in the
Upper Chalk at Locality 20.
The average cumulative (total) displacement on
bedding faults in a IO-m-thick unit is 0.3 m on the
Dorset coast and 0.1 m on the Isle of Wight.
,.
1cm
-
("
B
Zone 6: This covers the gently dipping up-
thrown part of the monocline (Figure 13). Most of
this part of the monocline has been removed by
Strain pattern in the Purbeck-Isle of Wight Monocline 575
erosion on the Dorset coast. However, limited ex-
posures of the Lower Chalk in the southern parts of
the Isle of Wight show the mesostructural pattern in
the upthrown part of the monocline.
The measured mesostructures (Figure 7B) and
related extensional strain (Figure 14B) show a
rather simple pattern. Faults and stylolites are
absent, and only joints are found in this part of the
structure (Figure 17). The joints fall into three
systems of conjugate (shear) joints: CnEd' CsEd, and
CnEs, and two extension joint sets: EdCsn and CsEnd,
respectively. Field evidence indicates that the
CnEd' CsEd, and EdCnd systems are younger than
the CnEs and CsEnd systems (Table 2). This implies
an early layer-parallel extension in a direction
perpendicular to the monoclinal axis, followed by
layer-parallel extension in a direction parallel to the
monoclinal axis (Figure 15).
DISCUSSION AND CONCLUSIONS
The present study shows that the Upper Cre-
taceous to Tertiary rocks in the Purbeck-Isle of
Wight region accommodated themselves to reverse
movement of a pre-existing fault in the basement, in
a predominantly ductile manner on a macroscopic
scale, forming the monoclinal fold.
The forced folding took place mainly by meso-
scopic and microscopic brittle and semibrittle flow
(fracture flow or cataclasis). This includes frac-
turing and readjustment of fracture-bounded blocks
by faulting, dilation, and/or pressure solution. Al-
though faulting and jointing were observed by pre-
vious authors (Strahan, 1895, 1898; Bristow, 1889;
White, 1921; Arkell, 1936a,b, 1947; Philips, 1964;
Bevan, 1985a,b,c), no previous mention was made to
the rejuvenation of these fractures by faulting, dila-
tion, and pressure solution. Stylolites in the Upper
Chalk of the Dorset coast were noted by Jones and
others (1984), but no measurements of these struc-
tures were given and their tectonic implications
were not commented upon.
Most mesostructures and their related exten-
sional strain pattern are symmetrically oriented
about the bedding rather than the horizontal. This
suggests that layering anisotropy played an impor-
tant role as a stress guide during folding and that
the principal stresses remained near normal or
parallel to bedding. During folding, considerable
bedding-parallel slip occurred. This is particularly
apparent in the steep limb and downthrown gentle
Figure 17. Joints in the Lower Chalk in the gently
dipping upthrown part of the monocline (Locality 27,
Figure 3S): A) conjugate CnEd joints (1 and 2); Bl EdCns
extension joint. '8' is bedding trace.
576
M.S.Ameen
flank of the monocline. The sense of dip is such that
upper layers slip away from the synclinal hinge of
the monocline.
The exposed part of the Purbeck-Isle of Wight
Monocline is divided into three major zones ac-
cording to the degree of deformation: (1) the most
deformed steep limb; (2) the less deformed, down-
thrown, gently dipping part; and (3) the least de-
formed, upthrown, gently dipping part.
The steep limb of the monocline, with more
than one generation of mesostructures, shows a
more complex mesostructural pattern and is more
deformed than the gently dipping upthrown and
downthrown parts. The structures within the steep
limb include faults, joints, and stylolites, and their
relative chronology varies in space. This indicates a
heterogeneous pattern of strain in both space and
time. The limb is divided into at least four different
zones. The lack of accessible, continuous exposure
makes it difficult to determine relative ages of
structures that fall in different zones. However bor-
ders ofthe different zones were located as accurately
as possible, and the sequence of development of
mesostructures in each zone was deduced. The
earliest mesostructures in the monocline indicate an
early dominant layer-parallel contraction in a direc-
tion perpendicular to the fold axis. As displacement
on the basement fault increased, the limb of the
monocline rotated until it reached a position where
layer-parallel extension at about 90 to the fold axis
dominated. Further displacement on the basement
fault led to part of the steep limb being overturned
locally and, in the inner arc of the syncline, to a
second stage oflayer-parallel contraction.
After the fold locked up, a strike-slip fault sys-
tem developed, with sense of movement indicating
contraction at about 90 to the fold axis. This was
followed or accompanied by macrofaulting of the
cover rocks. The macrofaulting is limited to the
Dorset part of the structure and was discussed
separately (Ameen, 1990).
The mesostructures and the deduced strain
patterns indicate that the monocline (effectively,
the cover of Upper Cretaceous to Tertiary rocks on
both the Dorset coast and on the Isle of Wight)
developed as a forced (drape) fold during reverse
movement on a S-dipping Variscan basement fault
during mid-Oligocene to early Miocene time. This
conclusion is supported by recent geophysical data
(Simpson and others, 1987). The suggestion that the
monocline is an "independent buckle fold" (see
Arkell, 1936a,b, 1947; Bevan, 1985a) is convincingly
disproved. Buckle folds owe their development to a
compressive stress that acts along the length of the
layers (Ramsay, 1967). Therefore, the strain pat-
tern in a buckled multilayer sequence is dominated
by early contraction parallel or subparallel to the
layering (tangential longitudinal strain) and/or
layer-parallel shear (flexural slip or flexural flow).
As a buckle fold grows, its inner arc experiences con-
traction approximately parallel to the layering
while the outer arc of the fold experiences extension.
These two zones are separated by a "neutral sur-
face" parallel or at a low angle to the layering
(Ramsay, 1967, Figs 7.63 and 7.65). As folding
progresses, the neutral surface moves toward the
inner arc so that areas which have experienced
layer-parallel contractions may begin to experience
layer-parallel extension. However in a forced fold
like the Purbeck-Isle of Wight Monocline there is a
lack of evidence of layer-parallel contraction in the
gently dipping upthrown limb of the monocline.
Furthermore, the extensional strain pattern is con-
siderably more complicated, with at least several
zones experiencing different patterns of extensional
strain. Although the directions of maximum exten-
sion or contraction in these zones are parallel or
subparallel to the bedding, the finite border between
the zones (which is equivalent to the finite neutral
surface in buckle folds) can be parallel to or crosscut
the layering at any angle. Some strain zones experi-
ence more than one change in strain pattern from
layer-parallel extension to layer-parallel contrac-
tion or vice versa.
Strain patterns similar to those of the Purbeck-
Isle of Wight have been observed in experimentally
produced forced folds due to reverse faulting in the
basement (Ameen, 1988). Although strain pattern
in the latter folds varies according to shape, angle of
dip and displacement of the basement fault, and
properties and thickness of the cover rock, all ex-
periments show a heterogeneous pattern of exten-
sional strain with both time and space. Therefore,
at each stage of the folding it is possible to divide the
fold into zones according to the pattern of finite
strain. Layer- parallel slip develops mainly in the
steep limb and, to a lesser extent, in the gently
dipping downthrown limb of the monocline. A mini-
mal amount of layer-parallel slip may develop in the
upthrown, gently dipping limb of the monocline.
The strain is highly concentrated in the vicinity of
the basement fault (the steep limb and synclinal
part) and dies away from it. This strain pattern
contrasts with that obtained in experiments with
Strain pattern in the Purbeck-Isle of Wight Monocline 577
normal basement faulting, where the cover rocks
show a qualitatively homogeneous pattern of strain
dominated by layer-parallel extension in a direction
perpendicular to the fold axis and contraction per-
pendicular to the bedding (for details, see Ameen,
1988). Further detailed field and experimental
studies are needed for a better understanding of
forced folding due to dip-slip basement faults. This
will further help in finding reliable criteria for
predicting the type of basement faults from strain
analysis in the cover rocks.
ACKNOWLEDGMENTS
The author wishes to thank Dr. J. W. Cosgrove,
Imperial College, and two anonymous reviewers for
reading the manuscript and their helpful comments,
Miss Gill Davies for typing the manuscript, and the
Arabian Gulf University, Bahrain, for its award.
REFERENCES
ALEKSANDROWSKI, P., 1985, Graphical determination of
principal stress directions for slickenside lineation popu-
lations: An attempt to modify Arthaud's method: Journal
of Structural Geology, v. 7, p. 73-82.
AMEEN, M. S., 1988, Folding of Layered Cover Due to
Dip-Slip Basement Faulting: PhD dissertation, Imperial
College, London, England, UK, 231 p.
__ , 1990, Macrofaulting in the Purbeck-Isle of Wight
monocline. Proceedings of the Geologists' Association, v.
101, p. 31-46.
__ , and J. W. COSGROVE, 1990a, A kinematic analysis
of the Ballard Fault, Swanage, Dorset. Proceedings of the
Geologists' Association, v. 101, p. 119-129.
__ , and J. W. COSGROVE, 1990b, A kinematic analysis
of mesofractures from Studland Bay, Dorset. Proceedings
of the Geologists' Association, v.101, p. 303-314.
ANGELlER, J., 1984, Tectonic analysis of fault slip data
sets: Journal of Geophysical Research, v. 89, p. 5835-5848.
ARKELL, W. J., 1936a, The tectonics of the Purbeck and
Ridgeway faults in Dorset: Geological Magazine, v. 73, p.
56-73.
__ , 1936b, The tectonics of the Purbeck and Ridgeway
faults in Dorset: Geological Magazine, v. 73, p. 97-118.
__ ,1947, The Geology of the Country Around Wey-
mouth, Swanage, Corfe and Lulworth: Memoir of the
Geological Survey, United Kingdom, 386 p.
BEVAN, T. G., 1985a, A reinterpretation offault systems in
the Upper Cretaceous rocks of the Dorset Coast, England.
Proceedings of the Geologists' Association, v. 96, p. 337-342.
__ , 1985b, Tectonic evolution of the Isle of Wight: A
Cenozoic stress history based on mesofractures: Proceed-
ings of the Geologists' Association, v. 96, p. 227-235.
__ , 1985c, A Cenozoic Stress History of Southern
England Inferred from Mesofractures: PhD disserta-
tion, University of Bristol, Bristol, England, UK, 158 p.
BRISTOW, H. W., 1889, The Geology of the Isle of Wight:
Memoir of the Geological Survey, England & Wales, 949 p.
CHATWIN, C. P., 1960, British Regional Geology, the
Hampshire Basin and Adjoining Areas: Her Majesty's
Stationery Office, London, England, UK, 99 p.
COLTER, V. S., and D. J. HOWARD, 1981, The Wytch Farm
oil field, Dorset; pp. 494-503 In L. V. Bling and G. D
Hobson (eds.), Petroleum Geology of The Continental
Shelf of North-West Europe: Proceedings of the 2nd
Conference (March 1980): Hayden & Son, London, Eng-
land, UK, 539 p.
EDMONDS, E. A., M. C. McKEOWN, and M. WILLIAMS,
1975, British Regional Geology, South-west England:
Institute of Geological Sciences, London, England, UK, 139
p.
HANCOCK, P. L., 1985, Brittle microtectonics, principles
and practice: Journal of Structural Geology, v. 7, p. 437-
457.
HARRIS, J. F. , L. G. TAYLOR, and J. L. WALPER, 1960,
Relation of deformational fractures in sedimentary rocks
to regional and local structures: American Association of
Petroleum Geologists Bulletin, v. 44, p. 1853-1873.
HINDE, P., 1980, The development of the Wytch Farm oil-
field. Institute of Gas Engineers (London), Communication
1133.
JONES, M. G., J. BEDFORD, and C. CLAYTON, 1984, On
natural deformation mechanisms in the Chalk: Journal of
the Geological Society of London, v. 141, p. 675-683.
KULANDER, B. R., C. C. BARTON, and S. L. DEAN, 1979,
The Application of Fractography to Core and Outcrop
Fracture Investigations: U.S. Dept. of Commerce, Na-
tional Technical Information Service, Springfield, Vir-
ginia, USA, 174 p.
LAMING, D. J. C., 1966, Imbrication, paleocurrents and
other sedimentary features in the lower New Red Sand-
stone, Devonshire, England: Journal of Sedimentary
Petrology, v. 36, p. 129-135.
578 M.S.Ameen
LEES G. M., and P. T. COX, 1937, The geological basis of
the present search for oil in Great Britian by the D'Arcy
Exploration Company Ltd: Quarterly Journal of the Geo-
logical Society of London, v. 93, p. 156-194.
MELVILLE, R. V., and E. C. FRESHNEY, 1982, British
Regional Geology, the Hampshire Basin and Adjoin-
ing Areas: Her Majesty's Stationery Office, London,
England, UK, 146 p.
NORRIS, D. K., 1958, Structural conditions in Canadian
coal mines: Geological Suruey of Canada Bulletin, v. 44, p.
1-53.
PHILLIPS, W. J., 1964, The structures in the Jurassic and
Cretaceous rocks of the Dorset coast between white N othe
and Mupe Bay: Proceedings of the Geologists' Association,
v. 75, p. 375-407.
PRICE, N. J., 1966, Fault and Joint Development in
Brittle and Semibrittle Rocks: Pergamon Press, Oxford,
England, UK, 176 p.
RAMSAY, J. G., 1967, Folding and Fracturing of Rocks:
McGraw-Hill, New York, New York, USA, 568 p.
SELLEY, R. C., and R. STONELEY, 1983, A Field Guide to
the Petroleum Geology of the Wessex Basin: Geology
Department, Imperial College, London, England, UK, 27 p.
SIMPSON, I. R., P. M. GRAVESTOCK, D. HAM, H. LEACH,
and S. D. THOMPSON, 1987, Inversion tectonics of the
Wessex Basin:. Abstracts of the Geological Society Con-
ference on Inuersion Tectonics (London, March 1987), p. 17.
STONELEY, R., 1982, The structural development of the
Wessex Basin: Journal of the Geological Society of London,
v. 139, p. 543-554.
STRAHAN, A., 1895, On overthrusting of the Tertiary date
in Dorset: Quarterly Journal of the Geological Society of
London, v. 51, p. 549-562.
__ , 1898, The Geology of the Isle of Purbeck and
Weymouth: Memoir of the Geological Survey, United
Kingdom, 278 p.
WEBBY, B. D., 1965a, The stratigraphy and structure ofthe
Devonian rocks in the Brendon Hills, West Somerset:
Proceedings of the Geologists' Association, v. 76, p. 39-60.
__ , 1965b, The stratigraphy and structure of the
Devonian rocks in the Quantock Hills, West Somerset:
Proceedings of the Geologists' Association, v. 76, p. 321-344.
WHITE, H. J. 0., 1921, Short Account of the Geology of
the Isle of Wight: Memoir of the Geological Survey, Great
Britain, 201 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Characteristics of the circum-Tyrrhenian Hercynian massifs
and their role in the Alpine-Apennine chains
NELLO MINZONI
Dipartimento di Scienze Geologiche e Paleontologiche dell'Universita, Corso Ercole d'Este, 32, 44100 Ferrara, Italy
(received and accepted November 18, 1988)
SARDINIAN-CORSICAN MASSIF
Southern Sardinia
In the southern part of Sardinia, the Cambrian-
Lower Ordovician succession is slightly deformed by
the "Sardic Phase." Large open folds run roughly
east-west and are never accompanied by penetrative
cleavage or metamorphism. The main deformation
and metamorphism were due to Hercynian orogene-
sis. The main Hercynian tectonics are characterized
by a subvertical strain-slip cleavage or, in the lower
structural levels, by isoclinal folds accompanied by a
subhorizontal slaty cleavage and syn-kinematic
metamorphism in the greenschist facies (Minzoni,
1980).
The Paleozoic succession begins with acidic
volcanic rocks and volcaniclastic rocks (the "Por-
phyroids"), which are conformably overlain by fos-
siliferous Lower Cambrian rocks. The Porphyroids
are very similar to the "OUo de Sapo" Formation
exposed at the base of the Cambrian in central
Spain. Therefore the Porphyroids, like the analo-
gous Spanish formation, may be considered of infra-
Cambrian age.
In southern Sardinia, the Cambrian-Lower Or-
dovician is characterized by a thick epicontinental
succession consisting of sandstone with archeocya-
thid-bearing dolomite overlain by dolomite and
limestone (Lower-Middle Cambrian) and, lastly, by
Paradoxides-bearing and Dictyonema flabelliforme-
bearing sandstone (Middle Cambrian-Lower Ordo-
vician).
Cambrian-Lower Ordovician rocks, deformed
by the Sardic Phase, are unconformably overlain by
a polygenic conglomerate (the "Puddinga") and
sandstone, passing at the top into thin (5-30 m)
fossiliferous siltite and slate of Late Ordovician age.
Silurian-Devonian rocks consist of graptolite-
bearing slate and limestone. Basic volcanic rocks
may be interbedded with the Silurian strata.
Middle Ordovician acidic and basic volcanic rocks
are scarce or absent.
Central Sardinia
In central Sardinia, the Sardic Phase is charac-
terized primarily by tensional tectonics (Minzoni,
1988). Volcanic rocks of Middle Ordovician age are
very abundant, and the Puddinga is replaced by
volcaniclastic rocks. Middle Ordovician sequences
show a marked lateral variability in both thickness
and lithology. Upper Ordovician rocks consist of a
very thick (500-1000 m) nonfossiliferous flysch
succession. Silurian-Devonian rocks consist of black
shales and limestones. Basic volcanic rocks of Silur-
ian age may be very abundant. The Cambrian-
Lower Ordovician succession consists of Porphyroids
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 579-582. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
579
580 N. Minzoni
that are overlain by Middle Cambrian-Lower Ordo-
vician terrigenous formations or, in some areas,
directly by Caradocian strata, Therefore, there are
marked stratigraphic hiatuses in central Sardinia.
Cambrian-Early Ordovician deposition was accom-
panied by syn-sedimentary normal faults.
Several Hercynian tectonic phases are recog-
nizable in central Sardinia, but only the first phase
produced isoclinal recumbent folds, accompanied by
a sub horizontal slaty cleavage and syn-kinematic
metamorphism in the greenschist facies (Carmig-
nani and others, 1978). The polarity of sequences
and the geometry of the isoclinal recumbent folds
indicate that the direction of the overturning is west
or southwest. The most important result is large
nappes of easterly or northeasterly provenance. The
emplacement of the allochthonous units in the more
internal zones of central Sardinia differs from that
of the more external zones. In the more external
zones, overthrust surfaces formed by utilizing
Middle Cambrian-Lower Ordovician syn-sedimen-
tary normal faults. Therefore the nappes consist,
above all, of Porphyroids and/or Middle Cambrian-
Lower Ordovician terrigenous successions. The
more internal zones display a structural style char-
acteristic of deeper levels. There are large recum-
bent folds with inverted limbs often preserved
(Carmignani and others, 1979a; Minzoni, 1988).
These folds consist primarily of very thick Upper
Ordovician flysch sequences and of Silurian and
Devonian rocks.
Northeastern Sardinia
The northeastern area of Sardinia is charac-
terized by high-grade metamorphism. Monotonous
pelitic sequences and abundant basic volcanic rocks
were metamorphosed during the Hercynian orogeny
in the amphibolite facies. It is generally accepted
that most of the original volcanic-pelitic rocks were
derived from a Middle Ordovician-Devonian succes-
sion, deposited in a rifted area.
The data indicate that during Middle Ordo-
vician times, the compressional tectonics of south-
western Sardinia formed contemporaneously with
tensional tectonics of central-northern Sardinia.
These different, but contemporaneous, phenomena
in adjacent areas suggest that during the Middle
Ordovician, a single geodynamic regime governed
all tectonic, sedimentary, and volcanic events. The
more coherent geodynamic model that unifies all
geological data is represented by "transcurrent
rifting." This model explains contemporaneous
formation of the rifted area in central-northern Sar-
dinia and of the compressional-tensional tectonics in
central-southern areas of the island.
During Cambrian to Early Ordovician times,
Sardinia was subjected only to tensional tectonics
that developed very slowly and produced (especially
during Middle Cambrian-Early Ordovician time)
similar sedimentary environments in widespread
areas. Similar successions crop out in central-
southern Sardinia and in the Calabrian-Peloritan
Massif. In other words there is no contrast of dif-
ferent paleogeographic domains throughout wide-
spread areas. On the contrary, beginning with the
Middle Ordovician, a clear contrast of different
environments began to materialize in adjacent
areas. The very thick nonfossiliferous flysch se-
quences of Late Ordovician age in the central-
northeastern area are opposed to very thin, fossili-
ferous sequences in central-southwestern areas.
This means that starting from the Middle
Ordovician, the Hercynian "Geosyncline" began to
develop and the true Hercynian cycle began to form.
The Middle Ordovician was a period of particular
instability of the crust, evidenced by deposition of
thick flysch sequences and outflow of abundant
volcanic rocks in central-northeastern Sardinia. In
this framework, the Cambrian-Lower Ordovician
rocks do not belong to the true Hercynian cycle but
represent a post-Assintyc molasse, and the Por-
phyroids may represent late to post-Assintyc
magmatism.
Summary
In conclusion, in central-northeastern Sardinia
the crust was deeply fractured starting from the
Middle Ordovician, thus forming an intracontinen-
tal rift. The compressional movements in south-
western Sardinia (Sardic Phase) must be considered
as a minor phenomenon of transcurrent movements.
The absence of ophiolite associations indicates the
Hercynian cycle involved only the continental crust.
This entirely ensialic evolution is not in contrast
with the structural style of the Hercynian belt and
is analogous to that of collisional chains along con-
tinental margins.
There are some unsolved questions: the pos-
sible existence of a pre-Hercynian basement and the
significance of Ordovician magmatism. The exis-
Characteristics of the circum-Tyrrhenian Hercynian massifs 581
tence of a pre-Cambrian basement is suggested by
the Porphyroids of southern Sardinia. These vol-
canic rocks show a crustal origin (87Sr/86Sr = 0.7122
0.0058) and they crop out at the base of Lower
Cambrian rocks. Moreover, at the base ofthe Lower
Cambrian succession there are conglomerate lenses
containing porphyroid clasts. However, the radio-
metric age (42734 Ma) conflicts with the above-
mentioned stratigraphic data.
The other question pertains to the geodynamic
significance of Middle Ordovician magmatites.
Their geochemical data suggest a late-orogenic
magmatic-anatectic association and their age
suggests a Caledonian orogenic cycle (Memmi and
others, 1982). However, there are some data
opposed to this interpretation. For example, the
absence of major tectonism and penetrative
deformation prior to Hercynian folding. The Sardic
Phase is very slight and it is limited to southwestern
Sardinia. At present, no proof has been obtained of
a regional metamorphism that can be referred to the
Caledonian orogenic cycle. The Sardic Phase
formed contemporaneously with transcurrent-
tensional tectonics that are related to the beginning
of the Hercynian geosyncline.
In Sardinia, a tectono-metamorphic division
into zones analogous to the "Alpine belt" can be
made. Slightly metamorphosed southwestern Sar-
dinia can be identified as a domain with the
structural characteristics of a foreland: the central
zone represents an area of accumulation of nappes
originating more internally, and northeastern
Sardinia is a "root zone" characterized by several
superimposed metamorphic episodes. There is a
parallel between paleogeographic domains and
Hercynian tectono-metamorphic zones, which mani-
fests itself in considering the Cambrian-Lower
Ordovician (post-Assintyc) domains in the foreland
areas and the Middle Ordovician-Devonian (true
Hercynian) domains in the more internal areas.
Northeastern Sardinia is a rifted area and corres-
ponds to the portion of the crust most deeply
subducted along an intracontinental NE-dipping
shear zone. In this way, Corsica is the backland that
overthrust the axial zone of the belt (Carmignani
and others, 1979b). The tectono-metamorphic
situation of northeastern Sardinia suggests a long
period of crustal thinning starting in the Middle
Ordovician. This zone of weakness in the crust was
deeply reactivated during Hercynian shortening
and became the site of the more important shears.
The large recumbent structures of the more internal
part of the central zone formed in correspondence to
thick flysch sequences of Late Ordovician age. This
means that in the more internal zones of central
Sardinia, as in northeastern areas, the Middle
Ordovician-Devonian paleogeographic domains
parallel the structural lineaments. In the more
external zones of central Sardinia, the alloch-
thonous units developed by utilizing syn-sedimen-
tary normal faults of Cambrian-Ealry Ordovician
age (post-Assintyc paleogeographic domains). In
fact, nappes in this area consist above all of
Cambrian-Lower Ordovician sequences.
CALABRIANPELORITAN MASSIF
In northern Calabria, the Middle Cambrian-
Lower Ordovician rocks show the same lithology as
those of central-southern Sardinia. Porphyroids can
be present in some areas. In Calabria, the Sardic
Phase is characterized by tensional tectonics and by
outflow of abundant acidic and basic volcanic rocks.
The Silurian-Devonian succession consists of black
shales and limestones. Carboniferous terrains are
present in some areas. The Hercynian orogeny
produced isoclinal recumbent folds accompanied by
syn-kinematic, low-grade metamorphism. Some
high-grade metamorphites probably represent
Precambrian basement. The tectono-metamorphic
characteristics of the Calabrian-Peloritan Massif
suggest it was originally situated near central-
southern Sardinia.
ROLE OF THE HERCYNIAN MASSIFS IN THE
ALPINEAPENNINE CHAINS
It is well known that Corsica and Sardinia
played the role of Alpine nappe foreland. In Corsica,
the Alpine chain is cut by the Tyrrhenian Sea.
According to Scandone (1979), originally the Corsica
Alpine nappes must have been connected with the
Calabrian Alpine Belt in the south. Therefore, the
Tyrrhenian Sea is a Neogene-Quaternary exten-
sional area that, from Corsisca to southern Italy,
separates fragments of the Paleogene Europe-
verging eo-Alpine chain. During the early-middle
Miocene, the eo-Alpine chain overthrust part of the
African continental margin, which later generated
the Appennines.
Structural studies carried out in recent years
confirm the above-mentioned geodynamic model. In
582 N. Minzoni
fact, Paleogene W-verging nappes are present in
several areas of Calabria (Minzoni, 1990). The
nappes consist of pre-Mesozoic crystalline rocks or
Mesozoic carbonate sediments, and/or Paleogene
terrigenous sequences. The following paleogeo-
graphic units occur from Tyrrhenian to the Ionian
Sea: a pre-Mesozoic crystalline massif; a Mesozoic
carbonate platform overlain by Paleogene terri-
genous sequences; the "Argille varicolori" overlain
by the "Numidian" sequences. According to Mostar-
dini and Merlini (1986), the "Argille varicolori"
must be considered as having been deposited east of
the Mesozoic platform. This means that the pre-
Mesozoic crystalline rocks must be considered as an
ancient Austroalpine domain that represents the
deformed backland of the Alpine chain.
REFERENCES
CARMIGNANI, L.,T. COCOZZA, N. MINZONI, and P.C.
PERTUSATI, 1978, Falde di ricoprimento erciniche nella
Sardegna a nord-est del Campidano: Memorie Societa
Geologica Italiana, v.19, p. 501-510.
CARMIGNANI, L. ,M. FRANCESCHELLI, P.C. PERTUSATI,
and C. A. RICCI, 1979a, Evoluzione tettonico-metamorfica
del basamento ercinico dell Nurra (Sardegna NW):
Memorie Societa Geologica Italiana, v. 20, p. 57-84.
CARMIGNANI, L., T. COCOZZA, N. MINZONI, P. C.
PERTUSATI, and C. A. RICCI, 1979b, E' la Corsica il
retropaese della catena ercinica della Sardegna?: Memorie
Societa Geologica Italiana, v.20, p. 47-55.
MEMMI, I., S. BARCA, L. CARMIGNANI, T. COCOZZA, F.
ELTER, M. FRANCESCHELLI, M. GATTIGLIO, C. GHEZZO,
N. MINZONI, G. NAUD, P. C. PERTUSATI, and C. A. RICCI,
1982, Further geochemical data on the pre-Hercynian
igneous activities of Sardinia and on their geodynamic
significance: IGCP Project No.5 Newsletter, v. 5, p. 87-91.
MINZONI, N., 1980, Precambriano nel Sulcis meridionale
(Sardegna): Mineralogica Petrographica Acta, v. 24, p. 51-
56.
__ , 1988, Geologia strutturale della zona di Gadoni-
Funtana Raminosa (Sardegna centrale): Memorie Societa
Geologica Padoua, v. 90, p. 195-201.
__ , 1990, Falde paleogeniche ovest-vergenti nella
Calabria meridionale: Accademia Nazionale Lincei, v. 82,
p.773-778.
MOSTARDINI, F., and S. MERLINI, 1986, Appennino
centro-meridionale: Sezioni geologiche e proposte di mod-
ello strutturale: Azienda Generale Italiania Petroli
Mineraria (Milano), p. I-56.
SCADONE, P., 1979, Origin of the Tyrrenian sea and
Calabrian Arc: Bollettino Societa Geologica Italiana, v. 98,
p.27-34.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Structure and tectonic evolution of the western continental
margin of India: Evidence from subsidence studies for a
25-20 Ma plate reorganization in the Indian Ocean
ABINASHAGRAWAL
1
and JOHN J. W. ROGERS
2
IBasin Research Division, Room 172, KDM Institute of Petroleum Exploration,
Oil and Natural Gas Commission, Kaulagarh Road, Dehra Dun: 248195, India
2Department of Geology, CB 3315, University of North Carolina, Chapel Hill, North Carolina, 27599-3315, USA
(received September 19, 1988; revision accepted March 31, 1989)
ABSTRACT
The western continental margin of India developed by Mesozoic rifting and has subsided and undergone
further tectonic modification during India's northward movement and collision with Asia. Segmentation of
the margin has apparently been controlled partly by inheritance of Precambrian structures, across one of
which different rates of subsidence and degrees of stretching have occurred. At about 25-20 Ma, following a
widespread erosional unconformity, the rate of subsidence greatly increased both on the continental shelf and,
in particular, seaward ofthe shelf edge (hinge line). This increased stretching occurred during reorganization
of the plates in the region of the Indian Ocean, coincided with greater resistance to underthrusting of India
beneath Asia, and may be related to enlargement of the Indian plate from the Owen fracture zone westward to
the Dead Sea-Gulf of Aqaba shear zone.
INTRODUCTION
The western continental margin of India (Fig-
ure 1) formed by rifting from Africa and has been
affected by numerous processes throughout the
Cenozoic. This preliminary paper outlines the
significant features of the continental margin and
discusses the importance of a major change in
subsidence rates at about 25-20 Ma.
STRUCTURE
The west coast of India represents a passive
margin initially developed by Mesozoic rifting (Fig-
ures 1 and 2). Three major EIW-trending features
(Narmada-Son lineament, Ratnagiri-Alibag fault,
and Vengurla arch) divide the western shelf into
four distinct blocks (from north to south): Kutch-
Saurashtra, Bombay, Ratnagiri, and Konkan-
Kerala (Mitra and others, 1983). The following dis-
cussion indicates that the Narmada-Son, Ratnagiri-
Alibag, and Vengurla structures may all have been
inherited from Precambrian structures.
The Narmada-Son lineament is an ENEIWSW-
trending strike-slip fault system that has shown
dextral, sinistral, and dip-slip movement at various
times, has been active episodically since the Pro-
terozoic, and is traceable in Madagascar (Crawford,
1978; Das and Patel, 1984). The fault forms a
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 583-590. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
583
584
A. Agrawal and J. J. W. Rogers
...
N
r

4J!.U1.A./II
';::
w l'
,.
o QUATERNARV IiU
""-WI ....
_dJ
o PROTEROZOIC
" "\,f
BAS
'r

LACCAOIVE
100
2'f' , , , DEPRESSION
KILOMETERS
.. 70 n 7C
Figure 1. Index map showing various tectonic features on the west coast ofIndia and the locations of wells studied (A to G).
tectonic boundary that separates the Bombay block,
on the south, from the offshore Kutch-Saurashtra
basin and on-land Kathiawar platform and Cambay
basin (Biswas, 1982; Bhattacharya and Subrah-
manyam, 1986; Kaila, 1986). Rifting occurred in
the Kutch-Saurashtra margin during the Triassic
and in the area south of the Narmada-Son
lineament during the Late Cretaceous. The Cam-
bay basin, a small intracratonic basin north of the
Narmada-Son lineament, formed by Cretaceous
rifting between the Kathiawar platform and the
Precambrian Aravalli province (Raju, 1968, 1983).
The Ratnagiri-Alibag fault is a transfer fault
that separates the Bombay basin from the Ratnagiri
basin. The wide shelf of the Bombay block and its
observed westward movement between the Nar-
mada-Son and Ratnagiri-Alibag bounding faults
(Rao and Talukdar, 1980) indicates major crustal
thinning in the Bombay block. The Ratnagiri-Ali-
bag fault also offsets N/S-trending features, includ-
ing a hinge-line (shelf edge) (Rao and Talukdar,
1980) and the Laxmi ridge. (Data from structural!
seismic transects, lines A to E in Figure 1, will be
published elsewhere.) The Ratnagiri-Alibag fault
trends eastward from the continental shelf toward
the broad Deccan plateau, making it difficult to
determine whether the fault is inherited from a
Precambrian structure or evolved wholly within the
Mesozoic extensional terrane. Two lines of evi-
dence, however, indicate an on-land extension of the
fault: (1) the drainage patterns in the Deccan pla-
teau are different north and south of the on-land
extension of the fault; and (2) ENE- and WSW-
trending fractures and dyke swarms in the plateau
show sinistral strike-slip movement along the on-
land extension (Rao and Talukdar, 1980).
The Vengurla arch is a broad region along the
coast with substantial basement relief (Ramaswamy
and Rao, 1980), particularly on its northern side. A
possible pre-existing structure is a (faulted) shelf
margin of the Late Proterozoic Bhima and Kaladgi
basins, which lie on the Precambrian craton of
southern India and disappear northward beneath
the Deccan basalt. Other Middle to Late Proterozoic
A 25-20 Ma plate reorganization in the Indian Ocean? 585
GENERALIZEP CROSS SECTION FROM BOMBAY COAST TO LAXMI RIDGE
LAXMI LACCADfYE
DEPRESSION
,.0
Figure 2. Generalized cross-section from Bombay coast to Laxmi ridge.
basins of India appear to have formed along rifted
margins (Rogers, 1986), but the evidence for the
Bhima and Kaladgi basins having developed in such
an environment is inferential because of the Deccan
cover.
In addition to the three features mentioned
above, a fourth possible Precambrian "structure"
(the "Tellicherry arch" of Biswas and Singh, 1988)
occurs between the Konkan and Kerala subbasins.
Different Paleogene sedimentary sequences in the
two subbasins (Mohan and Kumar, 1980; Rama-
swamy and Rao, 1980) may be related to the off-
shore extension of the 2500-Ma transition zone that
separates rocks of amphibolite facies (to the north)
from those of granulite facies (to the south) (Rogers,
1986).
Features parallel to the coast, in a NNW-SSE
direction (Figure 2), include horst-graben struc-
tures, a hinge line (shelf edge), the Kori-Comorin
ridge, and the Laxmi ridge. Magnetic anomaly
transects and estimates of crustal thickness from
subsidence curves indicate that thin, transitional
crust extends to the Laxmi ridge and may actually
underlie the Eastern Basin (Naini and Talwani,
1982) of the Arabian Sea, hitherto considered to be
oceanic. The Laxmi ridge may be a continental
sliver. The tectonic boundary between the thick,
block-faulted continental crust of the Indian shelf
and thinned, transitional crust underlying the
Eastern Basin is shown as a hinge line (Figures 1
and 2).
SUBSIDENCE STUDIES
Seven offshore deep wells (locations A to G in
Figure 1) were chosen along the western shelf to
calculate the depth of unloaded basement through
Paleogene and early Neogene time. Airy compensa-
tion was assumed, and the determinations followed
the procedures of Sclater and Christie (1980) and
Heller and others (1982). Wells A and B, north of
the Narmada-Son lineament, show slightly higher
subsidence rates prior to 25 Ma than wells farther
south. All wells, both on the present shelf and in the
basin seaward of it, show an increase in subsidence
rates immediately following the period 25-20 Ma
(Figures 3 and 4). The break in subsidence curves
generally correlates with the end of a major Oligo-
cene hiatus in the stratigraphic record (Mohan and
Kumar, 1980; Mohan and others, 1980, 1982; Mo-
han, 1985).
The subsidence rates have been used to esti-
mate stretching factors (j3) at the sites of the various
wells during the periods before and after 25-20 Ma.
The determinations were based on the instantane-
ous stretching model of McKenzie (1987); used the
relationship between lithospheric thickness and
586
A. Agrawal and J . J. W. Rogers
1500 1500 1500
LOCATION A LOCATION.C
Iii Iii Iii
a: a:
1000
a:
w 1000 w w
tu tu
I-
W

w w w
(J

(J u
z z z
w
500
w
500
w
0 0 0
iii iii iii
CD CD ..
:> :> :>
<II <II <II
0 0 0
60 40 20 60 40 20 60 40 20
TIME INMA TIME IN MA TIME IN MA
Figure 3. Subsidence curves for unloaded basement for wells A, B, and C. The period 25-20 Ma is shown by the stippled
pattern. Subsidence was calculated using the formulation of Heller and others (1982).
1500 1500
LQCATIQI'I-Q LQCATIQI'I-E
Iii Iii
a: a:
1000
\
w 1000 w
to
I-
W
!.
w w
(J U
Z Z
w w
500
0 0
iii iii
III III
:> :>
"'
<II
0 0
60 40 20 60 40 20
TIMEINMA TIME IN MA
1500 1500
LQCAIIQI'I- E LQCATIQt:!-G
Iii Iii
a:
w
l-
1000
a:
w
I-
1000
W W
!.
w w
(J u
z z
w
500
w
500
0 0
iii iii
.. III
iil
:>
"'
0
40 20 60 40 20
TIME IN MA TIME INMA
Figure 4. Subsidence curves for unloaded basement for wells D, E, F, and G. The period 25-20 Ma is shwon by the stippled
pattern. Subsidence was calculated using the formulation of Heller and others (1982).
subsidence rates proposed by McKenzie (1978) and
Sclater and Christie (1980); and assumed that litho-
spheric and crustal thinning were proportionate.
Assuming that initial crustal thicknesses were
equal at all sites, the stretching factors are in-
versely proportional to crustal thicknesses during
the period of subsidence. Because of the assump-
tions and uncertainties in the data, these estimates
A 25-20 Ma plate reorganization in the Indian Ocean? 587
are only approximate. Subsidence rates prior to the
period 25-20 Ma yield the stretching factors shown
in Table 1. Subsidence in all wells immediately
following that period is at the extremely high rates
characteristic of young oceanic crust. Oceanic (or
transitional) crust is possible for basinal areas
seaward of the shelf. Wells A, C, D, and F are on the
shelf, however, and although they clearly subsided
very rapidly after 20 Ma, either their actual subsi-
dence rates are lower than those calculated here or
they were accentuated by non-isostatic processes.
RELATIONSHIP OF INDIAN WEST COAST TO
EVENTS IN THE INDIAN OCEAN
The evolution of the west coast ofIndia provides
some information concerning the development of the
Indian Ocean (Figures 5 and 6). We call attention
to inferences that can be drawn from an increase in
subsidence rates at about 20 to 25 Ma along much of
the coastal area.
One inference results from the relationship
between the time of increased subsidence rates and
the development of a widespread, late Oligocene
erosional unconformity in the entire western shelf.
Such a regional unconformity might have resulted
from continued pushing by the Carlsberg ridge, in
the Arabian Sea, when the rate of underthrusting of
Asia by the Indian plate was considerably reduced
(Patriat and Achache, 1984). An increase in subsi-
dence rates after 25-20 Ma indicates that the uplift
which caused the unconformity was accompanied by
increased stretching along the continental margin.
A second inference results from the fact that an
age of 25-20 Ma is within the period of time in which
major plate reorganizations occurred in the Indian
Ocean and neighboring continents. Prior to about
45 Ma (magnetic anomaly 20), India probably was
on a plate bordered on the east by the Ninetyeast
ridge, on the south by precursors to the present
Carlsberg and Indian Ocean ridge system, on the
north by a subduction zone under the Asiatic main-
land, and on the west by the Owen fracture zone
(Figure 5; McKenzie and Sclater, 1971; Sclater and
Fisher, 1974; Norton and Sclater, 1979; Molnar and
others, 1988). The magnetic record in the Indian
Ocean is not complete in the middle Cenozoic, but
after about 35 Ma (in the range of magnetic ano-
malies 12 to 18), and before about 10 Ma (anomaly
5), some plate reorganization occurred by: (1) deve-
lopment of a diffuse, EfW -trending, compressional
Table 1. Pre-25-Ma stretching factor ([3) for well
locations A through G (assuming rifting at the
beginning of the Cenozoic).
Well
Pre-25 Ma Stretching Factor
Location (1l)
A >1.25- <1.5
B 1.25
C <1.25
D <1.25
E <1.25
F <1.25
G <1.25
plate margin across the central Indian Ocean (Stein
and Okal, 1978; Weins and others, 1985); and
(2) possible locking of the Owen fracture zone and
enlargement of the Indian plate westward to the
Dead Sea-Gulf of Aqaba shear zone (Figure 6;
Weins and others, 1985; Bohannon and others, 1989;
McGuire and Bohannon, 1989). Movement along
the Dead Sea-Gulf of Aqaba shear zone began at
about 20 Ma (summary by Rogers and others, 1989),
the same time as the initiation of enhanced sub-
sidence rates along the Indian west coast. Thus, the
age of plate reorganization in the Indian Ocean
probably can be constrained to about 20 Ma.
SUMMARY AND CONCLUSIONS
The evolution of the western continental mar-
gin of India has apparently been greatly affected by
pre-existing (Precambrian) structures and by plate
reorganizations in the western Indian Ocean.
Where known or inferred on-land Precambrian
structures intersect the coastline, the offshore ex-
tensions appear to have acted as transfer faults that
separate different terranes. This fragmentation of
the continental shelf, however, has not been so ex-
tensive as to prevent recognition of major changes
affecting the entire northwestern Indian Ocean re-
gion.
The principal change recognized in the Ceno-
zoic sections investigated by us is a relationship
588 A. Agrawal and J. J. W. Rogers
AFRICA
BEFORE
REORGANIZATION
36 Ma
Figure 5. Outline of the Indian plate, bordered by the Australian plate along the Ninetyeast ridge on the east and by the
Owen fracture zone on the west at about 36 Ma (after Norton and Sclater, 1979).
between a widespread late Oligocene unconfonnity
and an increase in subsidence rates beginning about
25-20 Ma. This period coincides with a time of in-
creased resistance to underthrusting of the Asiatic
mainland by India and a reorganization of plates in
the Indian Ocean. Our data confinn that the late
Oligocene is the most likely time for locking of the
Owen fracture zone, enlargement of the Indian plate
to the Dead Sea-Gulf of Aqaba transfonn fault, and
development of a compressional margin in the cen-
tral Indian Ocean.
ACKNOWLEDGMENTS
We thank two anonymous reviewers for com-
ments and criticism; this paper has benefited great-
ly from their reviews. We sincerely thank Member
(Exploration), Oil and Natural Gas Commission of
India for pennission to utilize unpublished infonna-
tion and without whose endorsement the study
could not have been undertaken. The financial sup-
port to A.A. for research and presentation, provided
from the Foreign Research FundIMcCarthy Fund of
the Department of Geology, University of North
Carolina at Chapel Hill, is gratefully acknowledged.
REFERENCES
BHATI'ACHARYA, G. C., and V. SUBRAHMANYAM, 1986,
Extension of the Narmada-Son lineament on the contin-
ental margin off Saurashtra, western India, as obtained
from magnetic measurements: Marine Geophysical Re-
search, v. 8, p. 329-344.
BISWAS, S. K., 1982, Rift basins of the western margin of
India and their hydrocarbon prospects, with special refer-
ence to the Kutch basin: American Association of Petro-
leum Geologists Bulletin, v. 66, p. 1497-1513.
A 25-20 Ma plate reorganization in the Indian Ocean? 589
AFRICA
SOMALIAN
PLATE
IN 0 I A N
AFTER REORGANIZATION
o Ma
BROt(;E.N fUDGE.
Figure 6. Present outline of the Indian plate, bordered to the east by the northern Ninetyeast ridge and to the southeast by a
diffuse boundary-zone. On the west, the plate is bordered by the Red Sea and the Dead Sea-Gulf of Aqaba shear zone.
__ , and N. K. SINGH, 1988, Western continental margin
of India and hydrocarbon potential of deep-sea basins:
Proceedings of the 7th Offshore Conference, South East
Asia Petroleum Exploration Society (Singapore), p. 170-
175.
BOHANNON, R. G., C. W. NAESER, D. L. SCHMIDT, and R.
A. ZIMMERMANN, 1989, The timing of uplift, volcanism,
and rifting peripheral to the Red Sea: A case for passive
rifting?: Journal of Geophysical Research, v. 94, p. 1683-
1701.
CRAWFORD, A. R., 1978, Narmada-Son lineament ofIndia
traced into Madagascar: Geological Society of India Jour-
nal, v. 19, p. 144-153.
DAS, B., and N. P. PATEL, 1984, Nature of the Narmada-
Son lineament: Geological Society of India Journal, v. 25,
p.267-276.
HELLER, P. L., C. M. WENTWORTH, and C. W. POAG, 1982,
Episodic post-rift subsidence of the United States Atlantic
continental margin: Geological Society of America Bulle-
tin, v. 93, p. 379-390.
KAlLA, K. L., 1986, Tectonic framework of Narmada-Son
lineament - A continental rift system in central India
from deep seismic soundings; pp. 133-150 in M. Barazangi
and L. D. Brown (eds.), Reflection Seismology: A Global
Perspective (Ithaca, New York, 1984): American Geo-
physical Union, Geodynamics Series 13, 311 p.
McGUffiE, A. V., and R. G. BOHANNON, 1989, Timing of
mantle upwelling: Evidence for a passive origin for the
Red Sea rift: Journal of Geophysical Research, v. 94, p.
1677-1682.
McKENZIE, D., 1978, Some remarks on the development of
sedimentary basins: Earth and Planetary Science Letters,
v. 40, p. 25-32.
__ , and J. G. SCLATER, 1971, The evolution of the
Indian Ocean since the Late Cretaceous: Geophysical
Journal of the Royal Astronomical Society, v. 24, p. 437-
528.
MITRA, P., P. L. ZUTSHI, R. A. CHOURASIA, M. L. CHUGH,
S. ANANTHANARAYANAN, and R. SHUKLA, 1983, Ex-
ploration in western offshore basins; pp. 15-24 in L. L.
Bhandari, B. S. Venkatachala, R. Kumar, S. N. Swamy, P.
590
A. Agrawal and J. J. W. Rogers
Garga, and D. C. Srivastava (eds.)., PetroUferous Basins
of India: Petroleum Asia Journal (Dehra Dun), v. 6, 189 p.
MOHAN, M., 1985, Geohistory analysis of Bombay High
region: Marine and Petroleum Geology, v. 2, p. 350-360.
__ , and P. KUMAR, 1980, Biostratigraphy of Kerala
Offshore: Geoscience Journal, v. 1, p. 1-9.
__ , P. KUMAR, V. NARAYANAN, A. GOVINDAN, K. S.
SOODAN, and P. SINGH, 1982, Biofacies, paleoecology, and
geological history of Bombay Offshore region: Oil and
Natural Gas Commission Bulletin (Dehra Dun), v. 19, p.
13-27.
__ , V. NARAYANAN, and P. KUMAR, 1980, Paleogene
and early Neogene biostratigraphy of Bombay Offshore
region; pp. 180-194 in Proceedings of the VII Indian
CoUoquium on Micropaleontology and Stratigraphy
(December 1978): Madras University, Madras, India.
MOLNAR, P., F. PARDO-CASAS, and J. STOCK, 1988, The
Cenozoic and Late Cretaceous evolution of the Indian
Ocean Basin: Uncertainties in the reconstructed positions
of the Indian, African and Antarctic plates: Basin Re-
search, v. 1, p. 23-40.
NAINI, B. R., and M. TALWANI, 1982, Structural frame-
work and evolutionary history of the continental margin of
western India; pp. 167-191 in J. S. Watkins and C. L.
Drake (eds.), Continental Margin Geology: American
Association of Petroleum Geologists, Memoir 34, 801 p.
NORTON, I. 0., and J. G. SCLATER, 1979, A model for the
evolution of the Indian Ocean and the breakup of Gond-
wanaland: Journal of Geophysical Research, v. 84, p. 6803-
6830.
PATRIAT, P., and J. ACHACHE, 1984, India-Eurasia
collision chronology has implications for crustal shorten-
ing and driving mechanism of plates: Nature, v. 311, p.
615-62l.
RAJU, A. T. R., 1968, Geological evolution of Assam and
Cambay Tertiary basins of India: American Association of
Petroleum Geologists Bulletin, v. 52, p. 2422-2437.
__ , 1983, Development of coastal and sedimentary
basins of Peninsular India in terms of plate tectonics
[abstract]; p. 378 in R. H. Gabrielsen, I. B. Ramberg, D.
Roberts, and O. A. Steinlein (ed.), Proceedings of the IV
International Conference on Basement Tectonics (Os-
lo, Norway, 1981): International Basement Tectonics Ass-
ociation, Inc., Salt Lake City, Utah, 382 p.
RAMASWAMY, G., and K. L. N. RAO, 1980, Geology of the
continental shelf of the west coast of India; pp. 801-821 in
A. D. Miall (ed.), Facts and Principles of World Petro-
leum Occurrence: Canadian Society of Petroleum Geolo-
gists, Memoir 6, 1003 p.
RAO, R. P., and S. N. TALUKDAR, 1980, Petroleum geology
of Bombay Highfield, India; pp. 487-506 in M. T. Halbouty
(ed.), Giant Oil Fields of the Decade 1968-1978: Ameri-
can Association of Petroleum Geologists, Memoir 30, 596 p.
ROGERS, J. J. W., 1986, The Dharwar craton and the as-
sembly of peninsular India: Journal of Geology, v. 94, p.
129-144.
__ , M. E. DABBAGH, B. M. WHITING, and S. A.
WIDMAN, 1989, Subsidence and origin of the northern Red
Sea and Gulf of Suez; in B. R. Rosendahl, J. J. W. Rogers,
and N. M. Rach (eds.), Rifting in Africa - Karoo to
Recent: Journal of African Earth Sciences, v. 8 (special
issue), p. 137-629.
SCLATER, J. G., and P. A. F. CHRISTIE, 1980, Continental
stretching - An explanation of the post-mid-Cretaceous
subsidence of the central North Sea basin: Journal of
Geophysical Research, v. 85, p. 3711-3739.
__ , and R. L. FISHER, 1974, Evolution of the east-
central Indian Ocean, with emphasis on the tectonic set-
ting of the Ninetyeast ridge: Geological Society of America
Bulletin, v. 85, p. 683-702.
STEIN, S., and E. A. OKAL, 1978, Seismicity and tectonics
of the Ninetyeast ridge area: Evidence for internal de-
formation of the Indian plate: Journal of Geophysical Re-
search, v. 83, p. 2233-2245.
WEINS, D. A., C. DEMETS, R. G. GoRDON, S. STEIN, D.
ARGUS, J. F. ENGELN, P. LUNDGREN, D. QUffiLE, C.
STEIN, S. WEINSTEIN, and D. F. WOODS, 1985, A diffuse
plate boundary model for Indian Ocean tectonics: Geo-
physical Research Letters, v. 12, p. 429-432.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Tectonics and regional unconformities of eastern India basins
J. J. GALLAGHER, Jr.
ARCO Oil and Gas Research, 2300 West Plano Parkway, Plano, Texas 75075, USA
Current address: 7140 Crooked Oak Drive, Dallas, Texas 75248-2231, USA
(received November 4, 1988; revision accepted March 6, 1989)
ABSTRACT
Synthems are high-rank stratigraphic units of sedimentary, igneous and metamorphic rock bounded by
interregional unconformities (Chang, 1975; ISSC, 1987). When synthems relate to large-scale tectonic events,
such as plate reorganizations, they are tectono synthems (Gallagher and Tauvers, 1992). Such units encom-
pass tectonic and sedimentary processes which occur on the same regional scale and which correlate with
similar unconformity-bounded units on other cratons.
The rifted basins of eastern India record the effects of plate separation and collision. Withjack and
Gallagher (1983) define three tectonostratigraphic units for the rifted basins of eastern India. At least one of
these, Tectonostratigraphic Unit 2, is also a tectonosynthem. Unit 2 is bounded above and below by regional
unconformities that are in the Cauvery, Palar, Godavri-Krishna, Mahanadi-Brahmani, and West Bengal
Basins. The lower unconformity is synchronous with Late Jurassic to Early Cretaceous continental breakup
as India broke away from Antarctica and Australia. The lower strata of Unit 2 are predominantly clastics
related to rifting. These strata are deposited unconformably on Precambrian crystalline basement or on
Paleozoic and Mesozoic sedimentary rocks of Tectonostratigraphic Unit 1.
The Rajmahal traps of the West Bengal Basin are more than 600-m-thick tholeiitic basalt flows with K-Ar
dates of about 100 Ma, corresponding to the onset of seafloor spreading (McDougall and McElhinny, 1970;
Withjack and Gallagher, 1983). Younger rocks in Unit 2 are predominantly marine, reflecting trailing
margin subsidence as India moved northward away from Australia and Antarctica and toward Asia. Unit 2 is
cut by normal faults related to rifting and continental margin subsidence. Tectonostratigraphic Unit 3
unconformably overlies Unit 2; this unconformity does not coincide precisely with a major drop in sea level
(Curray, 1984). Consequently, the Unit 2/Unit 3 unconformity could be the result of the Eocene collision of
India with Asia. During the collision, faults in Unit 2 were reactivated and accompanied by gentle folding.
The tectonosynthem represented by Unit 2 is a relatively simple example of the tectonosynthem concept,
which records rifting and drifting on a Mesozoic plate margin, followed by Tertiary collision. The collision
zone is 1000 km to the North and yet provides an overprint on previous structural and stratigraphic features.
INTRODUCTION
A common approach to the interpretation of
sequence stratigraphy is "seismic stratigraphy,"
which presumes that unconformities and inter-
vening strata are the product of sea-level rises and
falls (Vail and others, 1977). An alternative ap-
proach is suggested here: regional-unconformity-
bounded stratigraphic units and related structures
present in several basins of eastern India relate to
plate-tectonic events.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 591-597. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
591
592
J. J. Gallagher, Jr.
Before applying this approach to the eastern
India basins, a brief discussion of the tectono-
synthem concept is in order. It must be stated at the
outset that there is no intention to discredit in any
way the seismic stratigraphy approach, which is
accepted here as valid. The validity is restricted to
certain conditions - those of a passive continental
margin with moderate sediment load. The intention
here is to show that a single unconformity-bounded
unit on one continental margin is determined not by
sea-level fluctuations alone but by a sequence of
low-frequency tectonic events superimposed on the
subtler, higher frequency effects of sea-level rises
and falls.
SYNTHEM DISTINCTIONS
Synthems are major stratigraphic units of ig-
neous, metamorphic, and sedimentary rock bounded
by regional unconformities (Chang, 1975; ISG,
1976). The International Subcommittee on Strati-
graphic Classification (ISSC, 1987) restricts use by
stating that the term "synthem" could include no
suggestion about cause such as tectonics, or eustasy.
Gallagher and Tauvers (1992) described synthems
in northwestern South America which they related
to tectonic events. Honoring the intent of the ISSC
(1987), they used a causal prefix and call the units
tectonosynthems.
Similarly, synthems with other causes would
have an appropriate prefix - for example, eustati-
synthems. Major differences are expected to occur
which would help distinguish eustatisynthems from
tectonosynthems in the geologic record (Table 1).
Because sea-level changes happen very rapidly rela-
tive to tectonic events, and they affect all oceans and
continental margins at the same time, eustati-
synthems are in principle bounded by global uncon-
formities. Transgressions take time to occur, but
regressions are not shown to be time-transgressive
by Vail and others, (1977, p. 85). The time taken by
a transgression is short relative to the duration of a
tectonic event, such as subsidence of the margins of
a drifting continental plate or mountain building
along a plate collision zone.
Eustatisynthems contain similar sedimentary
transgressive or regressive sequences of rocks for a
given time interval. Provenance, climate, and other
variables affecting marine and marginal marine
deposition control differences of passive-margin,
moderate sediment-load eustatisynthems. There
are other possible settings for eustatisynthems, such
as active margins, but these have not been widely
studied and reported. If the eustatisynthem
sequence stratigraphy paradigm is to be applied to
nonmarine basins, it is necessary to first establish
how sea-level changes can affect nonmarine ero-
sional and depositional processes.
In contrast, tectonosynthems are marked by
regional to interregional (but not global) time-
transgressive and complex unconformities. Rock
sequences need not be similar from place to place,
and they typically include sedimentary, igneous,
and metamorphic rocks, whereas eustatisynthems
are almost exclusively sedimentary rocks. Further-
more, tectonosynthem sediments may be marine or
nonmarine in origm. Seismic stratigraphic se-
quences are defmed in terms of sea level and are
therefore marine (e.g., Vail and others, 1977).
Clearly, tectonosynthems on continental mar-
gins are combined in whole or in part with one or
more eustatisynthems. The nature of the rock
types, scale (whether global or regional), time trans-
gressiveness of unconformities, and time-space rela-
tions to known tectonic events all help distinguish
Table 1. Comparision of two types of synthems
Setting Duration Scale Rock Type Unconformities
Eustati- Passive Short Global Well-defined Not time-transgressive regression
synthem margin sequences of marine Relatively systematic
and marginal marine Erosional and depositional
sedimentary rocks
Tectono- All Long Regional Igneous, meta- Time-transgressive
synthem tectonic morphic, and Relatively complex
settings sedimentary rocks Erosional, depositional
Angular, tectonic emplacement
Tectonics and regional unconformities of eastern India basins 593
Table 2. Generalized correlation of tectonic events with unconformity-bounded units for the basins of eastern India.
Geological Rock Uncon- Tectono-
Age Plate-Tectonic Event UnitlType formity Structure synthem
Pleistocene Tectonic subsidence Shallowing marine Basement
detached
extensional TS3
faulting
Early Eocene Minor
Collision ofIndia with Asia Keratu Traps Regional
Paleocene Carbonates Minor Reactivated
Fine Clastics horsts and TS2
Coarse Clastics grabens
Late Cretaceous Drifting Deepening marine
Onset of sea-floor spreading Rajmahal traps Regional
Early Cretaceous Shallow marine
Continental clastics TSI
Early Jurassic Gondwana Series
Rifting Regional Onset of
horstand
graben faulting
Archean Igneous and
metamorphic complex
between the two synthem types. These characteris-
tics are the basic criteria identifying tectono-
synthems in the basins of eastern India.
EAST INDIAN BASINS
The three tectonostratigraphic units of With-
jack and Gallagher (1983) form the basis for defini-
tion of tectonosynthems in the basins of eastern
India. These authors used well and seismic data and
defined the units by stratigraphic sequences and
structures, but they did not consider unconformities.
The term "tectonosynthem" is broader in that it
includes the strata and the structure but, in
addition, includes the bounding unconformities.
These unconformities occur in the wells reported by
Cantwell and others (1978), Curray and others
(1982), and Curray (1984), and are seen in the
seismic sections (Withjack and Gallagher, 1983, p.
45-46). Two tectonosynthems are defined by these
unconformities: an older one, TS1, and a younger
one, TS2 (Table 2). If the present-day land and sea-
floor surfaces are considered to be unconformities,
then TS3 overlies TS2 and TSl.
This paper concentrates on TS2, which occurs in
the basins shown in Figure 1-the Cauvery, Palar,
Godavari-Krishna, Mahanadi-Brahmani, and West
Bengal Basins (Withjack and Gallagher, 1983).
Only the onshore extent of these basins is shown in
Figure 1; each of these basins extends offshore as
well.
The lower unconformity of TS2 typically sepa-
rates Early from Late Cretaceous rocks (Table 2).
There is agreement in the literature that this is a
regional unconformity and represents one of the
greatest hiatuses within the region (Cantwell and
others, 1978; Curray and others, 1982; Curray,
1984). The unconformity corresponds in time with a
major sea-level drop during Cenomanian time (Vail
and others, 1977). However, Curray and others
(1982) showed that magnetic anomalies in the sea
594 J. J. Gallagher, Jr.
o 400 KM

Figure 1. TS2 outcrop locations (after Withjack and
Gallagher, 1983). Cross-section A-B given in Figure 2.
floor and DSDP sedimentary data both suggest that
India was drifting away from Antarctica and
Australia by this time. Other evidence that the
unconformity is related to extensional tectonism is
suggested by tholeiitic basalt flows in the igneous
Rajmahal Traps of the West Bengal Basin with
Albian K-Ar age determinations (Withjack and
Gallagher, 1983). TS2 strata are cut by normal
faults that were active beginning in the Late Cre-
taceous, which supports the tectonic nature of this
unconformity. Thus the lower bounding unconfor-
A
Darn 11 Pass/s/ 11
PSlk Bay 11
NW
-9- -9-
mity of TS2 is a strong overprint of tectonic in-
fluence on a eustatic event.
The following discussion emphasizes data from
the Cauvery Basin. References in Withjack and
Gallagher (1983) provide data of a similar type for
the other eastern India basins, which show that the
structures, unconformities, and stratigraphic se-
quences in those basins relate to global tectonic
events in much the same way as in the Cauvery
Basin.
There is a marked depositional change across
the lower bounding unconformity of TS2. This
change is seen in the Delft-l well (Figure 2) in
which the TSI rocks beneath the unconformity are
Early Cretaceous continental and supratidal clas-
tics and carbonates, including reefs at the top.
Rocks above the unconformity are early Oligocene
clastics. The contrast across this unconformity is
also typified by the Pesalai-1 well in which Late
Cretaceous massive sandstone unconformably over-
lies Early Cretaceous claystone and siltstone.
TSI is missing on some basement highs, and
TS2 lies directly on an Archean igneous and meta-
morphic complex (Figure 2). In sharp contrast, the
rocks above the unconformity are Late Cretaceous
clastics. Coarse sandstone and conglomerate units
at the base of the TS2 sequence are shore zone to
inner shelf but rapidly grade upward into fine
clastic and carbonate rocks deposited in shelf and
slope environments. At the same time, India drifted
away from Antarctica and Australia (Figure 3).
This TSIITS2 unconformity and the contrasting
lithologies are observed in both wells mentioned
above and in the Palk Bay-l well, which is about
halfway between the other two (Figure 2).
B
SE
Figure 2. Schematic structural styles based on seismic and well data in the Cauvery Basin (modified from Withjack and
Gallagher,1983): TS1 white; TS2 slanted pattern; TS3 stippled pattern. Wells are highly schematic; the Delft well is 50 km
NW and the Pesalai well is 30 km SW of the Palk Bay well (well data from Cantwell and others, 1978).
Tectonics and regional unconformities of eastern India basins 595
A
130
o
53
Tethys
Sea


B
105
E
<53
>32

Figure 3. Plate-tectonic schematic for India and adjacent plates (modified from Curray and others, 1982; Withjack and
Gallagher, 1983). Numbers are millions of years before present. Double lines represent an active spreading ridge; single lines
denote a transform; sawteeth show subduction; asterisks show a volcanic arc; dashed lines are inferred and dotted lines are
precursor plate boundaries. TS1 basement faulting began by rifting (A) and continued by drifting (B). The TSlITS2
unconformity (Table 2) is between Band C; the TS2/TS3 unconformity (Table 2) is between C and D. Major fault reactivation,
sediment influx, and the TS2/TS3 unconformity resulted from collision (D, E, F).
The well data cited above reveal a systematic
change of sedimentary lithologies and environments
of deposition between the unconformities bounding
TS2 in the Cauvery Basin. There are two parts to
this systematic change. The first part begins at the
base with nonmarine to shallow marine clastics,
changing rapidly upward to deeper marine clay-
stones, siltstones, and carbonates. These sediments
were deposited on a subsiding margin as India
drifted away from Antarctica and Australia (Figure
3C). The second part, continuous with and over-
lying the first part, is shallowing upward fossil i-
ferous limestones. These sediments span the entire
Late Cretaceous and Paleocene in age. They were
deposited as India continued to drift and as tectonic
subsidence slowed, permitting basins to begin filling
up with sediment (Figure 3C).
The lower unconformity ofTS2 affected rocks as
old as Precambrian to as young as Oligocene and
possibly younger, but it is typically Early Creta-
ceous on Precambrian. The upper unconformity
ranges in age from oldest Late Cretaceous to Mio-
cene, but it is most commonly Eocene on Paleocene
or Eocene on latest Cretaceous.
596
J. J. Gallagher, Jr.
In the Palk Bay well, the upper bounding un-
conformity of TS2 is basal Eocene (Ypresian).
According to Curray (1984), this unconformity does
not correspond precisely with either of the two
regressions described for Ypresian by Vail and
others (1977), although it is generally regarded as a
major unconformity. Its tectonic origin is suggested
by events happening concurrently. According to
Curray and others (1982), India was colliding with
Asia (Figure 3D) and, simultaneously, the Paleo-
cene igneous Keratu Traps were extruding in the
Godavari-Krishna Basin. In addition, the direction
of sea-floor spreading was changing from N-S
(Figure 3D) to NE-SW (Figure 3E).
Withjack and Gallagher (1983) show that the
style of faulting changed from basement-involved
normal faulting in TS1 and TS2 to basement-
detached listric normal faulting in TS3 (Figure 2).
This reflects the change from rift-related faulting
(Figure 3A) and reactivation on the faults due to
intense early subsidence as drifting began (Figure
3B) to non-basement-involved extension. The ex-
tensional strain is related to sediment compaction
and gravity sliding from basin edges toward basin
centers as drift continued, as subsidence slowed, and
as basins began to fill (Figure 3C). Withjack and
Gallagher (1983) showed that above the uncon-
formity in TS3 there are folds associated with listric
faults and that there are folds over the gently re-
activated basement-involved faults. These are asso-
ciated with the collision of India with the continent
to the north (Figure 3E,F).
DISCUSSION
The perhaps cumbersome new term "tectono-
synthem" is necessary in this case because existing
terms are either poorly defined or have connotations
that are problematical. Still other terms would be
inappropriate. The term "tectonostratigraphic unit"
is poorly defined - i.e., "melange" (Bates and
Jackson, 1980) is loosely used in the literature and
is applied over a wide range of scales, from that of
syntectonic conglomerates associated with an indi-
vidual thrust sheet to the scale of continental mar-
gins or grander scales. The suffix "stratigraphic"
seems, by definition, to exclude unconformities,
whereas "synthem" is defined in terms of unconfor-
mities (lSG, 1976; ISSC, 1987). "Tectonosynthem" is
meant to be applied to all rock types bounded by
regional unconformities. Both the strata and the
unconformities in tectonosynthems are primarily
consequences of tectonics but, in addition, may be
influenced by subordinate causes such as eustasy.
The term "tectonosynthem" is intended to
emphasize the prefix "tectonic" as distinguished in
scale from structures - that is, the prefix denotes
the plate-tectonic scale. In the case of the basins of
eastern India, this is the scale of all the rifts on the
eastern India margin, taken together, which are
affected by plate-tectonic scale events. The prefix is
also meant to distinguish seismic expressions of
strata and unconformities related to tectonic causes
from those strata and unconformities related to
other causes.
Tectonosynthem 2 (TS2) in the basins of eastern
India may be a good example of an application of
this concept. The unconformities bounding the
strata are regional unconformities and occur in
basins of eastern India from the Bengal in the north
to the Cauvery in the south. The strata include both
igneous and sedimentary rocks. The unconformities
can be related to tectonic events - specifically to
the break-up of India from adjacent plates in the
case of the lower unconformity, and to the collision
of India with Asia for the upper unconformity. The
sedimentary strata are rift-related and primarily
clastic beneath the lower unconformity. Between
the unconformities bounding TS2, the strata are
drift-related and record rapid water-deepening up-
ward, followed by slowly shallowing-upward marine
strata. The igneous strata are related to extensional
crustal fracturing at the plate scale, which led to
rift-related extrusion of the Rajmahal and Keratu
Trap rock.
One problem encountered in relating unconfor-
mities to tectonic events is that well-developed cri-
teria to distinguish eustatic unconformities from
tectonic unconformities have not been defined.
Table 1 is an attempt, perhaps primitive, to define
those criteria used in this study of the basins of
eastern India. We hope that it will serve as a start-
ing point for better definition of the criteria.
CONCLUSION
TS2 can be observed on seismic sections and is
recorded in wells of the basins of eastern India. The
unconformities and strata of TS2 are a result of
plate separation, plate drift and plate collision.
These observations can be used to apply the tec-
tonosynthem concept to seismic and well data from
Tectonics and regional unconformities of eastern India basins 597
other parts of the world. It is possible that some
regional unconformity-bounded sequences inter-
preted as totally eustatic may instead be partly or
dominantly, or perhaps totally, tectonic. Separating
these two causes as seismic interpretations are
made for other basins should improve the quality of
interpretations as it has for the basins of eastern
India.
REFERENCES
BATES, R. L., and J. A. JACKSON (eds.), 1980, Glossary of
Geology, Second Edition: American Geological Insti-
tute, Falls Church, Virginia, USA, 749 p.
CANTWELL, T., T. E. BROWN, and D. G. MATHEWS, 1978,
Petroleum Geology of the Northwest Offshore Area of
Sri Lanka: SEAPAX: Proceedings of the _ Offshore
South East Asia Conference, 12 p.
CHANG, K. H., 1975, Unconformity-bounded stratigraphic
units: Geological Society of America Bulletin, v. 86, 1p.
544-1552.
CURRAY, J. R., 1984. Sri Lanka: Is it a mid-plate plate-
let?: Journal of the National Aquatic Resources Angency
(Sri Lanka), v. 31, p. 30-50.
__ , F. E. EMMEL, D. G. MOORE, and R. W. RAITT, 1982,
Structure, tectonics and geological history of the northeast
Indian Ocean; pp. 399-450 in A. E. M. Nairn and F. G.
Stehli (eds.), The Ocean Basins and Margins, Volume
6: The Indian Ocean: Plenum Press, New York, New
York, USA, 776 p.
GALLAGHER, J. J., Jr. and P. R. TAUVERS, 1992, The
tectonic evolution of northwestern South America; pp. 123-
137 in R. Mason (ed.), Proceedings of the 7th Interna-
tional Conference on Basement Tectonics (Kingston,
Ontario, 1984): Kluwer Academic Publishers, Dordrecht,
The Netherlands, 480 p.
ISG (International Stratigraphic Guide), 1976, A Guide to
Stratigraphic Classification, Terminology and Proce-
dure: International Union of Geological Sciences,
Commission on Stratigraphy, International Subcom-
mission on Stratigraphic Classification: Wiley, New
York, New York, USA, 200p.
ISSC (International Subcommission on Stratigraphic
Classification), 1987, Unconformity-bounded stratigraphic
units: Geological Society of America Bulletin, v. 98, p. 232-
257.
McDOUGALL, 1., and M. W. McELHINNY, 1970, The
Rajmahal Traps of India - K-Ar ages and paleomagne-
tism: Earth and Planetary Science Letters, v. 9, p. 371-378,
1970.
VAIL, P. R., R. M. MITCHUM, Jr., and S. THOMPSON III,
Seismic stratigraphy and global changes of sea level, Part
4: Global cycles of relative changes of sea level; pp. 83-97
in C. E. Payton (ed.), Seismic Stratigraphy - Appli-
cations to Hydrocarbon Exploration: American Asso-
ciation of Petroleum Geologists, Memoir 26,516 p.
WITHJACK, M. 0., and J. J. GALLAGHER, Jr., 1983, The
rifted basins of eastern India: SEAPEX: Proceedings of
the Offshore South East Asia Conference (Singapore), v. 6,
p.41-57.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Basement-cover relationships of the Peninsular Gneiss with the
greenstone belts of Karnataka, India
A.G.UGARKAR
Department of Geology, University of Po on a, Pune 411007, India
Current address: Department of Geology, Karnatak University, Dharwad 580003, India
(received and accepted November 14, 1988)
EXTENDED ABSTRACT
The Archaean terrain of Karnataka is com-
posed of greenstone belts and high-grade schists set
in a sea-of-gneiss complex (Peninsular Gneiss). The
existence of several greenstone belts and continuous
transitions between gneiss and greenstone terrains
in Karnataka provide unique insight into the
evolution of the Archaean crust. The older mafic
and ultramafic enclaves in the Peninsular Gneiss
probably represent relicts partially of early oceanic
crust and partially of reworked continental crust.
Over this polycyclic gneiss complex were deposited
supercrustal rocks in small linear basins, which are
now represented by high-grade schists. A major
tectonothermal event (Pan tectogenesis) around
3000 Ma ensued throughout the Karnataka Craton,
and involved the high-grade schist and gneiss
complex. Subsequent to the Pantectogenesis, green-
stone belts evolved on a basement of the gneissic
complex. Under a relatively unstable regime, the
older greenstones (i.e., Keewatin type of eugeo-
synclinal greenstone belts - Kolar, Hutti, and
Mangalur belts) comparable to the Archaean green-
stone belts evolved. These are also widely known as
the Kolar-type greenstone belts. A more stable
regime witnessed the evolution of younger green-
stones (Dharwar type of greenstone belts), the
Chitradurga Group and the Bababudan Group,
which are similar to the Early Proterozoic basins
and geosynclines. The Kolar-type greenstone belts
are mainly metavolcanic belts with predominant
mafic to andesitic volcanic rocks, subordinate acid
volcanic rocks, ultramafites, graphitic and sulfidic
schist, ironstones, cherts, and immature sediments
like wackes and polymictic conglomerates. The
Dharwar-type greenstone belts are predominately
meta volcano-sedimentary belts, with shallow water
"shelf facies" sediments at the base. The metavol-
canic rocks of the older greenstone belts are geo-
chemically comparable to oceanic tholeiites. A
marginal-basin tectonic setting is the most favour-
able environment for their formation. Among the
Dharwar-type greenstone belts, the Chitradurga
Group metavolcanic rocks are marked by ocean-
floor basalts followed by a volcanic-arc series made
up of low-K tholeiites and calcalkaline basalts
formed in a submarine tectonic environment. The
Bababudan Group metavolcanic rocks are of tho-
leiitic to calc alkaline affinity formed in a shallow
basin stable tectonic environment.
Basement-cover relations are clear in Dharwar-
type greenstone belts because: (1) there are clear-
cut unconformities between basement gneisses and
overlying conglomerate-quartzites, and (2) green-
stones unconformably rest on a gneiss basement. In
Kolar-type greenstone belts, the contacts are ob-
scured either by massive invasion of later granites
or by soil cover. There is neither evidence of an
oceanic crust nor evidence to infer that the Peninsu-
lar Gneiss forms the basement in Kolar-type green-
stone belts, although within these greenstone belts,
the presence of a conglomerate horizon, with well-
rounded pebbles of gneiss of granodioritic composi-
tion, probably suggests a pre-existing sialic crust.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), p. 599. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
599
CHAPTER 6
METALLOGENY AND
CONTINENTAL MARGINS
INTERNATION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Earthquake faulting, induced fluid flow, and fault-hosted
gold-quartz mineralization
RICHARD H. SIBSON
Department of Geological Sciences, University of California, Santa Barbara, California 93106, USA
Current Address: Department of Geology, University of Otago, P.O. Box 56, Dunedin, New Zealand
(received August 12, 1988; revision accepted May 1, 1989)
ABSTRACT
Gold-quartz vein systems with textures recording incremental histories of deposition are often hosted in
faults and shear zones. Epithermal deposits tend to develop at shallow depths 1-2 km) in the vicinity of
extensional or transtensional fault systems. In contrast, many mesothermal gold-quartz lodes are associated
with high-angle reverse or reverse-oblique shear zones of mixed brittle-ductile character, developed under
greenschist-facies metamorphic conq.itions at "" 10 km depth. These mesothermal deposits apparently repre-
sent the roots of brittle fault zones.
Although the mechanisms differ in detail, abrupt fluctuations in fluid pressure induced by earthquake
faulting seem likely to play a key role effecting mineral precipitation in both of these depositional
environments, which bracket the continental seismogenic regime. Within this regime, slip transfer across
dilational fault jogs and bends during rupture propagation leads to abrupt local reductions in fluid pressure
below ambient (hydrostatic?) values. At high crustal levels in geothermal systems, such pressure reductions
may induce boiling, triggering mineral deposition throughout the phase of aftershock activity. Dilational
fault irregularities thus act essentially as suction pumps.
In the mesothermal environment at the base of the seismogenic regime, high-angle reverse faults playa
different role as fluid-pressure-activated valves, capping over-pressured metamorphic and/or magmatic
systems and promoting cyclic fluctuations between litho static and hydrostatic levels of fluid pressure.
Frictional shear failure can only initiate when supralithostatic fluid pressures are attained. Rupturing
through the upper crust then allows sudden drainage of the over-pressured reservoir along the reactivated
fault, with fluid pressures dropping rapidly toward hydrostatic values. Discharge is accompanied by phase
separation of CO
2
, rapid mineral precipitation, and hydrothermal self-sealing. Fluid pressures then rebuild
toward the supralithostatic values needed to trigger the next episode offault slip, and the cycle repeats.
INTRODUCTION
Faults and allied fracture systems have long
been recognized as important hosts for hydro-
thermal mineralization at different structural levels
within the crust. Early workers (e.g., Buckland,
1836; Hulin, 1925; Knopf, 1929; McKinstry, 1948)
paid considerable attention to the textural charac-
teristics of hydrothermal cements in such deposits,
often recognizing evidence for the incremental
passage of hydrothermal fluids accompanying epi-
sodes of fault slip. For example, Newhouse (1942)
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 603-614. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
603
604 R. H. Sibson
noted: " ... epigenetic ore deposits ... appear rarely to
have been introduced into a dead or static structural
feature ... ," and went on to state: " ... evidence for
repeated faulting or intermineralization fracturing
is widespread." This raises the interesting question
as to whether episodic faulting drives intermittent
fluid flow by some mechanism or whether the cyclic
build-up of fluid pressure induces successive
faulting events (Sibson, 1981).
Possible interrelationships between faulting
and fluid flow are reviewed in this paper in terms of
simple models developed from our present under-
standing of the mechanics and depth distribution of
earthquake faulting in seismogenic crust. Empha-
sis is placed on the role of dynamic fault processes in
triggering episodes of hydrothermal precipitation by
provoking sudden local reductions in fluid pressure
at specific structural sites. Episodes of rapid
hydrothermal deposition are then induced through
boiling (phase separation). One mechanism, the
"suction-pump" model, appears dominant in the
epithermal environment; the alternative "fault-
valve" model plays an especially important role in
mesothermal, shear-zone-hosted mineralization.
CONTINENTAL SEISMOGENIC REGIME
Modern seismologic techniques have demon-
strated that seismicity in actively deforming con-
tinents away from areas of active subduction is
largely confined to the upper half of the crust, and
also that within this seismogenic regime, displace-
ments are mostly accomplished by seismic slip
increments on existing faults (Chen and Molnar,
1983; Sibson, 1983). In crust with moderate to high
heat flow (e.g., San Andreas fault system, Basin and
Range), background microearthquake activity is
largely restricted to the top 10-15 km. Larger
shocks (M > 5.5) tend to nucleate toward the base of
the seismogenic regime. The ruptures expand most-
ly upward and laterally over fault surfaces from the
nucleation site at rates approaching the shear-wave
velocity ("'=' 3 km/s), so that aftershocks are largely
restricted to the same zone (Figure 1). The maxi-
mum depth of seismic activity deepens in colder
cratonic crust but shallows locally to as little as 4-5
km in regions of active geothermal fluid upwelling.
Examination of the style and rock products of
shearing in ancient fault zones exhumed from differ-
ent crustal levels provides a geologic basis for inter-
preting these variations in the depth of seismic
activity. For quartzofeldspathic crust, a reconstruc-
ted depth sequence of dominant fault-rock types
within localized shear zones is:
mylonite
mylonitic gneiss.
The cataclasite-mylonite transition represents a
major change in the mechanical response of the
crust, with crystal plastic flow of quartz in mylo-
nites becoming significant at temperatures >300-
350C, corresponding to the onset of greenschist-
facies metamorphic conditions (Sibson, 1977). The
base of the seismogenic zone thus represents the
transition from unstable frictional (FR) behavior,
where shear resistance is dominantly pressure
dependent and increases with depth, to quasi-plastic
(QP), aseismic shearing in mylonite belts where
flow-shear resistance is temperature dependent and
decreases with depth. Peak shear resistance there-
fore occurs in the vicinity of the FR/QP transition,
the zone of large rupture nucleation, where mixed
continuous and discontinuous deformation must
occur over a large range of strain rates.
In summary, the seismogenic regime appears to
represent a zone of unstable frictional behavior, the
base of which is defined by the onset of greenschist-
facies metamorphic conditions at temperatures of
"'='300
0
-350C. Within this regime, in the vicinity of
major active faults, the accumulation of elastic
strain during interseismic periods and its sudden
release during earthquake rupturing must give rise
to a sawtooth oscillation in shear stress with an
amplitude of the order of 1-10 MPa - the range of
stress drops characterizing shallow crustal earth-
quakes (Figure 2).
FAULT-HOSTED MINERALIZATION
An important distinction exists between the
structural settings of fault- or shear-zone-hosted
gold-quartz deposits developed at different crustal
levels. Near-surface, epithermal deposits are al-
most invariably hosted by normal or strike-slip fault
systems developed in extensional or transtensional
tectonic regimes. Tectonic settings of mesothermal
gold-quartz lodes, which are inferred to have formed
at considerable depth (Roberts, 1987), are in distinct
contrast. These deposits tend to be hosted in high-
angle reverse or reverse-oblique shear zones of
mixed "brittle-ductile" character which were active
within compressional or transpressional regimes.
Earthquake faulting, induced fluid flow, and fault-hosted gold-quartz mineralization 605
W/////////////////////////////////////////////////////////////! ////////.
SHEAR
RESI STANCE
P
I
P
f
.
IU l d
P
h
Epit hermal
Epithermal Environment
FQ EQ
Suct ion Pump
TIME
Though replacement is undoubtedly signficant
in some instances, much hydrothermal deposition in
the epithermal environment occurs in cavity space
formed incrementally by episodic slip transfer
across fault irregularities, or in associated arrays of
extension fractures. I confine attention here to the
problem of incremental slip transfer around irregu-
larities transverse to the direction of rupture pro-
pagation on faults that are either purely dip-slip or
strike-slip, bearing in mind that the full three-
dimensional geometry, especially for oblique-slip
faulting or in the case of complex fault intersections,
Figure 1. Strike-parallel profile along a fault
surface illustrating propagation of a strike-slip
rupture within the seismogenic regime. FR and
QP denote, respectively, the frictional and quasi-
plastic fault regimes.
Figure 2. Local oscillations of fluid pressure and
shear stress ("L) within the seismogenic regime for
suction-pump and fault-valve mechanisms in
relation to successive earthquake ruptures (EQ)
(after Sibson and others, 1988). Ph and PI repre-
sent hydrostatic and lithostatic fluid pressures,
respectively. Note that the restoration of fluid
pressure to lithostatic values for the valve mech-
anism and the reaccumulation of shear stress
between successive failure events are not neces-
sarily simply related to each other as shown.
may be much more complex (cf. Bruhn and Parry,
1988; King, 1983).
Observations from Modern Earthquake
Rupturing
Recognition, from paleoseismic studies, that
segments of major active faults often exhibit charac-
teristic earthquake behavior (Schwartz and Copper-
smith, 1984) - that is, ruptures of about the same
size tend to repeat along particular fault segments
at fairly regular intervals - has recently focused
attention on the role of fault irregularities in rup-
606
R. H. Sibson
ture nucleation and arrest. Irregularities trans-
verse to the direction of rupture propagation are
recognizable in map view for strike-slip fault sys-
tems, allowing ready correlation with aftershock
epicentral distributions (Sibson, 1986). Apart from
complex fault intersections, two main types of
irregularity are discernible: (1) fault jogs linking en
echelon fault segments, and (2) isolated fault bends.
Each may be either dilational or antidilational,
depending on the tendency for areal increase or
reduction, respectively, accompanying incremental
slip transfer (Figure 3). Cross-strike dimensions of
the stepover regions in fault jogs may range up to 10
km or so. Precision aftershock studies suggest that
in at least some instances, these irregularities per-
sist through the seismogenic zone to depths ap-
proaching 10 km (e.g., Reasenberg and Ellsworth,
1982). Note that an important directivity effect is
associated with isolated fault bends. Incremental
slip transfer around a bend to a differently oriented
fault segment may involve dilation or enhanced
compression across the segment, depending on the
direction of rupture propagation into the bend
(Sibson, 1989).
The simple message from recent, well-studied
strike-slip earthquakes is that larger dilational
irregularities (cross-strike widths > 1 km) act as
preferential sites for rupture termination. Rupture
arrest is often followed by delayed slip transfer
across the structures, with a concentration of after-
shock activity in the local area of dilation (Sibson,
1986, 1987). Somewhat surprisingly, therefore, rup-
tures are terminating where slip transfer involves
local extensional opening at jogs or bends. This
contradicts expectations from quasi-static stress
analyses, which suggest that antidilational jogs and
restraining bends should form the most serious
impediments to slip transfer (Segall and Pollard,
1980).
Internal Structure of Dilational Fault Jogs
The mining literature (e.g., Newhouse, 1942;
McKinstry, 1948) provides many examples of differ-
ent types of dilational fault jogs (Figure 4). At high
crustal levels, jogs tend to be characterized by
crustiform fissure veins with textures recording his-
tories of incremental extensional opening, or by
high-dilation, hydrothermally cemented breccias of
wall-rock fragments resulting from multiple epi-
sodes of brecciation and cementation. Ladder-vein
arrays, cymoid loops, and horsetail veining may all
be viewed as different varieties of dilational jog ac-
commodating localized extension. Larger jogs may
consist of a mesh of extensional fractures and sub-
sidiary shears (Hill, 1977). In such cases, it may be
rather hard to envisage the form or size of the struc-
ture as a whole when only part is exposed. One im-
portant common characteristic is that away from
the jog the driving faults themselves often remain
largely barren. This has often led to a false infer-
ence that vein systems and faults are of different
ages.
The analyses of Segall and Pollard (1980) show
that local perturbations of the stress field should
favor extensional fracturing within the stepover
region of dilational jogs independent of depth. How-
ever, it seems clear that massive hydrothermal
deposition is restricted to the uppermost kilometer
or so. Two factors may be important here. First,
sustained boiling within geothermal systems can
occur only within the upper 1-2 km and is likely to
be most effective as a precipitation mechanism in
the top 500 m (Cathles, 1981; Henley, 1985). Se-
cond, in stress fields associated with normal and
strike-slip faults, vertical extensional veins may
form perpendicular to the least principal compres-
sive stress (03) by hydraulic fracturing under hydro-
static fluid pressures within perhaps a kilometer or
so of the surface, without any requirement of "ab-
normal" fluid pressure conditions (Sibson, 1981). In
contrast, supra litho static fluid pressures are gener-
ally required at all levels for the formation of sub-
horizontal extension veins in the vicinity of thrusts.
"Suction-Pump" Mechanism
From the internal structure of dilational jogs, it
is apparent that slip transfer across the structures
requires concomitant opening of the linking exten-
sion fractures. The arrest of propagating ruptures
at dilational fault jogs and bends, and the ensuing
time-dependent slip transfer, are both attributed to
the difficulties of opening such a linking extensional
fracture mesh within fluid-saturated crust in times
comparable to earthquake slip durations of a few
seconds (Sibson, 1985a). Forced opening of the ex-
tensional mesh at rates faster than fluids can flow in
to restore equilibrium leads to a fluid-pressure im-
balance and to a suctional force opposing the slip
transfer, which scales with the width of the jog.
JOGS
8E OS
Earthquake faulting, induced fluid flow, and fault-hosted gold-quartz mineralization 607
DI LATIONAL
RNTIDIL AT IO AL



. #'---
0;;;:-

0;;;:-


'"
/
0;;;:-

,/

0;;;:-
Ladder Vein Cymoid Loop Horsetail
Alternale Shear/ Exlens ional Mesh Mode l s
For Di lalional Jogs
(al Pre- rupt ure: uniform fluid pressur es
\

'--
,
(b) Post-rupture: fluid influx into jogs
--... ..
t
Perturbation
Figure 3. Seismotectonic sketchs of
dilational and antidilational jogs and
bends on a right-lateral strike-slip fault .
Solid circles represent epicenters, arrow-
heads the direction of rupture propa-
gation (after Sibson, 1989).
Figure 4. Varieties of dilational fault
jog: ladder-vein arrays, cymoid loops,
horsetails, and composite shear-exten-
sional mesh modes. Diagram represents
map view for strike-slip faults, cross-
sectional view for normal faults. Faults
represented by bold lines, extension veins
by cross-hatching.
Figure 5. Schematic representation of
dilational fault jogs (idealized as single
extension fractures) causing rupture
perturbation and arrest along a dextral
strike-slip fault. Rupture propagation
from left to right (after Sibson, 1987).
608
R. H. Sibson
The process of rupture perturbation and arrest
at small- and large-scale dilational jogs is illustra-
ted schematically (Figure 5). The passage of a
moderate or larger rupture will be only slightly
impeded by a small dilational jog (width 1 km),
but for a larger jog (width> 1 km), the suction in-
duced by incipient opening of the fracture mesh may
be sufficiently large for the jog to act as a major
energy sink and inhibit further rupture propaga-
tion. Following rupture arrest, the lowered fluid
pressure within the jog is restored by inward fluid
percolation from the surrounding rock. This allows
time-dependent opening of the extensional fracture
mesh and delayed slip-transfer across the jog, ac-
companied by aftershock activity, as residual strain
energy around the rupture tip is slowly released.
The suction-pump mechanism accounts for
many commonly observed features of fault-hosted
epithermal deposits. For example, at depths within
a dilational jog greater than a few hundred meters,
the induced fluid-pressure differential between an
extension fracture and its immediate surroundings
should be sufficiently large for wallrock brecciation
to occur by hydraulic implosion into the fast-formed
cavity. Downward passage of crustiform exten-
sional fissure veins into multiply recemented wall-
rock breccias has, in fact, been noted in descriptions
of some epithermal deposits (Wisser, 1941). More-
over, delayed slip-transfer and extensional opening
within the jog provide a plausible explanation for
the high-dilation "floating clast" breccias that char-
acterize such structural settings. However, in some
circumstances it may be difficult to distinguish
these fault-related implosion breccias from breccias
of magmatic-hydrothermal origin of the kind des-
cribed by Sillitoe (1985).
Pumps as a Precipitation Trigger
Active geothermal systems are commonly asso-
ciated with areas of active extensional or transten-
sional faulting. Measured fluid-pressure gradients
within the top kilometer or so of such areas, believed
equivalent to ancient epithermal mineralizing envi-
ronments, generally stay fairly close to hydrostatic
(Henley, 1985). Severe local perturbation of the
pressure gradient may occur if the ascending hydro-
thermal plume coincides with a dilational jog on a
seismically active fault. A hypothetical pressure-
time plot for the interior of a major dilational jog is
shown in Figure 2. The abrupt post-seismic reduc-
tion in fluid-pressure below ambient hydrostatic
values is progressively restored by inward percola-
tion. It seems likely that the period of restoration
may extend throughout the aftershock phase, which
for a moderate to large earthquake may last for
several months. Violent local boiling (flashing) may
be triggered by sudden pressure reduction, perhaps
giving rise to hydrothermal eruptions along the
high permeability pathway formed by the fresh rup-
ture. Another possibility is ingress of cold ground-
water from the surrounding rock into the jog. Both
are plausible scenarios for episodes of rapid hydro-
thermal precipitation (Drummond and Ohmoto,
1985; Henley, 1985).
It is clear in the case of much epithermal min-
eralization that faulting drives fluid flow. The sud-
den creation of cavity space in dilational jogs re-
duces fluid pressure, triggering boiling episodes and
inducing a concentrated fluid influx. Either flash-
boiling or cold-water mixing from sudden fluid in-
gress is capable of triggering major episodes of
hydrothermal deposition in such environments.
Importance of Strike-Slip Dilational Jogs as
Mineralizing Sites
Though not widely recognized in the mmmg
literature, major dilational jogs in strike-slip fault
systems seem likely to form particularly powerful
mineralizing engines for epithermal depositon. As
an instructive present-day example, we may con-
sider the southern portion of the San Andreas fault
system (Figure 6), where the right-lateral San
Andreas fault itself is linked through the Imperial
and Cerro Prieto strike-slip faults (by a series of
major dilational jogs) to spreading centers in the
Gulf of California. Mafic and felsic magmatism, up-
welling high-temperature (T>300C) geothermal
plumes (locally highly metalliferous brines), and
dense suites of subvertical fractures are all localized
within jog areas susceptible to fluid pressure pertur-
bation by strike-slip rupturing on flanking faults
(Sibs on, 1987). Indeed, there are historic records of
hydrothermal eruptions triggered by major ruptures
on the Cerro Prieto fault, with geysering continuing
for several months after the main shock and re-
intensifying after major aftershocks.
Because of the complexity of deformation with-
in the jogs themselves (a mixture of normal and
strike-slip faulting with extensional fracturing), the
likelihood of intense hydrothermal alteration, and
Earthquake faulting, induced fluid flow, and fault-hosted gold-quartz mineralization 609
Figure 6. Dilational fault jogs in southern San
Andreas fault system (from Sibson, 1987). Stipple
indicates areas of intense microearthquake acti-
vity (SSJ, Salton Sea jog; BJ, Brawley jog;
CPJ, Cerro Prieto jog). Circles mark epicenters of
major recent strike-slip ruptures; arrowS indicate
direction and extent of rupture propagation (note
termination at jogs); asterisks mark Quaternary
volcanic centers; triangles mark geothermal fields
(larger triangles are T> 300C fields) .
the often inconspicuous character of the driving
strike-slip faults, it is easy to see why recognition of
fossil strike-slip jogs may prove difficult. However,
increasing numbers of epithermal gold-quartz depo-
sits in possible strike-slip jog settings are being
documented. These include the Martha Mine in the
Hauraki gold field, New Zealand (Wellman, 1954;
Sibson, 1987), the Mahd adh Dhahab and other
deposits in western Saudi Arabia (Moore and Al-
Shanti, 1979; Huckerby and others, 1983), the Cino-
la deposit located on the Sandspit fault in the
northern Queen Charlotte Islands of British Colum-
bia (Champigny and Sinclair, 1982), the Cannon
Mine, Chelan County, Washington (Ott and others,
1986), and possibly the Mesquite Mine in south-
eastern California (Willis and others, 1987).
,
.
o 2S km
L.. !
\
.. """
'.
MESOTHERMAL ENVIRONMENT
Mesothermal gold-quartz lodes hosted in steep-
ly inclined reverse or reverse-oblique shear zones of
mixed "brittle-ductile" character account for a sig-
nificant proportion of global gold production. These
vein systems are best known from Archean granite-
greenstone belts in the Superior and Slave Pro-
vinces of Canada (Roberts, 1987), the Yilgarn Block
of Western Australia (Boulter and others, 1987),
and similar Archean terranes, where host rocks are
mainly mafic-ultramafic volcanic sequences but
include calcalkaline volcanics and volcano-clastic
rocks, turbidite sequences, and felsic intrusions.
Younger deposits are, however, known from com-
parable tectonic and lithologic settings (e.g., the
610 R. H. Sibson
Early Cretaceous Mother Lode, Grass Valley, and
Alleghany vein systems in California; Knopf, 1929;
Fergusson and Gannet, 1932; Johnston, 1940;
Bohlke and Kistler, 1986). The lower Paleozoic
Bendigo-Ballarat deposits of southeastern Austra-
lia, hosted in a turbidite-slate assemblage, also
share similar structural characteristics (Chace,
1949; Cox and others, 1986).
There are no unequivocal controls on the depth
of formation of these lode-gold deposits, but several
factors suggest that they formed near the base of the
seismogenic zone. The hosting shear zones deve-
loped mostly under low to middle greenschist-facies
metamorphic conditions, and the complex mix of
"brittle" and "ductile" deformation fits well with the
deformation style inferred for this structural level
in the crust. It is clear from the detailed structural
studies of Robert and Brown (1986), and others, that
formation of brittle vein fractures has occurred in-
termittently during semi-continuous ductile shear-
ing in the shear zones. Carbonate alteration is often
intense in the vicinity of veins that contain mixed
H
2
0-C0
2
fluid inclusions of relatively low salinity.
SEISMOGENIC
REGIME
-H4-- 0
I
Figure. 7. Inferred tectonic setting (not to scale) for
mesothermal gold-quartz lodes in relation to continental
seismogenic regime (after Sibson and others, 1988). Origin
of ascending fluids may be magmatic hydrothermal, or
from deep crustal metamorphism, or possibly by juvenile
derivation from the mantle.
Fluid inclusion and isotope studies suggest that
these deposits developed in the temperature range
250
0
-400C at pressures of 2-4 kb ("" 7 -14 km depth)
(Kerrich, 1986; Robert and Kelly, 1987). The origin
of the mineralizing fluids remains debatable; pos-
sible sources include metamorphic fluids from the
deep crust, magmatic hydrothermal fluids, or juve-
nile CO
2
-rich fluids from the upper mantle (Roberts,
1987). Whatever the source, it is clear that the
shear zones acted as major conduits for focused fluid
ascent.
The overall structural inference, however, is
that veining developed around the base of the seis-
mogenic regime within the roots of brittle reverse-
fault systems (Figure 7), in marked contrast to epi-
thermal gold-quartz mineralization.
Structural Characteristics
The following summary of the characteristics of
the hosting shear zones is based largely on the
Canadian lode-gold mining districts of Yellowknife
(Henderson and Brown, 1965), Kirkland Lake
(Watson and Kerrich, 1983), and Val d'Or (Robert
and Brown, 1986; Vu and others, 1987). However, it
is apparent from the literature that many of these
characteristics are shared worldwide by other shear-
zone-hosted lode-gold deposits of Archean and youn-
ger age. The gold-quartz vein-systems are hosted by
steeply dipping (50
0
-80
0
) schistose shear zones ex-
hibiting a mixed discontinuous-continuous ("brittle-
ductile") style of deformation, with an L-S tectonite
fabric locally disrupted by discrete shears and vein
fractures. Stretch lineations and fibrous slicken-
lines on discrete slip surfaces tend to rake at high
angles in the steeply inclined foliation. Whereas
many of the shear zones may have initiated as
strike-slip or normal fault systems (and have clearly
undergone a long, varied history of movement),
indicators of shear sense invariably show that min-
eralization developed during a phase of late high-
angle reverse or reverse-oblique slip (Sibson and
others, 1988). Where established, the total reverse
displacement is not great - of the order of a few
hundred meters at most. One notable aspect of the
vein systems is that they often extend to great
depth. Several existing mines are> 2 km deep.
Structural characteristics of vein systems
within a number of mines in the Val d'Or district of
Quebec have been studied in great detail (Ames,
1948; Wilson, 1948; Robert and Brown, 1986; Vu
Earthquake faulting, induced fluid flow, and fault-hosted gold-quartz mineralization 611
and others, 1987). Two main vein-sets occur: len-
ticular fault veins lying subparallel to schistosity
within steeply inclined shear zones across which
motion is almost purely reverse, and flats (sub-
horizontal extension veins). A third type of hybrid
shear-extensional vein occurs occasionally where
the dip of steep reverse shear zones locally shallows.
The relative development of different vein-sets may
vary from one mine to another within a particular
district, and from one district to the next.
Within each vein-set, composite vein textures
record histories of incremental deposition. The
purely extensional character of flats, revealed by
fiber growth and crack-seal textures, suggests that
they formed as hydraulic extension fractures per-
pendicular to the least compressive stress (03) when
it was exceeded by fluid pressure (i.e., P>03) (Rob-
ert and Brown, 1986). This, coupled with the pre-
sence of a steep conjugate set of reverse shear zones,
constrains orientations of the principal compressive
stresses (Figure 8) with a large angle of reactiva-
tion (a
r
) between the maximum stress (01) and the
shear zones. Following Guha and others (1983),
fault veins are interpreted as lenticular cavity-fills
formed by local overriding during slip episodes on
discrete irregular shears; in essence, they are a
variety of dilational jog. Robert and Brown (1986)
made the key structural observation that there is no
consistent cross-cutting relationship between the
different sets of veins; this has the important im-
plication that the two main vein-sets probably
developed at different stages of a repeating cycle
(Figure 9).
Reactivation of Unfavorably Oriented Faults
For existing cohesionless faults whose poles lie
in the 01/03 plane (Figure 8), the optimum angle for
frictional shear reactivation, at which the stress
ratio for reactivation is a minimum, is given by
a
r
* = 0.5 tan -l(lI}ls) where}ls is the static coefficient
of friction. For a typical rock friction-coefficient, }lS
= 0.85 (Byerlee, 1978), a
r
* = 25. At other than the
optimum angle, greater stress ratios are needed for
reactivation. Of particular relevance here is that for
reactivation angles, a
r
> 2a
r
* = 50, a necessary con-
dition for shear reactivation of faults is that P> 03
so that the least stress becomes effectively tensile
(Sibson, 1985b).
"Fault- Valve" Model
In view of the previous discussion, high-angle
reverse shear zones hosting lode-gold vein-systems
are clearly very unfavorably oriented with regard to
the prevailing stress field (50< a
r
< 80), but the
existence of flats testifies to the occasional presence
of fluid pressures exceeding the litho static load -
Figure 8. Schematic representation of different vein-sets
associated with shear-zone-hosted lode-gold deposits in the
Val d'Or mining district, shown in relation to inferred
stress field (principal compressive stresses, 01> 02 >03)
and fault reactivation angle, 8"
-+-
(J3
flat
/
5
(JI
Figure 9. Schematic evolutionary sequence (1-5) of pre-
failure flats alternating with post-failure discharge veins,
illustrating development of mutually cross-cutting vein-
sets in a lode-gold vein system (after Robert and Brown,
1986). Note problems in interpretation that would arise if
only a few cross-cutting relationships between veins were
exposed (after Sibson, 1989).
612
R. H. Sibson
the necessary condition for reactivation in shear.
Thus, in a compressional regime, the base of the
seismogenic zone appears to act as a largely im-
permeable cap to pressurized fluids ascending from
below (Figure 7), and the shear zones behave as
fluid-pressure-activated fault-valves, promoting re-
peated fluctuations between litho static and hydro-
static fluid pressure linked to the earthquake-stress
cycle (Sibson and others, 1988). The following
repeating sequence of events is envisaged (Figures
2 and 9).
Fluid pressure around the base of the seis-
mogenic zone builds up to supralithostatic values,
with flats opening up by hydraulic fracturing during
the pre-seismic phase of the stress cycle. Shear
failure then becomes possible and a reverse-slip
earthquake rupture nucleates in the shear zone
near the base of the seismogenic regime. The
rupture propagates mostly upward, relieving shear
stress and creating fracture permeability along the
fault. Post-failure discharge from the geopressured
reservoir at depth then occurs along the rupture
zone, with fluid pressures dropping rapidly toward
hydrostatic values. Hydrothermal deposition occurs
within rideovers to form lenticular fault veins dur-
ing the post-seismic phase of afterslip and after-
shock activity, while the discharge progressively
diminishes. Hydrothermal self-sealing then allows
the cycle to repeat. Mutual cross-cutting relation-
ships between fault-veins and flats (illustrated in
Figure 9) are thus a natural consequence of the
cyclic nature of the process.
In summary, therefore, the vein systems appear
to represent the nucleation sites of high-angle
reverse ruptures near the base of the seismogenic
zone, with the build-up and release of fluid pressure
linked to different phases of the earthquake-stress
cycle. It should be noted that repeated rupture
nucleation at the same site is an expectation of char-
acteristic earthquake behavior. If a moderate earth-
quake rupture involves slip of ""0.1-1.0 m, a total
displacement of, say, 100 m across a shear-zone sys-
tem would involve 10
2
_10
3
cycles of fluid pressure
accumulation and release.
Some form of fault-valve behavior may also be
expected for other faults that are less than optimally
oriented for reactivation, and perhaps for faults that
become strongly cemented during interseismic
periods. However, high-angle reverse faults provide
the optimum setting for this fluid-activated valve
behavior, in which the largest amplitude fluid-
pressure fluctuations can be expected.
Fault- Valve Behavior as a Precipitation
Mechanism
Precipitation mechanisms for gold and asso-
ciated gangue minerals are probably less well
understood for mesothermal than epithermal en-
vironments. However, some consideration has been
given in the literature to the effects of major fluid-
pressure reductions at this crustal level. For ex-
ample, Walther and Helgeson (1977) pointed out
that isoenthalpic reduction of fluid pressures from
lithostatic to hydrostatic values at "" 12 km depth
would precipitate as much as 2300 ppm silica from a
pure water-silica solution. Because the pressure
drops are potentially so large, increasing downward
at "" 17 MPa/km for typical crustal densities, phase
separation (boiling) of CO
2
becomes likely in a
mixed H
2
0-C0
2
fluid, even at considerable depth.
Evidence that this has occurred comes from wide-
spread carbonate alteration in the vicinity of depo-
sits and the presence of calcite as an important
gangue consituent. According to Robert and Kelly
(1987), gold entrapment in some of the Val d'Or
deposits appears associated with late calcite gangue,
which they attributed to CO
2
effervescence during
large pressure fluctuations.
CONCLUSIONS
This paper emphasizes the importance of con-
sidering the structural setting of fault-hosted gold-
quartz mineralization in relation to the seismogenic
regime in continental crust, and the interplay
between episodic earthquake faulting, induced flow
of fluids, and hydrothermal precipitation. Favored
structural sites for fault-hosted epithermal and
mesothermal gold-quartz deposits occur at the top
and bottom, respectively, of the seismogenic regime
- the former associated with extensional!transten-
sional fault systems, the latter with compressional!
transpressional systems.
In both cases, favored sites are such that abrupt
reductions in fluid pressure induced by seismic-slip
episodes seem likely to have played a major role in
hydrothermal precipitation. Two main mechanisms
are proposed, though others may exist for more
complicated fault geometries. In the epithermal
environment, pressure reduction below ambient
hydrostatic values is achieved by a suction-pump
effect at dilational fault jogs, bends, and intersec-
tions. In the mesothermal environment, the ability
Earthquake faulting, induced fluid flow, and fault-hosted gold-quartz mineralization 613
of unfavorably oriented high-angle reverse faults to
behave as fluid-pressure-triggered valves causes
fluid pressure to follow a sawtooth oscillation be-
tween litho static and hydrostatic values. There is
clearly much scope for exploring the geochemical
and thermodynamic implications for mineral pre-
cipitation of these different mechanical models.
Recognition of these dynamic fault mechanisms
for perturbing fluid pressure reinforces the com-
monly held belief of many early workers that epi-
sodes of fluid flow and hydrothermal deposition are
intimately related to intermittent earthquake fault-
ing. In most instances, the fault/fracture systems
cannot simply be regarded as passive fluid conduits.
ACKNOWLEDGMENTS
I thank Bob Mason of Queen's University for
review of the manuscript, Neville Price for en-
couraging my nascent interests in fluids and
faulting, and John McMahon Moore for beginning
and Howard Poulsen and Francois Robert (Geo-
logical Survey of Canada) for continuing and
deepening my education in mine geology. Work
leading to the development of these ideas was
supported by National Science Foundation Grant
EAR-86-07445 and U.S. Geological Survey Grant
14-08-0001-G1331.
REFERENCES
AMES, H. G., 1948, Perron Mine; pp. 893-898 in Struc-
tural Geology of Canadian Ore Deposits: Canadian
Institute of Mining and Metallurgy, Symposium Voume,
948p.
BOHLKE, J. K., and R W. KISTLER, 1986, Rb-Sr, K-Ar,
and stable isotope evidence for the ages and sources of fluid
components of gold-bearing quartz veins in the northern
Sierra Nevada foothills metamorphic belt, California:
Economic Geology, v. 81, p. 296-322.
BOULTER, C. A., M. G. FOTIOS, and G. N. PHILLIPS, 1987,
The Golden Mile, Kalgoorlie: A giant gold deposit loca-
lized in ductile shear zones by structurally induced in-
filtration of an auriferous metamorphic fluid: Economic
Geology, v. 82, p. 1661-1678.
BRUHN, R. L., and W. T. PARRY, 1988, Structural and fluid
characteristics of normal faults: U.S. Geological Survey,
Open-File Report 87-673, p. 374-384.
BUCKLAND, W., 1836, Geology and Mineralogy Con-
sidered with Reference to Natural Theology: William
Pickering, London, England, UK, 596 p.
BYERLEE, J. D., 1978, Friction of rocks: Pure and Applied
Geophysics, v. 116, p. 615-626.
CATHLES, L. M., 1981, Fluid flow and genesis of hydro-
thermal ore deposits: Economic Geology, 75th Anniversary
Volume, p. 424-457.
CHACE, F. M., 1949, Origin of the Bendigo saddle reefs
with comments on the formation of ribbon quartz: Econo-
mic Geology, v. 44, p. 561-597.
CHAMPIGNY, N., and A. J. SINCLAIR, 1982, The Cinola
gold deposits, Queen Charlotte Islands, British Columbia:
Canadian Institute of Mining and Metallurgy, Special
Volume 24, p. 243-254.
CHEN, W. P., and P. MOLNAR, 1983, Focal depths of intra-
continental and intraplate earthquakes and their implica-
tions for the thermal and mechanical properties of the
lithosphere: Journal of Geophysical Research, v. 88, p.
4183-4214.
COX, S. F., M. A. ETHERIDGE, and V. J. WALL, 1986, The
role of fluids in syntectonic mass transport, and the loca-
lization of metamorphic vein-type ore deposits: Ore Geo-
logy Reviews, v. 2, p. 65-86.
DRUMMOND, S. E., and H. OHMOTO, 1985, Chemical
evolution and mineral depostion in boiling hydrothermal
systems: Economic Geology, v. 80, p. 126-147.
FERGUSSON, H. G., and R. W. GANNET, 1932, Gold-
Quartz Veins of the AUeghany District, CaUfornia:
U.S. Geological Survey, Professional Paper 172,139 p.
GUHA, J., G. ARCHAMBAULT, and J. LEROY, 1983, A cor-
relation between the evolution of mineralizing fluids and
the geomechanical development of a shear zone as illu-
strated by the Henderson 2 Mine, Quebec: Economic Geo-
logy, v. 78, p. 1605-1618.
HENDERSON, J. F., and J. C. BROWN., 1965, Geology and
Structure of the YeUowknife Greenstone Belt, District
of McKenzie: Geological Survey of Canada, Bulletin 141,
87p.
HENELY, R W., 1985, The geothermal framework of epi-
thermal deposits; pp. 1-44 in B. R. Berger and P. M. Bethke
(eds.), Geology and Geochemisty of Epithermal Sys-
tems: Society of Economic Geologists, Reviews in Econo-
mic Geology 2, 298 p.
HILL, D. P., 1977, A model for earthquake swarms: Jour-
nal of Geophysical Research, v. 82, 1347-1352.
HUCKERBY, J. A., J. McM. MOORE, and G. R. DAVIS,
1983, Tectonic control of mineralization at Mahad adh
Dhabab gold mine, western Saudi Arabia: Transactions of
the Institute of Mining and Metallurgy, v. 92, p. Bl71-B182.
HULIN, C. D., 1925, Structural control of ore deposition:
Economic Geology, v. 24, p. 15-49.
JOHNSTON, W. D., 1940, The Gold-Quartz Veins of
Grass Valley, California: U.S. Geological Survey, Profes-
sional Paper 194,101 p.
KERRICH, R, 1986, Fluid infiltration into fault zones:
Pure and Applied Geophysics, v. 124, p. 225-268.
614
R. H. Sibson
KING, G. C. P., 1983, The accommodation of large strains
in the upper lithosphere of the earth and other solids by
self-similar fault systems - The geometrical origin of b-
value: Pure and Applied Geophysics, v.121, p. 761-815.
KNOPF, A., 1929, The Mother Lode System of Cali-
fornia: u.s. Geological Survey, Professional Paper 157,88
p.
McKINSTRY, H. E., 1948, Mining Geology: Prentice-Hall,
Englewood Cliffs, New Jersey, USA, 677 p.
MOORE, J. McM., and A. M. AL-SHANTI, 1979, Structure
and mineralization in the Najd fault system, Saudi Arabia;
pp. 17-28 in Evolution and Mineralization of the Ara-
bian-Nubian Shield, Volume 2: Institute of Applied
Geology, King Abdul Aziz University (Jeddah)lPergamon
Press, Oxford, England, UK, 225 p.
NEWHOUSE, W. H, 1942, Ore deposits as Related to
Structural Features: Princeton University Press, Prince-
ton, N ew Jersey, USA, 280 p.
OTT, L. E., D. GROODY, E. L. FOLLIS, and P. L. SIEMS,
1986, Stratigraphy, structural geology, ore mineralogy,
and hydrothermal alteration at the Cannon Mine, Chelan
County, Washington, U.S.A.; pp. 425-435 in A. J. Mac
Donald (ed.), GOLD '86: An International Symposium
on the Geology of Gold Deposits: Proceedings volume,
Toronto, Ontario, 517 p.
REASENBERG, P., and W. L. ELLSWORTH, 1982, Mter-
shocks of the Coyote Lake, California, earthquake of
August 6, 1979 - A detailed study: Journal of Geophysical
Research, v. 87, p.10,637-10,655.
ROBERT, F., and A. C. BROWN, 1986, Archean gold-
bearing quartz veins at the Sigma Mine, Abitibi green-
stone belt, Quebec, Part I: Geologic relations and forma-
tion of the vein system: Economic Geology, v. 81, p. 578-
592.
__ , and W. C. KELLY, 1987, Ore-forming fluids in
Archean gold-bearing quartz veins at the Sigma Mine,
Abitibi greenstone belt, Quebec, Canada: Economic Geo-
logy, v. 82, p. 1464-1482.
ROBERTS, R. G., 1987, Ore deposit models #11- Archean
lode gold deposits: Geoscience Canada, v. 14 p. 37-52.
SCHWARTZ, D. P., and K. J. COPPERSMITH, 1984, Fault
behavior and characteristic earthquakes - Examples from
the Wasatch and San Andreas fault zones: Journal of Geo-
physical Research, v. 89, p. 5681-5698.
SEGALL, P., and D. D. POLLARD, 1980, Mechanics of dis-
continuous faults: Journal of Geophysical Research, v. 85,
p.4337-4350.
SIBSON, R. H. 1977, Fault rocks and fault mechanisms:
Journal of the Geological Society of London, v. 133, p. 191-
213.
__ , 1981, Fluid flow accompanying faulting - Field evi-
dence and models; pp. 593-603 in D. W. Simpson and P. G.
Richards (eds.), Earthquake Prediction: An Interna-
tional Review: American Geophysical Union, Maurice
Ewing Series 4, 680 p.
1983 Continental fault structure and the shallow
source: Journal of the Geological Society of
London, v. 140, p. 741-767.
__ , 1985a, Stopping of earthquake ruptures at
dilational fault jogs: Nature, v. 316, p. 248-251.
__ , 1985b, A note on fault reactivation: Journal of
Structural Geology, v. 7, p. 751-754.
__ ,1986, Earthquakes and lineament infrastructure:
Philosophical Transactions of the Royal Society of London,
v. A317, p. 63-79.
__ , 1987, Earthquake rupturing as a mineralizing
agent in hydrothermal systems: Geology, v. 15, p. 701-704.
__ , 1989, Earthquake faulting as a structural process:
Journal of Structural Geology, v. 11, p. 1-14.
__ , F. ROBERT, and K. H. POULSEN, 1988, High-angle
reverse faults, fluid-pressure cycling, and mesothermal
gold-quartz deposits: Geology, v. 16, p. 551-555.
SILLITOE, R. H., 1985, Ore-related breccias in volcanoplu-
tonic arcs: Economic Geology, v. 80, p. 1467-1514.
VU, L., R. DARLING, J. BELAND, and V. POPOV, 1987,
Structure of the Ferderber gold deposit, Belmoral mines
Ltd., Val d'Or, Quebec: Canadian Institute of Mining and
Metallurgy Bulletin, v. 80, p 68-77.
WALTHER, J. V., and H. C. HELGESON, 1977, Calculation
of the thermodynamic properties of aqueous silica and the
solubility of quartz and its polymorphs at high pressures
and temperatures: American Journal of Science, v. 277, p.
1315-1351.
WATSON, G. P.,and R. KERRICH, 1983, Macassa Mine,
Kirkland Lake - Production History, Geology, Gold
Ore Types and Hydrothermal Regimes: Ontario Geo-
logical Survey, Miscellaneous Paper 110, p. 56-74.
WELLMAN, H. W., 1954, Stress pattern controlling lode
formation and faulting at Waihi Mine, and notes on the
stress pattern in the northwestern part of the North Island
of New Zealand: New Zealand Journal of Science and
Technology, v. 36B, p. 201-206.
WILLIS, G. F., R. M. TOSDAL, and S. L. MANSKE, 1987,
The Mesquite Mine, southeastern California - Epither-
mal mineralization in a strike-slip fault system [abstract):
Geological Society of America Abstracts with Programs, v.
19, p. 892.
WILSON, H. S., 1948, Lamaque Mine; pp. 882-891 in
Structural Geology of Canadian Ore Deposits: Cana-
dian Institute of Mining and Metallurgy, Symposium Vol-
ume,948p.
WISSER, E. 1941, Some observations in ore search: Trans-
actions of the American Institute of Mining and Metal-
lurgical Engineers, v. 144, p. 140-145.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Transpressive tectonics and the Archean gold deposits of
Superior Province, Canadian Shield
K. H.POULSEN,F. ROBERT, and K.D.CARD
Geological Survey of Canada, Energy, Mines and Resources Canada, 601 Booth Street, Ottawa, Ontario KIA DE8, Canada
(received August 12, 1988; revision accepted March 31, 1989)
ABSTRACT
Superior Province is an Archean craton surrounded by Proterozoic orogenic belts. Although some
Proterozoic deformation locally affected the craton, most of its structures are Archean. The craton comprises
alternate EIW -trending subprovinces of volcano-plutonic, metasedimentary, plutonic, and high-grade gneiss
types. Gold-quartz vein deposits are confined to the volcano-plutonic subprovinces and their margins and
occur in or adjacent to brittle-ductile faults that range from one to hundreds of kilometers long. Regardless of
size, gold-bearing structures and regional faults conform to a remarkably consistent province-wide pattern of
orientation and slip sense; EIW-trending faults are typically high-angle reverse or dextral strike-slip, and
oblique-slip faults of the NW- and NE-trending sets have dextral and sinistral components, respectively.
Rocks adjacent to faults are typically penetratively strained over large areas; this strain takes the form of
large folds subparallel to subprovince boundaries and steep metamorphic foliations and stretching lineations
with variations from subvertical to subhorizontal plunges. The reticulate to anastomosed patterns produced
by interaction of the fault sets, coupled with the large-scale variations in orientation of finite strain axes,
suggest a transpressive tectonic regime related to oblique accretion of subprovinces in the late Archean.
INTRODUCTION
Superior Province is an Archean craton which
forms the core of the North American plate. It is a
remnant of a once more extensive piece of Archean
crust that has been truncated by Proterozoic oro-
gens. Superior Province contains 120 sizable (> 3
tonnes) gold deposits and accounts for more than
half of past Canadian gold production. Although
geologic characteristics of individual deposits are
diverse, most comprise arrays and networks of
quartz veins of syntectonic origin. In this paper we
describe the structure of individual deposits and
districts and discuss alternate hypotheses concern-
ing tectonic regimes in which they developed.
SUPERIOR PROVINCE
Superior Province comprises alternate EIW-
trending subprovinces of volcano-plutonic, meta-
sedimentary, plutonic, and high-grade gneissic
types (Card and others, 1989). The dominant E-W
structural trends created by this alternation are
mainly attributable to accretion of oceanic island
arcs, turbiditic accretionary prisms, and older sialic
continental nuclei during the late Archean Kenoran
Orogeny (Langford and Morin, 1976). Gold deposits
are confined to the volcano-plutonic subprovinces
that have undergone similar orogenic histories
(Card and others, 1989) including the following:
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 615-623. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
615
616 K. H. Poulsen, F. Robert, and K. D. Card
Deposition of immature sedimentary rocks,
mainly conglomerate and arenite, which are loca-
lized in restricted basins adjacent to major fault
zones. These rocks are typically deposited above
unconformities on upturned volcanic and plutonic
units and are locally accompanied by volcanic rocks
of alkalic affinity.
Widespread but inhomogeneously developed
deformation and low- to medium-grade metamor-
phism resulting in steep EIW -trending metamorphic
foliation, locally axial planar to folds a few kilo-
meters in wavelength. Intense fabric is developed in
and adjacent to the major faults.
Syntectonic emplacement of large volumes of
granitoid rock in the form of batholiths and stocks.
Most authors agree that the timing of gold
deposition corresponds with the late development of
these orogenic features but there is still consider-
able debate about the relative importance of the
source of gold-bearing fluids: magmatic, metamor-
phic, and deep crustal sources have been proposed.
Nonetheless, all models of gold deposition ascribe
importance to the role of faults whether as deep
fluid conduits or as features that localize magmatic
activity.
STRUCTURE OF GOLD DEPOSITS
Most of the important gold deposits in Superior
Province comprise quartz-carbonate veins that form
three-dimensionally connected networks of variable
complexity. In the simplest cases, single veins, ty-
pically one meter wide, occupy narrow shear zones a
few meters wide. With increasing size, deposit
geometry is more complex, commonly as arrays of
veins in subparallel shear zones. In large deposits,
vein networks are common and are composed of two
elements: veins in shear zones and shear fractures,
and veins occupying extensional fractures compa-
tible with the prevailing shear sense. Two examples
serve to illustrate some recurring patterns.
The Sigma Mine at Val d'Or, Quebec, is deve-
loped on a structural network comprising EIW-
striking subvertical veins occupying narrow ductile
shear zones and flat-lying extensional veins that
connect with and typically form links between the
shear veins (Figure 1). The subvertical shear zones
consistently demonstrate reverse displacements and
these, coupled with flat extension veins, indicate
shortening about a sub horizontal N-S axis with co-
LEVEL
5-
10-
15-
20-
25-
30-
35-
40-
incident subvertical extension (Robert and Brown,
1986).
The San Antonio Mine at Bissett, Manitoba,
also exhibits two distinctive vein types that inter-
connect to form a three-dimensional network (Fig-
ure 2). Veins in steep NE-striking shear zones form
an array which transects a sill of metadiabase, and
these are connected by "stockwork" zones composed
of central NW-striking shear fractures, surrounded
in turn by extensional vein breccia and tabular
extensional spurs (Lau and others, 1986). Both vein
types are localized in structures that have oblique
reverse displacements (Lau, 1988).
These are but two of many possible examples
which serve to illustrate some recurring aspects of
the structure of Superior Province gold deposits:
In virtually all cases where documentation
exists, reverse faults or oblique reverse faults have
a:
<
w
1:
'"
1:
0-
a:
0
z
,200m,
SIGMA MINE
\
,
IQ
I
I
LEGEND
DUCTILE SHEAR ZONE.
SUB VERTICAL VEIN
FLAT VElN
LATE FAULTS
Figure 1. Structural cross-section through the Sigma
Mine (after Robert, 1986). The vein network is localized
between the North and South shears which, on a regional
scale, are second-order structures.
Transpressive tectonics and the Archean gold deposits, Canadian Shield 617
been shown to be the loci of gold deposition (San-
born-Barrie, 1987; Vu and others, 1987; Melling and
others,1986; Callan and Spooner, 1989; Tessier and
others, 1988). Only rarely have mineralized veins
in strike-slip zones been reported (Poulsen, 1984),
and these are of limited economic significance.
Regular orientation of shear and extensional
veins within a deposit suggests formation under
conditions of uniform directed stress, invariably in-
volving sub horizontal compression. It is well estab-
lished that steep structures striking either E-W,
NE-SW, or NW-SE, as illustrated by Sigma and San
Antonio, are common throughout Superior Pro-
vince.
x
y
,'--------z Shal t A
The structural networks of veins which con-
stitute many gold deposits are themselves impres-
sive structures that typically extend over a vertical
interval > 2 km in the crust and that undoubtedly
involved seismic failure over large rock volumes
(Sibson and others, 1988).
FAULTS AND GOLD DEPOSITS
In addition to an association with deposit-scale
shear zones, it is empirically well known that in
Superior Province gold deposits tend to cluster near
larger faults and shear zones. These are rarely
........
16Type (Shear) velrls
38' Type (StOcJ<work) veLl1s
Figure 2. Perspective view of
the San Antonio Mine (after
Poulsen and others, 1986) .
The two reference surfaces
represent the hanging-wall
and footwall contacts of a sill
of metadiabase. The heavy
dashed lines represent the
traces of veins in NE-striking
shear zones, which cut trans-
versely across the sill. The
NW-striking stockwork veins
cut the sill at a low angle, and
their projection onto the hang-
ing-wall surface is depicted.
The distance between each
level of the mine is '" 50 m.
8. CD Wnze Jl(Ol'lChOllS on the
hanglngwal scrlace
______________
XY N55W/ 90
Xl Homo",al plane
yZ N35E190
618
...
D
l[2]
o
LJ
K. H. Poulsen, F. Robert, and K. D. Card
...
1-
.j.
+
.j.
-\-
+ ...
<-
...
4-
+ ...
... -\-
>- +
+
... J. ... ...
+- +
+
... ...
-I- ...
... ..
+
..
4-

+ +
..
...-+
...
..
... r +
-I- ...
<- ...
...
...
...
"t
01-
10 Km
<-
r
...
10 Km
PLUTONIC ROCKS
-
F AU L T AND SHEAR ZONE
VOLCANIC ROCKS
GOLD DEPOSIT
S S IGMA MINE
CONG LOM ERATE, ARENITE, WACKE
SA SAN AN TONIO MINE
lC CADI L L AC FAUl T
WACKE-MUDSTONE
WR WAN I PIGOW RIVER FAUL T
MR MANIGO TAGAN RI VER FAUl T
S R S EINE RIV ER FAUL T
o OUE TI CO FA UL T
Figure 3, A comparison of district-scale fault patterns: A) Val d'Or, Quebec; B) Bissett, Manitoba; and C) Mine Centre,
Ontario. The Sigma (S) and San Antonio (SA) mines are located in A and B, respectively. Note the distribution of
sedimentary sequences containing conglomerate and arenite with respect to the faults.
Transpressive tectonics and the Archean gold deposits, Canadian Shield 619
mineralized but commonly share the effects of hy-
drothermal alteration in the form of chloritization,
carbonatization, and sericitization with the smaller
ore-bearing structures. The large faults vary in size
but commonly are observed at two scales. Con-
tinuous first-order faults extend up to a few hundred
kilometers in length and are prominent regional
features associated with gold districts (Figure 3).
Smaller faults, up to a few tens of kilometers long,
form discontinuous strands throughout the pro-
vince. The interplay of faults of different order im-
parts a reticulate to anastomosed pattern of fault-
ing, which is observed at all scales (Figures 3 and
4).
The first-order faults in Superior Province fall
into two general categories. The first of these,
termed "shear zone faults" by Norman (1946), are
characterized by parallelism with stratigraphic
contacts, development of a 20 to 200-m-wide zone of
metasomatic schist, and localized intensification of
regional metamorphic fabric. Examples include the
EIW-striking Larder-Cadillac and Porcupine-Destor
fault systems in eastern Superior Province and the
Seine River and Paint Lake faults in western
Superior Province (Figure 4). Because of their
phyllonitic character, they rarely lend themselves to
definitive kinematic analysis, but the presence of
steep lineation within and adjacent to them, as well
SCALE: 0 ....
Figure 4. Map of Superior Province faults and the distribution of major gold deposits. Val d'Or, Quebec (A), Bissett, Manitoba
(B), and Mine Centre, Ontario (C) are located for reference to Figure 3.
620 K. H. Poulsen, F. Robert, and K. D. Card
as their concordant map-view orientation with re-
spect to stratigraphic units, has led most workers to
regard them as steep thrusts (Dimroth and others,
1983). In some localities, minor dextral offsets are
also indicated (Dimroth and others, 1983; Williams,
1983; Park, 1981). The second type of first-order
faults is represented by mylonitic zones up to 1 km
wide, which at least locally are discordant to re-
gional trends. Examples include the Sydney Lake,
Manigotagan, Wanipigow, and Quetico fault zones
in western Superior Province (Figure 4) which, on
the basis of shallow lineation and observed offsets
and deflections, are interpreted to be dextral trans-
current structures (Schwerdtner and others, 1979;
McRitchie, 1971).
Second-order faults, although typically of a
district scale, possess similar orientations and styles
as the first-order structures, except that the former
are also represented by common NW or NE strikes.
In an assessment of a large number of such struc-
tures in northwestern Superior Province, Park
(1981) found that the NW-striking faults typically
have dextral components of displacement, whereas
northeasterly ones are sinistral. This is in accord
with our observations in gold districts, except that
some NE- and NW -trending structures contain
inclined lineations suggestive of oblique-reverse
shear. Early workers also recognized this pattern
(Norman, 1946) and cautioned that sinistral move-
ments on NE-trending structures may be of Protero-
zoic age, particularly near the margins of the craton
(Dimroth and others, 1984).
One of the important characteristics of many
large faults is their tendency to be the loci of coarse
clastic sedimentary rocks deposited above angular
unconformities (Figures 3 and 4). Many authors
have suggested that this results from a direct gene-
tic relationship between faulting and sedimentation
(Dimroth and others, 1983; Poulsen, 1985; Hodgson,
1986), but there is no consensus in the literature as
to the nature of that relationship.
TECTONIC REGIMES
The foregoing paragraphs describe the observa-
tions which suggest that there is a fundamental
relationship between faults and Superior Province
gold deposits. The inference is drawn largely from
the similarities of style, orientation, and displace-
ment sense of structures, from the scale of ore bodies
to that of the province as a whole. Virtually all
authors agree that the overall regime for the late
Archean is one of northerly directed shortening, to
account for the structural grain of Superior Province
and the apparently conjugate nature of the NE- and
NW-trending fault sets. There is considerable un-
certainty about the nature and exact timing of the
first-order phyllonitic faults that are the major
structures directly associated with major gold dis-
tricts (Figure 4). Interpretations of tectonic regime
variously involve reactivation of normal faults,
compressional (i.e., thrust) faulting, wrench fault-
ing, and transpression.
Normal Faults
The interpretation that major phyllonitic faults
such as the Porcupine-Destor and Larder-Cadillac
faults initiated as normal faults is rooted in their
proximity to marked sedimentologic transitions
from coarse clastic sedimentrary rocks to finer
grained turbidites (Dimroth and others, 1983). In
such interpretations, the faults initiated as growth
faults that were perhaps listric to account for the
steepening of underlying volcanic rocks (Hodgson,
1986) but that were reactivated as high-angle
thrusts during subsequent N -directed compression.
The existence of early normal faults is based
primarily on sedimentologic inference, as there is no
direct structural evidence for development of large-
scale normal faults in the late Archean Superior
Province.
Compressional Faults
Over large parts of Superior Province, particu-
larly in the eastern part where the influence oflarge
lateral displacements is difficult to demonstrate, the
first-order faults have been commonly regarded as
thrusts with local oblique-slip components (Kallio-
koski, 1968, 1987; Archambault, 1985). This inter-
pretation also involves the generation of NE- and
NW-trending faults as complementary strike-slip
structures (Figure 5a) that accommodate a coaxial
N-S shortening (Archambault, 1985; Wilson and
others, 1984; Park, 1981). This interpretation ap-
pears to be valid over large areas of Superior Pro-
vince but does not adequately account for the
common development of major strike-slip faults
parallel to the structural trend.
Transpressive tectonics and the Archean gold deposits, Canadian Shield 621
Wrench Faults
In western Superior Province in particular,
wrench faults have been conclusively recognized at
several localities (Schwerdtner and others, 1979;
Poulsen, 1985; McRitchie, 1971). These are typi-
cally EIW -striking and mylonitic but are commonly
coextensive with and subparallel to phyllonitic
varieties. Although they have been questioned,
wrench-fault models have also been applied to the
interpretation of the Larder-Cadillac and Porcu-
pine-Destor faults in eastern Superior Province
(Hubert and others, 1984) on the basis of patterns of
oblique en echelon folds adjacent to these structures.
Wrench-fault interpretations demand external
shortening along directions other than N-S
(Schwerdtner and others, 1979) and account for NE-
and NW-striking faults as complementary strike-
slip structures (Figure 5b). Wrench faulting has
also been invoked to account for the localized depo-
sition of coarse sedimentary sequences in strike-slip
basins (Poulsen, 1985). Although locally an im-
portant factor in the structural development of some
areas, extensive wrench-faulting is not likely be-
cause of the dominance of steep lineation in much of
Superior Province, the absence of demonstrable
large lateral dislocations in many areas, and the
lack of consistent attitudes of large folds with
respect to fault traces.
Transpression
Apart from research on gold deposits and major
fault zones, several structural geologists have
invoked transpressive models to explain large-scale
variations in strain and patterns of fold and fault
development (Hudleston and others, 1988; Stott and
others, 1987; Williams, 1987). This interpretation
(Figure 5c) best explains the dichotomy between
EIW -oriented faults of both strike-slip and thrust
characteristics and allows for considerable variation
from place to place in Superior Province. For ex-
ample, among the districts we have illustrated
(Figure 3), the Mine Centre area is dominated by
dextral wrenching, whereas there is only a minimal
indication of dextral wrenching at Val d'Or (Poulsen
and Robert, 1988). A transpressive model also ac-
counts for the tendency for NE- and NW -striking
faults to have oblique-slip histories although
retaining consistent sinistral and dextral strike-slip
components, respectively. A transpressive interpre-
(b) Transcurrent
(a) Compression
(e) Transpression
Figure 5. Schematic diagram illustrating the theoretical
implications of alternate interpretations of tectonic regime
in the Superior Province. Labeling of tectonic axes: e, axis
of extension; c, axis of shortening; and 1, no shortening or
extension. An initial unit reference block is illustrated by
dashed lines in (b).
tation also allows for localized zones of extension to
accommodate deposition of immature clastic rocks
and the emplacement of coeval magmas into the
upper part of the crust (Wyman and Kerrich, 1988).
CONCLUSIONS
In summary, although there are variations in
details of interpretation there exists a consensus
that the late Archean tectonic regime in Superior
Province involved considerable N-directed shorten-
ing in response to oblique convergence. Gold depo-
sits and districts reflect this structural style and are
predominately related to reverse or oblique-reverse
components of the regime. Given the end members
of crustal extension, strike-slip and compression
(with vertical extension), Superior Province gold
deposits are most strongly influenced by the latter.
622
K. H. Poulsen, F. Robert, and K. D. Card
REFERENCES
ARCHAMBAULT. G., 1985, Archean wrench fault tectonics
and the structural evolution of the Blake River Group,
Abitibi Belt: Discussion: Canadian Journal of Earth
Sciences, v. 22, p. 943-945.
CALLAN, N. J., and E. T. C. SPOONER, 1989, Archean Au
quartz vein mineralization hosted in a tonalite-
trondhjemite-granodiorite terrane, Renabie Mine area,
Wawa, North Ontario, Canada; pp. 9-18 in R. R. Keays, W.
R. H. Ramsay, and D. I. Groves (eds.), The Geology of
Gold Deposits: The Perspective in 1988: Economic
Geology Monograph 6, 667 p.
CARD, K. D., K. H. POULSEN, and F. ROBERT, 1989, The
Archean Superior Province of the Canadian Shield and its
lode gold deposits; pp. 19-36 in R. R. Keays, W. R. H.
Ramsay, and D. I. Groves (eds.), The Geology of Gold
Deposits: The Perspective in 1988: Economic Geology
Monograph 6, 667 p.
DIMROTH, E., and M. ROCHELEAU, 1985, Archean wrench
fault tectonics and the structural evolution of the Blake
River Group: Discussion: Canadian Journal of Earth
Sciences, v. 22, p. 941-942.
__ , G. ARCHAMBAULT, N. GoULET, J. GUHA, and W.
MUELLER, 1984, A mechanical analysis of the late
Archean Gwillim Lake Shear Belt, Chibougamau area,
Quebec: Canadian Journal of Earth Sciences, v. 21, p. 96'3-
968.
__ , L. IMREH, N. GOULET, and M. ROCHELEAU, 1983,
Evolution of the south-central part of the Archean Abitib
Belt, Quebec, Part II: Tectonic evolution and geomech-
anical model: Canadian Journal of Earth Sciences, v. 20, p.
1355-1373.
HODGSON, C. J., 1986, Place of gold ore formation in the
geological development of Abitibi greenstone belt, Ontario,
Canada: Transactions of the Institution of Mining and
Metallurgy, v. 95-B, p. 183-194.
HUBERT, C., P. TRUDEL, and L. GELINAS, 1984, Archean
wrench fault tectonics and structural evolution of the
Blake River Group, Abitibi Belt, Quebec: Canadian Jour-
nal of Earth Sciences, v. 21, p. 1024-1032.
HUDLESTON, P. J., D. SCHULTZ-ELA, and D. L. SOUTH-
WICK, 1988, Transpression in an Archean Greenstone Belt,
northern Minnesota: Canadian Journal of Earth Sciences,
v. 25, p. 1060-1068.
KALLIOKOSKI, J., 1968, Structural features and some
metallogenic patterns in the southern part of the Superior
Province, Canada: Canadian Journal of Earth Sciences, v.
5, p. 1199- 1208.
__ , 1987, The Pontiac problem, Quebec-Ontario, in light
of gravity data: Canadian Journal of Earth Sciences, v. 24,
p.1916-1919.
LANGFORD, F. F., and J. A. MORIN, 1976, The develop-
ment of Superior Province of northwestern Ontario by
merging island arcs: American Journal of Science, v. 276,
p. 1023-1034.
LAU, S., 1988, Structural Geology of the Vein System in
the San Antonio Gold Mine, Bissett, Manitoba, Cana-
da: MSc thesis, University of Manitoba, Winnepeg,
Manitoba, Canada, 154 p.
__ , W. C. BRISBIN, K. H. POULSEN, and D. AMES, 1986,
Structural geology of the San Antonio Gold Mine, south-
eastern Manitoba [abstract]; pp. 92-95 in A. M. Chater
(ed.), Gold '86: An International Symposium on the
Geology of Gold Deposits: Newmont Exploration Canada
(Toronto), Poster Paper Abstracts Volume, 181 p.
McRITCHIE, W. D., 1971, Geology of the Wallace Lake-
Siderock Lake area: A reappraisal; pp. 107-125 in Geo-
logy and Geophysics of the Rice Lake Region, S.E.
Manitoba: Manitoba Mines Branch, Publication 71-1,430
p.
MELLING, D. R., D. H. WATKINSON, K. H. POULSEN, L. B.
CHORLTON, and A. D. HuNTER, 1986, The Cameron Lake
gold deposit, northwestern Ontario, Canada: Geological
setting, structure, and alteration; pp. 149-169 in A. J. Mac
Donald (ed.), Gold '86: An International Symposium
on the Geology of Gold Deposits: Konsult Inc. (Willow-
dale, Ontario), Proceedings Volume, 517 p.
NORMAN, G. W. H., 1946, Major faults, Abitibi Region,
Quebec: Canadian Mining and Metallurgical Bulletin, no.
406, p. 129-144.
PARK, R. G, 1981, Shear-zone deformation and bulk strain
in granite-greenstone terrain of the Western Superior
Province, Canada: Precambrian Research, v. 14, p. 31-47.
POULSEN, K. H., 1984, Archean Tectonics and Minera-
lization at Rainy Lake, Northwestern Ontario: PhD
dissertation, Queen's University, Kingston, Ontario,
Canada, 338 p.
__ , 1986, Rainy Lake wrench zone: An example of an
Archean subprovince boundary in northwestern Ontario;
pp. 177-179 in Workshop on Tectonic Evolution of
Greenstone Belts: Lunar and Planetary Institute (Hous-
ton), LP/ Technical Report 86-10.
__ , and F. ROBERT, 1988, A comparison of structural
style and gold endowment of three Archean gold districts,
Superior Province, Canada [abstract]: Bicentennial Gold
88 (Melbourne, Australia): Geological Society of Austra-
lia, Extended Abstracts/Poster Programme Volume, v. 23,
no. 1-2, p. 36-38.
Transpressive tectonics and the Archean gold deposits, Canadian Shield 623
__ , D. E. AMES, S. LAU, and W. C. BRISBIN, 1986,
Preliminary report on the structural setting of gold in the
Rice Lake area, Uchi Subprovince, southeastern Manitoba;
p. 213-221 in Current Research, Part B: Geological Sur-
veyofCanada, Paper 86-1B.
ROBERT, F., 1986, The Sigma gold deposit: Joint Annual
Meeting of the Geological Association of Canada, Minera-
logical Association of Canada, and Canadian Geophysical
Union (Ottawa, Ontario), Field Trip Guidebook 14, p. 2-10.
__ , and A. C. BROWN, 1986, Archean gold-quartz veins
at the Sigma Mine, Abitibi greenstone belt, Quebec, Part I:
Geologic relations and formation of the vein system:
Economic Geology, v. 81, p. 578-592.
SANBORN-BARRIE, M., 1987, A volcanic-hosted lode gold
deposit in a shear zone at the Cochenour Willans Mine,
Red Lake, Ontario [abstractJ; in Yellowknife 87: Geo-
logical Association of Canada, Program with Abstracts for
the Summer Field Meeting [unpagedJ.
SCHWERDTNER, W. M., M. D. STONE, K. OSADETZ, J.
MORGAN, and G. M. STOTT, 1979, Granitoid complexes
and the Archean tectonic record in the southern part of
northwestern Ontario: Canadian Journal of Earth Sci-
ences, v. 16, p. 1965-1977.
SmSON, R. H., F. ROBERT, and K. H. POULSEN, 1988,
High angle reverse faults, fluid-pressure cycling and
mesothermal gold-quartz deposits: Geology, v. 16, p. 551-
555.
STOTT, G. M., M. SANBORN-BARRIE, and F. CORFU, 1987,
Major transpression events recorded across Archean Sub-
province boundaries in northwestern Ontario [abstractJ; in
Yellowknife 87: Geological Association of Canada, Pro-
gram with Abstracts for the Summer Field Meeting [un-
paged].
TESSIER, A., F. ROBERT, and C. J. HODGSON, 1988, Struc-
tural interpretation of the Perron-Beaufor-Pascalis Nord
Archean gold deposit, Val d'Or, Quebec: Preliminary re-
sults [abstractJ: Joint Annual Meeting of the Geological
Association of Canada, Mineralogical Association of Cana-
da, and Canadian Society of Petroleum Geologists (St.
John's, Newfoundland), Program with Abstracts, p. A124.
WILLIAMS, H., 1987, Structural Studies in the Wabi-
goon and Quetico Subprovinces: Ontario Geological
Survey, Open File Report 5668, 163 p.
WILSON, B. C., H. HELMSTAEDT, and J. M. DIXON, 1984,
Shear fracturing, dike and vein intrusion and gold min-
eralization in the Red Lake Belt; pp. 177- 180 in Summary
of Field Work, 1984: Ontario Geological Survey, Mis-
cellaneous Paper 119.
WYMAN, D., and R. KERRICH, 1988, Alkaline magmatism,
major structures and gold deposits: Implications for green-
stone belt gold metallogeny: Economic Geology, v. 83, p.
454-461.
VD, L., R. DARLING, J. BELAND, and V. POPOV, 1987,
Structure of the Ferderber gold deposit, Belmoral Mines
Ltd., Val d'Or, Quebec: Canadian Institute of Mining and
Metallurgy Bulletin, v. 80, p. 68-77.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Structural controls during formation and deformation of
Archean lode-gold deposits in the Canadian Shield
R. MASON and H. H. HELMSTAEDT
Department of Geological Sciences, Queen's University, Kingston, Ontario, K7L 3N6, Canada
(received December 31, 1988; revision accepted February 20, 1989)
INTRODUCTION
Archean lode-gold deposits of the Canadian
Shield have commonly been regarded as "struc-
turally controlled," although the precise nature of
such control has been the subject of much debate.
Prior to about 1975, these deposits were either
considered fault related or as a combination of
intrusion and fault related. The papers in Struc-
tural Geology of Canadian Ore Deposits, pub-
lished by the Canadian Institute of Mining and
Metallurgy (Wilson 1948), and the treatise Gold
Deposits of the World, by Emmons (1937), present
evidence for and reflect such interpretations.
Renewed interest in gold deposits as a result of
deregulation of the gold price in the 1970's led to
interpretations of these deposits based on the
prevailing syngenetic theories for volcanogenic
massive sulphide deposits, and these interpretations
were presented in a review by Hutchinson and
Burlington (1984).
The past decade has witnessed the rise of a new
theory of ore genesis based on the idea of meta-
morphic hydrothermal fluids being focused in major
shear zones. As a result, there is now a widespread
belief that Archean lode-gold deposits formed
contemporaneously with or post-dated development
of regional penetrative fabrics in Archean green-
stone belts (Roberts, 1987; Sibson and others, 1988;
Hodgson, 1989). It is commonly assumed that the
deposits formed relatively deep in the crust (10-15
km), in shear zones developed in the transition zone
between an upper brittle and a lower plastic crust.
The Canadian Shield has yielded a large num-
ber of gold deposit discoveries, which include every
known type of Archean lode-gold deposit in a wide
range of litho structural settings and metamorphic
conditions. A ubiquitous association of deposits
with early faulting and with episodes of felsic
magmatism can be noted (see also Mason, this vol-
ume) and, to a greater or lesser extent, most deposits
appear to have been penetratively deformed and
metamorphosed. A significant issue surrounding
the controversy about these gold deposits, therefore,
is their temporal relationship to regional meta-
morphism and penetrative deformation.
Structural characteristics of Archean gold de-
posits are examined here, along with the hierarchy
of structures found associated with them. We sug-
gest that some of the present confusion in genetic
interpretations for these deposits can be resolved by
a better distinction between pre-ore structures, syn-
ore structures, and post-ore structures. We propose
that a number of Archean lode-gold deposits are
metamorphic analogues of more recent hydrother-
mal gold deposits, and that they do not represent a
unique class of metamorphogenic hydrothermal ore
deposits.
In: Basement Tectonics 8: Characterizatwn and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 625-631. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
625
626 R. Mason and H. H. Helmstaedt
STRUCTURAL HIERARCHY
Most Archean gold deposits occur in environ-
ments characterized by structural complexity and,
in particular, by repeated episodes of fracturing -
including pre-mineralization faulting, fracturing
related to mineralization events, and post-minera-
lization penetrative deformation and faulting.
Mineralizing events may be related to hydrothermal
activity associated with anyone or more of a num-
ber of volcanic and intrusive episodes in a particular
greenstone belt. Thus, the hierarchy of pre- and
post-mineralization structures will vary from dis-
trict to district, depending on the relative position of
mineralization in the geologic history of each
district.
A number of separate gold mineralizing events
associated with 2.73-2.67 Ga volcanism and intru-
sion can be identified in the Abitibi belt, and similar
relationships can be observed elsewhere in the
Canadian Shield. Volcanogenic massive sulphide
(VMS) systems are commonly associated with felsic
magmatic centers (usually bimodallcalcalkalic) in
the Canadian Shield, and some of these are enriched
in gold, including the Horne, Quemont, Dumagami,
and Bousquet deposits in the Abitibi greenstone belt
(Kerr and Mason, 1990). Syn-volcanic felsic intru-
sions related to these magmatic centers were also
associated with auriferous hydrothermal systems
(Goldie, 1982; Valliant and Hutchinson, 1982). All
of these gold deposits are related to the main
episodes of Keewatin volcanism and subvolcanic
intrusion characteristic of Canadian greenstone
belts. Moreover, they pre-date uplift, erosion, and
deposition of Timiskaming-type conglomerates and
other clastic sedimentary rocks that occupy local
fault-bounded basins and lie unconformably on the
Keewatin volcanics. Gold deposits at Kirkland
Lake are associated with syn-volcanic alkalic felsic
intrusions associated with local Timiskaming alka-
lic volcanism (Cooke and Moorhouse, 1969).
The final (and widespread) episodes of gold
mineralization across the Canadian Shield post-date
the Timiskaming-type events and are commonly
associated with late-stage felsic intrusions of so-
called "Algoman-type," as in the Porcupine gold
camp at Timmins (Ferguson, 1968; Mason and Mel-
nik 1986; Mason and others, 1988).
Apart from early synvolcanic VMS and intru-
sion-related hydrothermal systems, which were
emplaced prior to folding, most other auriferous
hydrothermal systems were emplaced into folded
strata. However, there is no record of early planar
fabrics related to folding, which appears to have
been caused by buckling without penetrative defor-
mation. Throughout the shield, the first penetrative
fabrics recognized in late Archean (2.73-2.67 Ga)
rocks appear to have developed locally around
granitoid intrusions of the Algoman suite or, on a
shield-wide basis, associated with crustal shorten-
ing during the subsequent "Kenoran" tectono-
thermal event. As a result of this last event, all gold
deposits, of whatever age, have been penetratively
deformed, with metamorphic recrystallization of
alteration mineral assemblages and mineralization
under greenschist- to amphibolite-facies conditions.
Local remobilization of gold is commonly associated
with this event. The currently exposed metamor-
phic state of a particular area - in the Superior
Province, for instance - is interpreted as a function
of the burial-uplift-erosion history of various parts
of the province during the Kenoran tectonothermal
event.
Thus, there are several gold mineralizing
events associated with late Archean felsic mag-
matism (Table 1), and intrusion-related hydrother-
mal systems are similar to one another irrespective
of the age of these intrusions. The older systems are
clearly not associated with penetrative fabric deve-
lopment. The only deposits that could be considered
as having been genetically associated with regional
metamorphism and penetrative fabric development
are those associated with terminal phases of the
Algoman intrusive suite and early stages of the
Kenoran tectonothermal event. However, the
characteristics of Archean lode-gold deposits are
still very similar, irrespective of their age relative to
penetrative deformation.
PRE-ORE STRUCTURES
Pre-ore structures include a variety of primary
and tectonic structural elements developed prior to
hydrothermal activity and gold mineralization. It is
clear (Table 1) that gold deposits related to the
Algoman intrusive suite will normally be associated
with more complex structural environments than
older gold deposits related to syn-volcanic intrusions
of Keewatin age. Mason (this volume) suggests the
existence of a spectrum of intrusion versus wallrock-
hosted gold deposits in which hydrothermal activity
may be related to intrusive or fault-related activity,
or to combinations of both. Within this scheme are
Structural controls on Archean lode-gold deposits, Canadian Shield
Table 1. Time and place of gold mineralization in the Abitibi belt.
Keewatin Volcanic Sequences One or more of several volcanic-sedimentary sequences developed 2.75-2.70 Ga.
Timiskaming Sedimentary
Sequences
Algoman Intrusive Suite
Au associated with VMS deposits and with syn-volcanic felsic intrusions in
terminal phases of above volcanic sequences - especially the youngest of these
"=2.70 Ga.
(Corfu and others, 1989)
Buckle folding and faulting
UPLIFT AND EROSION (regional unconformities)
Local fault-bounded basins developed 2.70-2.69 Ga.
Au associated with synvolcanic alkalic felsic intrusions related to alkalic
volcanism.
Felsic intrusions (diorite-tonalite-granodiorite-quartz-monzonite, etc. + albitites) -
mainly small epizonal plugs and plutons with a few batholithic complexes exposed at
present level of erosion 2.69-2.67Ga.
Emplacement of intrusions usually fault-lineament controlled.
627
Au associated with variety of small epizonal intrusions mainly emplaced about
2.69 and 2.67 Ga.
(Corfu and others, 1989)
Isolated intrusions may show development of marginal pentrative fabrics associated
with late ballooning.
All intrusions and gold deposits affected by subsequent Kenoran deformation.
Kenoran Tectonothermal
Event
Regional metamorphism and penetrative fabric development related to crustal
shortening. Partial melting and granitoid intrusions in lower mid-crustal sections
2.67-2.35 Ga.
Uplift - widespread faulting
Early Proterozoic Mafic Dikes Dikes orthogonal to regional strikes of supracrustal formations and penetrative
fabrics.
Widespread faulting
endless combinations of stratigraphic and intrusive
contacts, unconformity surfaces, fold culminations,
stratigraphic units with significant primary per-
meability control (volcanic flow breccias, hyaloclast-
ite breccias, conglomerates, etc) and, of course,
faults as local controls on hydrothermal fluid flow
and emplacement of mineralization.
Many felsic intrusions and associated auri-
ferous hydrothermal systems discovered in the
Canadian Shield are related to lineaments and/or
fault zones, reflecting their common need for assis-
tance in emplacement into the upper crust (Mason,
this volume). Such fault zones and lineaments are
regarded as more regional-scale structural controls
on felsic intrusions and related hydrothermal gold
systems. In addition there are numerous examples
in which anticlinal fold culminations have provided
loci for intrusion and containment of hydrothermal
systems associated with felsic intrusions.
The common assumption that gold mineraliza-
tion was synchronous with penetrative deformation
is not supported by careful analysis of field rela-
tionships. Most gold deposits (Table 1), irrespective
of their age relative to the sequence of geologic
events, formed in environments characterized by
lack of penetrative fabrics. Fracture-induced per-
628
R. Mason and H. H. Helmstaedt
meability is typical in wall rock environments
around ore structures; in addition, there is often
evidence for primary permeability at the time of
mineralization in volcanic breccias, tuff units, and
conglomerates.
ORE STRUCTURES
Ore structures in Archean lode-gold deposits
are all typical of structures formed by brittle failure
in the upper crust. In particular, they are charac-
terized by stockwork fracturing within and sur-
rounding individual deposits. Major zones of
carbonate-chlorite alteration that surround most
gold deposits in the Canadian Shield are closely
related to stockwork fracture-damaged wallrocks
(Mason and others,1988; Melnik, 1992; Mason, this
volume). Many ore bodies at Timmins, Kirkland
Lake, Malartic, and Yellowknife formed as quartz-
pyrite stockworks associated with felsic intrusions,
fault zones, and chimney or pipe-like breccia zones.
Veins and veins systems (en echelon arrays, steep
flat vein arrays, etc), vein swarms, and sheeted
veins exhibit all the features typical of brittle cracks
formed in a near-surface environment. There is
widespread evidence that fractures were either open
or permeable at the time of mineralization. Fur-
thermore, wallrock margins adjacent to ore struc-
tures characteristically exhibit intense micro-
stockwork fracture damage to wallrocks, which dies
off away from ore structures (as does hydrothermal
alteration). Wallrock alteration was controlled by
secondary permeability created by microfracturing.
The ore structures and accompanying hydro-
thermal alteration are very often localized about
felsic intrusive centers, which suggest a cause and
effect relationship, as pointed out at Yellowknife
(Helmstaedt and Padgham, 1986) and at Timmins
(Mason and others, 1988). The early fault system at
Yellowknife is a major regional system, whereas ore
structures that occur within the fault zone are very
localized and appear to be centered on a small por-
phyry stock (Helmstaedt and Padgham, 1986). The
relationships at Yellowknife are similar to those at
Kirkland Lake and Malartic in the Abitibi green-
stone belt.
Hydrothermal alteration associated with ore
structures usually includes early albitization and/or
K-sparlhiotite alteration followed by sericitic
alteration, with mineralization introduced late into
the densely fractured and altered environment.
Sericitization of wallrocks in Archean lode-gold
deposits is usually a good guide to mineralization,
but it is also important because it predisposes such
wallrocks to subsequent penetrative deformation.
Thus, shear zones are formed preferentially along
zones of phyllosilicate alteration associated with
hydrothermal activity and mineralization in pre-
existing structures.
POST-ORE STRUCTURES
Post-ore structures are easy to recognize at
most gold deposits because they affect ore structures
and associated hydrothermal alteration. The ear-
liest identifiable cleavages and linear elongation
fabrics in almost all gold deposits in the Canadian
Shield deform and recrystallize hydrothermal alter-
ation mineral assemblages under regional green-
schist- to amphibolite-facies metamorphic condi-
tions. Regional metamorphism was associated with
Kenoran tectonothermal events, which followed
emplacement of the Algoman intrusive suite across
the Canadian Shield and pre-dated emplacement of
the earliest (Matachewan-type) Proterozoic dikes. A
few deposits (e.g., Kirkland Lake) are preserved as
subgreenschist-facies zones in which penetrative
deformation is very weak or absent. Many other
deposits show correlation of high-strain zones with
sericitic alteration, and there are numerous ex-
amples where high-strain zones are discordant with
respect to ore structures and non-sericitic alteration
(Mason and Melnik, 1986; Melnik and Mason, 1986;
Kuhns, 1988; Helmstaedt and Padgham, 1986;
Mason and others, 1988).
An important aspect of penetrative deformation
and associated metamorphic recrystallization is a
late-stage local remobilization of quartz and gold
within ore shoots and in quartz-filled tension cracks
formed normal to cleavage. These features are
frequently, and uncritically, cited as evidence that
hydrothermal activity outlived penetrative defor-
mation (Wood and others, 1986; Roberts, 1987;
Hodgson, 1989).
Post-ore faulting is ubiquitous in gold deposits
of the Canadian Shield. Strike faults, which com-
monly exhibit high-angle reverse and, locally,
strike-slip movements (Poulsen and others, this
volume), follow the same zones as major pre-ore
regional faults. However, post-ore faulting clearly
disrupts ore structures and penetrative fabrics
alike, and the faults are typically brittle and usually
Structural controls on Archean lode-gold deposits, Canadian Shield 629
SEISMOGENIC
REGIME
- 10 km
j
DlsUtnce. km
"om C.W. Bdrnhern , 1$7 0
Mesothermal ./
GOld-auartZ
J

ACTIVE
METAMORPHISM
:"--fluids
Figure 1. A crustal perspective of Burnham's (1979) model for a porphyry-type hydrothermal system in which explosive
exsolution of a vapour phase from the water-saturated carapace of an epizonal felsic intrusion causes brittle failure of
wallrocks, leading to stockwork fracturing and development of breccia pipes. This contrasts with Sibs on's (1988) fault-valve
model for development of Archean mesothermal gold deposits associated with high-angle reverse faults at lower crustal levels
below the seismogenic regime and in the zone of active penetrative deformation and metamorphism. Common patterns of
stockwork fracturing, hydrothermal alteration, and development of hydrothermal breccias in gold deposits linked to both
models suggest common origins that are incompatible with contemporaneous penetrative deformation and metamorphic
recrystallization.
gouge-filled. They are not usually reactivated early
planar features but are newly created fractures that
follow the same general zone as early planar
features. The best example of this distinction is at
Kirkland Lake, where early faulting clearly con-
trolled Timiskaming sedimentation, volcanism, and
intrusion. Here, ore structures generally followed
the same trends but were localized about a small
felsic intrusive plug (Thompson, 1948). The Kirk-
land Lake fault system consists of brittle, gouge-
filled faults that form part of a major regional fault
zone following the same trend as earlier faulting.
The Kirkland Lake faults are offset by brittle cross
faults normal to the trend of the main fault zone,
and they are cut by Early Proterozoic dikes with the
same trends as cross faults (Nemcsok and others,
1989). Such relationships are common and can be
observed elsewhere across the shield.
DISCUSSION
The interpretation of complex fault zones and
penetrative deformation in hydrothermally altered
and mineralized environments as interrelated,
penecontemporaneous phenomena associated with
brittle-ductile shear zones (Groves and Phillips,
1987; Roberts, 1987; Hodgson, 1989) may be mis-
leading. The evidence suggests gold mineralization
occurred in several discrete episodes during the late
Archean volcanic-sedimentary-intrusive history of
the Canadian Shield, and this mineralization was
not related to a single regional metamorphic event
(the Kenoran tectonothermal event) that post-dated
the Algoman intrusive suite.
Archean lode-gold deposits in the Canadian
Shield exhibit features similar to those of Mesozoic
to Recent gold deposits that are associated with a
variety of epithermal and porphyry-type hydrother-
630 R. Mason and H. H. Helmstaedt
mal systems in recently active and fossil magmatic
arcs (Sillitoe, 1988, and this volume). Models for
such deposits are constrained by the physical neces-
sity for the most productive ore-forming processes to
develop in the top part (0-4 km) of the brittle crust
(Burnham, 1979; Burnham and Ohmoto, 1980).
These models are based on the concept of hydro-
thermal fluids being ex solved during felsic magma-
tism or evolving from meteoric water circulating in
convective cells above a heat source, with or without
a magmatic fluid input. In either case, mineralized
environments are characterized by brittle struc-
tures such as widespread stockwork fracturing and
phreato-magmatic breccia pipes and dikes (Burn-
ham and Ohmoto, 1980; Sillitoe, 1985).
Recognition of brittle structures and hydrother-
mal breccias and their association with hydrother-
mal alteration related to felsic intrusive centers in
most Archean lode-gold deposits of the Canadian
Shield invites comparisons with more recent
hydrothermal gold systems. Regional metamor-
phism and penetrative deformation associated with
the Kenoran tectonothermal event appear to post-
date and modify many gold deposits and their
hydrothermally altered environments. Ore struc-
tures and mineralization were products of typical
hydrothermal systems that developed close to the
surface in the brittle crust. Subsequent tectonic
activity depressed mineralized zones into various
deeper levels of the crust, below the threshold of
quartz plasticity, and in some cases such as Hemlo,
below the threshold offeldspar plasticity.
Figure 1 illustrates problems associated with
contemporaneous development of typical hydrother-
mal deposits, stockwork fracturing and hydrother-
mal breccias (Burnham 1979), and regional meta-
morphism in the transitional zone between the
seismogenic brittle upper crust and the plastic lower
crust (Sibson and others, 1988). Thus, the meta-
morphogenic "shear zone" model summarized by
Roberts (1987) is not appropriate for many Archean
lode-gold deposits. It does not address the incompa-
tibility of brittle structures, widespread fracture-
controlled permeability, and primary permeability,
with plastic features, metamorphic
recrystallization, and development of penetrative
fabrics. Nor does it explain mineralization in the
context of superposition of pre-ore, ore, and post-ore
structures.
REFERENCES
BURNHAM, C. W., 1967, Hydrothermal fluids at the
magmatic stage; pp. 34-76 in H. L. Barnes (ed.), Geo-
chemistry of Hydrothermal Ore Deposits: Holt,
Rinehart and Winston, New York, New York, USA, 670 p.
__ , and H. OHMOTO, 1980, Late-stage processes offelsic
magmatism: Mining Geology, Special Issue, no. 8, p. 1-11.
COOKE, D. L., and W. W. MOORHOUSE, 1969, Timiskam-
ing volcanism in the Kirkland Lake area, Ontario,
Canada: Canadian Journal of Earth Sciences, v. 6, p. 117-
132.
CORFU, F., T. E. KROGH, Y. Y. KWOK, and L. S. JENSEN,
1989, U-Pb zircon geochronology in the south-western
Abitibi greenstone belt, Superior Province: Canadian
Journal of Earth Sciences, v. 26, no. 9, p. 1747-1763.
EMMONS, W. H., 1937, Gold Deposits of the World:
McGraw-Hill, New York, New York, USA, 562 p.
FERGUSON, S. A., 1968, The Geology and Ore Deposits
of Tisdale Township, District of Cochrane: Ontario De-
partment of Mines, Geological Report 58, 177 p.
GoLDIE, R., 1982, Lithostratigraphy and the Distribution
of gold in the south-central Abitibi belt of Quebec; pp. 15-
26 in R. W. Hodder and W. Petruk (eds.), Geology of
Canadian Gold Deposits, Proceedings of the CIM Gold
Symposium: Canadian Institute of Mining and Metal-
lurgy, Special Volume 24,286 p.
HELMSTAEDT, H., and W. A. PADGHAM, 1986, Strati-
graphic and structural setting of gold-bearing shear zones
in the Yellowknife greenstone belt; pp. 322-346 in L. A.
Clark (ed.), Gold in the Western Shield: Canadian
Institute of Mining and Metallurgy, Special Volume 38, 537
p.
HODGSON, C. J. 1989, The structure of shear-related, vein-
type gold deposits - A review: Ore Geology Reviews, v. 4,
p.231-273.
HUTCHINSON, R. W., and J. L. BURLINGTON, 1984, Some
broad characteristics of greenstone belt gold lodes; pp. 339-
372 in R. P. Foster (ed.), Gold 82: The Geology, Geo-
chemistry and Genesis of Gold Deposits (Geological
Society of Zimbabwe, Special Publication 1): A. A.
Balkema, Rotterdam, The Netherlands, 753 p.
KERR, D. J., and R. MASON, 1990, A re-appraisal of the
geology and ore deposits of the Horne Mine Complex at
Rouyn-Noranda, Quebec; pp. 153-165 in G. Riverin and A
Simard (eds.), Northwestern Quebec PolymetaUic Belt:
Canadian Institute of Mining and Metallurgy, Special
Volume 43.
Structural controls on Archean lode-gold deposits, Canadian Shield 631
KUHNS,R J., 1988, The Golden Giant Deposit Hemlo,
Ontario - Geologic and Geochemical Relationships
Between Mineralization, Alteration, Metamorphism,
Magmatism and Tectonism: PhD dissertation, Univer-
sity of Minnesota, Minneapolis, Minnesota, USA, 458 p.
MASON, R, and N. MELNIK, 1986, The anatomy of an
Archean gold system - The McIntyre-Hollinger Complex
at Timmins, Ontario, Canada; pp. 40-55 in A. J. Macdonald
(ed.), Gold '86: An International Symposium on the
Geology of Gold Deposits (Toronto): Konsult Int. (Wil-
lowdale, Ontario), Proceedings Volume, 517 p.
__ , D. 1. BRISBIN, and S. AITKEN, 1988, The geological
setting of gold deposits in the Porcupine mining camp:
Ontario Geological Survey, Miscellaneous Paper 140, p.
133-145.
MELNIK, N., 1992, Hydrothermal Alteration, Minera-
lization, Metamorphism and Deformation at the Mc
Intrye and Hollinger Mines, Timmins, Ontario: PhD
dissertation, Queen's University, Kingston, Ontario,
Canada, 400 p.
__ , and R MASON, 1986, Deformation of mineralized
structures in the McIntyre-Hollinger Complex, Timmins,
Ontario [abstract]; pp. 112-113 in A. M. Chater (ed.), Gold
86: An International Symposium on the Geology of
Gold Deposits (Toronto): Konsult Int. (Willow dale, On-
tario), Poster Paper Abstracts Volume Volume, 181 p.
NEMCSOK, G., K. RATTEE, K. CATER, and T. N. LACKEY,
1989, Macassa Division Mine Geology: Lac Minerals
Ltd, Kirkland Lake, Ontario, Canada, Unpublished Report,
11 p., 8 figures.
ROBERTS, R. G., 1987, Ore deposit models - Archean lode
gold deposits: Geoscience Canada, v. 14, no. 1, p. 37-52.
SmSON,.R. H., F. ROBERT, and H. K. POULSEN, 1988,
High angle reverse faults, fluid pressure cycling and meso-
thermal gold-quartz deposits: Geology, v. 16, p. 551-555.
SILLITOE, RH., 1985, Ore-related breccias in volcano-
plutonic arcs: Economic Geology, v. 80, p. 1467-1514.
__ , 1988, Gold and silver deposits in porphyry systems;
pp. 233-2571n R W. Schafer, J. J. Cooper, and P. G. Vikre
(eds.), Proceedings of the Symposium on Bulk Mine-
able Precious Metal Deposits of the Western United
States: Geological Society of Nevada, Reno, Nevada, USA,
755p.
THOMSON, J. E., 1948, Geology of Teck Township and
Kenogami Lake area, Kirkland Lake gold belt: Ontario
Department of Mines, Annual Report, v. 57, pt. 5, p. 1-53.
VALLIANT, R 1., and R W. HUTCHINSON, 1982, Strati-
graphic distribution and genesis of gold deposits, Bousquet
region, northwestern Quebec; pp. 27-40 in R W. Hodder
and W. Petruk (eds.l, Geology of Canadian Gold De-
posits, Proceedings of the CIM Gold Symposium:
Canadian Institute of Mining and Metallurgy, Special
Volume 24,286 p.
WOOD, P. C., D. R BURROWS, A. V. THOMAS, and E. T. C.
SPOONER, 1986, The Hollinger-McIntyre Au-quartz vein
system, Timmins, Ontario, Canada: Geologic character-
istics, fluid properties and light stable isotopes; pp. 56-80
in A. J. Macdonald (ed.), Gold 86: An International
Symposium on the Geology of Gold Deposits (Toronto):
Konsult Int. (Willowdale, Ontario), Proceedings Volume,
517p.
WILSON, M. E. (ed.), 1948, Structural Geology of Can a-
dian Ore Deposits [Canadian Institiue of Mining and
MetaUurgy Special Volume]: Mercury Press, Ltd.,
Montreal, Quebec, Canada, 948 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Tectonic setting of Mesozoic gold deposits in the
Canadian Cordillera
WILLIAM J. McMILLAN and ANDREJS PANTELEYEV
British Columbia Ministry of Energy, Mines, and Petroleum Resources, Geological Survey Branch,
Victoria, BC, V8V lX4, Canada
(received September 19, 1988; revision accepted April 3, 1989)
ABSTRACT
Gold mineralization took place when plate tectonic activity initiated large-scale, deeply circulating,
hydrothermal flow regimes in convergent continental margin settings. Mineralization during Middle
Jurassic to mid-Cretaceous time also resulted when Paleozoic and early Mesozoic oceanic and island-arc rocks
of Superterrane I collided with the continental margin clastic wedge to produce the Omineca Crystalline Belt.
Similar processes took place outboard of the newly formed Intermontane Belt in mid-Cretaceous to early
Tertiary time when similar oceanic and island-arc rocks of Superterrane II were accreted to form the Coast
(Plutonic) Belt. The gold deposits formed are mesothermal-type veins similar to Archean greenstone-hosted
auriferous veins and California Mother Lode-type deposits. They developed in regional shear zones and
related dilational structures deep within deformation zones. Most deposits are near terrane boundaries that
are transcurrent or thrust faults. Examples are deposits in the Jurassic to Early Cretaceous Cassiar, Cariboo-
Barkerville, and Atlin Camps and in the Late Cretaceous Bridge River Camp, historically the largest gold-
producing district in British Columbia.
Numerous other, generally smaller gold deposits are controlled by local or district-scale structures and
thermal zoning related to intrusive or volcanic centres. Most are in Jurassic island-arc terranes. The host
rocks are calcalkaline to alkaline (shoshonitic), suggesting formation in island-arc and backarc tectonic
settings. Environments of ore deposition range from mesozonal through transitional to epizonal. The wide
variety of auriferous deposits formed includes vein, skarn, replacement, porphyry copper, breccia, and
hotspring-related deposits. Deposit distributions have been modified by Cenozoic tectonic events, particularly
large, right-lateral fault movements that segment the older terranes.
INTRODUCTION
The Canadian Cordillera contains a wealth of
mineral deposits, and gold deposits in particular
evoke both exploration and research interest.
Understanding the distribution patterns of Mesozoic
gold deposits in this geologically and structurally
complex area (Figure 1) requires knowing the tec-
tonic settings in which they formed and under-
standing the tectonic make-up of the Cordillera
(McMillan and others, 1987).
The Canadian Cordillera is a collage of allocht-
honous terranes welded onto the western edge of the
North American craton. Many of the terranes ori-
ginated far to the south of their present position
(Irving and others, 1980); some amalgamated en-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 633-651. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
633
634
W. J. McMillan and A. Panteleyev
route (Monger and others, 1982). Since docking,
large-scale strike-slip displacements at terrane
boundaries have juxtaposed unrelated terranes and
units within terranes have been disaggregated by
movements along other major faults.
Mesozoic gold-bearing deposits formed in three
main environments: at terrane boundaries, in is-
land arcs, and in backarc settings. Deposit types
range from volcanogenic massive sulphide to epi-
thermal and mesothermal vein types to skarns and
porphyries.
Gold Production from Mesozoic Deposits
More than half the recorded value of gold
produced in British Columbia has been byproduct
gold from base metal deposits of both volcanogenic
and, more importantly, porphyry types. Vein, re-
placement, and placer deposits are the most signi-
ficant primary gold producers. The major lode-gold
producing camps in British Columbia are at terrane
boundaries like Bralorne (130 tonnes gold), near the
basal thrust of the ophiolitic Sylvester allochthon at
Cassiar (about 5 tonnes), in island-arc settings like
Rossland (85 tonnes) and Stewart (58 tonnes), and in
backarc settings like Hedley (54 tonnes) and per-
haps Greenwood (34.5 tonnes).
Geographic Distribution of Gold Deposits
Two Mesozoic metamorphic-granitic belts run
the length of the Canadian Cordillera. These are
interpreted as suture zones where the leading edges
of major composite allochthonous terranes - Super-
terranes I and II - crumpled into the North Ameri-
can plate (Figure 2). Although this conclusion is
not universally accepted (see Price and Carmichael,
1986), Irving and others (1980) and Tipper (1984)
argued from paleomagnetic and paleontologic data
that individual elements of the superterranes ori-
ginated far to the south of the present location.
Monger and others (1982) further argued that, as
they moved northward, the terranes amalgamated
into two distinct superterranes that subsquently
"docked" with North America. Internal elements of
Superterrane I amalgamated in Late Triassic time
and consist of the oceanic Slide Mountain and Cache
Creek terranes and the island-arc Quesnel and
Stikine Terranes. Superterrane I collided with the
North American craton in Early Jurassic time,
MAIN DEPOSITS __ _
PFtOSPi:CTS
EPITHERMAL. DEPOSITS
BLACKOOME
Figure 1. Distribution of important gold deposits in the
Canadian Cordillera.
"" 190 Ma (Figure 3), giving rise to the Omineca
Crystalline Belt. By Middle Jurassic time, internal
elements of Superterrane II - mainly the composite
island-arc Alexandria and Wrangellia terranes -
were together. In the mid-Cretaceous, "" 100 Ma,
Superterrane II collided with the west side of Super-
terrane I, forming the Coast Plutonic Complex.
Presumably, the oceanic Bridge River terrane was
separate and apparently docked against Superter-
rane I, probably as an accretionary wedge, in ear-
liest Cretaceous time, "" 140 Ma (Monger and
others, 1982).
It is likely that relative collision vectors were
oblique to terrane boundaries during these amalga-
mation and docking events. Consequently, timing
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 635
,...--- r ----,
0EASfERN
o OUESro.:U.LIA
0=" CREE.
GSTJKltA
o VUKO CRYSTALliNE
o 8RIOGERIVR
GCASCAOtA
....---- 0 ----,
r-;I ALEMNOER GRAVINA
L.:..J NUTIOflN IGN
G ",.mAI\GLLtA
o PACifIC RIM
G!J CttUGACH

Figure 2. Diagram showing the distribution of alloch-
thonous terranes and Superterranes I and II in the
Canadian Cordillera (simplified after Monger and others,
1982).
of collision and related plutonic events can be
expected to have varied along the length of the
Cordillera. The curved to uneven shapes of the
rafting island arcs would be a major local influence
on variations in stress-regimes generated, and
resultant movement vectors would have produced
compressional, extensional, or transverse forces.
Resulting consuming and collisional events pro-
vided ground preparations for forming gold deposits
by producing deeply penetrating fracture systems
and by generating heat and magmas that favor
formation oflarge hydrothermal systems.
AGES AND TECTONIC SETTINGS OF MAJOR
CORDILLERAN MESOZOIC GOLD DEPOSITS
The earliest Mesozoic gold deposits are largely
related to consuming margin, island-arc processes.
Volcanogenic massive sulphide, gold-bearing por-
phyry, propylitic gold (Panteleyev, 1988a), gold
skarn, mesothermal to transitional vein and re-
placement, and epithermal vein deposits formed in
the Late Triassic and Jurassic island arcs of Ques-
nellia, Stikinia, and Wrangellia (Figure 4).
Docking events produced granite-related and
terrane boundary-localized precious metal deposits.
Younger terrane-boundary deposits are characteri-
zed by Mother Lode-type mesothermal gold deposits
hosted in disrupted ophiolite or mixed oceanic rock
assemblages. Mineralization is generally related to
collision-generated plutonism that is significantly
younger than the host rocks.
Post-docking granite and volcanic-related min-
eralization includes polymetallic veins, mantos, por-
phyry molybdenum, skarn, and epithermal deposits.
Deposits in Oceanic and Island-Arc Settings
Volcanogenic Massive Sulphides. Mesozoic
deposits in oceanic and island-arc settings formed
mainly before or during amalgamation. During
extensional events, some volcanogenic cupriferous
pyrite-type massive sulphide deposits formed on or
close to the sea floor, either at ocean ridges or in
backarc settings with or without generation of
oceanic crust. In the Alexander terrane, examples
are Anyox (Alldrick, 1986) and Windy Craggy (Mac
Intyre, 1986). Island arc-related volcanogenic mas-
sive sulphide deposits of Kuroko type also formed at
Britannia in Wrangellia (Stone and Payne, 1982)
636
W. J . McMillan and A. Panteleyev
COAST
Approximate time of accretion to North America and
present location relative to tectonic belts ... _ - __
North America . . .. .. .... , ....... , . . . . . . .
INSULAR BELT PLUTONIC
COMPLEX

--------
K
GJ
.... --
----- --- -
GN
J
w -
A
SUPERTERRANE I
Ca
'--
o
-
------SUPERTERRANE n-
p
Figure 3. Stratigraphic sections showing amalgamation and docking events for allochthonous terranes in the Canadian
Cordillera (after Monger and others, 1982).
and at Granduc (Grove, 1986) and Kutcho Creek
(Bridge and others, 1987) in Stikinia.
Gold-Bearing Porphyry Copper Deposits.
Gold-bearing porphyry copper deposits have been
the source of more than 40% of the gold produced in
British Columbia. Porphyries carrying significant
gold fall within both the alkaline and calcalkaline
classes (Drummond and Godwin, 1976, p. 59). Gen-
erally, important gold values occur only in porphyry
deposits in volcanic settings or with high-level por-
phyry intrusions (McMillan and Panteleyev, 1988).
Thus, the alkaline Copper Mountain, Afton, Galore,
Cariboo Bell, and QR deposits and the calcalkaline
Fish Lake and Schaft Creek deposits contain eco-
nomically important amounts of gold. Deeper,
plutonic-hosted porphyry deposits, like those in
Highland Valley have small but economically
important amounts of silver but little gold. This
accords well with observations in the southwestern
Pacific (Sillitoe, this volume) that only porphyries
related to high-level plutons (1.5-4 km depth) have
high gold signatures.
Porphyry deposits characteristically form from
large, granite-related hydrothermal systems, and
mineralization is mainly fracture controlled. Cop-
per Mountain and Afton (Barr and others, 1976) are
producing alkalic, gold-bearing porphyries in Super-
terrane I, and Island Copper (Cargill and others,
1976) in Superterrane II is a subvolcanic, calcalka-
line example. Island Copper gold values are highest
at the top of the mineralized zone in an area of
advanced argillic alteration but are also important
in the core of the deposit and in the propylitic al-
teration zone.
Mesozoic Gold Skarns. There are more than
100 gold-enriched skarns (those that carry> 1 ppm
Au) in island-arc settings in British Columbia
(Figure 5). Invariably, they are related to calcalka-
line, I-type intrusions (Ettlinger and Ray, 1988,
1989) emplaced into limy sediment-bearing succes-
sions. Both the intrusions and the country rock may
be skarnified. Gold in these deposits is generally
localized in areas of retrograde alteration contained
in larger zones of pyroxene and pyroxene-garnet
skarn. With few exceptions, garnet and pyroxene
minerals in the auriferous skarns are low in man-
ganese (Meinert, 1983) - a potential exploration
guide. Permeability contrasts within the impure
calcareous host rocks focused hydrothermal flow
and localized skarn alteration and mineralization.
The most productive gold skarn deposit known
to date in the Canadian Cordillera is the Hedley
deposit in south-central British Columbia. This
skarn deposit is within a backarc succession of Late
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 637
AGE OF GOLD DEPOSITS BY TERRANE - CANADIAN CORDILLERA
WEST
WRANGELLIA COAST BELT
STIKINIA
BRIDGE R.
QUESNELLI A
SLIDE
MTN
EAST
NORTH
AMERICAN

CINOLA
MT. WASHINGTON
ZEBELLOS
HARRISON LK
Minto
SKUKUM
- 50 + - ---- -1-------t- BLACKDOME -l----DUSTY-__ i-------+_ GREW CREEK -
EQUITY MAC
o
6
N
o
(f)
llJ
2
CAROLIN
Figure 4. Radiometric ages of selected precious metal deposits of the Canadian Cordillera (after Dawson and others, 1991) .
Triassic clastic and calcareous sediments that are
cut by slightly younger diorite to gabbro sills, dikes,
and stocks and by much younger post-mineral
granodiorite. Historical production from the mine
was almost 54 tonnes of gold; published reserves for
the new open pit operation are 8.9 million tonnes at
4.56 g/T gold (Ray and others, 1988).
The sedimentary succession formed in a tec-
tonically active backarc basin. Triassic mafic intru-
sions are most abundant at the eastern, fault-
bounded edge of the basin (Ray and others, 1988).
Skarn is extensively developed in and adjacent to
the diorites, and gabbros and ore bodies consist of
pyrrhotite, arsenopyrite with lesser pyrite, chalco-
pyrite, gersdorffite, sphalerite, and gold. Hedley
mineralization occurs in retrograde alteration zones
with the host garnet-pyroxene endo- and exoskarns.
Apparently, the diorites and the ore deposits are
localized at a hinge zone that marks the transition
from the eastern shallow water facies to the western
shelf facies in the backarc basin (Ray and others,
1988).
Mesothermal Gold Deposits in IslandArc
and Backarc Settings. Mesothermal gold deposits
are also common in island-arc environments. The
vein mineralogy and types of alteration are similar
to those in oceanic host rocks (see following), but
base metal concentrations in the veins tend to be
higher and mixed volcanic and sedimentary rocks
comprise the wallrocks. The Rossland Camp in
Quesnellia (Superterrane I) is an example of this
638 w. J. McMillan and A. Panteleyev
GOLD ENRICHED SKAANS IN BRITISH COLUMBIA
IAFTER BC GEOLOGICAL SURVEY MINFllE)
100 l'OO
1lO...
(EHlinger and Ray. 1967)
Acapulco
OF THE 340 SKARN
OCCURRENCES IN THE
PROVINCE.
APPROXIMATEl.Y \f,
ARE GOLD ENRICHED
Figure 5. Distribution of auriferous skarns in British Columbia (after Ettlinger and Ray, 1988).
kind of deposit. There, copper-gold veins consist of
chalcopyrite and pyrrhotite with lenses of quartz
and calcite. They fonned by replacing wallrock
along well-defined fractures and by filling fracture
and fault zones (Fyles, 1984). Veins in the camp
fonn steeply N-dipping conjugate sets that strike
mainly 60_70 and 120, respectively. The major
veins are related to N-trending structures and lie
north of the NE- to E-trending 'Rossland Break'
(Figure 6). The Rossland Break is a Jurassic zone of
weakness separating structural domains within the
Rossland Group; it was also a locus for intrusions
and later faulting.
The deposits are localized in an area that is
characterized by large scale faults with associated
serpentinite bodies, and by extensive areas of horn-
felsic alteration within a succession of Jurassic
shoshonitic volcanics, related monzonite and syenite
intrusions, and many dikes (Fyles, 1984). Tertiary
molybdenite porphyry deposits on Red Mountain,
and bonanza epithennal gold vein deposits like
those in Little Sheep valley to the south, overprint
the mesothermal system. Production from gold
deposits in the camp was 72.8 tonnes gold, 95.2
tonnes silver, and 56 000 tonnes copper from 5.6
million tonnes of ore.
Epithermal Gold Deposits. Epithennal depo-
sits in the Canadian Cordillera are associated with
subaerial to submarine, felsic to intermediate
volcanic rocks in areas of block faulting or associa-
ted with caldera structures. Typically gold to silver
ratios are low (Nesbitt, and others, 1986) and veins
have related low pH alteration zones. Mineraliza-
tion is structurally controlled and best developed
near volcanic centers where high-level intrusions
cut subaerial volcanic rocks (Panteleyev, 1988b).
Veins and vein stock works fonn under hydrostatic
conditions and tend to be characterized by sym-
metrical zonation and open space-filling textures.
Typical vein minerals are quartz, chalcedony, cal-
cite, hematite, adularia, barite, pyrite, chalcopyrite,
argentite, electrum, and acanthite. Gold-silver
ratios range from 1:5 to 1:20. Related alteration
includes silicification, pyritization, and propylitic,
potassic, or advanced argillic types. Fluid inclusion
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 639
IUTlA.U
Ktrron JorlMtlon
Volcan.c roclS
JUAASSIC
CJ Co,-yell Syetutt'
c:::J
iiiranltlC phJtOfti
lDCII"Llonitll!
1::::::1 troup
St<ll IIIC'fIU , Mile
Figure 6. Geologic setting of the Rossland mesothermal
gold deposits, south central British Columbia (after Fyles,
1984).
studies suggest that ore-fonning fluids had low
salinity, moderate temperature, and relatively low
CO
2
contents (Hedenquist and Henley, 1985).
Extensional environments, in either continental,
island-arc, or backarc settings, are likely target
areas for epithennal deposits.
Deposits in the Toodoggone area (Figure 7) in
the Stikine terrane of north-central British Colum-
bia are Early Jurassic but have all the characteris-
tics of younger epithennal deposits. The Jurassic
Toodoggone Volcanic Belt is 100X20 km. Dacitic to
andesitic flows and pyroclastics in the belt are cut
by subvolcanic stocks and feeder dikes. Typically,
Toodoggone volcanics are relatively potassic and
range in character from feldspar-porphyry andesitic
and dacitic flows through ash-flow tuffs to epiclastic
rocks.
The Toodoggone volcanics are dominantly sub-
aerial and apparently underlie marine volcanics of
the Jurassic Hazelton Group. Both these sequences
unconfonnably overlie Late Triassic Takla Group
basalts and andesites. Cretaceous seccessor basin
sediments with potential for paleoplacer deposits
(Eisbacher, 1974) onlap the beltfrom the west.
The best of the known deposits in the Toodog-
gone Belt lie along its volcanic-plutonic axis. Baker
Mine, which is hosted by Takla basalts, operated
from 1980 until 1983 and produced 1.17 tonnes gold
and 23.1 tonnes silver; Lawyers began production in
1989 with reserves in excess of 1.9 million tonnes,
and the AL deposit had pilot-test production in
1986.
Isotopic age detenninations from the Takla
basement average 210 Ma. The Toodoggone vol-
canics erupted intennittently over a 20 million year
time span, with two main pulses at 204 and 182 Ma,
and mineralization occurred between 190 and 170
Ma (Diakow, 1985).
On the Silver Pond property, local zones of
intense argillic alteration and smaller areas of silica
flooding lie within broad zones of propylitic altera-
tion. Ore at Baker Mine is in quartz stockworks
within zones of silica alteration, and Lawyers has
good grades in quartz-amethysts stockworks. Ore,
gangue, and alteration minerals are similar to those
in typical epithennal deposits, with the exception
that native silver occurs locally.
A large vertical range within the hydrothennal
systems is exposed by block faulting in the Toodog-
gone area. The shallowest systems are siliceous,
alunite-clay cappings and fossil hot spring acid-
sulphate alteration zones (low pH acid caps); some,
like AL, have associated gold-barite mineralization.
The main epithennal deposits are those with either
amethystine quartz and few base metals or base
metal-bearing quartz-carbonate veins. Pyritic, seri-
cite-altered, precious metal-enhanced, intrusive-
related porphyry copper-molybdenum type deposits,
like the Porphyry Pearl prospect, formed at deeper
levels. Arsenic and antimony-rich cupriferous py-
ritic gold-silver deposits (hot and/or deep acid-
sulphate type) with enargite-alunite assemblages
are apparently rare in the spectrum of gold deposits
in the Canadian Cordillera.
640
W. J . McMillan and A. Panteleyev
LEGEND
CRETACEOUS
SUSlut
Congbmerllt. sands10nt
JURASSICS
Hazehon Group
Toodoggone Vo!ca";cs
'Gray Dacite' - Ash flows
f: :: j Ou.li1.lOSli plagiOdaSe porphyry !lows
c:J Ash 11ows, cryslat lUff, epidasflCS
TRI ASSIC
Takla
1:: :::::1 Basalt, anc:lll.Ut
PALEOZOIC
Asilka
Limestone. Chen, argilllle. metavolcarics
JURASSIC
GranodiOtU. qullU tl"CNonh
Gold Deposit .. , )(
o "

... -
Figure 7. Geologic setting of the Toodoggone epithermal gold deposits (after Diakow and others, 1985).
Transitional Epithermal to Mesothermal
Deposits. Transitional deposits in island-arc set-
tings have aspects of both epithermal and meso-
thermal or even porphyry copper deposits. Ex-
amples are found in mining camps such as Stewart
and Sulphurets in the Stikine terrane of north-
western British Columbia. Production from Stew-
art, which was continuous from 1918 to 1954 and
intermittent until 1968, yielded 58 tonnes of gold
and 1270 tonnes of silver. The two main historical
producers, Silbak-Premier, with present reserves
(mineable by open pit) of 5.9 million tonnes at 1.69
grI' gold and 80.23 grI' silver, and Big Missouri, with
reserves of 1.8 million tonnes at 3.6 grI' gold and and
29.49 grI' silver, were put back into production in
1989 (Randall, 1988). Mineralization is Jurassic
and, within this consuming margin setting, is rela-
ted to subvolcanic plutons and dikes cutting sub-
marine to subaerial island-arc volcanic and epi-
clastic rocks.
The Hazelton Group volcanic pile in the Stew-
art area (Figure 8) consists mainly of andesitic
pyroclastics and flows that are overlain by andesitic
epiclastic rocks and capped by a felsic volcanic unit
(Alldrick, 1985). The pile is intruded by the K-
feldspar megacrystic Texas Creek granodiorite,
dated at 200 Ma from zircon analyses, and Premier
Feldspar Porphyry dikes dated at 190 to 185 Ma
(Alldrick and others, 1986). Much of the ore is
probably related to the larger intrusions at 190 Ma
(Alldrick and others, 1986). There are also Tertiary
silver-lead-zinc deposits, like Silverado, that are
related to emplacement of the Coast Plutonic
Complex. Plotted lead-isotope analyses form
separate clusters for Jurassic and Tertiary events.
Lead in the Jurassic deposits is from a volcanic
source, whereas that in the Tertiary deposits has a
crustal component (Alldrick and others, 1987).
Ore zones at Silbak-Premier are generally at
Premier Porphyry dike-Hazelton andesite contacts
near terminations of the dikes. Wall rocks are
variably andesitic fragmental volcanics to acid
pyroclastics. Auriferous veins occur within dense,
multicyclic networks of reticulate quartz veins or in
silica-flooded zones. Gangue minerals are typically
quartz, K-feldspar, and calcite as stockworks and as
open-space fillings in breccias. Quartz-sericite and
carbonate alteration zones halo mineralization. Ore
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera
STRATIGRAPHIC COLUMN
STEWART AREA
SEDIMENTARY SEQUENCE
FELSIC VOLCANIC SEQUENCE
1. PremIer Porphyry Flow, ond Tuff,
Premier Porphyry Dykes
(194 Ma)
- Upper Siltslone Member
Texas Creek Hornblende
Granodiorite Stock
(195Ma)
Lower Siltstone Member
weST
P., ETRE$
E-W CROSS-SECTION THROUGH THE SILBAK-PREMIER MINE
STEWART AREA
MOUNT STEVf:NSOU
M c. SHOWlNO
EAST
CASCADE CREEK
TOPS
MEASURED OI PS
JURASSIC
Hazelton
m Epiclastic sequence
\'lOODBSNE DEPOSIT
INTERNATIONAl BORDER
Andesit ic flows and pyroclastics
I2J Lower and upper siltstone members
FOAl< PROSPEOT
TERTIARY
SILBAK- PREMIER
MINE

'I'll'll \IV'll v"v v v VV V'll II
VVVVVVY'IIYVVVVVVV
VVVVVVY'IIVVVVVVVYVVVVVVVVVVVyy
VVVYVVY'II v VVYVYVVVYV"'VVVVVYVVY
VVy vvv VVVVY V'll '11'11'11 'IVY V'll V'll 'IV IIVY V
VVVvVVYVVVVVVVVVVVVVYVVYVVVVY
VVY V'IIV VVVYV V'll V'll V 'I'll'll VV vvvv vvv V
VVVvYVY'IIVVVVVVVVVYVVVYVVVVVYV
VVVVVVY'IIVVYVVVVVVvvvvvvvvvvvv
VVVVVYVVVVVVVVVYVVVVVVvvvvvvv
vvvvvv VVVVYVVVVYVYVVVVVVvvvvv
VVVVVVVVVVYVVVVVVYVVVVVVvvvvv
VVVVVYVVVVYVVVVVVVVVVVVVYVVVV
GJ Hornblende quartz monzonite
JURASSIC
ITiKl Texas Creek hornblende granodiorite
o Premier Porphyry - dykes and flows
641
Figure 8. General geological setting of the Stewart and Sulphurets areas (after Alldrick, 1985; Britton and Alldrick, 1988).
minerals are electrum, tetrahedrite, polybasite, and
pyrargarite, with local galena and sphalerite and
rare arsenopyrite (Wodjak and Randall, 1986). The
ore shoots are variably interpreted to be syngenetic
and stratiform (Meade, 1986) or to be epigenetic and
cut stratigraphy (Alldrick, 1985).
642 W. J. McMillan and A. Panteleyev
Depth zoning is a characteristic of the camp.
Deposits formed at the shallowest depths are rela-
tively silver-rich and epithermal, whereas the deep-
est formed are gold-rich and mesothermal; other
deposits are transitional in character. Gold/silver
ratios vary from < 1:200 in shallow deposits, to 1:50
in transional deposits, to 1:1 in the deepest, meso-
thermal deposits (Alldrick, 1985). The Silbak-
Premier deposits mined to a depth of "'" 100 m were
true bonanza ore, with gold, electrum argentite,
polybasite, and some native silver (Bacon, 1978).
Mesothermal deposits include the deepest Silbak-
Premier orebodies and Scottie Gold. At Scottie,
gold-silver mineralization occurs in a series of para l-
leI veins of massive pyrrhotite and pyrrhotite-
pyrite. Disseminated base metals within zones of
intense chlorite alteration and hematitic siliceous
flooding fringe the veins. Production was from three
main veins within a series ofladder veins in a more
steeply dipping shear or fault zone. Andesitic lapilli
tuff and tuff breccia host these deposits.
In the Sulphurets area, interest in the 1960's
and 1970's was focused on local areas of porphyry
copper-molybdenum mineralization within exten-
sive gossans. The porphyry mineralization occurs in
quartz-sericite-pyrite alteration zones near syenitic
to feldspar porphyry intrusions. Alteration adds
potassium, barium, and sulphur and removes sod-
ium, calcium, maganese, and carbon dioxide. Path-
finder elements for ore at Sulphurets are gold, lead,
zinc, antimony, copper, barium, and cadmium (Le-
febure and others, 1987). The hydrothermal system
that produced the alteration and mineralization was
huge, "'" 10 X 3 km, and exploration target definition
is made difficult by both destruction of original
textures and minerals by alteration and by later
oxidation.
In the 1980's, mining companies discovered in-
teresting gold and silver mineralization within the
hydrothermal system and the focus shifted. The
best precious metal grades occur in quartz veins in
silica-flooded zones along NW-trending faults
(Drown and other, 1987). Alteration zoning extends
outward from quartz through quartz-sericite-pyrite
to sericite-pyrite to chlorite-carbonate or sericite-
carbonate with minor pyrite (Lefebure and others,
1987). Where recognizable, the wall rocks are ande-
sitic volcanics and epiclastics.
At the Sulphurets property quartz veins and
stockworks that host the mineralization are multi-
stage. Quartz-pyrite veins carry sphalerite with or
without electrum, tetrahedrite, sulphosalts, pyrar-
gyrite, polybasite, and some galena. Ruby silver
minerals are relatively late in the paragenetic
sequence. In the West Zone (Figure 9) geologic re-
serves are ""'850,000 tonnes of 15.7 g/T gold and 689
g/T silver (Mining Review, May/June, 1988); the
gold/silver ratio is "'" 1:40.
CoUision and Accretion-Related Deposits:
Mesothermal Gold Deposits in Oceanic Host
Rocks
Deposits in oceanic host rocks do not appear to
be arc-related; often they are situated (Figure 9) at
or near major, Cordilleran-scale tectonic boundaries
that may also be terrane boundaries. Ore zones are
structurally controlled and mineralization is gener-
ally in veins and tension gashes generated by trans-
verse faulting. The host rocks, which can be dis-
membered ophiolites or assemblages consisting of
mixed island-arc and marine clastic rocks, are
generally intruded by felsic to mafic plutonic rocks
and serpentinites. Veins consist mainly of quartz,
carbonate, and albite with lesser pyrite, arseno-
pyrite, scheelite, graphite, base metal sulphides,
native gold, and gold tellurides. They are ribboned
to massive and multicyclic. Replacement ore occurs
locally. As Taylor (this volume) described for the
Mother Lode District, alteration-related gold values
occur in wallrock adjacent to veins but are of a
different generation than gold in the veins, which is
generally late stage. Gold is high relative to silver,
with ratios ranging from 1:1 to 10:1. Wallrock and
wallrock inclusions in the veins are variably car-
bonatized, pyritized, silicified, or sericitized; other
typical alteration minerals are albite, chlorite, fuch-
site (mariposite), talc, and graphite. Listwanitic
alteration, where present, generally replaces ultra-
mafic rocks. Alteration adjacent to veins is zoned,
and alteration selvages are narrow to several
meters wide.
Veins in mesothermal ore deposits register
episodic passage of hydrothermal fluids. Veins are
ribboned, sealed then re-opened, brecciated, and
recemented, and there is evidence of open space
porosity (Taylor, this volume). Gold is deposited in
the altered country rock adjacent to veins but late-
stage, fracture-controlled gold is more economically
important. For the Mother Lode District, Taylor
argues that the host fault system persisted for at
least 30 million years and orginated 10-15 million
years before gold-mineralized veins were deposited.
In deep-seated systems, where rock to water ratios
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 643
o 100 200 300
KILOMETRES
LEGEND
OCEANIC ASSEMBLAGES
CACHE CREEK __ Cc
QUESNE LLiA ___ Q
EASTERN ________ E
WRANGELLI A _____ W
ULTRAMAFIC ROCKS _ -
Figure 9. Distribution of oceanic terranes and related mesothennal gold deposits in the Canadian Cordillera (modified after
Monger, 1972).
are high, waters become isotopically heavy and do
not resemble seawater, magmatic, or meteoric
water. In mesothermal systems, fluid chemistry is
rock buffered. Oxygen fugacity and pH do not vary
significantly, and thermal gradients are weak. The
fluids are CO
2
rich, and deposition of gold is ap-
parently related to either cooling, rapid drops in
pressure, CO
2
effervescence, or mixing with cooler,
possibly meteoric waters. Effervescence, which
would destablilize gold bisulfide complexes (Sibson,
this volume), may be triggered by pressure drops
caused by fault movements. Alternatively, Kerrick
(1983) argued that ultramafic rocks act as CO
2
'sinks' and mediate gold precipitation, raising the
possibility that gold is transported as carbonyl or
related complexes.
The Bridge River Gold Camp, as an example,
produced "" 130 tonnes (4.1 million ounces) of gold,
644 W. J. McMillan and A. Pantel eye v
predominantly from the Pioneer and Bralorne
mines along the Cadwallader-Fergusson-Empire
fault system; there is a possibility of renewed
production from the Bralorne Mine. Present pub-
lished reserves are just under 1 million tonnes of "'" 9
g/T gold. The Bridge River terrane (Tipper, 1981) is
a fault-bounded slice of oceanic rocks that was
obducted onto Superterrane I. It consists of basalts,
ribbon cherts, and argillites of the Permo-Triassic
Bridge River Group; calcareous argillite, argillite,
and generally pillowed greenstones of the Late
Triassic Cadwallader Group; and the Shulaps ophio-
TERTIARY
D Mainly volcanic t OC
J RA I TO PPER CRETA EO S
D Mainly volanic and roc
D Graywacke:. arkose:. conglomerate. andc lie. ba:wlh
UPPER TRIASSIC
D yaughton Gro\Jp limestone
D Cadwallader Group argillite. grccruuonc,limt5.IOne,
diorite
D Metamorphosed e.quivalent or Cad"",-alladc:r Group
PERMIAN TO TRIASSIC
D Bridge River (Fergusson) Group chert. argillite.
bas3h. phylli.e
TRIASSIC OR OLD R
D Uhramaric: rocks
CRETACEOUS TO ECCE 'E
D Plutonic rocks mainly granodiorite and quartz
monzonite
METAL ZONING
gold,jlver (AuAgl ___
on.imony(Sb) ___ _
Mercury (Hg) ____
Figure 10. Geologic setting and metal zoning of the Bridge River mining camp, southwest British Columbia (modified after
Woodsworth and others, 1977).
60
Mf:1RfS
t At( olu.a 10lUIU'-,
r J..,t*ro y eno ,,..,,
I.PJ'fIl'TI' UUIC
( ... lllMltr "0041
;,.:<"'4 ""' .... Ul' ;. lnOwold

1>(1011 "'" 10 ULAIIle
kidte t", .. r (fti"",IoOI"I) "0\41 - 1,., HLh.,
boIl. II, pb.,. lIl1.
U!lUI( C. Q;Oft
ulttlllM'tc rO(h
..,U1
"-QU," ,,.,.,,.,.. dl""" . , fIoI'rCHW1dIU
CJ $.OCe,rll'l t t.

Figure 11. Fault and vein systems in the Bralorne mine area (after Leitch and Godwin, 1987).
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 645
lite (T. Calon, personal commun., 1988) complex.
Conflicting isotopic age data, stratigraphic rela-
tionships, and tectonic interpretations of the area
are a source of much debate at this time (see Leitch
and Godwin, 1986, 1987; Church, 1987; Church and
others, 1988, Glover and others, 1988).
In the camp (Figure 10), Leitch and Godwin
(1986,1987) and Church (1987) describe the general
geologic setting and Leitch and Godwin (1987) the
results of isotopic dating and analyses. The Bra-
lorne Diorite, a Paleozoic quartz diorite (270 Ma),
was cut by a large, dike-like body of Soda Granite of
similar age, it was subsequently intruded by Cre-
taceous, pre-mineral albitite bodies (91 Ma) and
then by post-ore porphyry dikes (85 Ma). Thus,
mineralization took place in Late Cretaceous time,
shortly after Superterrane II collided with North
America, and apparently is related to granites
emplaced during formation of the Coast Plutonic
Complex (Woodsworth and others, 1977).
Igneous rocks in the camp are characteristically
low in potassium and relatively high in sodium
(Leitch and Godwin, 1988). Presumably as a result
of greenschist facies metamorphism or alteration
associated with mineralization, plagioclase in even
the most basic intrusives is albite (Ano to AnIO).
Near veins, alteration causes a sharp increase in
calcium, sodium is leached, and potassium is added.
A huge hydrothermal system was generated in
the large-scale, deeply penetrating Cadwallader-
Fergusson-Empire fault system. Quartz veining
and gold mineralization formed along 6 km of strike
length and persist for > 1.5 km down dip. Depth of
formation of the mineralization is not known with
certainty, but related deformation was brittle.
Mineralization is fracture controlled and veins are
parallel to, and in splays off, the NW -striking
Cadwallader-Fergusson-Empire fault system. The
major veins are in the splays, with a dominant trend
of 100
0
and an average dip of 70
0
NW (Figure 11).
Woodsworth and others (1977) concluded that min-
eralization post-dated major transverse movement
on the fault system, and the system is pinned by the
younger 70 Ma Coast intrusions (Glover and
Schiarizza,1987). Perhaps some of the elongation of
Bridge River intrusions along the Cadwallader
system is related to shearing during transverse
fault offsets.
In the ore deposits, quartz veins are ribboned to
massive, with sulfide-rich wall rock septae contain-
ing arsenopyrite, pyrite, gold, pyrrhotite, and local
tetrahedrite, chalcopyrite, sphalerite, and galena.
Sulfide in the veins averages 1-3%.
Throughout the depth-range mined, there was
no change in alteration or vein mineralogy. Altera-
tion a.round veins is zoned from inner quartz-
sericite-carbonate-pyrite through quartz carbonate
to outer zones with carbonate-altered then chlorite-
altered amphibole. Presumably when ultramafic
rocks are altered, fuchsite is also present and the
alteration is listwanitic. Alteration selvages range
from a few centimeters to several metres wide.
The camp is characterized by large-scale metal
zoning outward from the Coast intrusives (Woods-
worth and others, 1977). The metal zones define a
NE trend that represents a pre-mineral fracture
zone. A core gold-silver-arsenic-tungsten-molyb-
denite zone gives way to fringing antimony-silver-
gold-arsenic, mercury-antimony with or without
gold, then mercury zones (Figure 10). Nesbitt and
others (1986) relate formation of stibnite and
cinnabar to successively shallower depths and cooler
areas in the vein system. Mineralization apparently
spanned a considerable time interval. The main
gold veins formed at 91-85 Ma, but some veins cut
Tertiary dikes. Northeast of the Bridge River
Camp, along the zoning trend, deposits become
progressively younger. The Poison Mountain por-
phyry copper-molybdenum-gold deposit is dated as
= 58 Ma (Glover and others, 1988) and, further to
the northeast, the Blackdome epithermal gold
deposit is dated as 51 Ma (Faulkner, 1986). Con-
sequently, the zoning may be more complex than it
first appears; it could represent overlapping of
several events along a long-lived NE-trending
structural zone.
Mesothermal gold deposits in the Coquihallah
area, near Hope, British Columbia, are spatially
related to the major Hozameen fault system and
associated with serpentine, gabbro, and diorite of
the Coquihallah Serpentine Belt (Ray and others,
1983). The Hozameen Group country rocks are
oceanic cherts, graywackes, and greenstones, and
most veins occur in graywackes near greenstone
contacts. Related carbonate and albite alteration
zones fringe the veins. A tectonic reconstruction by
Monger (1985) juxtaposes Hozameen and Bridge
River rocks, and right-lateral offset was apparently
70-90 km.
At Cassiar, in north-central British Columbia,
the general lithologic setting of gold deposits at and
around the Erickson Mine (Figure 12) is similar to
that of the Bridge River Camp. Unlike the Bridge
646
W. J. McMillan and A. Panteleyev
Berube,
"JIll.

N
I
CRETACEOUS
Quaru: ... ein S)'$lerns
SYLVESTER All..OCHTI10N
TRIASSIC
Sil[.S[ont. argi ll ile. graywa(:ke. conglomerate
limestone
PEN SYLVANIAN
Augite porphyry, minor limeMone and rebie
pyroclastics
MISSISSIPPIAN TO PERMIAN
o chen. argillite 11SS(:mblagc, limc-slone,
diabase 3Ild andesite sills
AUTOCHTI10 OUS ROCKS
CAMBRIAN TO OEVO IAN
Sedimentary rocks
Figure 12. Geologic setting of the Cassiar mesothermal gold deposits (after Panteleyev and Diakow, 1982).
BRITISH COLUMBIA EPITHERMAL MODEL
=0
0 0
o SI NTER $UULIMA.TE5 BROADLANDS. WAJOTAPU

,OA'"N''''' \ /
iIIIASt!
- 111;",
MET AL.S 1 NII.Av
IftIIx.:$ lkm
\ I
..
F,e ..
II.L8[IH$ HUMP
SllVERPOHD
MOOSIE
PORPHYRY
PEARL
FIN. KEMESS
McLAUGHlIN. I!IOR!ALIS
HEARTPfAJ(.S
ROUHDM'fN.
SAM GOOSlY (EQUITYI
SCOTTIE
ROSSL.lHD
TOODOGGONE OTHER DEPOSITS
AREA OEPOS'TS
HOt$F>RI NG$
S SIUeA
CAP'PINGS,
.. L CLAY
100"C ( $Ow pH. del sulpiMla l
us'C
.J. EPITHERMAL V1;I N$
,
ADIIANCIED AAGILUC
TRANSITION lONE

SKARN
PORPHYRY Cu. Mo
DEPOSIT TYPES
Figure 13. Depth zoning model for precious metal deposits in the Canadian Cordillera (after Panteleyev, 1988b).
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 647
Table 1. Characteristics and differences of Cordillera gold deposits.
Depth
Circulation pattern
Related plutonism
Ore fluid
Mesothermal
Deep
Tectonically driven
Deep circulation
Mesozonal to epizonal
Alkaline to neutral
Epithermal
Shallow
Shallow circulation
Subvolcanic
Alkaline to acidic
Mixed source, equilibrated with meta-
morphic rocks, reduced, high CO
2
, dilute
Effervescence likely
Meteroric, not equilibrated with host
rocks, oxygenated, less CO
2
, dilute
Boiling likely
River deposits, Cassiar gold veins occur within
thrust sheets near the base of the obducted Sylves-
ter allochthon. In Cassiar, most gold mineraliza-
tion occurs within late Paleozoic to Triassic rocks of
the Slide Mountain terrane within the allochthon.
Gold is mainly in small shoots within quartz veins
filling persistent fracture systems in greenstones
and at greenstone-argillite contacts (Panteleyev and
Diakow, 1982). Normal mesothermal alteration
zoning occurs in country rock adjacent to veins and,
in addition, some veins are highly graphitic (Sketch-
ley and Sinclair, 1987). Listwanite alteration zones
are not generally hosts for veins but, at least locally,
are spatially related to them and grades in veins
improve close to listwanite zones. The native gold
mineralization is associated with pyrite, sphalerite,
tetrahedrite, chalcopyrite, and arsenopyrite. Ske-
tchley and others (1986) dated the mineralizing
event at 129 Ma.
The setting of gold mineralization in the Atlin
Camp is generally similar to that of Bridge River;
however, at 160 Ma, mineralization in Atlin is older.
The veins are hosted in oceanic rocks of the Cache
Creek terrane, and listwanite alteration is associa-
ted with gold mineralization in the Atlin area (Ash
and Arksey, 1990).
SUMMARY AND CONCLUSIONS
Mesozoic gold deposits in the Canadian Cordi-
llera are related to island-arc development and to
plate collision and accretionary processes combined
with more localized heating and magmatic events.
Island-arc deposits are generally related to subduc-
tion at consuming margins in volcanic-arc or more
distal backarc settings, where porphyry copper-gold,
gold skarn, mesothermal, transitional, and epither-
mal deposits and volcanogenic massive sulphides
form. Cupriferous pyrite-type volcanogenic massive
sulphides are deposited at oceanic spreading cen-
ters, but also in continental border rifts or possibly
backarc basins.
Periodic accretionary events welded alloch-
thonous plates to the western edge of the Canadian
Cordillera, and mesothermal to epithermal gold
deposits were formed. The accreted terranes that
contain mesothermal gold deposits with associated
listwanitic alteration are generally oceanic, and
some are ophiolites obducted over the continental
margin during collision. Deep fractures and per-
haps mantle hotspots or other thermal events pro-
duced post-accretion auriferous vein, skarn, and
manto deposits.
Mesozoic tectonic processes are not unique in
the history ofthe Cordillera. Consequently, many of
the gold-bearing deposit types are also found in both
older and younger rocks. Erosional levels are
critically important to the types of deposits exposed
(Figure 13). Some deposit types, such as porphy-
ries, skarns, and epithermal deposits, are not known
in Paleozoic rocks, whereas others, such as meso-
thermal deposits with associated listwanite altera-
tion, are unknown in Tertiary rocks. Local geology
is one key factor that localizes mineral deposits.
Metal sources, the chemistry of the magma source
rock, permeability, faults, a heat source-driven
hydrothermal convection system, and the chemistry
of the ore fluid and the country rock all playa role.
Tectonic setting is important in creating areas
favorable for mineralization, but the other factors
648 W. J. McMillan and A. Panteleyev
control the locations and types of the ore deposits
formed.
Collisional events, both before and during
docking of Superterrane I, particularly those that
juxtaposed oceanic rocks with either North Ameri-
can or island-arc assemblages, caused faulting and
plutonism. Terrane boundary faults provided per-
meability and tapped deeply into the crust. Heat
and plutonism generated large-scale hydrothermal
systems that gave rise to mesothermal gold deposits.
These ores were deposited above the brittle-ductile
transition zone in the crust from low-salinity, CO
2
-
charged ore fluids that were isotopically modified by
wallrock interaction. Consuming margin island-arc
processes also gave rise to auriferous skarn and
porphyry deposits and to transitional to epithermal
gold deposits.
Early Cretaceous mesothermal gold deposits lie
along the length of the North American Cordillera
- from northern British Columbia to California.
This major metallogenic event may represent a sig-
nificant shift in movement vectors between the
Pacific and North American plates.
Later, docking of Superterrane II caused forma-
tion of the Bridge River mesothermal gold veins.
Mineralized veins there are related to transcurrent
faulting, suggesting docking vectors oblique to the
plate boundary. Metals in the camp are zoned out-
ward from the collision-generated Coast Range in-
trusions.
In island-arc settings, most gold-bearing depo-
sits are associated with either I-type, subduction-
related intrusions or high-potassium alkalic to calc-
alkalic backarc or marginal basin volcanics and in-
trusives.
Nesbitt and others (1986) pointed out two fun-
damental types of gold deposits in the Cordillera -
epithermal and mesothermal, whose characteristics
and differences are illustrated in Table 1.
A spectrum of deposit types exists between
porphyry, mesothermal, and epithermal deposits.
In the Stewart Camp, for example, epithermal depo-
sits are unusually base-metal rich and silver values
are high, which are general characteristics of meso-
thermal deposits. Perhaps this indicates that the
ore fluids were saline and derived from seawater,
not meteoric water. The depth-zoning model deve-
loped by Panteleyev (1988b) displays probable
settings and characteristics of Cordilleran deposits
(Figure 13). It illustrates variations that can be
expected in a vertical section through a major
composite hydrothermal system, and we recognize
these variations in gold deposits in British Colum-
bia. Part of the ore story is controlled by tempera-
ture and pressure of formation, part by the source
and composition of the hydrothermal fluids, and
part by related faulting, plutonism, and volcanism.
REFERENCES
ALLDRICK, D. J., 1985, Stratigraphy and petrology of the
Stewart mining camp: Geological fieldwork 1984: British
Columbia Ministry of Energy, Mines and Petroleum Re-
sources, Paper 1985-1, p. 316-342.
__ ,1986, Stratigraphy and structure in the Anyox area:
Geological fieldwork 1985: British Columbia Ministry of
Energy, Mines and Petroleum Resources, Paper 1986-1, p.
211-216.
__ , D. A. BROWN, J. E. HARAKAL, J. K. MORTENSEN,
and R. L. ARMSTRONG, 1987, Geochronology of the Stew-
art mining camp: Geological fieldwork 1986: British
Columbia Ministry of Energy, Mines and Petroleum Re-
sources.,Paper 1987-1, p. 81-92.
__ , J. K MORTENSEN, and R. L. ARMSTRONG, 1986,
Uranium-lead age determinations in the Stewart area:
Geological fieldwork 1985: British Columbia Ministry of
Energy, Mines and Petroleum Resources, Paper 1986-1, p.
217-218.
ASH, C. H., and R. L. ARKSEY, 1990, The Atlin ultramafic
allochthon: Ophiolite basement within the Cache Creek
terrane - Tectonic and metallogenic significance (l04NI
12): British Columbia Ministry of Energy, Mines and
Petroleum Resources, Paper 1990-1, p. 359-364.
BARR, D. A, P. E. FOX, K. E. NORTHCOTE, and V. A
PRETO, 1976, The alkaline suite porphyry deposits: A
summary; pp. 359-367 in Porphyry Deposits of the
Canadian CordUlera: Charles S. Ney Volume: Cana-
dian Institute of Mining and Metallurgy, Special Volume
15,510p.
BACON, W. R., 1978, Lode gold deposits in Western Cana-
da: Canadian Institute of Mining and Metallurgy Bulletin,
v. 71, p. 96-104.
BOHKLE, J. K., and R. W. KISTLER, 1986, Rb-Sr, K-Ar,
and stable isotope evidence for the ages and sources of fluid
components of gold-bearing quartz veins in the northern
Sierra Nevada Foothills metamorphic belt, California:
Economic Geology, v. 81, p. 296-322.
BRIDGE, D. A., J. A MARR, K. HASHIMOTO, M. OBARA,
and R. SUZUKI, 1987, Geology of the Kutcho Creek vol-
canogenic massive sulphide deposits, northern British
Columbia; pp. 115-128 in J. A. Morin (ed.), Mineral
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 649
Deposits of the Northern Cordillera: Canadian Institute
of Mining and Metallurgy, Special Volume 37, 377 p.
BRI'ITON, J. M., and D. J. ALLDRICK, 1988, Sulphurets
map-area (l04A/5W, 12W; 104B/8E, 9E), Geological field-
work 1987: Brititsh Columbia Ministry of Energy, Mines
and Petroleum Resources, Paper 1988-1, p. 199-210.
CARGILL D. G., J. LAMB, M. J. YOUNG, and E. S. RUGG,
1976, Island Copper; pp. 206-218 in Porphyry Deposits of
the Canadian CordiUera: Charles S. Ney Volume:
Canadian Institute of Mining and Metallurgy, Special
Volume 15, 510 p.
CHURCH, B. N., 1987, Geology and Mineralization of the
Bridge River mining camp: Geological fieldwork 1986:
British Columbia Ministry of Energy, Mines and Petroleum
Resources,Paper 1987-1, p. 23-29.
__ , R. B. GABA, M. J. HANNA, and D. A. JAMES, 1988,
Geological reconnaissance in the Bridge River mining
camp: Geological fieldwork 1987: British Columbia Minis-
try of Energy, Mines and Petroleum Resources,Paper 1988-
1, p. 93-100.
DAWSON, K. M., A. PANTELEYEV, A. SUTHERLAND
BROWN, and G. J. WOODSWORTH, 1991, Regional metallo-
geny; pp. 707-768 in H. Gabrielse and C. J. Yorath (eds.),
Geology of the CordiUeran Orogen in Canada: Geologi-
cal Survey of Canada, Geology of Canada 4.
DIAKOW, L. J., 1985, Potassium-argon age determinations
from biotite and hornblende in Toodoggone volcanic rocks:
Geological fieldwork 1984: British Columbia Ministry of
Energy, Mines and Petroleum Resources, Paper 1985-1, p.
299-300.
DROWN, T. J., K. E. HICKS, and F. G. HEWE'IT, 1987,
Sulphurets - An emerging gold-silver camp [abstract):
CIM Bulletin, v. 80, no. 904, p. 30.
DRUMMOND, A. D., and C. I. GODWIN, 1976, Hypogene
mineralization: An empirical evaluation of alteration
zoning; pp. 52-63 in Porphyry Deposits of the Canadian
CordiUera: Charles S. Ney Volume: Canadian Institute
of Mining and Metallurgy, Special Volume 15, 510 p.
EISBACHER, G. H., 1974, Sedimentary History and Tec-
tonic Evolution of the Sustut and Sitton Basins, North
Central British Columbia: Geological Survey of Canada,
Paper 73-31, 45 p.
ETTLINGER, A. D., and G. E. RAY, 1988, Gold-enriched
skarn deposits of British Columbia: Geological fieldwork
1987: British Columbia Ministry of Energy, Mines and
Petroleum Resources, Paper 1988-1, p. 263-280.
__ , and G. E. RAY, 1989, Precious Metal Enriched
Skarns in British Columbia - An Overview and Geo-
logical Study: British Columbia Ministry of Energy,
Mines and Petroleum Resources, Paper 1989-3, 128 p.
FAULKNER, E. L., 1986, Blackdome deposit: Geological
fieldwork 1985: British Columbia Ministry of Energy,
Mines and Petroleum Resources, Paper 1985-1, p. 107-109.
FL YES, J. T., 1984, Geological Setting of the Rossland
Mining Camp: British Columbia Ministry of Energy,
Mines and Petroleum Resources, Bulletin 74,61 p.
GLOVER, J. K., and P. SCHIARIZZA, 1987, Geology of the
Warner Pass map sheet: Geological fieldwork 1986: Bri-
tish Columbia Ministry of Energy, Mines and Petroleum
Resources, Paper 1987-1, p.157-169.
__ , P. SCHIARIZZA, and I. J. GARVER, 1988, Geology of
the Noaxe Creek map-area: Geological fieldwork 1987:
British Columbia Ministry of Energy, Mines and Petroleum
Resources, Paper 1988-1, p. 105-123.
GROVE, E. W., 1986, Geology and Mineral Deposits of
the Unuk River-Salmon River-Anyox Area: British
Columbia Ministry of Energy, Mines and Petroleum Re-
sources, Bulletin 63, 152 p.
HEDENQUIST, J. W., and R. W. HENLEY, 1985, The impor-
tance of CO
2
on freezing point measurements of fluid
inclusions: Evidence from active geothermal systems and
implications for epithermal ore deposition: Economic Geo-
logy, v. 80, p. 1379-1406.
IRVING, E., J. W. H. MONGER, and R. W. YOLE, 1980, New
paleomagnetic evidence for displaced terranes in British
Columbia; pp. 441-456 in D. W. Strangeway (ed.), The
Continental Crust and Its Mineral Deposits: Geological
Association of Canada, Special Paper 20, 804 p.
KERRICH, R., 1983, Geochemistry of Gold Deposits in
the Abitibi Greenstone Belt: Canadian Institute of Min-
ing and Metallurgy, Special Volume 27,75 p.
LEFEBURE, D. V., S. B. BALLANTYNE, and K. E. HICKS,
1987, Wall rock alteration at the Sulphurets gold-silver
property, British Columbia [abstract): CIM Bulletin, v. 80,
no. 904, p. 30.
LEITCH, C. H. B., and C. I. GODWIN, 1986, Geology of the
Bralorne-Pioneer gold camp: Geological fieldwork 1985:
British Columbia Ministry of Energy, Mines and Petroleum
Resources, Paper 1986-1, p. 311-316.
__ , and C. I. GODWIN, 1987, The Bralorne gold vein
deposit - An update: Geological fieldwork 1986: British
Columbia Ministry of Energy, Mines and Petroleum Re-
sources, Paper 1987-1, p. 35-38
__ , and C. I. GODWIN, 1988, Isotopic ages, wallrock
chemistry and fluid inclusion data from the Bralorne gold
vein deposit: Geological fieldwork 1987: British Columbia
Ministry of Energy, Mines and Petroleum Resources, Paper
1988-1, p. 301-324.
650
W. J. McMillan and A. Panteleyev
MacINTYRE, D. G., 1986, The geochemistry of basalts
hosting massive sulphide deposits, Alexander terrane,
northwest British Columbia: Geological fieldwork 1985:
British Columbia Ministry of Energy, Mines and Petroleum
Resources, Paper 1986-1, p. 197-210.
McMILLAN, W. J., and A. PANTELEYEV, 1988, Porphyry
copper deposts; pp. 59-66 in R. G. Roberts and P. A. Shee-
han (eds.), Ore Deposit Models: Geoscience Canada Re-
print Series 3, 194 p.
__ , A. PANTELEYEV, and T. HOY, 1987, Mineral
deposits in British Columbia: A review of their tectonic
settings; p. 1-18 in I. L. Elliott and B. W. Smee (eds.),
GEOEXPOI86: Proceedings: Association of Exploration
Geochemists, Rexdale, Ontario, Canada, 220 p.
MEADE, H. D., 1986, Problems in ore controls and genesis
in volcanic-hosted precious-base metal mineralization,
Silbak Premier and Big Missouri deposits [abstract]:
Abstracts of the Annual Meeting of the Northwest Mining
Association, Spokane, Washington, USA,
MEINERT, L, D., 1983, Variability of skarn deposits:
Guides to exploration; pp. 301-316 in S. J, Boardman (ed,),
Revolution in the Earth Sciences - Advances in the
Past Half-Century: Kendall/Hunt Publishing Company,
Dubuque, Iowa, USA, 385 p.
MONGER, J. W. H., 1972, Oceanic crust in the Canadian
Cordillera [Contribution #4 in The Ancient Oceanic
Lithosphere: Canadian Contributions Numbers 1-11
to the Geodynamics Project]: Energy, Mines and Re-
sources Canada, Publications of the Earth Physics Branch,
v. 42, no, 3, p. 59-64.
__ , 1985, Structural evolution of the southwestern
Intermontane Belt, Ashcroft and Hope map-areas, British
Columbia; pp. 349-358 in Current Research, Part A:
Geological Survey of Canada, Paper 85-1A,
__ , R. A. PRICE, and D. J. TEMPELMAN-KLUIT, 1982,
Tectonic accretion and the origin of the two major meta-
morphic and plutonic welts in the Canadian Cordillera:
Geology, v. 10, p. 70-75,
NESBITT, B. E., J. B. MUROWCHICK, and K. MUEHLEN-
BACHS, 1986, Dual origins of lode gold deposits in the
Canadian Cordillera: Geology, v. 14, p, 506-509.
PANTELEYEV, A., 1988a, Quesnel mineral belt - The
central volcanic axis between Horsefly and Quesnel Lakes
(93A/05E, 06W): Geological fieldswork 1987: British
Columbia Ministry of Energy, Mines and Petroleum
Resources, Paper 1988-1, p. 131-138.
__ , 1988b, A Canadian Cordilleran model for epither-
mal gold-silver deposits; pp. 31-44 in R. G. Roberts and P.
A. Sheehan (eds.), Ore Deposit Models: Geoscience
Canada Reprint Series 3, 194 p.
__ , and L. J, DIAKOW, 1982, Cassiar gold deposits
McDame map-area: Geological fieldwork 1981: British
Columbia Ministry of Energy, Mines and Petroleum Re-
sources, Paper 1982-1, p. 156-161.
PRICE, R. A., and D. M. CARMICHAEL, 1986, Geometric
test for Late Cretaceous-Paleogene intracontinental
transform faulting in the Canadian Cordillera: Geology, v.
14, p, 468-471.
RANDALL, A. W., 1988, Premier Gold Project - Geo-
logical Setting and Mineralization of the SUbak
Premier and Big Missouri Deposits: Society of Econo-
mic Geology Field Trip Guidebook to Major Gold
Silver Deposits of the Northern Canadian Cordillera
(September 22-26, 1988): Geololl:ical Survey of British
Columbia.
RAY, G. E., G. L. DAWSON, and R. SIMPSON, 1988, Geo-
logy, geochemistry and metallogenic zoning in the Hedley
gold-skarn camp: Geological fieldwork 1987: British
Columbia Ministry of Energy, Mines and Petroleum
Resources, Paper 1988-1, p. 59-80.
__ , J. T, SHEARER, and R. J. E. NIELS, 1983, Carolin
gold mine: Field trip no. 4; pp. 40-64 in G, E. Ray, L. W.
Carlyle, R. Simpson, L. W. Saleken, J. Bellamy, J, T,
Shearer, and R. J. E. Niels (contributors), Some Gold
Deposits in the Western Canadian CordUlera: Geo-
logical Association of Canada, Victoria Section, Victoria,
British Columbia, Canada, GAC-MAC-CGU Field Trip
Guidebook, v. 1, 64 p.
SCH}WETER, T, G., and A. PANTELEYEV, 1985, Lode gold-
silver deposits in northwestern British Columbia; pp, 178-
190 in J. A. Morin (ed,), Mineral Deposits of the North-
ern Cordillera: Canadian Institute of Mining and Metal-
lurgy, Special Volume 37, 377 p.
SKETCHLEY, D. A" and A. J. SINCLAIR, 1987, Multi-
element lithogeochemistry of alteration associated with
gold-quartz veins of the Erickson Mine, Cassiar district
(104P): Geological fieldwork 1986: British Columbia
Ministry of Energy, Mines and Petroleum Resources, Paper
1987-1, p, 57-63,
__ , A. J, SINCLAIR, and C. 1. GoDWIN, 1986, Early
Cretaceous gold-silver mineralization in the Sylvester
allochthon, near Cassiar, north central British Columbia:
Canadian Journal of Earth Sciences, v, 26, p, 1455-1458,
STONE, B. G., and J. G. PAYNE, 1982, Deformed Mesozoic
volcanogenic Cu-Zn sulphide deposits in the Britannia
district, British Columbia: Economic Geology, v. 77, p.
712-714,
TIPPER, H, W" 1981, Offset of an Upper Pleinsbachian
geographic zonation in the North American Cordillera by
transcurrent movement: Canadian Journal of Earth
Sciences, v. 18, p. 1788-1792.
Tectonic setting of Mesozoic gold deposits in the Canadian Cordillera 651
__ ,1984, The allochthonous Jurassic-Lower Cretaceous
terranes of the Canadian Cordillera and their relation to
correlative strata of the North American craton; pp. ll3-
120 in G. E. G. Westermann (ed.), Jurassic-Cretaceous
Paleogeography of North America: Geological Associa-
tion of Canada, Special Paper 27,315 p.
WODJAK, P. W., and A. W. RANDALL, 1986, The Silbak
Premier gold-silver deposit [abstract): Abstracts of the
Canadian Institute of Mining and Metallurgy District 6
Meeting, Victoria (October 1986), p. 24.
WOODSWORTH, G. J., D. E. PEARSON, and A. J. SINCLAIR,
1977, Metal distribution patterns across the eastern flank
of the Coast Plutonic Complex, south central British
Columbia: Economic Geology, v. 72, p. 170-183.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Tectonic setting of Au-Ag deposits hosted by Proterozoic strata
along the Lewis and Clark line, west-central Montana
LEEA. WOODWARD
Department of Geology, University of New Mexico, Albuquerque, New Mexico 87131, USA
(received August 12, 1988; revision accepted December 21, 1989)
ABSTRACT
The Lewis and Clark line is a major lineament trending WNW from central Montana to the Coeur d'Alene
mining district of northern Idaho. The Lewis and Clark line consists mainly of Cretaceous to Tertiary thrusts
and steep strike-slip and dip-slip faults that are partly superimposed on Middle Proterozoic growth faults
related to extension of the western continental margin.
Lode Au-Ag deposits are clustered along the Lewis and Clark line in the Big Belt uplift and comprise the
Confederate Gulch, York, and Beaver Creek mining districts. These lodes are mostly epigenetic quartz veins
with minor sulfides and are hosted by the Greyson Shale and Newland Limestone (of the Middle Proterozoic
Belt Supergroup) and dioritic intrusions. Some mineralization, along bedding planes in the Greyson and
Newland formations, may be synsedimentary or syndiagenetic, however. At least 2,581 kg (83,000 oz) Au
were produced from lodes and 28,612 kg (920,000 oz) Au from placers locally derived from similar lodes. The
average Au:Ag ratio for lodes is "" 1.5:1. Most lode Au occurrences along the Lewis and Clark line in the Big
Belt uplift are interpreted to be remobilized from Archean basement rocks and deposited in Belt strata, in
some cases syngenetically, near basement growth faults active during deposition of the Greyson and Newland
formations. Metals were probably transported by aqueous solutions along faults. Au was probably re-
mobilized during widespread but volumetrically minor magmatism in the Late Proterozoic and again in the
Cretaceous-Tertiary .
INTRODUCTION
The Lewis and Clark line (Figure 1), a major
structural discontinuity extending ESE from the
Coeur d'Alene mining district of northern Idaho to
west-central Montana (Harrison and others, 1974),
has been also called the Montana lineament (Weid-
man, 1965). This linear zone is ;::: 550 km long and
consists of steeply dipping strike-slip and dip-slip
faults, thrust, and en echelon folds. Some investiga-
tors consider the Lewis and Clark line to extend
even farther to the southeast, to include the Lake
Basin zone of en echelon faults in southern Montana
(Lorenz, 1984). The width of the Lewis and Clark
line may be as much as 95 km where it is marked by
parallel faults near its west end (Weidman, 1965);
elsewhere, this is a broadly transitional zone =::; 45
km wide (McMannis, 1959).
As noted by McMannis (1959), the Lewis and
Clark line separates two markedly different ter-
ranes. To the north there is a notable lack of large
granitoid intrusions, extensive Cretaceous and Ter-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 653-663. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
653
654
L. A. Woodward
'-'---'
I
I
I
\\
\-\ M 0 N TAN A
Eastern limit \ \
of outcrop of \ \
Figure 1. Generalized tectonic map of
western Montana and northern Idaho
showing location of mining districts of the
northern Big Belt uplift (BB), the Coeur
d'Alene district (CA), and the Willow
Creek fault (WCF). Numbered localities
are noted in text.
I
Salt r ocks Eastern edge 01
\ \ \..........----Cordilleran foldbelt
CA \ \-\
..... \\
'- .....
Falls
.... ."\.E:WIS a. ....... , "-
\ ..... C
L
AR/( ....... \ ')",. r,1
IDAHO
tiary volcanic rocks, and abundant lower Cenozoic
basin deposits. In contrast, the terrane to the south
is characterized by the Idaho, Boulder, Philipsburg,
and Tobacco Root batholiths, in addition to numer-
ous other granitoid intrusions. Also, volcanic rocks
and lower Cenozoic basin deposits are widespread.
Structural features along the Lewis and Clark
line cannot be attributed to a single event (Weid-
man,1965). Many of the faults along the line can be
shown to have Tertiary and (or) Cretaceous move-
ment; however, some faults and folds.along the line
formed in Late Proterozoic time (Harrison and
others, 1974, p. 11). Stratigraphic and sedimento-
logic studies by Winston (1983, 1986) and his col-
leagues suggest the presence of major growth faults
along parts of the Lewis and Clark line in Middle
Proterozoic time. Reynolds (1984) noted recurrent
tectonism along WNW and NE trends during depo-
sition of some Middle Proterozoic strata, with facies
and thickness changes across the Lewis and Clark
line. Movement along the line appears to have be-
gun prior to deposition of Middle Proterozoic rocks
(Reid, 1984). Thus, the Lewis and Clark line has
Precambrian ancestry, at least along some parts of
the line.
. 2" '. LINE: \
\'\', .. ... " .. 4 ..... ,/
) HELENA" "/1
( EMBAYMENT / J
) Butt e:ll:
.
) /'\ -< WCF
,.. I DILLON
\ ,'1 BLOCK
\
-'-'-' _.-.- -
"> \)
,-.1.(. 1 0 100km
'." --r....",..-. I
'"
Middle Proterozoic strata - the Belt Super-
group in the United States and the Purcell Super-
group in Canada - were deposited between "-' 1,500
and "-'900 Ma (Reynolds, 1984) in an epicontinental
basin and on an adjoining continental margin in
Montana, northern Idaho, northeastern Washing-
ton, and adjacent parts of Canada (Harrison and
Reynolds, 1976).
The Coeur d'Alene mining district, with cumu-
lative production of over 31.1 billion g (1 billion oz)
Ag, marks the western end of the Lewis and Clark
line (Billingsley and Locke, 1941). The Confederate
Gulch, York, and Beaver Creek mining districts,
occurring near the eastern end of the Lewis and
Clark line in west-central Montana, have an esti-
mated production of "-' 31,100 kg (1,000,000 oz) Au
and are the subjects of this paper. Nearly 30 other
mining districts, not discussed further here, also are
located along the Lewis and Clark line. It is signifi-
cant that these mineral deposits occur along a zone
where basement faults appear to have been re-
currently active since Precambrian time, even
though many of the deposits probably have different
geneses.
Au-Ag deposits hosted by Proterozoic strata along the Lewis and Clark Line 655
Western Montana Central Montana
Cambrian Flathead Sandstone
Pilcher QuartzIte

Garnet Range Fm.
Missoula McNamara formation
Group

Bonner Quartzite
Mt. Shields Formation
u
Shepard Format 100

"-<
o 0-
N
Snowslip Formation
o 0
" "
Middle Belt
QJ bIJ
w
Carbonate \-.Tallace Formation Helena Limestone
o QJ
P.
P-<
(f)
Empire Formation Empire Formation
QJ
W
'0--<
'0 OJ
St. Regis Formation Spokane Format ion
"-< '"
<:
Ravalli
Group Revett Formation
Burke Formation Greyson Shale
---
Newland Limestone
Lower
Belt Prichard Formation Chamberlain Shale
Neihart Quartzite
- - - --
Early Proterozoic Basement not Crystalline
and Late Archean exposed rocks
Figure 2. Generalized correlation chart for units within
the Belt Supergroup.
ROCK UNITS
The Belt Supergroup is composed mainly of
fine-grained clastic rocks (shale, argillite, and sil-
tite), with subordinate quartzite and carbonates and
minor conglomerate and breccia. Stable-isotope stu-
dies by Rye and others (1984) show deposition of
Belt sediments in a marine environment, support-
ing sedimentologic interpretations by many other
workers. These strata thicken westward from a zero
edge due to erosional truncation in the Little Belt
Mountains of central Montana (Figure 1, location 1)
to a maximum of ""20,400 m near Alberton in west-
ern Montana (Figure 1, location 2) where neither
the top nor base of the section is exposed (Harrison,
1972, p. 1219). An unconformity at the top of the
Belt Supergroup (Figure 2) cuts stratigraphically
lower eastward, truncating all the Belt rocks in
central Montana. The only major exposure of the
base of the Belt Supergroup is in the Little Belt
Mountains of central Montana (Figure 1, location
1); here, crystalline igneous and metamorphic base-
ment rocks are =1,700-3,500 Ma, with the last re-
gional synkinematic metamorphism occurring at
""1,900 Ma (Catanzaro and Kulp, 1964; Woodward,
1970). Crystalline basement rocks south of the He-
lena embayment (Figure 1) were metamorphosed
"" 2,750 Ma (James and Hedge, 1980).
In general, facies changes across the Belt basin
are, with a few exceptions, relatively minor. The
lowest unit of the Belt Supergroup on the west is the
Prichard Formation, composed principally of quart-
zite, siltite, and laminated argillite, generally con-
sidered to correlate with a transgressive sandstone,
shale, and limestone sequence in central Montana
(Figure 2). The fine-grained lower Belt lithologies
of central Montana change southward to extremely
coarse-grained conglomerate of the LaHood Forma-
tion adjacent to the syndepositional Willow Creek
fault (Figure 1) that bounds the south side of the
Helena embayment (McMannis, 1963). Carbonate
sedimentary breccia in the Wallace Formation and
the Newland Limestone occur as discontinuous in-
traformational lenses that are interpreted as mar-
king syndepositional faults (Wallace and others,
1976; Zieg, 1981).
Belt rocks along the eastern edge of the Belt
basin are essentially non-metamorphosed, but there
is a gradual increase in metamorphic grade to bio-
tite zone in western Montana, and locally to amphi-
bolite facies near major plutons (Harrison, 1972).
In the Big Belt Mountains, the Belt rocks that
host Au-Ag deposits are the Newland Limestone
and the Greyson Shale. The Newland consists of
yellow-brown, blue-gray, red-gray, red-brown, olive-
gray, and dark-brown, calcareous and non calc are-
ous, thinly laminated argillite, and dark blue-gray
densely crystalline, thin- to medium-bedded lime-
stone. This unit is "" 1,680 m thick near the York
mining district (Shaffer, 1971). The Greyson Shale
consists of yellow-brown to olive-gray, wavy lamina-
ted, noncalcareous argillite, dark yellow-brown,
unevenly laminated shale, and subordinate dark
yellow-brown, very fine-grained, thin-bedded quart-
zitic sandstone and is = 1,615 m thick near the York
mining district (Shaffer, 1971).
Two mafic sills intrusive into Belt strata in the
northern Big Belt uplift were radiometrically dated
by K-Ar methods as 82641 Ma and 74437 Ma
(Marvin and Dobson, 1979) and thus are of Late
Proterozoic age. Granodiorite in the Big Belt uplift
yielded K-Ar dates of 62.8 2.4 Ma from biotite and
656 L. A. Woodward
69.73.2 Ma from hornblende (Daniel and Berg,
1981, p. 80).
TECTONIC FRAMEWORK
Belt (Purcell) rocks were deposited along the
western edge of the North American craton on
oceanic and tectonically attenuated continental
crust that probably was rifted during Middle
Proterozoic time (Price, 1984). The Precambrian
Belt basin is principally depositional (Harrison,
1972) but was slightly modified by later tectonism
and erosion (Reynolds, 1984). The zero isopach of
Belt rocks on the east side of the Belt basin was
produced by pre-Middle Cambrian erosion. How-
ever, judging from eastward thinning of units with-
in the Belt Supergroup, the depositional edge was
probably subparallel to, but slightly east of, the
present zero isopach. The southern margin of the
Helena embayment is the syndepositional Willow
Creek fault (Figure 1), along which coarse debris
(LaHood Formation) was fed into the Belt basin
(McMannis, 1963). Reynolds (1984) suggested that
the depositional edge of the basin was farther to the
south, but pre-Middle Cambrian erosion removed
the Belt rocks from the high area (Figure 1, Dillon
block) south of the fault. Harrison and others (1974)
suggested that the western edge of the Dillon block
was also bounded by a high-angle fault during Belt
time.
Several basement blocks within the basin
appear to have been active during Belt deposition
(Winston, 1983; Reynolds, 1984), with differential
subsidence resulting in differing thicknesses of Belt
strata from one block to another (Figure 3). Rey-
nolds (1984) noted recurrent tectonism along WNW
and NE trends during deposition of several lower
Belt units, with facies and thickness changes in the
Newland Limestone and Helena Limestone-Wallace
Formation across the Lewis and Clark line. The
Newland is thin north of the line, whereas the
Helena-Wallace interval thickens markedly into the
Lewis and Clark zone and thins abruptly to the
south. Winston (1983) suggested the presence of
three W-trending fault zones and a NW-trending
fault in the basin during Belt deposition (Figure 3);
two of the EIW-trending faults are adjacent to syn-
genetic mineral deposits hosted by Belt strata. The
eastern end of Winston's (1986) Garnet line (Figure
3) corresponds to the Lewis and Clark line in gen-
eral and to the Volcano Valley thrust (Figure 1,
location 4) in particular (Zieg and Godlewski, 1986).
This latter fault probably represents a major crustal
break that had recurrent movement during the
Proterozoic, Devonian, Mississippian, Jurassic, Cre-
taceous, and Tertiary (Zieg and Godlewski, 1986).
This growth fault was close to the northern margin
of the Helena embayment and was down-thrown on
the south side during Middle Proterozoic time (Win-
ston, 1986). Significantly Au-Ag lode deposits hos-
ted by Belt strata occur south of the Volcano Valley
116
0
I 114
0
112
0
"00
I
I
i
l

f'vl 0 N TAN A
Figure 3. Hypothetical fault zones (Jocko,
Garnet, Townsend, and Perry lines) sepa-
rating Proterozoic crustal blocks of the Belt
basin (modified from Winston, 1986), and
York (Y), Confederate Gulch (C), Beaver
Creek (B), and Coeur d'Alene (CA) mining
districts.
CA \
-1 \ JOCKO
"
GARNET
I D A H 0
PER RY
i.\../\
C 100 km
! ,
1Y
LINE
'"
0", C 1S
U
D
1--1l' HELENA
"'1-
0
EMBAYMENT
DILLON
BLOCK
L J NE
Au-Ag deposits hosted by Proterozoic strata along the Lewis and Clark Line 657
Figure 4. Tectonic map of western
Montana showing Au-Ag lode occur-
rences in York (Y) Confederate
Gulch (C), and Beaver Creek (B)
mining districts and other precious-
metal occurrences hosted by Belt
rocks.
fault, which in turn marks the northern edge of the
Lewis and Clark line in west-central Montana.
Widespread but volumetrically minor magma-
tism that occurred in the Belt basin during Middle
and Late Proterozoic time may have been important
in mineralization processes. Mafic sills, dated at
1,075-1,200 Ma, occur in the northern and east-cen-
tral part of the Belt basin (Reynolds, 1984). Dia-
basic dikes and sills that are "" 7 50-830 Ma (Rey-
nolds, 1984) and clearly post-date deposition of the
Belt rocks are present at numerous localities.
During the Cretaceous and early Tertiary, the
Belt rocks were telescoped along E-directed thrusts
within the basin (Figure 4) and were faulted along
the Precambrian block boundaries (Winston, 1983).
Harrison and others (1974) gave a very good descrip-
111 -
s
o 8Qlt
o ks
fault.
........,....- on uPP<ilr
or mol 101,11\ . tl(IiI!o on
- dowl"ltnrow f"l StdQ:
_ HI Qn-onglQ: IOuit 'W' ltn
--- movomQnt
-t- Anllcll nCl
- S y nCI.n"
- y -Ov (21- turneo sync !.no
o
""
X Other prClCIOuS-IT'Ultol occurr Q: ncQS
nOSlQ(J Dy Belt rocks
, 00 km
----- -
tion and analysis of the effect of Belt basin struc-
tures on the later tectonics, particularly along the
complicated Lewis and Clark line.
The York, Confederate Gulch, and Beaver
Creek mining districts occur in the upper plate of
the Scout Camp thrust (Figure 5). Regionally, the
Scout Camp thrust plate consists of a stratigraphic
section ranging from the Newland Limestone (Pro-
terozoic) to Cretaceous strata. This thrust plate is
bounded on the southwest by the Eldorado over-
thrust (Figure 5). Minimum N-S separation on the
Scout Camp thrust, as determined from map pat-
terns, is "" 12 km. The Scout Camp thrust is one of
several faults defining the northern margin of the
Helena salient, a Late Cretaceous-early Tertiary
658
L. A. Woodward

volC;II-n,e;.



.e(ll."", .,u,
:1

- - .. .. Thf\IM 'MItt . tNth on upper pi.ll.-
-=- - - .. .. S.nl<.1Iop .......
--- - . Hogh onvIo ' .uh
-+- --
-
O hed where apprOXimate, dotted wher. conc led
Figure 5. Tectonic map of the Big Belt Mountains and adjacent areas, showing York (Y), Confederate Gulch (C), and Beaver
Creek (B) mining districts.
structural feature that coincides with the Protero-
zoic Helena embayment (Figure 1).
East-west crustal extension during middle and
late Cenozoic time resulted in further disruption of
Belt rocks into horsts, grabens, and tilted fault
blocks. Reynolds (1979) emphasized that the Lewis
and Clark line was a zone of right-slip, as it sepa-
rates stable crust on the north from extensional
features on the south. This strike-slip movement is
translated into dip-slip displacement on normal
faults that splay southward from the line.
Thus, the Belt basin has had a complex tectonic
history characterized by rifting along a continental
margin and development of fault blocks within the
basin during Belt deposition in the Middle Pro-
terozoic. Most Belt rocks are now allochthonous,
having been telescoped and thrust eastward during
the Cretaceous and early Tertiary (Woodward,
1981). Middle and late Cenozoic extension resulted
in uplifts that are separated by basins filled with
Tertiary and Quaternary sediments.
LODE AU-AG DEPOSITS
Lode Au-Ag deposits are clustered along the
Lewis and Clark line in the Big Belt uplift of west-
central Montana and comprise the York, Confeder-
ate Gulch, and Beaver Creek mining districts
(Figures 4 and 5). These lodes are mainly epigene-
tic quartz veins with minor amounts of sulfides.
Some of the precious-metal mineralization occurs
along bedding planes in the Greyson Shale and
Newland Limestone and may be synsedimentary
and (or) syndiagenetic. Contact metamorphic depo-
sits adjacent to Cretaceous-Tertiary intrusions in
the Confederate Gulch district were also noted by
Johnson (1973). The vein deposits are mostly hosted
by the Greyson Shale and the Newland Limestone of
the Belt Supergroup and by dioritic intrusions.
Pardee and Schrader (1933) tentatively considered
the dioritic intrusions to be of Oligocene age, but
they could be Precambrian (McClernan, 1980) or
Cretaceous, as are similar rocks near Helena
Au-Ag deposits hosted by Proterozoic strata along the Lewis and Clark Line 659
(Knopf, 1963). At least 2,581 kg (83,000 oz) Au were
produced from lodes, and an estimated 28,612 kg
(920,000 oz) Au came from placers locally derived
from similar lodes. This region is therefore one of
the major producers in the United States, with total
production of;:=: 31,100 kg (1,000,000 oz) Au.
The minimum production of Au from lode depo-
sits in the York district is 2,177 kg (70,000 oz) and
;:=: 8,241.5 kg (265,000 oz) were recovered from
placers (Koschmann and Bergendahl, 1968, p. 157).
Between 1933 and 1943, the Golden Messenger lode
produced 1,600 kg (51,440 oz) Au; thus, total produc-
tion of the district is probably significantly larger.
In addition to the Golden Messenger, the Old Amber
and Little Dandy produced significant amounts of
Au from lodes (McClernan, 1983).
The Golden Messenger ore body consists of
replacement veins in a diorite dike and in the Grey-
son Shale. The mineralized zone is 120 m wide and
210 m long, containing veins of quartz, pyrite, and
galena that are ::510 m thick and 60-90 m in their
other dimensions (Pardee and Schrader, 1933). Au
occurs with pyrite and galena in the primary ore,
which averaged"" 10.3 glT (0.3 oz/T) Au with little
or no Ag. The Old Amber mine produced ""311 kg
(10,000 oz) Au from three veins that were 7-10 cm
wide in a zone"" 1 m wide. The veins are parallel to
bedding in the Greyson Shale and contain quartz
and pyrite. The ore contained"" 13.7 glT Au (0.4
ozlT). The Little Dandy mine produced "" 186.6 kg
(6,000 oz) Au from a quartz vein ::51.3 m wide that
cuts a diorite dike and the Greyson Shale.
Baitis (1988) reported that stratiform ortho-
clase-rich sandstones and siltstones with carbonates
and minor amounts of pyrite that occur in the Grey-
son Shale in the York district contain 0.34 to 1.7 glT
(0.01 to 0.05 ozlT) Au. These deposits were inter-
preted by Baitis (1988) to be syngenetic or dia-
genetic, with local remobilization during Laramide
(Late Cretaceous-early Tertiary) deformation.
The Confederate Gulch district produced
"" 18,600 kg (600,000 oz) Au from exceedingly rich
placer deposits, mostly between 1864 and 1869, and
"" 311 kg (10,000 oz) Au has been produced from lode
deposits (Koschmann and Bergendahl, 1968, p. 146).
The lode deposits are mostly gold-quartz veins along
fractures in quartz diorite and bedding planes in
shale zones of the Newland Limestone that were
intruded by the quartz diorite.
Two principal lode producers in the district are
the Miller (Slim Jim) and Satellite (Baker) pro-
perties; total Au production from each is not known
but presumably amounted to "" 100 kg. The Miller
has been a consistent, small-scale producer of rich
ore since its discovery in 1900. Reed (1951, p. 33)
reported that the ore averaged ""275 g/T Au and
158 g/T Ag (8 ozlT Au and 4.6 oz/T Ag, respectively),
although it is highly erratic in tenor. The lode is a
zone 0.3-1.3 m wide and is composed of irregular
quartz veinlets along the contact between shale of
the Newland Limestone and a small quartz diorite
stock; native Au also occurs in the host rocks, but
mostly is in the quartz veinlets. The Satellite
consists of a quartz vein 1.0-1.5 m wide in shale of
the Newland Limestone near a small quartz diorite
stock. The ore is reported to average 18.87 gm Au
and 37.7 glT Ag (0.55 ozlT Au and 1.1 ozlT Ag).
Another Au deposit, the Humming Bird, has minor
production averaging 42.9 glT Au and 58.3 gm/T Ag
(1.25 oz/T Au and 1.7 oz/T Ag) from fracture fillings
and bedding plane replacements of quartz with py-
rite and chalcopyrite in veins 10-45 cm thick in the
Newland Limestone.
The Beaver Creek district is estimated to have
produced ""2,083.7 kg (67,000 oz) Au from placers
and;:=: 93.3 kg (3,000 oz) Au from lodes (Roby, 1950).
Quartz veins with sparse pyrite, chalcopyrite, and
galena are"" 10 cm thick and occur in shales of the
Newland Limestone near dioritic intrusions. Princi-
pal lode deposits include the Snowbank, Bigler, and
Porcupine. Only the Snowbank has recorded pro-
duction; since 1932 the ore averaged 24 g/T Au and
16 g/T Au (0.7 ozlT Au and 0.47 oz/T Ag). This de-
posit consists of a narrow quartz vein nearly parallel
to bedding of the host rock.
DISCUSSION
Concentration of precious- and base-metal veins
hosted by Belt rocks along the western end of the
Lewis and Clark line was recognized by Billingsley
and Locke (1941), but the influence of the eastern
end of the line on mineralization has received little
attention. Although the origin of ore-bearing fluids
and timing of ore deposition is not totally clear
along the Lewis and Clark line, it is significant that
the mineral deposits occur in a strongly deformed
area where basement faults have been recurrently
active since Middle Proterozoic time.
Most Au-Ag lodes in the York, Confederate
Gulch, and Beaver Creek districts of the Big Belt
uplift are epigenetic, and many post-date the small
and scattered intrusions that appear to be of Pre-
660
L. A. Woodward
cambrian and (or) Cretaceous-Tertiary age, al-
though some of the bedded deposits hosted by
Proterozoic Belt rocks may be syngenetic and (or)
syndiagenetic. Leach and others (1988) presented
compelling evidence that the vein deposits of the
Coeur d'Alene district resulted from regional
metamorphism of Belt strata "" 850 Ma. Chan-
nelization of ore fluids occurred along major faults,
and the principal control on localization of the veins
is generally agreed to be complex structures of the
Lewis and Clark line (Hobbs and others, 1965).
Thus, the timing of mineralization along the line is
of different ages.
It has been suggested that metals were derived
from Belt strata in many parts of the Belt basin.
McClernan (1984, p. 39) speculated that lode de-
posits in the Big Belt uplift contain remobilized Au
initially trapped in algal mats of the Newland
Limestone, with ultimate derivation of Au from the
Archean terrane south of the Willow Creek growth
fault that bounds the south side of the Helena
embayment (Figure 1). I am not aware of any work
to indicate that algal growths of the Newland con-
tain anomalous amounts of Au, however.
Hershey (1916) proposed that Pb and Zn in
veins in the Coeur d'Alene district were remobilized
from disseminations in the Prichard Formation.
Leach and others (1988) indicated that Ag in the
Coeur d'Alene district was derived largely from
Revett-St. Regis rocks of the Belt Supergroup. Pb
isotope studies of deposits along the western end of
the Lewis and Clark line show that the Pb is of
Precambrian age (Zartman and Stacey, 1971; Zart-
man, this volume). Widespread and very large
strata-bound Cu-Ag deposits in Belt rocks of north-
western Montana (Harrison, 1972) were interpreted
to be syngenetic (Garlick, 1987). Lange and Sherry
(1983) suggested that upward movement of metal-
bearing solutions through syndepositional, base-
ment-controlled faults formed the Cu-Ag sulfide-
bearing zones. Hayes and Einaudi (1986) proposed
that ore emplacement in these deposits occurred
during diagenesis from solutions that migrated up-
ward and laterally through the Belt sediments.
Massive sedimentary exhalative Fe and Zn sul-
fide deposits hosted by Belt rocks occur along the
fault-bounded southern margin of the Belt basin
(Thorson, 1984) (Figure 1, location 3). Bedded iron
oxide and fine-grained silica deposits in the New-
land Formation at the Black Butte-Sheep Creek
deposit (McClernan, 1969) were interpreted as an
intensely oxidized gossan of a massive sulfide de-
posit. This deposit occurs near the Volcano Valley
thrust (Figure 1, location 4), a growth fault with the
south side down during deposition of the Newland
Limestone (Godlewski and Zieg, 1984). Himes and
others (1988) reported that the mineralization at
Sheep Creek consists of thinly laminated to massive
pyrite, chalcopyrite, cobaltiferous pyrite, sphalerite,
and galena interbedded with clastic sediments and
barite; the sulfide bodies are associated with debris-
flow conglomerates and are interpreted to have
formed from exhalative activity near a growth fault
(Himes and others, 1988). Thus, there appears to
have been widespread syngenetic to diagenetic de-
position of Cu, Ag, Zn, Pb, and Co, and possibly Au,
in Belt rocks.
I interpret the Au-Ag lodes in the Greyson
Shale and Newland Limestone along the Lewis and
Clark line in the Big Belt uplift to be localized by
basement faults active during deposition of the
Greyson and Newland formations. The metals may
have been remobilized from crystalline basement
rocks by deep convection of meteoric water along
basement faults (Nesbitt, 1988) that were active
during sedimentation of the Belt strata. If so, some
of the deposits were remobilized (perhaps multiply)
into epigenetic veins. Expulsion of metal-bearing
brines during dewatering of the Belt sediments
(Sawkins, 1984) with fluid release from geopres-
sured zones during faulting may have occurred.
Derivation of ore fluids during Late Proterozoic
regional metamorphism, as suggested for the Coeur
d'Alene district (Leach and others, 1988), does not
seem as likely for the Big Belt area because the Belt
strata are essentially non-metamorphosed. Wide-
spread intrusion of volumetrically minor mafic
dikes and sills ""1,075-1,200 and 750-830 Ma (Rey-
nolds, 1984; McGrew, 1977) could have provided
heat sources for deep circulation of meteoric waters
through the Belt strata. Cretaceous-Tertiary intru-
sions were also involved, as the veins in several
cases are adjacent to and within these igneous
bodies.
Although the source of metals, the origin of the
ore-bearing fluid, and the details of the timing of ore
deposition are problematic, it is clear that Au-Ag
lode deposits hosted by Proterozoic Belt rocks are
concentrated along the Lewis and Clark line in
west-central Montana. If basement growth faults
that were active during Belt sedimentation are the
major control in localization of Au-Ag deposits, as
suggested here, then the trends of these faults in the
Belt strata are favorable for mineral exploration.
Au-Ag deposits hosted by Proterozoic strata along the Lewis and Clark Line 661
These growth faults mark the margins of differen-
tially subsided basement blocks and are subtle fea-
tures that can be located by detailed stratigraphic
and sedimentologic studies, as discussed by Winston
(1986).
ACKNOWLEDGMENTS
I thank Fess Foster, Ian Lange, and Gerald A.
Zieg for reviewing the manuscript and suggesting
major improvements.
REFERENCES
BAITIS, H. W., 1988, York-A Proterozoic "shale"-hosted
gold system in Montana [abstract]: Abstracts of the 94th
Annual Convention of the Northwest Mining Association,
Spokane, Washington, USA, p. 7.
BILLINGSLEY, P., and A. LOCKE, 1941, Structure of ore
districts in the continental framework: American Institute
of Mining and Metallurgical Engineers Transactions, v.
144, p. 9-64.
CATANZARO, E. J., and J. L. KULP, 1964, Discordant
zircons from the Little Belt (Montana), Beartooth (Mon-
tana), and Santa Catalina (Arizona) Mountains: Geo-
chi mica et Cosmochimica Acta, v. 28, p. 87-124.
DANIEL, F., and R. B. BERG, 1981, Radiometric Dates of
Rocks in Montana: Montana Bureau of Mines and Geo-
logy, Bulletin 114, 136 p.
GARLICK, W. G., 1987, Genesis of the Spar Lake strata-
bound copper-silver deposit, Montana, Part I: Controls
inherited from sedimentation and preore diagenesis - A
discussion: Economic Geology, v. 82, p. 1967-1971.
GODLEWSKI, D. W., and G. A. ZIEG, 1984, Stratigraphy
and depositional setting of the Precambrian Newland
Limestone; pp. 2-4 in S. W. Hobbs (ed.), The Belt: Ab-
stracts with Summaries - Belt Symposium II, 1983:
Montana Bureau of Mines and Geology, Special Publication
90,117p.
HARRISON, J. E., 1972, Precambrian Belt basin of
northwestern United States: Its geometry, sedimentation,
and copper occurrences: Geological Society of America
Bulletin, v. 83, p. 1215-1240.
HARRISON, J. E., and M. W. REYNOLDS, 1976, Western
U.S. continental margin: A stable platform dominated by
vertical tectonics in the late Precambrian [abstract]:
Geological Society of America Abstracts with Programs, v.
8, p. 905.
HARRISON, J. E., A. B. GRIGGS, and J. D. WELLS, 1974,
Tectonic Features of the Precambrian Belt Basin and
Their Influence on Post-Belt Structures: U.S. Geo-
logical Survey, Professional Paper 866, 15 p.
HAYES, T. S., and M. T. EINAUDI, 1986, Genesis of the
Spar Lake strata-bound copper-silver deposit, Montana:
Part I, Controls inherited from sedimentation and preore
diagenesis: Economic Geology, v. 81, p. 1899-1931.
HERSHEY, O. H., 1916, Origin and Distribution of Ore
in the Coeur d'Alene Mining District: Washington
Mining Scientific Press, Spokane, Washington, USA, 32 p.
HIMES, M. D., F. T. BOURNS, H. C. GOLDEN, J. A.
PERELLO, S. A. TAYLOR, R. J. WINDER, P. W. RANKIN, S.
JENNINGS, and G. A. ZIEG, 1988, The Sheep Creek project,
a Proterozoic sediment-hosted massive sulfide deposit,
Meagher County, Montana [abstract]: Abstracts of the
94th Annual Convention of the Northwest Mining Asso-
ciation, Spokane, Washington, USA, p.17.
HOBBS, S. W., A. B. GRIGGS, R. E. WALLACE, and A. B.
CAMPBELL, 1965, Geology of the Coeur d'Alene Dis-
trict, Shoshone County, Idaho: U.S. Geological Survey,
Professional Paper 4 78, 139 p.
JAMES, H. L., and C. E. HEDGE, 1980, Age of the basement
rocks of southwest Montana: Geological Society of America
Bulletin, v. 91, p. 11-15.
JOHNSON, E. A., 1973, Geology and Gold Deposits of
the Confederate Gulch-White Gulch Area, Broad-
water County, Montana: MS thesis, Montana College of
Mineral Science and Technology, Butte, Montana, USA, 53
p.
KNOPF, A., 1963, Geology of the Northern Part of the
Boulder Bathylith and Adjacent Area, Montana: U.S.
Geological Survey, Miscellaneous Geologic Investigations
Map 1-381, scale 1:48,000.
KOSCHMANN, A. H., and M. H. BERGENDAHL, 1968,
Principal Gold-Producing Districts of the United
States: U.S. Geological Survey, Professional Paper 610,
283p.
LANGE, 1. M., and R. A. SHERRY, 1983, Genesis of the
sandstone (Revett) type of copper-silver occurrences in the
Belt Supergroup of northwestern Montana and north-
eastern Idaho: Geology, v. 11, p. 643-646.
LEACH, D. L., G. P. LANDIS, and A. H. HOFSTRA, 1988,
Metamorphic origin of the Coeur d'Alene base- and pre-
cious-metal veins in the Belt basin, Idaho and Montana:
Geology, v. 16, p. 122-125.
LORENZ, J. C., 1984, The function of the Lewis and Clark
fault system during the Laramide orogeny; pp. 221-230 in
1984 Guidebook, Northwestern Montana: Montana
Geological Society, Billings, Montana, USA.
662
L. A. Woodward
MCCLERNAN, H. G., 1969, Geology of the Sheep Creek
Area. Meagher County, Montana: MS thesis, Montana
College of Mineral Science and Technology, Butte, Mon-
tana, USA, 75 p.
__ , 1980, Metallogenic Map of the White Sulphur
Springs Quadrangle, Central Montana: Montana Bur-
eau of Mines and Geology, Geologic Map 7, scale 1:250,000.
__ , 1983, Metallic Mineral Deposits of Lewis and
Clark County, Montana: Montana Bureau of Mines and
Geology, Memoir 52,73 p.
__ , 1984, Gold deposits of the Big Belt Mountains:
Montana Bureau of Mines and Geology Bulletin, v. 121, p.
33-40.
McGREW, L. W., 1977, Geologic Map of the Sixteen
Quadrangle, GaUatin and Meagher Counties, Mon-
tana: U.S. Geological Survey, Geologic Quadrangle Map
GQ-1383, scale 1:24,000.
McMANNIS, W. J., 1959, Salient tectonic features of
western Montana; pp. 71-75 in GSA Rocky Mountain
Section: Guidebook to Field Trips: Geological Society
of America, Boulder, Colorado, USA.
__ ,1963, LaHood Formation - A coarse facies of the
Belt Series in southwestern Montana: Geological Society of
America Bulletin, v. 74, p. 407-436.
MARVIN, R. F., and S. W. DOBSON, 1979, Radiometric
ages: Compilation B: U.S. Geological Survey, Isochron!
West, no. 26, p. 3-32.
NESBITI', B. E., 1988, Gold deposit continuum: A genetic
model for lode Au mineralization in the continental crust:
Geology, v.16, p. 1044-1048.
PARDEE, J. T., and F. C. SCHRADER, 1933, Metalliferous
Deposits of the Greater Helena Mining Region, Mon-
tana: U.S. Geological Survey, Bulletin 842, 318 p.
PRICE, R. A., 1984, Tectonic setting of the Purcell (Belt)
rocks of the southeastern Canadian Cordillera and adja-
cent parts of the United States; pp. 47-48 in S. W. Hobbs
(ed.), The Belt: Abstracts with Summaries - Belt
Symposium II, 1983: Montana Bureau of Mines and
Geology, Special Publication 90, 117 p.
REED, G. C., 1951, Mines and Mineral Deposits (Except
Fuels), Broadwater County, Montana: U.S. Bureau of
Mines, Information Circular 7592, 62 p.
REID, R. R., 1984, Structural control of Coeur d'Alene ore
deposits; pp. 49-51 in S. W. Hobbs (ed.), The Belt: Ab-
stracts with Summaries - Belt Symposium II, 1983:
Montana Bureau of Mines and Geology, Special Publication
90, 117p.
REYNOLDS, M. W., 1979, Character and extent of Basin-
Range faulting, western Montana and east-central Idaho;
pp. 185-193 in Proceedings of the Basin and Range
Symposium: Rocky Mountain Association of Geologists!
Utah Geological Association, Salt Lake City, Utah, USA.
__ , 1984, Tectonic setting and development of the Belt
basin, northwestern United States; pp. 44-46 in S. W.
Hobbs (ed.l, The Belt: Abstracts with Summaries -
Belt Symposium II, 1983: Montana Bureau of Mines and
Geology, Special Publication 90, 117 p.
ROBY, R. N., 1950, Mines and Mineral Deposits (Except
Fuels), Meagher County, Montana: U.S. Bureau of
Mines, Information Circular 7540, 40 p.
RYE, R 0., J. F. WHELAN, J. E. HARRISON, and T. S.
HAYES, 1984, The origin of copper-silver mineralization in
the Ravalli Group as indicated by preliminary stable
isotope studies; pp. 104-107 in S. W. Hobbs (ed.), The Belt:
Abstracts with Summaries - Belt Symposium II,
1983: Montana Bureau of Mines and Geology, Special
Publication 90, 117 p.
SAWKINS, F. J., 1984, Ore genesis by episodic dewatering
of sedimentary basins: Application to giant Proterozoic
lead-zinc deposits: Geology, v. 12, p. 451-454.
SHAFFER, W.L., 1971, Geology of the Hogback
Mountain Area. Northern Big Belt Mountains, Mon-
tana: MS thesis, University of New Mexico, Albuquerque,
New Mexico, USA, 66 p.
THORSON, J. P., 1984, Suggested revisions of the lower
Belt Supergroup stratigraphy of the Highland Mountains,
southwestern Montana; p. 10-12 in S. W. Hobbs (ed.), The
Belt: Abstracts with Summaries - Belt Symposium
II, 1983: Montana Bureau of Mines and Geology, Special
Publication 90, 117 p.
WALLACE, C. A., J. E. HARRISON, M. R. KLEPPER, and J.
D. WELLS, 1976, Carbonate sedimentary breccias in the
Wallace Formation (Belt Supergroup), Idaho and Mon-
tana, and their paleogeographic significance [abstract):
Geological Society of America, Abstracts with Programs, v.
8, p.1159.
WEIDMAN, R. M., 1965, The Montana lineament; pp. 137-
143 in 16th Field Conference Guidebook: Billings Geo-
logical Society, Billings, Montana, USA, 169 p.
WINSTON, D., 1983, Middle Proterozoic Belt Basin
syndepositional faults and their influence on Phanerozoic
thrusting and extension: American Association of Petro-
leum Geologists Bulletin, v. 67, p. 1361
__ , 1986, Middle Proterozoic tectonics of the Belt basin,
western Montana and northern Idaho; pp. 245-257 in S. M.
Roberts (ed.l, Belt Supergroup: A Guide to Proterozoic
Rocks of Western Montana and Adjacent Areas: Mon-
tana Bureau of Mines and Geology, Special Publication 94,
311 p.
WOODWARD, L. A, 1970, Time of emplacement of the Pinto
Diorite, Little Belt Mountains, Montana: Wyoming Geo-
Au-Ag deposits hosted by Proterozoic strata along the Lewis and Clark Line 663
logical Association, Earth Science Bulletin, v. 3, no. 3, p.
15-26.
__ , 1981, Tectonic framework of disturbed belt of west-
central Montana: American Association of Petroleum Geo-
logists Bulletin, v. 65, p. 291-302.
ZARTMAN, R. E., and J. S. STACEY, 1971, Lead isotopes
and mineralization ages in Belt Supergroup rocks, north-
western Montana and northern Idaho: Economic Geology,
v. 66, p. 849-860.
ZIEG, G. A., 1981, Stratigraphy, Sedimentology, and
Diagenesis of the Precambrian Upper Newland Lime-
stone, Central Montana: MS thesis, University of Mon-
tana, Missoula, Montana, USA, 182 p.
__ , and D. W. GODLEWSKI, 1986, Road log no. 1: A
traverse across the eastern Belt Basin from Neihart to
Townsend, Montana; pp. 1-16 in S. M. Roberts (ed.), Belt
Supergroup: A Guide to Proterozoic Rocks of Western
Montana and Adjacent Areas: Montana Bureau of
Mines and Geology, Special Publication 94, 311 p.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Geotectonic setting of western Pacific gold deposits
RICHARD H. SILLITOE
27 West Hill Park, Highgate Village, London, N6 6ND, England, UK
(received October 11, 1988; revision accepted April 26, 1989)
ABSTRACT
Fifty-six principal gold deposits in the western Pacific region were generated in multiple volcanoplutonic arcs
during the last 25 million years. At least 75% of the contained gold was introduced in association with
intrusive stocks, commonly porphyries. The gold-bearing magmatic arcs were all constructed in response to
subduction of oceanic lithosphere. However, the nature of the magmatism and the style of associated gold
mineralization depend to some degree on the arc setting. Calcalkalic suites of mainly andesitic to dacitic
composition are typified by intrusion-related gold deposits, especially those of porphyry and acid-sulfate
epithermal types, and were constructed along neutral to compressional arcs. Some of these arcs are paralleled
by transcurrent fault zones. Weakly bimodal suites dominated by silicic volcanic rocks of calcalkalic
composition and associated adularia-sericite type epithermal gold deposits characterize volcanotectonic
depressions within extensional arcs. In the D'Entrecasteaux Islands, Papua New Guinea, the extension and
consequent detachment faulting are attributed to the on-land propagation of an oceanic spreading center.
Magmatic centers of alkalic petrochemistry also contain adularia-sericite type epithermal gold deposits and
may be linked to more advanced extensional events in back-arc settings. Determination of the precise
geotectonic settings of gold-bearing arcs is controversial even in the youthful western Pacific region, and the
difficulties involved are compounded in older island-arc terranes. Moreover, Mesozoic and older arcs are
likely to have lost most of their epithermal and sub volcanic gold deposits as a result of uplift and erosion, as
witnessed by the accreted Mesozoic arc terranes of west em North America.
INTRODUCTION
Gold deposits in western Pacific island arcs are
known from Japan in the north through Taiwan, the
Philippines, New Guinea, the Solomon Islands, and
Fiji to North Island, New Zealand in the south
(Figures 1, 2, and 3). A recent analysis of this gold
province (Sillitoe, 1989) reveals the presence of 56
principal deposits, defined as those with> 10 metric
tons of contained gold: a total for the province of
> 7000 metric tons of gold (Table 1; Figures 1, 2,
and 3).
None of the principal gold deposits is older than
25 million years, all but one are < 20 million years,
and 75% of them are < 10 million years old (Sillitoe,
1989). The youthfulness of deposits maximizes the
chances of relating gold mineralization to the com-
plex and rapid tectonic evolution of the western
Pacific region. This evolution is the result of conver-
gence of the major Eurasian and Indo-Australian
plates with the Pacific plate. Subduction of oceanic
lithosphere, expansion and closure of back-arc
basins, transform faulting, and collisions involving
continents, continental fragments, remnant and
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 665-678. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
665
666
R. H. Sillitoe
Table 1. Principal gold deposits in the western Pacific region
(keyed to Figures 1, 2, and 3).
Japan 1 Konomai 9 Bajo
2 Teine 10 Taio
3 Chitose 11 Okuchi
4 Nurukawa 12 Hishikari
5 Sado 13 Yamagano
6 Takatama 14 Kushikino
7 Toi 15 Kasuga
8 Seigoshi 16 Iwato
Taiwan 17 Chinkuashih
Philippines 18 Lepanto 27 Bulawan
19 Marian 28 Placer
20 Baguio 29 Siana
21 Thanksgiving 30 Diwalwal
22 Santo Tomas II 31 Compostela
23 Dizon 32 Masara
24 Nalesbitan 33 Hijo
25 Paracale 34 Kingking
26 Masbate
Irian Jaya, Indonesia 35 Ertsberg-Grasberg
Papua New Guinea 36 Ok Tedi 44 Woodlark
37 Nena 45 Umuna
38 Porgera 46 Wild Dog
39 Wafi River 47 Sorowar
40 Wau-Edie Creek 48 Ladolam
41 Kerimenge 49 Kabang
42 Hidden Valley 50 Panguna
43 Wapolu
Solomon Islands 51 Gold Ridge
Fiji 52 Emperor
New Zealand 53 Golden Cross 55 Thames
54 Waihi
active island arcs, and oceanic plateaux have all
characterized the region's Miocene to present-day
development. As a consequence, western Pacific
island arcs are products of the coalescence of many
discrete and commonly far-traveled tectono-
stratigraphic terranes.
This geotectonic overview highlights the funda-
mental metallogenic role played by the subduction
process but stresses the variety of subduction-
related settings in which western Pacific gold
deposits are generated. Finally, selected settings
are compared briefly with gold-bearing regions
along the eastern Pacific rim.
56 Karangahake
GOLD DEPOSITS
Western Pacific gold deposits comprise porphy-
ry-type stockworks, contact-metasomatic skarns,
and the unique Porgera deposit - all related in-
timately to intrusions, volcanic-hosted epithermal
deposits of acid-sulfate and adularia-sericite types
and a single Kuroko-type massive sulfide deposit
(Figures 1, 2, and 3; Sillitoe, 1989).
The porphyry-type stockworks are confined to
K-silicate-altered porphyry stocks of dioritic, quartz
dioritic, granodioritic, or monzonitic composition
and their immediate wall-rocks, with gold accom-
40"
20"
es,
o
I
Geotectonic setting of western Pacific gold deposits
667
120"
160"
1000 km.
SEA OF JAPAN
PHILIPPINE
SEA
PALAU 0;
150"
Figure 1. Principal gold deposits and selected geotectonic elements of the Japan-Taiwan-Philippines-Halmahera region.
Numbers are keyed to deposit names in Table 1. See Figure 3 for legend and data sources.
668 R. H. Sillitoe
PALAU 0;
PACIFIC
0
0
OCEAN
/\
r .

I I
1000 km.
150
0 160
0
E.
Figure 2. Principal gold deposits and selected geotectonic elements of the New Guinea-Solomon Islands region. Numbers are
keyed to deposit names in Table 1. See Figure 3 for legend and data sources.
panied by economic amounts of copper. Auriferous
skarns containing copper or zinc abut stocks but are
not widespread (three only) in the region because of
the scarcity of shelf carbonate sequences. The Por-
gera deposit occurs as a stockwork in dioritic stocks
and contiguous calcareous mudstones, but it lacks
K-silicate alteration and appreciable skarn develop-
ment and contains zinc and lead rather than copper
(Fleming and others, 1986).
Adularia-sericite type epithermal deposits car-
rying carbonates and small amounts of base metals
constitute 65% of western Pacific gold deposits
(Sillitoe, 1989). Many of them are single or multiple
veins controlled by extensional faults or fractures,
but the largest deposit of this type, Ladolam in Lihir
Island, comprises flat-lying, tabular breccias and
stockworks (Davies and Ballantyne, 1987). Acid-
sulfate type epithermal deposits tend to be more
varied in geometry (veins, breccias, massive re-
placements) and generally contain copper as en-
argite and/or other high-sulfidation phases. Sillitoe
(1989) calculated that at least 75% of the gold in
principal western Pacific deposits occurs in porphy-
ry-type stockworks or is related closely to minera-
lized (mainly porphyry) stocks. Many epithermal
deposits occupy distal sites in porphyry systems -
those of adularia-sericite type commonly around
stocks, and those of acid-sulfate type above them.
These observations suggest strongly that a major
proportion of western Pacific gold was supplied by
magmas as a component of magmatic-hydrothermal
fluids, although cooler and more dilute meteoric
fluids were commonly instrumental in the final
concentration of at least some of the gold, including
that classified as epithermal. The gold contents of
meteoric-hydrothermal fluids were either contribu-
ted directly during admixture with magmatic brines
or volatiles, and/or indirectly as a result of partial
remobilization of magmatic-hydrothermal gold ore/
protore (Sillitoe, 1989).
Geotectonic setting of western Pacific gold deposits 669
lEGEND
ELEMENTS
..A.-...A- Active subduction zone
A....A- Inactive subduction zone
Active spreading center
+ l' " d"
Inactive sprea mg center
T ronscurrent fault (incl. transform)
Thrust fault
PACIFIC OCEAN
SOUTH FIJI
X
BASIN
Au DEPOSIT TYPE
... Porphyry capper
Skarn
COPOMANDEL
o Porgero
o Acid-sulfate } E "th I
pI erma
Adularia-sericite
Kuroko-type massive sulfide
160'
I
1000 km"
170' W. 40'
Figure 3. Principal gold deposits and selected geotectonic elements of the Solomon Islands-Vanuatu-Fiji-North Island, New
Zealand, region. Numbers are keyed to deposit names in Table 1. Gold deposits from Sillitoe (1989); geotectonic elements
slightly modified from Circum-Pacific Council for Energy and Mineral Resources (1984) and Hamilton (1979).
GEOTECTONIC SETTING
Preamble
From the standpoint of gold metallogeny, sub-
duction of Pacific marginal basin lithosphere and
construction of I-type volcanoplutonic arcs were the
most important geodynamic processes since the
beginning of the Miocene in the western Pacific
region. All of the region's principal gold deposits are
hosted by subduction-related arcs. Individual sub-
duction episodes were generally short-lived 20
670
R. H. Sillitoe
million years) and were commonly terminated by
changes in plate boundary configurations in re-
sponse to collisional events. However, a variety of
subduction-related arc settings are distinguishable
in the western Pacific region, and many of them
may be related confidently to the precursors of cur-
rently active trenches. A description of some of the
different arc settings constitutes much of this
report.
The distribution of the various types of princi-
pal gold deposits in the western Pacific region does
not appear to bear any relationship to the age of the
oceanic lithosphere that was being subducted at the
time of, or immediately preceding, mineralization.
Nor does gold mineralization seem to have been
influenced in any obvious way by the nature and
thickness of the island-arc crust upon which the host
volcanoplutonic arcs were constructed. For ex-
ample, gold-rich porphyry deposits occur in ensi-
matic (Panguna) and ensialic (Ok Tedi) arcs as well
as in arcs with intermediate crustal foundations
(Santo Tomas II). However, the type of gold deposit
does display a measure of dependency on the char-
acter of the associated magmatic suite which, in
turn, seems to be diagnostic of the arc setting.
Simple Arcs
Simple arcs are defined as those marked by
linear belts of andesitic to dacitic volcanic rocks and
subjacent intrusions of intermediate composition,
all dominantly characterized by ca1calkalic petro-
chemistry. Simple arcs seem to have been construc-
ted in response to roughly orthogonal plate con-
vergence. Gold-bearing arcs of this type charac-
terize the Luzon Central Cordillera, built during
late Miocene to Holocene eastward subduction from
the Manila trench (Figure 1), and the Solomon
Island chain, underlain by a Pliocene to Holocene,
N-dipping subduction zone (Figure 2). The New
Guinea Mobile Belt, along the central line of New
Guinea as far east as the Woodlark and Umuna
(Misima) gold deposits (Figure 2), is thought to be of
the same type albeit more deeply eroded. The arc is
believed to have been constructed in the Miocene
during southward subduction of marginal basin
lithosphere, an event terminated by late Miocene-
early Pliocene collision between the Australian
craton and an island arc (Hamilton, 1979; Kroenke,
1984; Hill and Hegarty, 1987).
In contrast to the subduction of normal mar-
ginal basin lithosphere, the central portion of the
Solomons arc (New Georgia Group) has been under-
thrust northeastward since the Pliocene by the
active Woodlark spreading center (Figure 2), a
situation that has given rise to eruption of un-
usually abundant high-Mg picritic and basaltic
rocks as well as to other distinctive features
(Dunkley, 1983; Ramsay and others, 1984). How-
ever, detailed geochemical studies by Johnson and
others (1987) suggest strongly that magmatism in
the New Georgia Group is linked more closely to a
remnant SW -dipping slab than to the subducted
spreading center. No principal gold deposit has
been defined yet in the New Georgia Group, al-
though epithermal gold mineralization is under
exploration.
Simple arcs are typified by intrusion-related
gold deposits, especially those of porphyry type, and
by acid-sulfate type epithermal deposits. Where
adularia-sericite type epithermal deposits are pre-
sent, they commonly exhibit sericitic rather than
adularia-sericite alteration, a greater than normal
abundance of base metals, and a close association
with intrusions, especially porphyry copper stocks
(Sillitoe, 1989).
Arcs Associated with Transcurrent Faulting
The western Pacific region provides instructive
examples of simple volcanoplutonic arcs coincident
with, paralleled by, and in proximity to transcur-
rent fault zones, which result from oblique con-
vergence, including collision. One of the most clear-
cut of these is in the eastern part of the Philippine
archipelago (Figure 1), where Pliocene to Holocene
arc activity is closely associated with the Philippine
fault zone, a major sinistral strike-slip system (Al-
len, 1962). The evolutionary history of the Philip-
pine fault zone was complex and may date back to
the early Tertiary (Wolfe, 1983). Karig (1983)
favored some 200 km of transcurrent displacement
on the northern part of the fault zone in central
Luzon during the last 15 million years, perhaps in
response to oblique subduction. The southern part
of the fault zone, in eastern Mindanao (Figure 1), is
inferred to have been active at least during the
Pliocene to Holocene interval and, by some workers
(Cardwell and others, 1980), to coincide with the
surface trace of a Pliocene collisional suture.
Geotectonic setting of western Pacific gold deposits 671
Arc magmatism along the Philippine fault zone
is andesitic to dacitic in composition and, in common
with the associated gold deposits, is essentially
indistinguishable from that characteristic of simple
arcs. The control of gold deposits (and porphyry
copper deposits; Sillitoe and Gappe, 1984) by the
fault zone is particularly evident in eastern Min-
danao (Figure 4), where second-order splays rather
than first-order breaks localized ore deposition. It is
concluded that oblique westward subduction from
the Philippine trench (Figure 1) instigated both
magmatism and transcurrent faulting, and that the
fault zone, especially extensional domains within it,
facilitated crustal ascent of gold-bearing magmas.
Transcurrent faulting, following continent-
island arc collision, also characterized post-Miocene
times in northern New Guinea (e.g., Hill and Hegar ..
ty, 1987) and the Plio-Pleistocene in Taiwan (e.g.,
Ernst and Jahn, 1987). In the New Guinea High-
lands, Pleistocene shoshonitic volcanism and em-
placement of the Ok Tedi porphyry copper-gold
deposit took place along the southern side of the
earlier magmatic arc of the New Guinea Mobile
Belt, in a position south of the main zone of left-
lateral strike-slip faulting (including the Ramu-
Markham fault; Figure 2). The magmatic activity
is attributed to reactivation of a residual oceanic
slab or to renewed post-collisional oblique subduc-
tion (Hamilton, 1979; Kroenke, 1984; Sillitoe, 1987;
Warnaars, 1987) rather than to lower crustal partial
melting triggered by collision-induced uplift (e.g.,
Mason and Heaslip, 1980). Magmas and contained
metals intruded an actively developing foreland
thrust belt (Figure 2). In Taiwan, generation of the
Chinkuashih acid-sulfate type gold deposit may be
linked to northward subduction beneath the north-
eastern part of the island at the same time as trans-
current faulting was active only 140 km to the south
(Figure 1).
Arcs Associated with Rifting
Many of the principal gold deposits in Japan
and North Island, New Zealand, were generated
during intervals of subduction-related intra-arc
rifting, which may be considered as unsuccessful
attempts to open marginal ocean basins. The exten-
sional events may be linked to oceanward retreat of
trenches.
The Kuroko-type gold deposit at Nurukawa and
the associated base-metal deposits in the Hokuroku
basin of northeastern Honshu were emplaced in a
short-lived, mid-Miocene rift zone which failed to
open sufficiently to result in a Japan Sea-type
marginal basin (Sillitoe, 1982; Cathles and others,
1983; Urabe, 1987).
Similar intra-arc rifts, albeit partly subaerial
rather than entirely submarine, have been active
since the late Miocene in the gold provinces of
Kyushu (Figure 5; Izawa and Urashima, 1987) and
North Island (Figure 3; Skinner, 1986). Rifts along
the eastern side of the Coromandel Peninsula in-
clude the Waihi graben and others detected offshore.
The Taupo-Rotorua Depression (Figure 3), the zone
of currently active subduction-related extension in
North Island, is well known because of its volcanism
and high-temperature geothermal systems.
The ensialic intra-arc rifts of Kyushu and the
Coromandel Peninsula may be taken as the on-land
continuations of marginal basins in the Okinawa
Trough (Figure 1; Letouzey and Kimura, 1986) and
Havre Trough (Figure 3; Cole, 1984), respectively.
The obliquity of subduction beneath the Taupo-
Rotorua Depression is the cause of strike-slip fault-
ing in the fore-arc region (Figure 3).
Gold-bearing magmatic suites in zones of ensia-
lie extension are characterized by an abundance of
rhyolitic to rhyodacitic rocks, and in a regional
sense commonly exhibit a bimodal character be-
cause of the presence of small volumes of coeval
high-alumina basaltic rocks. Bimodality was a
salient feature of the rift in northeastern Honshu
during Kuroko generation (Sillitoe, 1982; Cathles
and others, 1983; Urabe, 1987). However, these
extensional areas also underwent normal andesitic-
dacitic magmatism, as shown by Kyushu during the
late Miocene to Quaternary period (Izawa and
Urashima, 1987), and the Taupo-Rotorua Depres-
sion during Quaternary times (e.g., Cole, 1984).
Gold deposits associated with intra-arc ex-
tension are typified by the volcanic-hosted adularia-
sericite type epithermal vein systems of Kyushu and
North Island and, under submarine conditions, by
Kuroko deposits. It is noteworthy though that
andesitic-dacitic magmatism in these rift zones
tends to be associated with the same types of gold
deposits as in simple volcanoplutonic arcs. For ex-
ample, andesitic volcanic centers in southern Kyu-
shu host the Nansatsu-type acid-sulfate gold de-
posits, such as Iwato and Kasuga (Figure 1; Izawa
and Urashima, 1987). In North Island, however,
late Miocene generation of porphyry copper-type
mineralization and the associated acid-sulfate type
672
00
o
0 km 200
I I I

R. H. Sillitoe
!J
t:Jv
0
0
PARACALE
/
/
High-angle faults
Principal gold deposits 0
Limit of Pre-Jurassic .,."
continental crust
Q
Figure 4. Relations between principal gold deposits in the Philippines (from Sillitoe, 1989) and major high-angle faults,
including the Philippine strike-slip fault zone (from Philippine Bureau of Mines and Geo-Sciences, 1982).
Geotectonic setting of western Pacific gold deposits 673
G) Late Cenozoic volcanic centres
c:z:; Shimanto Supergroup
@ Principal gold deposit
---
Fault zone
Graben
Figure 5. Relations between Pliocene-
Quaternary gold deposits, extensional fea-
tures, and basement terranes in Kyushu,
southwestern Japan. MTL, Median tec-
tonic line; BTL, Batsuzo tectonic line.
Taken from Izawa and Urashima (1987).
gold deposit in the Thames district (Figure 3)
probably pre-dated the extensional event in the
Coromandel Peninsula (Sillitoe, 1989).
The extensional setting for the adularia-sericite
type gold deposit at Wapolu in the D'Entrecasteaux
Islands, eastern New Guinea (Figure 2), is attribu-
ted to a rather different geodynamic scenario. Rift-
ing is associated with rhyolitic magmatism and the
development of metamorphic core complexes
flanked by detachment faults. Extension is the
result of the westward propagation into continental
crust of the post-Miocene spreading center in the
W oodlark marginal basin (Figures 2 and 6; Davies
and others, 1984; Davies and Warren, 1988). South-
ward subduction from the Trobriand trench (Fig-
ures 2 and 6) during the extensional phase was
suggested first by Hamilton (1979).
Back-Arc Extension
In the western Pacific region, several principal
gold deposits of adularia-sericite type are associated
with alkalic rocks rather than calcalkalic suites
displaying signs of bimodality (Sillitoe, 1989) and
are inferred to occupy back-arc extensional settings.
The precise geotectonic positions of these deposits -
Marian, Emperor, and Sorowar, Ladolam, and Ka-
bang in the Tabar-to-Feni island chain (Figures 1,
2, and 3) - remain imperfectly known but are
worthy of comment.
Marian and the associated syenitic intrusive
suite were emplaced at about 25 Ma along the west-
ern boundary of the Cagayan Valley graben, which
occupies a back-arc position with respect to the vol-
canoplutonic arc along the Luzon Central Cordillera
674
R. H. Sillitoe
Figure 6, Relations between the
Wapolu gold deposit in the D'Entre-
casteaux Islands, Papua New Guinea,
and rifting related to on-land propaga-
tion of the Woodlark spreading center.
Taken from Davies and others (1984).
148" 1SO' IS2' 1>1'
PLATE / <f>
/ S'
/ '1.
/
,/
/"
PENINSULA
IDn \ m WOODlARK 10'
'+ SASIN
INDO-AUSTRALIAN PLATE
+ Spreading Center
"'"T- Active (1) trench
Normal fault, showing
down thrown si de
Plio-Quaternary rift
w. 0 Wopolu Au deposit
"$- Strike slip fault or transform
(Figure 1; Knittel, 1983). Arc construction and
back-arc extension are best linked to a late Oligo-
cene to mid-Miocene episode of eastward subduction
from a proto-Manila trench (Stephan and others,
1986).
The basic-alkalic rocks and gold deposit at Em-
peror were emplaced during the final stages of
westward subduction and the early opening of the
contiguous North Fiji marginal basin (Figure 3;
Whelan and others, 1985).
Gold deposits in the Tabar-to-Feni chain, in-
cluding the giant Ladolam deposit, present a greater
geotectonic enigma. Wallace and others (1983) pro-
posed that emplacement of the associated magmas
was controlled by a NW -striking fault but divorced
from the influence of subduction. However, Sillitoe
(1987) and Warnaars (1987) suggested that the de-
posits occupy a back-arc position with respect to the
New Britain trench and came under the influence of
the down-dip extremity of the subducted slab when
New Ireland was located farther south with respect
to New Britain than today; its present more north-
erly position resulted from Plio-Pleistocene opening
of the Manus marginal basin (Figure 2; Kroenke,
1984).
GENERAL CONCLUSIONS
It is clear from the preceding section that most,
if perhaps not all, principal gold deposits in western
Pacific island arcs were emplaced during active
subduction. Where some doubt exists, as in the case
of Ok Tedi, Wapolu, Ladolam, and Emperor, resort
can be made to Johnson's (1987) hypothesis of
delayed partial melting of upper mantle material
modified geochemically by previous subduction
events. In this regard, the alkalic host rocks at
Emperor have been shown to possess arc-like trace-
element and isotopic signatures (Gill and others,
1984).
Although subduction, or its earlier geochemical
imprint on the mantle wedge, seems to constitute a
common denominator to all western Pacific gold
deposits, the state of stress in the overriding plate,
upon which volcanoplutonic arcs were constructed,
varied in both time and space. Overriding plates at
times of gold-deposit formation ranged from exten-
sional to compressional, and a transcurrent com-
ponent may be added in some cases.
Uyeda and Nishiwaki (1980) concluded that
arcs characterized by compressional and neutral
stress regimes, the latter commonly containing arc-
parallel transforms, are appropriate for porphyry
copper emplacement, whereas arcs subject to exten-
sional stress act as sites for Kuroko deposit genesis.
In terms of gold metallogeny, their scheme seems to
translate into intrusion-related and acid-sulfate
type gold deposits under compressional (and neu-
tral) stress regimes, and adularia-sericite type and
Kuroko gold deposits under extensional stress
regimes. It is hypothesized that adularia-sericite
type gold deposits associated with alkalic magma-
tism are hallmarks of more advanced crustal
Geotectonic setting of western Pacific gold deposits 675
extension in back-arc rather than intra-arc settings.
However, the precise state of stress that existed in
overriding plates during gold mineralization in
many western Pacific island arcs may only be sur-
mised.
Using the youthful western Pacific region as an
example, it is evident that exact relationships
between geotectonic setting and gold mineralization
are not always easy to decipher and, furthermore,
that such relationships are varied and complex in
detail. A measure of circumspection should there-
fore be applied to geotectonic reconstructions of
older gold provinces, in particular those of Precam-
brian age such as the Archean greenstone belts.
SELECTED COMPARISONS
The Western Pacific Region Compared with
Mesozoic Western North America
Several investigators, including Karig (1983)
and Saleeby (1983), concluded that eventual oro-
genic collapse of the complex fringing arc system
along the western Pacific rim (Figures 1, 2, and 3)
will result in a continental margin with many simi-
larities to Mesozoic western North America. There,
a series of exotic tectonostratigraphic terranes,
including oceanic arcs, were accreted to the North
American continental margin during mid-Jurassic
through Cretaceous time (e.g., Coney and others,
1980; Monger and others, 1982; Saleeby, 1983).
The Mesozoic collage of western North America
has undergone relatively deep erosion, partly in re-
sponse to collision-induced uplift. Shallow epither-
mal and subvolcanic environments have therefore
been removed from most of the accreted terranes. A
comparison of the region's gold metallogeny with
that of the more shallowly eroded Miocene to Holo-
cene western Pacific region is therefore not very
meaningful. However, it is worth pointing out that
a number of intrusion-related gold deposits (e.g.,
Hedley skarn, British Columbia) as well as a few
epithermal gold deposits (e.g., Toodoggone district,
British Columbia) were accreted to North America
in the Mesozoic as components of exotic arc terranes
(McMillan and others, 1987).
Most western North American gold deposits,
however, were generated along major terrane-
bounding faults during or following Mesozoic sutur-
ing events or in widespread post-collisional arcs
during the Tertiary. The former style of gold min-
eralization - the mesothermal or Mother Lode type
- was generated at depths of 5-10 km (e.g., Weir
and Kerrick, 1987), and is represented by the Moth-
er Lode belt in California, the Bralorne camp in
British Columbia, and the Juneau belt in Alaska. If
Mother Lode-type gold deposits were emplaced
along late Tertiary sutures in the western Pacific
region, they have yet to be unroofed.
Gold and Silicic Magmatism in Extensional
Settings
Many gold deposits along the eastern Pacific
rim were generated in neutral to compressional arcs
of dominantly andesitic-dacitic composition. Ex-
amples are provided by the Miocene gold-silver
province of the Chilean Andes and gold minera-
lization in Panama and Costa Rica. However, in
much of the western USA Cordillera, mid- to late
Cenozoic gold metallogeny accompanied rhyodacitic
to rhyolitic magmatic rocks emplaced in a complex
array of extensional settings linked to waning sub-
duction and the eventual imposition of a transform
regime.
As reviewed by Lipman (1980) and others,
silicic magmatism first flared up in the western
USA in the late Oligocene, when intra-arc extension
was promoted by decrease of subduction rate and
concomitant steepening and westward migration of
the underlying subduction zone. Epithermal gold
deposits formed during this stage are not only of the
adularia-sericite type (e.g., Round Mountain, Com-
stock, and Rawhide in Nevada) but also of the acid-
sulfate type (e.g., Goldfield in Nevada and Summit-
ville in Colorado).
A further change in the volcanotectonic asso-
ciation of parts of the western USA began in the
early to mid-Miocene when dominantly bimodal,
basalt-rhyolite volcanism was initiated in response
to transform motion on the San Andreas fault
system and increased extension farther east. This
bimodal volcanism, specifically the rhyolitic com-
ponent, was linked to epithermal gold minera-
lization of adularia-sericite type at Sleeper and Hog
Ranch in Nevada, Quartz Mountain in Oregon,
Delamar in Idaho, McLaughlin in California, and
Cinola in British Columbia. Indeed, Noble and
others (1988) proposed a fundamental genetic con-
nection between the basaltic component and gold
mineralization. In contrast to the western Pacific
region, the basalt-dominated bimodal magmatism
676
R. H. Sillitoe
in the western USA was not accompanied by the
eruption of andesitic-dacitic magmas. However, a
subduction-modified mantle [ct., Johnson's (1987)
hypothesis] was the source of at least some of the
earliest basaltic magmas (Ormerod and others,
1988).
Gold and Alkalic Magmatism in Extensional
Settings
In the western USA, a number of gold deposits
were emplaced along the eastern edge of the Cordi-
llera in association with magmatic rocks of alkalic
composition (Mutschler and others, 1985). The
Paleocene-Eocene (e.g., Zortman-Landusky in Mon-
tana and Gilt Edge in South Dakota) and Oligocene
(e.g., Cripple Creek in Colorado) ages of many of
these gold-bearing magmatic centers accord with an
origin during back-arc extension, which in Colorado
coincided approximately with early stages of exten-
sion along the Rio Grande rift zone. The back-arc
position of these western North American examples
is shared with the gold-bearing alkalic centers in
the western Pacific region.
Detachment-Related Gold Deposits
The Wapolu gold deposit in the D'Entrecas-
teaux Islands (Figures 2 and 6) is located on and
immediately beneath a detachment fault that con-
stitutes the surface of an actively rising metamor-
phic core complex (Billington, 1987). The deposit is
of the adularia-sericite type and probably related to
rhyolitic magmatism nearby on the same detach-
ment structure.
Davies and Warren (1988) pointed out that
metamorphic core complexes of the D'Entrecasteaux
Islands are closely comparable to mid-Tertiary
examples documented from the western USA, al-
though the causes of regional extension appear to be
different: a spreading center in the D'Entrecas-
teaux and waning subduction in the southwestern
USA. Notwithstanding this difference, however, it
is tempting to conclude that detachment-related
gold deposits in the southwestern USA, such as
Picacho and possibly Mesquite in southeastern
California, are, like Wapolu, conventional albeit
bulk-tonnage epithermal deposits (cr, Drobeck and
others, 1986). If so, then an association with de-
tachment faults does not signify a special mode of
genesis for gold mineralization. It implies simply
that detachments provided suitable structural loci
for epithermal gold deposition.
REFERENCES
ALLEN, C. R., 1962, Circum-Pacific faulting in the
Philippine-Taiwan region: Journal of Geophysical Re-
search, v. 67, p. 4795-4812.
BILLINGTON, W. G., 1987, The Wapolu gold prospect,
D'Entrecasteaux Islands, Papua New Guinea; pp.51-55 in
Proceedings of the Pacific Rim Congress 87 (Gold
Coast, Queensland): Australasian Institute of Mining
and Metallurgy, Parkville, Victoria, Australia, 949 p.
CARDWELL, R. K., B. L. ISACKS, and D. E. KARIG, 1980,
The spatial distribution of earthquakes, focal mechanism
solutions, and subducted lithosphere in the Philippine and
northeastern Indonesian islands; pp. 1-35 in D. E. Hayes
(ed.), The Tectonic and Geologic Evolution of South-
east Asian Seas and Islands: American Geophysical
Union, Monograph 23, 326 p.
CATHLES, L. M., A. L. GUBER, T. C. LENAGH, and F. O.
DUNDAS, 1983, Kuroko-type massive sulfide deposits of
Japan: Products of an aborted island-arc rift; pp. 96-114 in
H. Ohmoto and B. J. Skinner (eds.), The Kuroko and
Related Volcanic Massive SUlfide Deposits: Economic
Geology, Monograph 5, 604 p.
CmCUM-PACIFIC COUNCIL FOR ENERGY AND MINERAL
RESOURCES, 1984, Plate-Tectonic Map of the Circum-
Pacific Region: Pacific Basin Sheet, 1:17,000,000-
Scale: American Association of Petroleum Geologists,
Tulsa, Oklahoma, USA.
COLE, J. W., 1984, Taupo-Rotorua depression: An ensialic
marginal basin of North Island, New Zealand; pp. 109-120
in B. P. Kokelaar and M. F. Howells (eds.), Marginal
Basin Geology: Geological Society of London, Special
Publication 16, 322 p.
CONEY, P. J., D. L. JONES, and J. W. H. MONGER, 1980,
Cordilleran suspect terranes: Nature, v. 288, p. 329-333.
DAVIES, H. L., and R. G. WARREN, 1988, Origin of
eclogite-bearing, domed, layered metamorphic complexes
("core complexes") in the D'Entrecasteaux Islands, Papua
New Guinea: Tectonics, v. 7, p. 1-21.
__ , P. A. SYMONDS, and 1. D. RIPPER, 1984, Structure
and evolution of the southern Solomon Sea region: BMR
Journal of Australian Geology and Geophysics, v. 9, p. 49-
68.
DAVIES, R. M., and G. H. BALLANTYNE, 1987, Geology of
the Ladolam gold deposit, Lihir Island, Papua New
Guinea; pp. 943-949 in Proceedings of the Pacific Rim
Congress 87 (Gold Coast, Queensland): Australasian
Geotectonic setting of western Pacific gold deposits 677
Institute of Mining and Metallurgy, Parkville, Victoria,
Australia, 949 p.
DROBECK, P. A., F. L. HILLE MEYER, E. G. FROST, and G.
S. LIEBLER, 1986, The Picacho mine: A gold mineralized
detachment in southeastern California: Arizona Geologi-
cal Society Digest, v. 16, p. 187-221.
DUNKLEY, P. N., 1983, Volcanism and the evolution ofthe
ensimatic Solomon Islands arc; pp. 225-241 in D. Shimo-
zuru and 1. Yokoyama (eds.), Arc Volcanism: Physics
and Tectonics: Terra Scientific Publishing Company,
Tokyo, Japan, 263 p.
ERNST, W. G., and B. M. JAHN, 1987, Crustal accretion
and metamorphism in Taiwan, a post-Palaeozoic mobile
belt: Philosophical Transactions of the Royal Society of
London, v. A321, p. 129-161.
FLEMING, A. W., G. A. HANDLEY, K. L. WILLIAMS, A. L.
HILLS, and G. J. CORBETT, 1986, The Porgera gold deposit,
Papua New Guinea: Economic Geology, v. 81, p. 660-680.
GILL, J. B., A. L. STORK, and P. M. WHELAN, 1984,
Volcanism accompanying backarc basin development in
the southwest Pacific: Tectonophysics, v. 102, p. 207-224.
HAMILTON, W., 1979, Tectonics of the Indonesian
Region: U.S. Geological Survey, Professional Paper 1078,
345p.
HILL, K. C., and K. A. HEGARTY, 1987, New tectonic
framework for PNG and the Caroline plate: Implications
for cessation of spreading in back-arc basins; pp. 179-182 in
Proceedings of the Pacific Rim Congress 87 (Gold
Coast, Queensland): Australasian Institute of Mining
and Metallurgy, Parkville, Victoria, Australia, 949 p.
IZAWA, E., and Y. URASHIMA, 1987, Geologic and tectonic
setting of the epithermal gold and geothermal areas in
Kyushu; pp. 1-12 in Y. Urashima (ed.), Gold Deposits and
Geothermal Fields in Kyushu: Society of Mining Geo-
logists of Japan, Guidebook 2, 59 p.
JOHNSON, R W., 1987, Delayed partial melting of sub-
duction-modified magma sources in western Melanesia:
New results from the late Cainozoic; pp. 211-214 in
Proceedings of the Pacific Rim Congress 87 (Gold
Coast, Queensland): Australasian Institute of Mining
and Metallurgy, Parkville, Victoria, Australia, 949 p.
__ , A. L. JAQUES, C. H. LANGMUIR, M. R PERFIT, H.
STAUDIGEL, P. N. DUNKLEY, B. W. CHAPPELL, S. R
TAYLOR, and M. BAEKISAPA, 1987, Ridge subduction and
forearc volcanism: Petrology and geochemistry of rocks
dredged from the western Solomon arc and Woodlark
Basin; pp. 155-226 in B. Taylor and N. F. Exon (eds.),
Marine Geology, Geophysics, and Geochemistry of the
Woodlark Basin-Solomon Islands: Circum-Pacific
Council for Energy and Mineral Resources, Earth Science
Series, v. 7,365 p.
KARIG, D. E., 1983, Accreted terranes in the northern part
of the Philippine archipelago: Tectonics, v. 2, p. 211-236.
KNITTEL, U., 1983, Age of the Cordon syenite complex and
its implication on the mid-Tertiary history of north Luzon:
Philippine Geologist, v. 37, no. 2, p. 22-31.
KROENKE, L. W., 1984, Cenozoic Tectonic Development
of the Southwest Pacific: United Nations ESCAP,
CCOPISOPAC Technical Bulletin 6, 122 p.
LETOUZEY, J., and M. KIMURA, 1986, The Okinawa
Trough: Genesis of a back-arc basin developing along a
continental margin: Tectonophysics, v. 125, p. 209-230.
LIPMAN, P. W., 1980, Cenozoic volcanism in the western
United States: Implications for continental tectonics; pp.
161-174 in Continental Tectonics, U.S.: National Re-
search Council, Geophysics Study Committee, Wash-
ington, DC, USA, 197 p.
McMILLAN, W. J., A. PANTELEYEV, and T. HOY, 1987,
Mineral deposits in British Columbia: A review of their
tectonic settings; pp. 1-18 in L. L. Elliot and B. W. Smee
(eds.), GEOEXPOI86: Exploration in the North Ameri-
can Cordillera: Association of Exploration Geochemists,
Rexdale, Ontario, Canada, 220 p.
MASON, D. R, and J. E. HEASLIP, 1980, Tectonic setting
and origin of intrusive rocks and related porphyry copper
deposits in the western Highlands of Papua New Guinea:
Tectonophysics, v. 63, p. 125-137.
MONGER, J. W. H., R A. PRICE, and D. J. TEMPELMAN-
KLUIT, 1982, Tectonic accretion and the origin of two
major metamorphic and plutonic welts in the Canadian
Cordillera: Geology, v. 10, p. 70-75.
MUTSCHLER, F. E., M. E. GRIFFEN, D. S. STEVENS, and S.
S. SHANNON, Jr., 1985, Precious metal deposits related to
alkaline rocks in the North American Cordillera - An
interpretative review: Transactions of the Geological
Society of South Africa, v. 88, p. 355-377.
NOBLE, D. C., J. K. McCORMACK, E. H. McKEE, M. L.
SILBERMAN, and A. B. WALLACE, 1988, Time of min-
eralization in the evolution of the McDermitt caldera
complex, Nevada-Oregon, and the relation of middle Mio-
cene mineralization in the northern Great Basin to coeval
regional basaltic magmatic activity: Economic Geology, v.
83, p. 859-863.
ORMEROD, D. S., C. J. HAWKESWORTH, N. W. ROGERS,
W. P. LEEMAN, and M. A. MENZIES,1988, Tectonic and
magmatic transitions in the western Great Basin, USA:
Nature,v. 333,p. 349-350.
PHILIPPINE BUREAU OF MINES & GEO-SCIENCES, 1982,
Geology and Mineral Resources of the Philippines,
Volume 1: Geology: Manila, Philippines, 406 p.
678 R. H. Sillitoe
RAMSAY, W. R. H., A. J. CRAWFORD, and J. D. FODEN,
1984, Field setting, mineralogy, chemistry, and genesis of
arc picrites, New Georgia, Solomon Islands: Contributions
to Mineralogy and Petrology, v. 88, p. 386-402.
SALEEBY, J. B., 1983, Accretionary tectonics of the North
American Cordillera: Annual Review of Earth & Planetary
Sciences, v.n, p. 45-73.
SILLITOE, R. H., 1982, Extensional habitats of rhyolite-
hosted massive sulfide deposits: Geology, v. 10, p. 403-407.
__ , 1987, Copper, gold and subduction: A trans-Pacific
perspective; pp. 399-403 in Proceedings of the Pacific
Rim Congress 87 (Gold Coast, Queensland): Austral-
asian Institute of Mining and Metallurgy, Parkville,
Victoria, Australia, 949 p.
__ , 1989, Gold deposits in western Pacific island arcs:
The magmatic connection; pp. 274-291 in R. R. Keays, W.
R. H. Ramsay, and D. 1. Groves (eds.), The Geology of
Gold Deposits: The Perspective in 1988: Economic
Geology Monograph 6, 667 p.
__ , and 1. M. GAPPE, Jr., 1984, Philippine Porphyry
Copper Deposits: Geologic Setting and Characteris-
tics: United Nations ESCAP, CCOP Technical Publication
14,89p.
SKINNER, D. N. B., 1986, Neogene volcanism of the
Hauraki volcanic region; pp. 21-47 in 1. E. M. Smith (ed.),
Late Cenozoic Volcanism in New Zealand: Royal
Society of New Zealand, Bulletin 23, 365 p.
STEPHAN, J. F., R. BLANCHET, C. RANGIN, B. PELLETIER,
J. LETOUZEY, and C. MULLER, 1986, Geodynamic evo-
lution of the Taiwan-Luzon-Mindoro belt since the late
Eocene: Tectonophysics, v. 125, p. 245-268.
URABE, T., 1987, Kuroko deposit modeling based on mag-
matic hydrothermal theory: Mining Geology, v. 37, p. 159-
176.
UYEDA, S., and C. NISHIWAKI, 1980, Stress field, metal-
logenesis and mode of subduction; pp. 323-339 in D. W.
Strangeway (ed.), The Continental Crust and Its Min-
eral Deposits: Geological Association of Canada, Special
Paper 20,804 p.
WALLACE, D. A., R. W. JOHNSON, B. W. CHAPPELL, R. J.
ARCULUS, M. R. PERFIT, and 1. H. CRICK, 1983, Caino-
zoic Volcanism of the Tabar, Lihir, Tanga, and Feni
Islands, Papua New Guinea: Geology, Whole-Rock
Analyses, and Rock-Forming Mineral Compositions:
Australian Bureau of Mineral Resources, Report 243, 142 p.
WARNAARS, F. W., 1987, Magma chemistry and minera-
lization of Papua New Guinea in a framework of Neogene
plate tectonic reconstruction [abstract): Geological Society
of America, Abstracts with Programs, v. 19, p. 882.
WEIR, R. H., Jr., and D. M. KERRICK, 1987, Mineralogic,
fluid inclusion, and stable isotope studies of several gold
mines in the Mother Lode, Tuolumne and Mariposa
Counties, California: Economic Geology, v. 82, p. 328-344.
WHELAN, P. M., J. B. GILL, E. KOLLMAN, R. A. DUNCAN,
and R. E. DRAKE, 1985, Radiometric dating of magmatic
stages in Fiji; pp. 415-435 in D. W. Scholl and T. L. Vallier
(eds.), Geology and Offshore Resources of Pacific
Island Arcs - Tonga Region: Circum-Pacific Council
for Energy and Mineral Resources, Earth Science Series 2,
488p.
WOLFE, J. A., 1983, Origin of the Philippines by accumu-
lation of allochthons: Philippine Geologist, v. 37, no. 3, p.
17-33.
INTERNA TION AL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Felsic magmatism and hydrothermal gold deposits:
A tectonic perspective
ROBERT MASON
Department of Geological Sciences, Queen's University, Kingston, Ontario, K7L 3N6, Canada
(received December 21, 1988; revision accepted February 20, 1990)
ABSTRACT
Hydrothermal gold deposits are considered to have formed as a result of exsolution of volatile phases from
I-type magnetite series felsic magmas generated in subduction-related plate margin situations as a result of
hornblende melting. Hot inflated cracked crust characteristic of plate margins overriding active subduction
zones in island arcs and continental lip arcs of the present circum-Pacific rim has provided ideal tectonic
settings for emplacement of epizonal felsic magmas and associated hydrothermal systems. Felsic plutons are
emplaced more easily from magma transferred through hot, weakened crust via major fracture zones and sub-
volcanic conduits. Less commonly, they may create their own conduits where volatile constituents in hotter
magma permit focused and rapid ascent through the crust. Vapour phase separation in apical regions of felsic
plutons, accompanied by boiling and decompression, has resulted in extensive damage to environments of
such plutons. Intense fracturing at scales from several kilometres down to microscopic are typical around
hydrothermal gold systems, as is development of phreatomagmatic and phreatic breccia dikes and pipes.
These features are also typical of porphyry Cu-Mo systems. Not surprisingly, many major gold deposits are
associated with porphyry-type hydrothermal systems, and most epithermal gold deposits can be considered as
offspring from such systems. This connection can also be drawn for most Archean lode gold deposits. Archean
hydrothermal gold deposits have much in common with their Cenozoic counterparts, especially their
association with felsic magmatism, and they are regarded here as metamorphosed equivalents of the Cenozoic
deposits. The concentration of productive gold systems in the Cenozoic is surpassed only by that of the late
Archean. It is suggested that this reflects similarities in the tectonic settings of Cenozoic magmatic arc
complexes and Archean greenstone belts.
INTRODUCTION
Cenozoic and Mesozoic hydrothermal gold and
copper deposits are associated with a wide variety of
porphyry-, epithermal-, and volcanogenic-type hy-
drothermal systems. These systems can usually be
overtly or covertly linked to felsic magmatism and
particularly to the generation of magnetite series/I-
type intrusive phases associated with arc magma-
tism related to subduction of hydrated oceanic crust
(Ishihara, 1981; Sillitoe, 1987). In cases where the
felsic magmatic connection is not obvious, the char-
acteristics of ore, hydrothermal alteration, and
structural damage to the surrounding environment
are identical, or very similar to, those associated
with deposits of clear felsic magmatic association.
There is widespread acceptance of the elegant
magmatic hydrothermal models for the genesis of
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (BuUe, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 679-687. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
679
680
R. Mason
porphyry copper deposits (Burnham, 1979; Burn-
ham and Ohmoto, 1980; Whitney, 1975; Gustafsen
and Hunt, 1975; Sillitoe, 1973) and for porphyry
molybdenum deposits (White and others, 1981; Car-
ten and others, 1988), in spite of lingering reserva-
tions summarized by Titley and Beane (1981).
These models emphasize direct genetic links be-
tween (a) separation of hydrous magmatic solutions
from felsic intrusive bodies; (b) development of
fractures in and around the felsic intrusions as a
result of emplacement of these intrusions, and
energy released by processes related to resurgent
boiling of hydrothermal fluids in a near surface
environment; and (c) fracture controlled hydro-
thermal alteration and mineralization developed as
a consequence of (b). Magmatic and magmatic-
hydrothermal sources and processes provide the
ingredients and driving mechanisms for ore concen-
tration and deposition, without requiring abnormal
enrichment of starting materials, and with modi-
fying influences provided by interaction with
meteoric waters (Burnham, 1981). I see similar
causes and effects in hydrothermal gold deposits of
all ages.
Apart from large zones of hydrothermal altera-
tion (potassic, phyllic, argillic, and propylitic) which
define mineralized areas, the most striking char-
acteristic of hydrothermal gold deposits is the extent
of both mesoscopic and microscopic fracture damage
to wall rocks surrounding them. Indeed, the extent
and intensity of this fracturing often appears to be
directly related to the size of an individual hydro-
thermal system and its ore deposits. If these
characteristics are combined with the common
association of epizonal felsic intrusions, breccia
pipes and breccia dikes, we clearly have a variation
on the porphyry copper theme. This should en-
courage us to examine hydrothermal gold deposits
in terms of magmatic hydrothermal models for
porphyry deposits in conjunction with, or rather
than, current geothermal models.
The association of many major gold deposits in
the western American Cordillera, and the western
Pacific region, with porphyry systems centered on
felsic intrusive centers (Sillitoe, 1979, 1987, 1988,
this volume; McMillan and Pantelayev, this vol-
ume), necessitates a re-evaluation of current geo-
thermal models for genesis of these deposits. Such
models suggest that epithermal gold systems could
evolve in any tectonic setting given heat, permeable
rocks, water, and time. However, the typical spec-
trum of gold deposits of mesothermal to epithermal
affiliation, which is associated with porphyry Cu-Mo
stockwork mineralization, and related felsic intru-
sions, suggests that similar epithermal gold deposits
which do not display this association at the present
level of erosion, and/or the present level of geologi-
cal knowledge of the environment of a particular
deposit, may well be manifestations of concealed
porphyry systems.
More than fifty years ago, Emmons (1937) used
what he saw as the very common and worldwide
association of hydrothermal gold deposits with a
variety of felsic intrusions as a basis for proposing
that there was a genetic connection between these,
and that emplacement of these intrusions and
related magmatic hydrothermal fluids was, in many
cases, responsible for creation of fractures which
were then mineralized. Emmons noted the ubiqui-
tous association of sericitic alteration, quartz, and
pyrite with hydrothermal gold deposits. He drew
attention to the fact that Precambrian (what later
turned out to be mainly Archean in age) lode gold
deposits, and particularly the major deposits, were
associated with felsic intrusive centers. I make no
apology for returning to the theme composed by
Emmons and now so well supported by a much
larger body of field evidence and the models cited
previously.
In this contribution I briefly review the petro-
genetic and tectonic controls on intrusive felsic
magmatism and associated hydrothermal activity
in terms of hydrothermal gold deposits, and I discuss
the implications.
TECTONIC ENVIRONMENT OF FELSIC
MAGMATISM
Felsic magmatism is predominantly a product
of subcrustal and crustal melting associated with
magmatic arcs at convergent plate margins (Pit-
cher, 1982). The generation of granitoid melts
requires several weight percent water, and the
source and amount of water available will affect
(1) melting temperatures, (2) subsequent magma
compositions and characteristics, and (3) subse-
quent intrusion crystallization histories and depths
of emplacement (Pitcher, 1979; Hyndman, 1981;
Whitney, 1988). All of these factors influence the
potential for any given felsic magma to develop a
productive ore-forming hydrothermal system (Burn-
ham, 1967, 1979, 1981; Burnham and Ohmoto,
1980; Whitney, 1975, 1977, 1988). The only tectonic
Felsic magmatism and hydrothermal gold deposits: A tectonic perspective 681
setting which allows the introduction of large
amounts of water (as hydrated minerals) into the
mantle is a convergent margin where hydrated
oceanic crust is being subducted. Dehydration of
such crust in a descending slab would release a
hydrous fluid containing incompatible and radio-
genic elements + silica (the IRS fluid proposed by
Gill, 1981). This fluid is thought to trigger partial
melting in an actively convecting, warm mantle
wedge, and prolonged activity of this nature results
in arc magmatism, crustal melting, and magmatic
inflation of the crust in overriding plate margins of
magmatic arcs (Gill, 1981; Wyllie, 1984; Pitcher,
1979; Whitney, 1988).
The nature and thickness of the crust in the
overriding plate appears to have a major influence
on felsic magma compositions (Hamilton, 1988),
although many intrusions emplaced in magmatic
arcs seem to have been derived from sub crustal
sources with or without subsequent crustal contami-
nation and are predominantly I-type granitoid
intrusions (Chappell and White, 1974). Sillitoe
(1987) showed that there is no relationship between
crustal thickness and hydrothermal copper and gold
deposits in magmatic arcs of the circum-Pacific rim.
Furthermore, the occurence of gold deposits associa-
ted with a wide spectrum of felsic magma com-
positions (Sillitoe, 1987, this volume) suggests that
specific magma composition is not a critical factor in
the genesis of such deposits. This is also true for
hydrothermal copper deposits (Gustafsen, 1979;
Sillitoe, 1987). The evidence from Archean gold de-
posits associated with a complete spectrum of felsic
magma types supports the conclusion that magma
composition is not a critical factor in evolution of
hydrothermal gold deposits.
The source of metals, sulphur, and chlorine as
ingredients of porphyry-type hydrothermal systems
is most likely the IRS fluid from subducted slab
(Sillitoe, 1987). However, Burnham (1979) pointed
out that at least the metals and sulphur do not need
any pre-concentration such as might be expected in
hydrated oceanic crust, and could just as easily be
derived from melted mantle material. Thus, al-
though one could postulate crustal sources in the
overriding plate for metal and sulphur, the evidence
presented by Sillitoe (1987), and Burnham (1979)
for sub-crustal sources, is compelling and suggests
that crustal sources for gold, copper, and sulphur are
unnecessary and unlikely in explaining concentra-
tions developed in hydrothermal copper and gold
deposits.
Figure 1 summarizes attempts to classify the
granitoid suite in terms of source rocks, hydrated
mineral phases involved in melting, oxidation state
of magma, and tectonic setting. There is a clear-cut
association of chalcophile hydrothermal ores of
copper and gold with compositionally expanded 1-
type, magnetite series, meta-aluminous-peralkaline
felsic magmas generated from hornblende melting
..- SUBDUCTION-RELATED MAGMATIC ARCS .......
CONTINENTAL COLLISION .......
.......... __ I-TYPE ........... __ .......... S-TYPE .......... -t. __
..-- MAGNETITE SERIES- - .......
..--- - -ILMENITE SERIES- -- - .......
..- HORNBLENDE ....... ..- BIOTITE - .... .......... - MUSCOViTE .......
>900
o
C KINa <1 750-850
o
C K/Na .... 1 650-750
o
C KINa >1
..-- - META ALUMINOUS - - ....... ..-- - PERALUMINOUS - - - .......
HYDROTHERMAL


GOLD SYSTEMS
..- - - - -Cu- Mo- - - --.... .......... -- ---Sn-W- - - - .......
Figure 1. Hydrothermal gold systems in the context of tectonic setting, characteristics of felsic magma generation and
emplacement of the granitoid suite From Pitcher (1982), Chappell and White (1974), Ishihara (1981), Whitney (1988), and
Sillitoe (1979, 1987).
682
R. Mason
in island arcs and continental-lip arcs of the present
circum-Pacific rim. This association was linked
specifically to subduction of hydrated oceanic litho-
sphere at compressional convergent margins by
Sillitoe (1987). Ishihara (1981) and Burnham (1981)
discussed the potentially enhanced metal- and
sulphur-carrying capacity of such magmas and asso-
ciated hydrothermal phases. On the other hand,
there is a notable absence of significant copper and
gold deposits of hydrothermal affinity where
muscovite melting and ilmenite series peralumin-
ous magmas have evolved in continental collisional
tectonic environments.
Hornblende melting occurs at temperatures sig-
nificantly above those for muscovite and biotite
(Figure 1) and, as Burnham (1979) and Hyndman
(1981) pointed out, this gives granitoid magmas
associated with hornblende melting the potential for
routine epizonal crustal emplacement (Figure 2),
whereas those associated with muscovite melting
have limited potential for ascent into upper parts of
the crust. In any case, magmas usually require
assistance from structural weaknesses in the crust
to enable them to reach the near-surface environ-
ment. The extent to which such weaknesses are
inherent, or induced as a consequence of magma
generation in the upper mantle and lower crust, is
debateable, but Shaw (1980) suggested that "frac-
ture mechanisms playa globally intrinsic role in the
transportation of magma from the melting sources
to the surface." Intrusion of felsic magmas, in
particular, would appear to demand some assistance
from pre-existing lineaments and fault zones in
most cases (Pitcher, 1979, 1982; Hutton, 1988), and
Pitcher also stressed the importance of prolonged
magmatic activity and hot crust along such fracture
zones in facilitating epizonal emplacement of grani-
toid magmas. Development of hydrous phases
during evolution and emplacement of granitoid
magmas in the crust, provides additional magma
mobility as well as the potential, in special circum-
stances, for generating ore-forming hydrothermal
systems (Burnham, 1967, 1979; Whitney, 1975;
Pitcher, 1979). Figure 2 illustrates a variety of
fault-related situations for emplacement of grani-
toid magma after Hutton, (1988).
The hot, inflated and cracked, transitional-
continental crust, typical of overriding plate mar-
gins above subducted oceanic crust in island arcs
and continental-lip arcs of the present circum-
Pacific rim (Hamilton, 1988; Pitcher, 1982), has
provided ideal environments for emplacement of
felsic magmas and related hydrothermal ore
systems into the upper crust (Sillitoe, 1981, 1987;
Pitcher, 1982). However, not all subduction-related
convergent margins hold potential for significant
mineralization, and even ones which do have poten-
tial are characterized by periods of magmatic
quiescence punctuated by discrete episodes of mag-
matism which themselves may be productive or
barren with respect to mineralization (Clark and
others, 1976; Sillitoe, 1987).
Uyeda and Nishiwaki (1980) proposed a rela-
tionship between compressional convergent mar-
gins (in which the overriding plate advances faster
than the retreat of the subduction hinge) and
development of porphyry-type magmatic hydrother-
mal systems. Such margins could be transpres-
sional (resulting from oblique convergence) or they
could result from orthogonal convergence. They are
associated with shallow to moderately steep sub-
duction, and major episodes of magmatism and
mineralization appear to be associated with periods
of more energetic convergence, magmatic inflation,
and uplift (Titley, 1975; Clark and others, 1976;
Pitcher, 1979; Sillitoe, 1987). Episodic changes in
convergence rates, slab segmentation, and varia-
tions in the angle of convergence can each con-
tribute to the complexity of convergent margins and
influence the structural style of the overriding
plate.
Hansen (this volume) points out that oblique
convergence and margin-parallel translation are
especially important where thermally or mechani-
cally weakened overriding plates advance over
shallow subduction zones - a condition that ap-
pears to have been characteristic of Mesozoic
convergent plate margins. Irregular geometry of
plates and plate margins would itself support
Hansen's suggestion that oblique convergence has
probably been more the norm throughout earth
history than orthogonal convergence. The potential
for trans-crustal faulting and controlled access for
magmas and hydrothermal systems would also be
enhanced in transpressional margins.
Sillitoe (1987) pointed out that episodes of
magmatic activity and associated mineralization in
the Andes are related to changes in the mode and
energetics of subduction, whereas such episodes in
the southwestern Pacific are related to fundamental
relocation of subduction zones. He also commented
on the possibilities of allochthonous mineralization
in accreted arc terranes, as opposed to the autoch-
thonous situation exemplified by the Andes. Of
Felsic magmatism and hydrothermal gold deposits: A tectonic perspective 683
A B c D E
." :::::::: :::::::::At.::
... :....:..:.. '-...: ... : ... :' ':: ..
5 Z BRITTLE
++++++
10 _ -y-:__ +:!!! __ ...;)..
I ++

11 \ Iii
20
30km
r
.L, T
i .2
.. ..
",&:I
M B H
III
V II
---Jf+l-
++ + +
+
11.+
Figure 2. Modes of emplacement of potentially mineralized (Au) granitoid intrusions based on Hutton (1988).
SZ, seismogenic zone; M,B,H, depth limits for emplacement of granitoids generated by muscovite, biotite, and hornblende
melting, respectively (after Hyndman,1981). The stippled area (Au) indicates the optimum zone for development and
deposition of hydrothermal gold mineralization and for occurrence of stockwork fracturing and hydrothermal breccias
(Burnham and Ohmoto, 1980; Burnham, 1985). A) Plutons emplaced diapirically without assistance from fault zone, usually
in a linear zone of arc magmatism, often exhibiting late ballooning. Intrusion and hydrothermal system create local fractures
centered on intrusion. Tectonic influence much reduced in these systems. B) Cauldron-caldera structures in major vertical
trans-crustal extensional fault zones. High level magma emplacement and hydrothermal activity associated with caldera
development. Many opportunities for development of ore structures related to fracturing and hydrothermal breccias. Tectonic
influences subordinate to intrusion-caldera emplacement-related structures. C) Trans-crustal fracture zone (transcurrent
fault, reactivated terrane boundary, etc.): ascent of magma and hydrothermal fluids encouraged and controlled by dilational
jogs, pull aparts, etc.; tendency for elongate plutons. D) Intra-crustal strike-slip fault zone: diapiric ascent of magma to mid-
crustal encounter with fault zone; thence controlled by dilational jogs, etc., tendency for elongate plutons; late ballooning
shown in figure. C and D involve situations in which there will be strong tectonic control on ore structures, modified to a
greater or lesser extent by hydrothermal activity. E) Listric normal fault zone (extensional regime of Basin and Range type):
diapiric ascent of magma intercepts shear zone/listric fault and may be emplaced as sheet-like bodies or in near surface
asymmetric cauldrons and calderas. Numerous possibilities for development of ore structures and for strong tectonic control
on these structures. NB: Ore-forming hydrothermal systems develop in top 5 km of crust. Only magmas derived from
hornblende melting can routinely enter the uppermost crust. Penetrative fabrics develop during tectonism below brittle
(seismogenic) zone (beyond threshold of quartz plasticity in the transition zone between brittle and plastic crust). Very local
fabrics may result from late ballooning of diapiric intrusions. Regional penetrative fabrics usually result from crustal
shortening.
course, the ultimate complexity arises where
continued subduction has occurred beneath accreted
arc terranes and terrane boundary faults become
trans-crustal weaknesses to be exploited by magma-
tism and hydrothermal systems. The contribution
by McMillan and Pantaleyev (this volume) em-
phasises some of this complexity.
It would appear that there are links between
more energetic convergence (most frequently ob-
lique), convective disturbance of the mantle wedge,
an increased heat budget in both wedge and the
overriding crust above the wedge, crustal inflation
and uplift, more prolific intermediate to felsic arc
magmatism, more trans-crustal fracturing and
cracking of the upper crust, and hydrothermal ore
systems containing copper and/or gold. This has
special implications for Archean metallogeny,
assuming the increased crustal heat budget and
plate motion energetics that appear to have char-
acterized the Archean.
684
R. Mason
STRUCTURAL ENVIRONMENT OF MAGMATIC
HYDROTHERMAL SYSTEMS
Two structural-magmatic associations seem to
recur in hydrothermal gold districts of all ages.
Firstly, there is a strong association between epi-
zonal felsic magmatism and major fault zones; se-
condly, there is a common association between large
magmatic-hydrothermal systems and regional- to
local-scale doming (Emmons, 1937; Wisser, 1960).
Both associations are further characterized by a
tendency for regional uplift prior to and during
magmatism and hydrothermal activity. The loca-
lization of magmatic and hydrothermal centers
associated with dilational jogs (Sibson, 1986) in
fault zones was a theme developed by Newhouse
(1942) and McKinstry (1948; 1955) and given
renewed impetus and a seismic perspective by Sib-
son (1986, this volume). This theme can be orches-
trated around both transpressional and orthogonal
compressional convergent situations. Whether dila-
tional jogs occur in strike-slip or dip-slip fault zones
is immaterial, as long as these structures have
connections with magma seeking access to the upper
crust.
Within this context, magmatic-hydrothermal
activity centres are characterized by intense frac-
turing, at microscopic to localized regional scales,
and development of pipe and dike-like hydrothermal
breccias centred on small 2 km diam.) epizonal
felsic intrusions. Fracture damage usually extends
in an aureole about the intrusion and affects a much
larger volume of rock than the intrusion itself.
Where faulting is a dominant control on both mag-
ma emplacement and evolution of related hydro-
thermal systems, the fracture damage tends to form
systematic, fault-related structural patterns and to
be concentrated in linear or elongated zones. Where
magma emplacement and hydrothermal activity
have created their own environment, without in-
fluence from fault structures, fracture damage tends
to form patterns related to emplacement of the
intrusion and forms an aureole around the intru-
sion. Hydrothermal alteration is controlled by the
fracturing, and both fracture density and intensity
of alteration tend to die out away from the intrusion.
There are striking similarities between structural
environments of hydrothermal gold deposits and
those described for porphyry-type hydrothermal sys-
tems (Rehrig and Heidrick, 1972; Titley, 1975;
Titley and Heidrick, 1978; Titley and Beane, 1981;
Titley and others, 1986).
Figure 3 illustrates the potential roles of
lithology, faulting, and magmatic hydrothermal
activity in creating the structural environment for
ore deposition. Whether or not faulting and fract-
uring was more or less related to, or influenced by,
tectonism or magmatic hydrothermal activity, the
development of fracture patterns and hydrothermal
alteration around or adjacent to epizonal intrusions
is itself substantial evidence of a genetic link.
In the Timmins district (Abitibi greenstone
belt) of Ontario, Canada, we showed (Mason and
others, 1988) that there is a close relationship
between felsic intrusive centers, fracturing, and
hydrothermal alteration patterns around these
centers and gold orebodies. I have observed similar
relationships in the Val d'Or, Malartic, Kirkland
Lake, Red Lake, and Yellow Knife districts, and
Kuhn (1986, 1988) documented the same rela-
tionships in the Hemlo complex. The intense and
widespread distribution of stockwork fracturing
surrounding gold mineralized felsic intrusions in
the Abitibi belt has not previously been recognised.
The stockwork fracturing, related hydrothermal
alteration, and gold-bearing veins and stockworks
have all been modified to some extent by subsequent
regional metamorphism and penetrative deforma-
tion.
Many epithermal gold deposits in the Great
Basin not only have the intrusive connection but
also exhibit fracture patterns, hydrothermal breccia
development, and an intensity of stockwork frac-
turing that cannot be explained simply by tec-
tonism. The most spectacular example of all is the
fracture-damaged environment of the Carlin Trend,
where local superimposition of dense stockwork
fracturing on already complexly fractured and im-
bricated rocks appears to have localized deposits in
one of the largest hydrothermal gold systems in the
world. On several recent visits to view the gold
deposits of the Great Basin, I have been struck by
the dual roles of faulting and intrusion in creating
mineralized environments and the much more
intensely fractured environments that enclose the
major gold deposits. This is often coupled with pri-
mary permeability in tuffs and secondary permeabi-
lity caused by hydrothermal dissolution of carbon-
ates and carbonate cements. Whilst the role offelsic
intrusions in evolution of gold deposits in the Great
Basin is in dispute (Percival and others, 1988), their
association with many deposits is striking, and their
relationship to specific episodes of magmatism in an
active magmatic arc is unlikely to be fortuitous. I
Felsic magmatism and hydrothermal gold deposits: A tectonic perspective 685
t
...
o
II:
...
Z
CONTROL ......
Figure 3. Schematic representation of magmatic versus
tectonic influences explaining common association of
hydrothermal gold deposits with centers of felsic
magmatism. Sector A covers those deposits in which ore
structures were generated by emplacement of intrusions
and related hydrothermal systems with very little tectonic
influence (Figure 2A,B). Sector D covers those deposits
associated with fault zones but without obvious magmatic
connections. Sectors Band C cover the spectrum of gold
deposits that show greater or lesser elements of magmatic
versus tectonic influence (Figure 2C,D,E). The wallrock-
hosted deposits are influenced by a whole range of primary
litho-structural features which hold potential for
controlling primary permeability. Secondary permeability
is related to hydrothermal fluid -wallrock interaction,
hydrothermal breccias, and stockwork fracturing centered
on small epizonal intrusions. Fault-hosted deposits (sector
D) exhibit all the characteristics of other hydrothermal
gold deposits (sectors A-C) apart from the lack of known
intrusive association.
would again point out that the disposition of stock-
work fracturing about felsic intrusions and minera-
lized centers suggests a magmatic hydrothermal
cause. This has implications for ore genesis and
mineral exploration.
Radial, concentric, and irregular (crackle brec-
cia) stockwork fracturing and hydrothermal breccia
pipes and dikes associated with emplacement of
magmatic hydrothermal systems (Koide and
Bhattacharji, 1975; Burnham, 1979; Knapp and
Norton, 1981; Carten and others, 1988) all form
important sites for emplacement of gold minera-
lization. The downwardly propagated concentric
fractures described by Carten and others (1988) at
the Henderson porphyry molybdenum deposit in
Colorado are remarkably similar to major vein
systems that project downward from intrusive
porphyry contacts at the McIntyre and Dome mines
at Timmins (Mason and Melnik, 1986; Mason and
others, 1988).
At Timmins we have also noted that litho-
structural anisotropies in layered volcanic and
sedimentary rocks around felsic intrusions have
partly controlled subsequent fracturing during
intrusion and hydrothermal activity. Major centers
of felsic intrusion and mineralization are located in
domal culminations. Furthermore, the widespread
association of hydrothermal breccia dikes with
porphyry intrusions at Timmins (Mason and others,
1988) provides evidence for porphyry-type hydro-
thermal systems similar to those of the more recent
past. Crustal shortening has subsequently modified
the geometry of existing structures and imposed
metamorphic recrystallization and penetrative
fabrics on the mineralized environment (Mason and
Melnik, 1986; Mason and others, 1988). Mason and
Helmstaedt (this volume) suggest that Archean lode
gold deposits are metamorphosed analogues of
Cenozoic hydrothermal gold deposits in the same
way that Archean volcanogenic massive sulphide
deposits are metamorphosed analogues of Phanero-
zoic counterparts.
CONCLUSIONS
Hydrothermal gold systems exhibit characteris-
tics, and occur in environments similar to those
which surround porphyry copper deposits. The mag-
matic hydrothermal model for I-type, magnetite
series magmas in compressional convergent mar-
gins, where hydrated oceanic crust is subducted
energetically and at a shallow angle beneath an
overriding plate of continental or transitional crust,
appears to best explain the characteristic features
and associations observed for these deposits world-
wide, and of whatever age. Variations on the theme
are created by variations in the local environment of
felsic intrusive emplacement, and particularly by
the extent to which intrusive processes, versus
faulting, control emplacement of intrusions and
related hydrothermal systems.
686
R. Mason
REFERENCES
BURNHAM, C. W., 1967, Hydrothermal fluids at the
magmatic stage; pp. 34-76 in H. L. Barnes (ed.), Geo-
chemistry of Hydrothermal Ore Deposits: Holt, Rine-
hart and Winston, New York, New York, USA, 670 p.
__ ,1979, Magmas and hydrothermal fluids; pp. 71-136
in H. L. Barnes (ed.), Geochemistry of Hydrothermal
Ore Deposits, 2nd Edition: John Wiley and Sons, New
York, New York, USA, 798 p.
__ , 1981, Physicochemical constraints on porphyry
mineralization; pp. 71-77 in W. R. Dickinson and W. D.
Payne (eds.), Relations of Tectonics to Ore Deposits in
the Southern Cordillera: Arizona Geological Society
Digest, v. 14,288 p.
__ , and H. OHMOTO, 1980, Late-stage processes of
felsic magmatism: Mining Geology, Special Issue, no. 8, p.
1-11.
CARTEN, R B., E. P. GERAGHTY, B. M. WALKER, andJ. R
SHANNON, 1988, Cyclic development of igneous features
and their relationship to high-temperature hydrothermal
features in the Henderson porphyry molybdenum deposit,
Colorado: Economic Geology, v. 83, p. 266-296.
CHAPPELL, B. W., and A. J. R WHITE, 1974, Two con-
trasting granite types: Pacific Geology, v. 8, p. 173-174.
CLARK, A. H., E. FARRAR, J. C. CAELLES, S. J. HAYNES,
R B. LORTI, S. L. McBRIDE, G. S. QUIRT, R. C. R ROBERT-
SON, and M. ZENTILLI, 1976, Longitudinal variations in
the metallogenic evolution of the central Andes: A pro-
gress report; pp. 24-58 in D. F. Strong (ed.), Metallogeny
and Plnte Tectonics: Geological Association of Canada,
Special Paper 14, 660 p.
EMMONS, W. H., 1937, Gold Deposits of the World, with
a Section on Prospecting: McGraw-Hill, New York, New
York, USA, 562 p.
GILL, J. B., 1981, Orogenic Andesites and Plate Tec-
tonics: Springer-Verlag, Berlin, Germany, 401 p.
GUSTAFSEN, L. B., 1979, Porphyry copper deposits and
calc-alkaline volcanism, pp. 427-468 in M. W. McElhinney
(ed.), The Earth: Its Origin, Structure and Evolution:
Academic Press, London, England, UK, 590 p.
__ , and J. P. HUNT, 1975, The porphyry copper deposit
at EI Salvador, Chile: Economic Geology, v. 70, p. 857- 912.
HAMILTON, W. B., 1988, Plate tectonics and island arcs:
Geological Society of America Bulletin, v. 100, p. 1503-
1527.
HYNDMAN, D. W., 1981, Controls on source and depth of
emplacement of granitic magma: Geology, v. 9, p. 244-249.
HUTTON, D .H. W., 1988, Granite emplacement mechan-
isms and tectonic controls: Inferences from deformation
studies: Transactions of the Royal Society of Edinburgh:
Earth Sciences, v. 79, p. 245-255.
ISHIHARA, S., 1981, The granitoid series and minera-
lization: Economic Geology, 75th Anniversary Volume, p.
458-484.
KNAPP, R, and D. NORTON, 1981, Preliminary numerical
analysis of processes related to magma crystallization and
stress evolution in cooling pluton environments: American
Journal of Science, v. 281, p. 36-68.
KOIDE, H., and S. BHATTACHARJI, 1975, Formation of
fractures around magmatic intrusions and their role in ore
localization: Economic Geology v. 70, p. 781-799.
KUHNS, R J., 1986, Alteration styles and trace element
dispersion associated with the golden giant deposit, Hemlo,
Ontario, Canada; pp. 340-354 in A. J. Macdonald (ed.),
Gold '86: An International Symposium on the Geology
of Gold Deposits (Toronto): Konsult Int. (Willow dale,
Ontario), Proceedings Volume, 517 p.
__ , 1988, The Golden Giant Deposit Hemlo, Ontario
Geologic and Geochemical Relationships Between
Mineralization, Alteration, Metamorphism, Magma-
tism and Tectonism: PhD dissertation, University of
Minnesota, Minneapolis, Minnesota, USA, 458 p.
McKINSTRY, H. E., 1948, Mining Geology: Prentice-Hall,
NewYork,NewYork, USA,680p.
__ , 1955, Structure of hydrothermal ore deposits:
Economic Geology, 50th Anniversary Volume, p. 170-225.
MASON, R, and N. MELNIK, 1986, The anatomy of an
archean gold system - The McIntyre-Hollinger complex at
Timmins, Ontario, Canada; pp. 40-55 in A. J. Macdonald
(ed.), Gold '86: An International Symposium on the
Geology of Gold Deposits (Toronto): Konsult Int.
(Willowdale, Ontario), Proceedings Volume, 517 p.
__ , D. I. BRISBIN, and S. AITKEN, 1988, The geological
setting of gold deposits in the Porcupine mining camp:
Ontario Geological Survey, Miscellaneous Paper 140, p.
133-145.
NEWHOUSE, W. H., 1942, Ore Deposits as Related to
Structural Features: Princeton University Press, Prince-
ton, New Jersey, USA, 280 p.
PERCIVAL, T. J.,W. C. BAGBY, and A. S. RADTKE, 1988,
Physical and chemical features of precious metal deposits
hosted by sedimentary rocks in the western United States;
pp. 11-34 in R. W. Schafer, J. J. Cooper, and P. G. Vikre
(eds.), Bulk Mineable Precious Metal Deposits of the
Western United States: Proceedings of the Sympo-
sium: Geological Society of Nevada, Reno, Nevada, USA,
755p.
Felsic magmatism and hydrothermal gold deposits: A tectonic perspective 687
PITCHER, W. S., 1979, The nature, ascent and emplace-
ment of granitic magmas: Journal of the Geological Society
of London, v. 136, p. 627-662.
__ , 1982, Granite type and tectonic environment; pp.
19-40 in K. J. Hsu (ed.), Mountain Building Processes:
Academic Press, London, England, UK, 263 p.
REHRIG, W. A., and T. L. HEIDRICK, 1972, Regional
fracturing in Laramide stocks of Arizona and its relation-
ship to porphyry copper mineralization: Economic
Geology, v. 67, p. 198-21.
SHAW, H. R., 1980, The fracture mechanisms of magma
transport from the mantle to the surface; pp. 201-264 in R.
B. Hargraves (ed.), Physics of Magmatic Processes:
Princeton University Press, Princeton, New Jersey, USA,
565p.
SmSON, R. H., 1986, Earthquakes and lineament infra-
structure: Philosophical Transactions of the Royal Society
of London, v. A317, p. 63-79.
SILLITOE, R. H., 1973, The tops and bottoms of porphyry
copper deposits: Economic Geology, v. 68, p. 799-815.
__ , 1979, Some thoughts on gold-rich porphyry copper
deposits: Mineralium Deposita, v. 14, p. 161-174.
__ , 1981, Ore deposits in Cordilleran and island-arc
settings; pp. 49-69 in W. R. Dickinson and W. D. Payne
(eds.), Relations of Tectonics to Ore Deposits in the
Southern Cordillera: Arizona Geological Society Digest,
v.14,288p.
__ ,1987, Copper, gold and subduction: A trans-Pacific
perspective; p. 399-403 in E. Brennan (convenor), Pacific
Rim Congress '87: Proceedings: Australian Institute of
Mining and Metallurgy, Parkville, Victoria, Australia, 949
p.
__ , 1988, Gold and silver deposits in porphyry systems;
pp. 233-257 in R. W. Schafer, J. J. Cooper, and P. G. Vikre
(eds.), Bulk Mineable Precious Metal Deposits of the
Western United States: Proceedings of the Sympo-
sium: Geological Society of Nevada, Reno, Nevada, USA,
755p.
TITLEY, S. R., 1975, Geological characteristics and en-
vironment of some porphyry copper occurrences in the
southwestern Pacific: Economic Geology, v. 70, p. 499-514.
__ , and T. L. HEIDRICK, 1978, Intrusion and fracture
styles of some mineralized porphyry systems of the
southwestern Pacific and their relationship to plate
interactions; pp. 891-903 in L. B. Gustafson and S. R. Tit-
ley (eds.), Porphyry Copper Deposits of the South-
western Pacific Islands and Australia: Economic Geo-
logy, Special Issue, v. 73, no. 5,985 p.
__ , and R. E. BEANE, 1981, Porphyry copper deposits,
Part 1: Geologic settings, petrology and tectonogenesis:
Economic Geology, 75th Anniversary Volume, p. 214-235.
__ , R. C. THOMPSON, F. M. HAYNES, S. L. MANSKE, L.
C. ROBINSON, and J. L. WHITE, 1986, Evolution of
fractures and alteration in the Sierrita-Esperanza hydro-
thermal system, Pima County, Arizona: Economic Geo-
logy, v. 81, p. 343-370.
UYEDA, S., and C. NISHIWAKI, 1980, Stress field, metallo-
genesis and mode of subduction; pp. 323-340 in D. W.
Strangeway (ed.), The Continental Crust and Its Min-
eral Deposits: Geological Association of Canada, Special
Paper 20, 804 p.
WHITE, W. H., A. A. BOOKSTROM, R. J. KAMILLI, M. W.
GANSTER, R. P. SMITH, D. E. RANTA, and R. C. STEININ-
GER, 1981, Character and origin of Climax-type molyb-
denum deposits: Economic Geology, 75th Anniversary
Volume, p. 270-316.
WHITNEY, J. A., 1975, Vapor generation in a quartz mon-
zonite magma: A synthetic model with application to por-
phyry copper deposits: Economic Geology, v. 70, p. 346-
358.
__ , 1977, A synthetic model for vapor generation in
tonalite magmas and its economic ramifications: Econo-
mic Geology, v. 72, p. 686-690.
__ , 1988, The origin of granite: The role and source of
water in the evolution of granitic magmas: Geological
Society of America Bulletin, v. 100, p. 1886-1897.
WISSER, E. H., 1960, Relation of Ore Deposition to
Doming in the North American Cordillera: Geological
Society of America, Memoir 77, 117 p.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Two-stage basement fault-block deformation in the
development of the Witwatersrand goldfields, South Africa
R. E. MYERS, 1. G. STANISTREET, and T. S. McCARTHY
Department of Geology, University of the Witwatersrand, Johannesburg 2050, South Africa
(received May 15, 1989; accepted June 6,1990)
ABSTRACT
A relationship between a convergent continental margin on the northern edge of the Kaapvaal Craton
(Limpopo Belt) and development of the late Archaean Witwatersrand Basin has recently been proposed by
several authors. The actual tectono-sedimentary style within the basin and that to be expected from such a
relationship have not been properly assessed and, consequently, have been misrepresented. If "thin-skin"
structures (folds, thrusts, and regional cleavage) resulting from emplacement of the Middle Proterozoic
Vredefort Dome into the Witwatersrand Basin are backstripped, a two-stage tectonic model for the basin can
be recognized: (1) a syn-Witwatersrand ("'" 2800 Ma) cratonwide compressional event culminating in the
outpouring of Klipriviersberg flood basalts; and (2) a post-Witwatersrand ("'" 2700 Ma) extensional event that
relaxed the major oblique-slip reverse fault zones of the earlier event to develop grabens and half-grabens in
which Platberg Group sediments accumulated. A reconstruction of the Witwatersrand Basin shows an EfW-
trending through-going set of left-lateral faults interacting with N-trending right-lateral faults to define
individually moving fault blocks consistent with regional NE-SW compression. Syn-sedimentary monoclines
developed during deposition of the auriferous Central Rand Group as the surficial expression of the basement
block faulting. This was associated with development of gold-placer-based, unconformity-bounded
tectonostratigraphic packages, comprising mainly braided fluvial sediments that repeatedly overstep older
packages and basement across the block-fault margins. Deformation culminated in thick alluvial-fan
sequences developed against fault scarps, which caused prominent topographic highs. The Klipriviersberg
flood basalts abruptly terminated sedimentation. The craton wide extensional event associated with
deposition of the Platberg Group followed. This sequence of events may have Phanerozoic parallels in the rise
and collapse of basement blocks in the Laramide province of North America.
INTRODUCTION
Several authors have recently suggested that
the Witwatersrand Basin formed in response to the
convergence and collision of the Zimbabwe Craton
with the Kaapvaal Craton (Burke and others, 1986;
Stanistreet and others, 1986; Winter, 1987; Clen-
denin and others, 1988). Geophysical studies (Fair-
head and Scovell, 1976) suggest that the Zimbabwe
Cratonic crust was partially subducted beneath
Kaapvaal cratonic crust resulting in the Limpopo
orogenic belt.
Most authors view the Witwatersrand Basin as
a sag response to the developing orogenic belt,
emphasising an orogenic belt decollement-style of
deformation on a Kaapvaal Craton foreland. Our
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 689-697. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
689
690
R. E. Myers, 1. G. Stanistreet, and T. S. McCarthy
......... REVE RSE OBLIQUE - SLIP
FAULTS
a..
:>
o
a:
C)
C)
a: _
....
CD
>-
<t
-1
a..
a..
:>

C)
C)
a:
....
CD
m
a:
....
5
iii:
a..
:::;
l<'
C
z
<t
a: a..
:>
-10
eta:
a:C)
>-
z
....
U
c
z
<ta..
a::>

mC)
....

a..
;:)
0
a:
(!)
a:
w
a..
;:)
til
a..
a:
0
0
til
a:
W

Z
W
>
a..
::J
0
a:
(!)
a:
w
a..
::J
til
0
Z

a:
til
a:
w




Figure 1. Location of the Witwatersrand Basin and stratigraphy of the late Archean within the Kaapvaal Craton.
recent studies of the Witwatersrand Basin and
synthesis of available structural, stratigraphic, and
sedimentologic information, derived mainly from
mining and exploration activities, suggest that such
a sag response is not compatible with the tectonic
style of the Witwatersrand Basin at all. We suggest
that the basement became intricately involved with
sedimentation through development of large fault-
bounded basement blocks that controlled sedimen-
tation, particularly in the last stages of the basin
history (Myers and others, 1990).
GEOLOGIC SETTING OF THE
WITWATERSRAND BASIN
The late Archaean Witwatersrand Basin (3050-
2708 Ma; Figure 1) is preserved as a triangular-
shaped structural entity (Pretorius, 1986) covered to
a great extent by Proterozoic and Phanerozoic
sequences. During its early history, the area of the
basin was covered by a widespread epicontinental
sea, which was much more widespread than the
present basin itself and in which the West Rand
Group of the Witwatersrand Supergroup was depo-
sited. The sediments represent sub-tidal deposits
(Eriksson and others, 1981) in which thick super-
mature sub-tidal sand sheets, up to tens of metres
thick, were intercalated with offshore basinal
ferruginous mud deposits and pelagic ironstones.
In contrast, the Central Rand Group was de-
posited in a shrinking basin whose geometry
evolved during deposition toward its present con-
figuration. Such basin shrinkage has been proposed
previously (Pretorius, 1976; Vos, 1975) but not to
the extent we envisaged here. The sedimentary
Two-stage basement fault-block deformation, Witwatersrand goldfields 691
response to this tectonic evolution was a less mature
suite of sediments, which now comprise conglomer-
ates and meta-arkoses of fluvial origin. These are
intercalated with several transgressive sequences,
represented in one extreme case by a thick inter-
bedded mudstone and siltstone unit, the Booysens
Shale Formation. The finale in the Witwatersrand
Basin was the outpouring of the Klipriviersberg
Group lavas - a >2 km thickness of tholeiitic flood
basalts. Zircons from these lavas are dated at 2708
Ma (Armstrong, pers. commun., 1989).
After the eruption of the Klipriviersberg Group
lavas, Platberg Group (Middle Ventersdorp) rocks
were deposited in small fault-related depositories
within the area of the basin itself and, more sub-
stantially, in a large extensional rift basin just to
the west - which we have named the Lichtenburg
Basin (Figure 2B).
STRUCTURAL BACKSTRIPPING
Central to our recent studies is the philosophy
that in order to understand the development of the
Witwatersrand Basin, subsequent structural events
that have had a major effect on the basin must first
be stripped off in order to "see through" to the syn-
sedimentary tectonism. Figure 2 summarises this
process. Figure 2A shows major structural features
associated with the emplacement of the Vredefort
Dome (Simpson, 1977; McCarthy and others, 1986).
These include the Potchefstroom Synclinorium and
the Rand Anticline, along with numerous circum-
ferentially disposed thrusts, such as the Westonaria
thrust (Killick, 1986) and the Master Bedding Plane
Fault (Fletcher and Gay, 1972; McCarthy and
others, 1986). This phase of deformation represents
a thin-skinned tectonic response to the radially
directed deformation of Middle Proterozoic age out-
ward from the Vredefort Dome (McCarthy and
others, 1986).
When this youngest event has been stripped off,
two other tectonic phases can be recognised, which
are both intimately related with the development
and termination, respectively, of the Witwatersrand
Basin itself. The earlier of the two tectonic phases
involved the fault-controlled development of the
main Witwatersrand Basin, as discussed below. The
later phase controlled the development of the
Lichtenburg Basin in which fault-controlled sedi-
mentation comprised debris-flow-dominated fans
that prograded into lakes (Buck, 1980). The deve-
A
B
.. \
..... \
..... \
:.:.?\,
MIDDLE .' > : .... , WELKOM _
PROTEROZOIC '.' -7." ...................... .
DEFORMATION ' ' \
..oJ 1 40 '1) >0
.:- > >i
.',.: /:- .' D .... ..... 1.. SE Ul::o.cr
.. ' . c::J
m IIIII0000t VENT ERSOOftP
PL.&TIIEItG
c
LATE ARCHAEAN - 2800.00
0 ..,. Al[ft!R4HD SlDl .. E"',I,,..ON
o eaS wEHT
.. PItOw,HE"fIIY St OI .. EHl
Figure 2. Backstripping of structural events affecting the
Witwatersrand Basin. A) Proterozoic "thin-skinned"
deformation associated with emplacement of the Vredefort
Dome. B) Extensional rift phase that gave rise to the
Lichtenburg rift basins. C) Syn-Witwatersrand compres-
sional deformation.
692 R. E. Myers, I. G. Stanistreet, and T. S. McCarthy
lopment of felsic volcanic centres in the Lichtenburg
basin appears to have been controlled by the same
faults that controlled alluvial-fan sedimentation
(Myers, 1990). Fault zones such as the Rietfontein
and Ireton fault zones within the area of the Wit-
watersrand Basin underwent left-lateral strike-slip
movement (Stanistreet and others, 1986) at this
stage, and small half-graben pull-apart basins
formed where similar alluvial-fan type sedimentary
environments were developed. Very little volcanic
material accumulated in these collapse basins
(Stanistreet and McCarthy, 1990) compared with
the main Lichtenburg Basin. The Lichtenburg
Basin developed as an extensional rift basin that
accepted large volumes of mafic and felsic vocanics
and sediments of the Platberg Group (Ventersdorp
Supergroup).
To "see through" this tectonic phase during
deposition of the Platberg Group to the earlier syn-
Witwatersrand deformation is relatively straight-
forward. Most fault zones that had a major normal
component during the syn-Platberg Group basin
development had a major reverse component during
Witwatersrand deposition. Thus, subsiding deposi-
tional areas during development of the Platberg
Group (Lichtenburg Basin, Ireton half-graben,
Bezuidenhout Valley half-graben) acted as topo-
graphically positive source areas during deposition
of the earlier Central Rand Group. This example of
tectonic inversion has an opposite polarity to other
cases of tectonic inversion cited in the literature
(e.g., Stone ley, 1982; Etheridge, 1986). Most re-
ported cases involve situations in which normal
faults in extensional basins invert to reverse faults
during subsequent compressional tectonics. This
area appears to be characterized by what we refer to
as "negative tectonic inversion," compared to the
more common situation of positive tectonic inver-
sion.
THE SYN-SEDIMENTARY TECTONIC STYLE
OF THE WITWATERSRAND BASIN
The map shown in Figure 3 was compiled by
Myers and others (1990). It is based primarily on
magnetic and gravity maps of the Witwatersrand
Basin published by Corner and others (1986). It
identifies major syn-Witwatersrand, mainly com-
pressional structural features recognized by their
control on sedimentary facies and stratigraphic
thickness. Geologic resolution is high where mining
and exploration activity have been at a maximum.
This resolution decreases markedly into areas
where depth has constrained such activity, parti-
cularly in the area surrounding the Vredefort Dome
and to the southeast of this structure.
Information in high-resolution areas is ade-
quate, however, to indicate that the sedimentation
of the basin was controlled by rising fault-bounded
basement blocks, of which nineteen have so far been
identified although more will undoubtedly be recog-
nised in the future. The Meyerton Block, Block 5,
together with its neighbours and some features of
Block 16 (Welkom Block), along with a neigh-
bouring block, serve as illustrations of the way in
which these blocks interacted with sedimentation.
The Meyerton Block is located south of the city
of Johannesburg. It is bounded on the north by the
Rietfontein fault and on the west by the West Rand
fault. Figure 4A is a structural cross-section across
the Rietfontein Fault, showing Witwatersrand
Supergroup rocks folded upward into a monoclinal
structure by reverse motion on the fault. That this
structure formed during sedimentation is shown by
overstepping of the Elsburg conglomerate across the
upturned monoclinal limb and also by the way the
Klipriviersberg Group lavas overstep the structure
to rest on top of lower Witwatersrand and basement
rocks north of the fault. Subsequently, the Riet-
fontein fault behaved as a normal fault, producing
the Bezuidenhout Valley half-graben which filled
with Platberg Group sedimentary rocks.
The syn-Witwatersrand nature of the Rietfon-
tein fault is confirmed by the stratigraphic profile of
Figure 4B. This shows that the Johannesburg Sub-
group maintains an almost constant thickness
within the Meyerton Block along a section parallel
to the Rietfontein fault, but it thins dramatically
along stratigraphic profiles measured toward this
fault into the monocline. Myers and others (1990)
have also shown that the West Rand and Bank
faults defining the western margin of the block were
syn-Witwatersrand reverse faults downthrown to
the east. This displacement resulted in erosion and
elimination of the entire Central Rand Group from
related monoclinal structures prior to their behavior
as normal faults during deposition of middle Ven-
tersdorp (Platberg) strata. The Meyerton Block was
therefore tilted and rotated under the compressive
stresses affecting the entire basin during Wit-
watersrand sedimentation, but it returned to its
original position during the extensional tectonics
that controlled deposition ofVentersdorp sediments.
Two-stage basement fault-block deformation, Witwatersrand goldfields 693
+ + +
+
+
+
+
+
+
+
+
+
+
+ NUMBERS INDICA TE FAUL T
BLOCKS RECOGNIZED
+ + +
+
+ +
+
2 HALFWAY HOUSE BLOCK
4 CARLETONVILLE BLOCK
5 MEYERTON BLOCK
6 EAST RAND BLOCK
JOHANNESBURG
o 25 50km
I I
--- SYNSEDIMENTARY FAUL TS



8 EVANDER BLOCK
10 KLERKSDORP BLOCK
1 1 PARYS BLOCK
16 WELKOM BLOCK
KLIPRIVERSERG
SOUTH HILLS B
LINKSFIELD LANGERMANSKOP
A
O:=J kUPRllIlERSeURG LA ....A
CENTRAL RANO GROUP
[:::=J WEST RAND GROUP
E:!J BASEMENT

----- :::02 :-- 0 __ 0:_- : _-__ - _:-_ 2 --- 0 '"''"''' ' , ,
...... ...
Figure 3. Structural map of the Wit-
watersrand Basin, showing the major syn-
sedimentary fault (or deformation) zones
defining the edges of basement fault blocks.
Key to selected fault blocks discussed in the
text:
2.
4.
5.
6.
8.
Halfway House Block
Carletonville Block
Myerton Block
East Rand Block
Evander Block
10. Klerksdorp Block
11. Parys Block
16. Welkom Block
Figure 4. Structures and strati-
graphic changes associated with the
Riefontein fault zone. A) Structural
cross-section showing the syn-sedi-
mentary monocline associated with
syn-Witwatersrand reverse move-
ment on fault; note the half-graben
filled by Platberg Group sediments
during the later relaxation phase.
B) Variations in stratigraphic thick-
ness parallel and perpendicular to
block-margin fault structure.
694
R. E. Myers, 1. G. Stanistreet, and T. S. McCarthy
Similar events affected Block 16, the Welkom
Block. During deposition of the Central Rand
Group, reverse motion on the Border Structure, its
western bounding fault, caused monoclinal folding
(locally overturning) analogous in scale and type to
that just described for Block 5. The structure is now
covered by Phanerozoic rocks but has been de-
lineated by gold exploration boreholes. At the
Loraine gold mine, Olivier (1965) documented
overstepping of the Johannesburg Subgroup by the
younger Turffontein Subgroup reefs across the
developing monocline. At the Beisa mine, Tweedie
(1984) documented the same overstepping relation-
ship across the monocline, which is overturned in
this area (Figure 5A). The Border Structure also
collapsed following eruption of Klipriviersberg
Group lavas, with normal fault movement trun-
cating the closure of the monocline (Figure 5). It
was during this collapse that the large Lichtenburg
grabens developed to the west of the then-defunct
Witwatersrand Basin (Figure 2B).
SYNTHESIS AND DISCUSSION
Both examples described above show that block
faulting in the Witwatersrand Basin during its late-
stage evolution had a two-stage history. The com-
pressional stage occurred during Central Rand
Group deposition in the Witwatersrand Basin when
blocks were inclined and rotated, causing block-
marginal monoclines exemplified by successive eli-
mination and overstepping of strata into and onto
the structure. This was followed by a Platberg
Group extensional stage during which blocks col-
lapsed back in association with the newly develop-
ing Lichtenburg Basin, forming localised half-
grabens along block edges.
This style of basement uplift during the com-
pressional phase has its Phanerozoic analogue in
Laramide (Cretaceous) block faulting during deve-
lopment of the Rocky Mountains in the western
USA. Fault-bounded basement blocks were also
inclined and rotated (Matthews, 1978) during a
compressional stage associated with the Laramide
orogeny, which developed in a foreland setting
during convergence of the Farallon and North
American plates. Like the Witwatersrand ex-
amples, the Laramide blocks also underwent a later
collapse stage, during the Tertiary extensional
tectonic regime, which resulted in formation of
grabens and half-grabens of the Basin and Range

KAROO SEQUENCE -
'"
REFERENCE =<
'"
PHANEROZOIC COVER
KliPRIVIERSBERG BASALTS
r::7J TURFFONTEIN SUBGROUP
c::=:::J JOHANNESBURG SUBGROUP
E::J WEST RANO GROUP "
m
" g)
g
"
VVVVVV VY
VVVVVVVV
VVVVVVVV
METRES
Figure 5. A) Overturned monocline developed on the
Border Structure in the southern Welkom Goldfield during
the syn-Witwatersrand compressional phase; note the
backward collapse of the structure truncating the fold
hinge. B) A comparable syn-sedimentary Laramide struc-
ture of Cretaceous age from the western USA; note the
backward collapse ofthis structure during the Tertiary.
Province. Sales (1983) has described how such half-
grabens formed as structures between two neigh-
bouring blocks. One illustration of this phenomeon
Two-stage basement fault-block deformation, Witwatersrand goldfields 695
from the Seminoe Mountains of Wyoming is repro-
duced in Figure 5B. The fault-related fold struc-
ture is remarkably similar in both size and style to
that reproduced in Figure 5A and to those described
by Tweedie (1984) from the Beisa area of the Wel-
kom goldfield. The Bezuidenhout Valley half-
graben collapse that developed along the Rietfon-
tein fault zone (Figure 4A) is analogous in its style
and timing to the Tertiary half-graben that deve-
loped from the Seminoe Mountains structure.
The two-stage development of the Witwaters-
rand Basin fault-blocks is therefore paralleled by
the Laramide structures. Even the size of the main
Witwatersrand Basin is more consistent with that of
basins between Laramide basement uplifts, such as
the Green River and Washakie Basins (Matthews,
1978), than it is with that of foreland basins
suggested by Burke and others (1986) and Winter
(1987).
The only element not so far discussed is the
outpouring of Klipriviersberg Group flood basalts
between the compressional stage and the exten-
sional stage. This phenomen also has its parallel in
the evolution of the Rocky Mountains. In both the
northern and southern Rocky Mountains, flood
basalts were extruded (i.e., Columbia River Group
and lateral equivalents) after compressional tec-
tonics that caused the Laramide orogeny and before
extensional tectonics associated with development
of the Basin and Range Province (Figure 6). With
KAAPVAAL CRATON
2699 Ma
PLATBERG GROUP
RELAXATION
{{
10Ma
V
v vyvvv
v v
the newly discovered preCIsIOn of zircon ion-
microprobe dating techniques, comparisons can also
be made between the relative timing of the three
events. The Laramide basement-block-fault style
developed over a period of at least 80 million years.
Although the Dominion, West Rand, and Central
Rand Groups together developed over a period of
350 million years, the Central Rand Group reflects
the Laramide-style of sedimentation related to base-
ment block faulting, it perhaps developed over a
span of "" 100 million years (Figure 6). Outpouring
of the Columbia River basalts and lateral equiva-
lents occurred during a period of about "" 7 million
years before Basin and Range extension, compared
with a period of "" 9 million years (Figure 6)
between the start of the Klipriviersberg Group
extrusion and development of the Platberg Group
grabens (Armstrong and others, 1986; Armstrong,
pers. commun., 1989).
Definite similarities are apparent between the
stages in evolution of the Kaapvaal Craton during
the late Archaean and the evolution of the western
United States during the Cretaceous and Tertiary
with respect to the occurrence of flood basalts. The
important feature about the later extensional phase
in the case of the Kaapvaal Craton was that it
collapsed with normal movement along the same
fault zones that had acted as reverse faults during
the compressional phase. Such a collapse involving
all the previously compressive structures would not
WESTERN U.S.A
- O-IOMa
+ BASIN a RANGE
STRUCTURES
17 Ma
COLUMBIA RIVER
Figure 6. Diagram showing the
possible comparable develop-
ment of late Archaean Kaapvaal
Craton and the western USA:
earlier compressional phase
(bottom); extrusion of flood ba-
salts (middle); and later exten-
sional phase (top).
2708 Ma
KLiPRIVIERSBERG
FLOOD BASALTS vv VVVVVVVYVVVYV
FLOOD BASALTS
-,2800 Ma
WITWATERSRAND
FAULT BLOCKS
-
v v v v v v v v v v v v V v AND EQUIVALENTS
Ma
LARAMIDE
FAULT BLOCKS
696
R. E. Myers, 1. G. Stanistreet, and T. S. McCarthy
be expected if fold-and-thrust-belt decollement-style
tectonics had dominated the cratonic response to
orogeny. It is more likely that the Witwatersrand
Basin was dominated by tectonic movement of deep-
seated basement block faults during sedimentation.
Similar responses to Phanerozoic compression have
recently been detected by deep seismic studies
within cratons in Australia (Goleby and others,
1989).
ACKNOWLEDGMENTS
Financial support from the C.S.I.R., A.E.C., Jim
and Gladys Taylor Trust, and the University of the
Witwatersrand is gratefully acknowledged. Judy
Wilmot, Adele Worman-Davies, and Una Ryan are
thanked for their assistance with manuscript
preparation. We thank Anglo American Corpora-
tion for permission to use unpublished data.
REFERENCES
ARMSTRONG, R. A., W. COMPSTON, E. A. RETIEF, and H.
J. WELKE, 1986, Ages and isotopic evolution of the Ven-
tersdorp volcanics [abstract]: Geocongress '86, University
of the Witwatersrand, p. 89-92.
BUCK, S. G., 1980, Stromatolite and ooid deposits within
the fluvial and lacustrine sediments of the Precambrian
Ventersdorp Supergroup of South Africa: Precambrian
Research, v. 12, p. 311-330.
BURKE, K., W. S. F. KIDD, and T. M. KUSKY, 1986,
Archaean foreland basin tectonics in the Witwatersrand,
South Africa: Tectonics, v. 5, p. 439-456.
CLENDENIN, C. W., E. G. CHARLESWORTH, and S.
MASKE, 1988, Tectonic style and mechanism of Early Pro-
terozoic successor basin development, southern Africa:
Tectonophysics, v.156, p. 275-291.
CORNER, B., R J. DURRHEIM, and L. O. NICHOLAYSEN,
1986, Structural framework of the Witwatersrand Basin as
revealed by gravity and aeromagnetic data [abstract):
Geocongress '86, University of the Witwatersrand, p. 27-30.
ERIKSSON, K. A., B. R TURNER, and R G. VOS, 1981,
Evidence of tidal processes from the lower part of the
Witwatersrand Supergroup, South Africa: Sedimentary
Geology, v. 29, p. 309-325.
ETHERIDGE, M. A., 1986, On the re-activation of exten-
sional fault systems: Philosophical Transactions of the
Royal Society of London, v. A317, p.179-194.
FAIRHEAD, J. D., and P. D. SCOVELL, 1976, Gravity study
of the Limpopo Belt, southern Africa: University of Leeds,
Research Institute of African Geology, Annual Report, v.
20, p. 31-35.
FLETCHER, P., and N. C. GAY, 1972, Analysis of gravity
sliding and orogenic translation: Discussion: Geological
Society of America Bulletin, v. 82, p. 2677-2682.
GOLEBY, B. R, RD. SHAW, C. WRIGHT, B. L. N. KENNET,
and K. LAMBERT, 1989, Geophysical evidence for "thick-
skinned" crustal deformation in central Australia: Nature,
v. 337, p. 323-330.
KILLlCK, A. M., A. M. THWAITES, A. E. SCHOCH, and G. J.
B. GERMS, 1986, A preliminary account of the tectonites
near the interface between the Ventersdorp and Wit-
watersrand Supergroups, west Rand area, South Africa
[abstract]: Geocongress '86, University of the Witwaters-
rand, p. 35-38.
McCARTHY, T. S., E. G. CHARLESWORTH, and 1. G.
STANISTREET, 1986, Post-Transvaal structures across the
northern portion of the Witwatersrand Basin: Trans-
actions of the Geological Society of South Africa, v. 89, p.
311-323.
MANGAN, M. T., T. L. WRIGHT, D. A. SWANSON, and G. R
BYERLY, 1986, Regional correlation of Grande Ronde
basalt flows, Columbia River Basalt Group, Washington,
Oregon, and Idaho: Geological Society of America Bulletin,
v.97,p.1300-1318.
MATTHEWS, V. (ed.), 1978, Laramide Folding Associa-
ted with Basement Block Faulting in the Western
United States: Geological Society of America, Memoir
151,370p.
MYERS, J. M., 1990, The Stratigraphy and Geochem-
istry of the Klipriviersberg and Platberg Groups of
the Ventersdorp Supergroup at Buffelsfontein, Wes-
tern Transvaal: MSc thesis, University of the Wit-
watersrand, Johannesburg, South Africa, 294 p.
MYERS, R E., T. S. McCARTHY, and 1. G. STANISTREET,
1990, A tectono-sedimentary reconstruction of the develop-
ment and evolution of the Witwatersrand Basin, with
particular emphasis on the Central Rand Group; pp. 180-
201 in C. R Anhaeusser (ed.), Origin and Evolution of
the Witwatersrand Basin and Its Mineralization:
South African Journal of Geology, Special Issue, v. 93, 310
p.
OLIVIER, H. J., 1965, The tectonics of the Witwatesrand
System in the Loraine area of the Orange Free State
Goldfield: Transactions of the Geological Society of South
Africa, v. 68, p. 143-177.
PRETORIUS, D. A., 1976, The nature of the Witwatersrand
gold-uranium deposits; pp. 29-88 in K. H. Wolf (ed.),
Handbook of Strata-Bound and Stratiform Ore De-
Two-stage basement fault-block deformation, Witwatersrand goldfields 697
posits, II: Regional Studies and Specific Deposits,
Volume 7: Elsevier, New York, New York, USA, 656 p.
__ ,1986, The geometry of the Witwatersrand Basin and
its structural setting in the Kaapvaal Craton [abstract):
Geocongress '86, University of the Witwatersrand, p. 47-52.
SALES, J. K., 1983, Collapse of Rocky Mountain basement
uplifts, pp. 79-97 in J. D. Lowell and R. Gries (eds.), Rocky
Mountain Foreland Basins and Uplifts: Rocky Moun-
tain Association of Geologists, Field Conference, 392 p.
SIMPSON, C., 1977, A Structural Analysis of the Rim
Synclinorium of the Vredefort Dome: MSc thesis, Uni-
versity of Witwatersrand, Johannesburg, South Africa,
257p.
STANISTREET, 1. G., and T. S. McCARTHY, 1990, Middle
Ventersdorp graben development north of the Carleton-
ville Goldfield and its relevance to the evolution of the
Witwatersrand Basin: South African Journal of Geology,
v. 93, p. 272-288.
__ , T. S. McCARTHY, E. G. CHARLESWORTH, R. G.
MYERS, and R. A. ARMSTRONG, 1986, Pre-Transvaal
wrench tectonics along the northern margin of the Wit-
watersrand basin, South Africa: Tectonophysics, v. 131, p.
53-74.
STONELEY, R., 1982, The structural development of the
Wessex Basin: Journal of the Geological Society of London,
v. 139, p. 543-554.
TWEEDIE, K. A. M., 1984, The discovery and exploration of
Beisa and Beatrix gold and uranium mines in the southern
extension of the Orange Free State goldfield, South Africa:
Proceedings of the 27th International Geological Congress
(Moscow), Section C.04, p.1-14.
VOS, R. G., 1975, An alluvial plain and lacustrine model
for the Precambrian Witwatersrand deposits of South
Africa: Journal of Sedimentary Petrology, v. 45, p. 480-
493.
WINTER, H. de la R., 1987, A cratonic foreland model for
Witwatersrand Basin development in a continental back-
arc plate tectonic setting: South African Journal of Geo-
logy, v. 90, p. 409-427.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Archean crustal lead in the Helena Embayment of the
Belt Basin, Montana
ROBERT E. ZARTMAN
U.S. Geological Survey, Mailstop 963, Federal Center, Denver, Colorado 80225, USA
(received August 12, 1988; revision accepted April 7,1989)
ABSTRACT
An extensive database of lead isotopic analyses for ore deposits throughout the Middle Proterozoic Belt
basin reveals two isotopic populations that represent Mesozoic/Cenozoic and Precambrian ages of
mineralization. The population of Precambrian age, which comprises lead originally introduced more or less
contemporaneously with deposition of the host sedimentary rock, offers the greatest promise of identifying
provenances for Belt sedimentation. Most of these deposits show a restricted range in isotopic composition-
including 207PbPo4Pb - which must characterize much of the source rocks giving rise to the Coeur d'Alene-
type vein deposits.
New isotopic analyses of the Precambrian lead reveal unusually high 207PbPo4Pb for galena as well as
whole-rock conglomeratic LaHood Formation. Unlike most of the Coeur d'Alene-type ore deposits previously
studied, these occurrences represent samples from the Helena Embayment near the southeastern margin of
the basin, close to outcrops of Archean crystalline rock. Both ore (Little Belt Mountains, Bridger Range, and
Highland Mountains) and some rock lead (LaHood Formation from Jefferson Canyon and Highland
Mountains, but not Newland and Greyson Formations from Highland Mountains) display a large departure
from the more tightly grouped 'normal' Coeur d'Alene- type lead.
The elevated 207PbPo4Pb and steep slope of this more extended isotopic array identify the presence of a
significant component of lead derived from early Archean (with Stacey and Kramers model ages of 3600-4100
Ma) continental crust. By contrast, the virtual absence of this Archean signature elsewhere in the Belt basin
means that another, clastic or dissolved component of lead - possibly derived from Early Proterozoic rock
sources - usually dominates the sedimentary budget. Only in certain tectonically controlled, nearshore sites,
such as the Helena Embayment, can the influx of lead from Archean sources be clearly identified.
INTRODUCTION
Lead isotope data now exist for galena hosted by
1500-1200 Ma rocks in the Middle Proterozoic Belt
Supergroup from approximately 100 mines and pro-
spects in northwestern Montana, northern Idaho,
northeastern Washington, and southeastern British
Columbia (Figure 1; see compilations of Marvin and
Zartman, 1984; Godwin and others, 1988). High-
precision analyses reveal two distinct populations
when plotted on 208Pb/
204
Pb us 206PbPo4Pb and 207Pb/
204Pb us 206PbPo4Pb diagrams (Figure 2). The two
populations were interpreted by Zartman and
Stacey (1971) and LeCouteur (1973) to reflect fun-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 699-710. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
699
700
/lIllll'>ll
-;
<::
I
/ .
\
____ I
\
o
\
\
/
100
MIDDLE
PROTEROZOI
\
\. -,
200
R. E. Zartman
\
\
I

EXPLANA TION
D
Missoula Groupand
!
Middl e Beh .,bono,.
Ravall i Group
Lower Beh
ARCHEAN A D {D
EARLY PROT ROZOIC Crystalline b.sement rocks
112"
1
" III U I \
r
1
---- -- - -- - - ---
'1 (1 "\ , .\, \
1
SAMPLE LOCALITIES
Galen.
o Meso.tolc Ceno7oic
Precam brian
Rock
... Lower Be-II
Figure 1. Geologic map of the Belt-Purcell basin showing sample location for lead isotope data of Middle Proterozoic (Coeur
d'Alene-type) and Mesozoic/Cenozoic ore deposits. Labelled localities represent newly analyzed ore deposits and Lower Belt
sedimentary rocks. Simplified from Winston (1986).
Archean crustal lead in the Helena Embayment 701
A
40
o
o
o 0
'"0
0
%
00
o 0
o

o 6'6' 0
o
Figure 2. 208
Pb
/204
Pb
us 206pb;204
Pb
(A)
and 207Pb;204Pb us 206Pb/ 204Pb (B) for pub-
lished Belt galenas. The two isotopic popu-
lations of ore deposits (see text) are re-
presented by: solid circles, Precambrian;
open circles, Mesozoic/ Cenozoic. Model
lead growth curves on this and subsequent
figures from Stacey and Kramers (1975).

206 Pb/ 204 Pb
15.8
15.2
B
ro
o
9cc-J---'--J-----;:20
206 Pb/204 Pb
damentally different sources of the lead and ages of
mineralization. Because lead isotope patterns are
useful in placing temporal and spatial constraints
on source materials, it is particularly instructive to
look for such information in large data sets. The
earlier studies within the Belt basin have already
been productive in this respect and encouraged an
additional search for lead capable of further con-
straining provenances.
One of the two isotopic varieties of Belt-hosted
lead (Mesozoic/Cenozoic; shown by open symbols on
Figures 1 and 2) occurs ubiquitously in hydro-
thermal ore deposits in many different host rocks
throughout the Rocky Mountain region of United
States and Canada. The mineralization is generally
associated with Mesozoic and Cenozoic plutonism
and/or volcanism, and the lead isotopes imply
derivation at depth from the Early Proterozoic
("" 1700 Ma) or Archean ("" 2700 Ma and older)
crystalline basement. Extensive studies of such de-
posits have led to the recognition of basement age-
correlated, regional lead isotope provinces (Zart-
702
R. E. Zartman
Figure 3. 208PbP04Pb us 206PbP04Pb
(A) and 207pbP04
Pb
us 206
Pb
/ 204
Pb
(B)
for Middle Proterozoic (Coeur d'Alene-
type) lead. Host rock of ore deposit
represented by: open circles, Lower
Belt; solid squares, Ravalli Group;
solid triangles, Middle Belt carbonate.
Lines a' and an are upper and lower
bounding secondary isochrons for
instantaneous isotopic evolution at
1300 Ma. The trend of lines a' and an
defines the elongate array,; heavy
arrows point to the alignment of most
isotopic data along the 'primary' array.
.D
a..

37.0
36.5
t
36.0
A
o

16.0 16.5 17.0 175
206Pb/ 204 Pb
15.60r--c--,---,---,---,---,---,---,---r--r--r---,
B
1555
1550
1545
1540
15.35
man, 1974), which otherwise usually show little
relationship with the supracrustal rocks imme-
diately hosting the deposit.
The other isotopic variety of lead (Precambrian;
shown by solid symbols on Figures 1 and 2) occurs
only in deposits hosted by rocks of the Middle
Proterozoic Belt Supergroup or coeval mafic dikes
and sills. Mineralization is presumed to have taken
place originally in Lower Belt strata more or less
contemporaneously with deposition of the enclosing
sedimentary rocks. The metals, which were some-
206 Pb/204 Pb
times localized as stratabound occurrences, have
been later commonly remobilized into fracture-
filling veins. This second variety (referred to as
Coeur d'Alene-type) of lead has a greater potential
than the Mesozoic/Cenozoic variety for containing
information on the provenances for Belt sedimenta-
tion and, consequently, it provides the stimulus for
the present study.
The Coeur d'Alene-type of lead, especially as it
is found in most of the producing mines of the area,
is characterized by a rather restricted range in iso-
Archean crustal lead in the Helena Embayment 703
topic composition. Several workers, however, have
called attention to discernible differences in isotope
ratios and speculated that this variability might
provide information about the source of the lead and
the mineralization process (Zartman and Stacey,
1971; LeCouteur, 1973). In detail, the analyses
define elongate arrays on both 208Pb/
204
Pb us 206Pb/
204Pb and 207Pb/
204
Pb us 206Pb/
2o4
Pb diagrams (Fig-
ure 3), which are best explained as arising from a
combination of inherited and superimposed pheno-
menon. Locations of individual data points relative
to the overall array on the 207Pb/
204
Pb us 206Pb/
204
Pb
diagram (Figure 3B) correlate to some extent with
stratigraphic position, suggesting either more evol-
ved isotopic composition with decreasing age of
mineralization or, more likely, incorporation of
indigenous radiogenic lead by ascending fluids.
Attempts to model the linear trend of data for
Ravalli Group and Middle Belt carbonate samples
produce secondary isochron ages of 1300 to 825 Ma
with large uncertainties but supportive of Precam-
brian remobilization (Zartman and Stacey, 1971).
If this interpretation is correct, the isotopic
composition of all samples of Coeur d'Alene-type of
lead on Figure 3B - whether or not they had ac-
quired an additional component of younger radio-
genic lead - should initially have lain between the
bounding secondary isochrons a' and a". Those
samples lying at the left extremity of the bounded
field are considered less likely to have been later
modified isotopically and more likely to reflect the
real isotopic heterogeneity of lead incorporated into
these Belt rocks at the time of sedimentation.
LeCouteur (1973) recognized the distinctively steep
slope of this 'primary' array on a 207Pb/
204
Pb us
206PbPo4Pb diagram and concluded that it might re-
present a mixing line between low 207Pb/
204
Pb mar-
ine and high 207PbPo4Pb continental components of
lead within the Belt basin of sedimentation. The
'primary' array might also serve to constrain the age
of the continental component, but the apparent
dominance of 'normal' marine lead allowed only
speculation that the array's steep slope was due to
the presence of Archean material. Indeed, the
marine lead itself could represent a clastic or dis-
solved component derived locally from Early
Proterozoic rocks rather than from some global(?)
oceanic dispersal oflead. [Hereafter, when mention
is made of marine lead, it will only be in regard to its
low and, presumably, more normal 207Pb/
204
Pb -
without inference about genesis.]
Interestingly, one of the highest 207PbPo4Pb in
either the Zartman and Stacey (1971) or LeCouteur
(1973) study was found on lead from the Sullivan
mine (Kimberley, British Columbia), which is the
largest lead-zinc deposit hosted in Purcell (= Belt)
Supergroup rocks yet discovered. The importance of
confirming LeCouteur's interpretation of the 'pri-
mary' array and possibly establishing a relationship
between the source(s) of Coeur d'Alene-type lead
and its location both geographically and stratigra-
phically prompted further investigation.
HIGH 207Pb/
204
Pb IN THE HELENA
EMBAYMENT
Thus began the search for more convincing
evidence of lead from recycled Archean crustal
rocks, such as are known to crop out today in the
Wyoming province immediately to the east and
south of the Belt basin (Figure 1). The premise was
that detritus derived from Archean upper crust with
a substantially elevated 11 (= 238U/204Pb) would im-
part a distinctively high 207PbPo4Pb to the rock.
Likewise, some dissolved component in an intersti-
tial fluid, either entering the basin independent of
the sediment or having isotopically equilibrated
with the sediment, should also be recognizable by its
high 207PbPo4Pb. Initially, the study focused only on
lead from ore minerals, but, because of questions
about the association between ore and sediment,
five whole-rock samples from the Belt Supergroup
were also investigated.
Most of the previously analyzed galena samples
came from the central and western parts of the Belt
basin at distances >300 km from exposed Archean
crystalline rocks of the Wyoming province (Figure
1). For the present study, lead-bearing samples
were obtained from occurrences within the Helena
Embayment close to the basin's present eastern and
southern borders. The original perimeter of the
basin, of course, has been modified by tectonic and
erosional processes; however, demonstratable un-
conformities and near-shore facies rocks provide
convincing evidence that the present and original
border of the basin have changed little in this area
(Schmidt and Garihan, 1986).
Isotopic analyses of samples from three loca-
lities (JS-24, -30, -36, Little Belt Mountains; PCD-l,
-2, Bridger Range; P 2063, P 2067, Highland Moun-
tains; Figure 1, Table 1) containing Coeur d'Alene-
type lead within the Helena Embayment clearly
704 R. E. Zartman
Table 1. Lead isotopic composition of anomalously high 206Pb/2o4Pb galena from the Belt Supergroup, Montana.
206
Pb
207
Pb
208
Pb
Sample Location Host Rock
204
Pb
204
Pb 204Pb
Tuc Puc
PCD-1 Bridger Range, LaHoodFm 17.134 15.810 37.109 3.80 11.15
Gallatin Co
PCD-2 Bridger Range, LaHoodFm 16.980 15.811 36.888 4.11 10.74
Gallatin Co
P2063 Highland Mtns, NewlandFm 16.817 15.671 36.466 3.97 10.52
Silver Bow Co
P2067 Highland Mtns, NewlandFm 16.805 15.700 36.497 4.09 10.47
Silver Bow Co
JS-24 Little Belt Mtns, NewlandFm 16.739 15.590 36.495 3.56 10.55
Meagher Co
JS-30 Little Belt Mtns, NewlandFm 16.745 15.597 36.522 3.60 10.54
Meagher Co
JS-36 Little Belt Mtns, NewlandFm 16.700 15.573 36.467 3.57 10.46
Meagher Co
Samples: PCD-1 and PCD-2, shear-controlled fracture-filling veins in LaHood Formation associated with the Cross
Range fault (4558'N, 111 03'W); P 2063 and P 2067, disseminated to massive sulfides along bedding planes in
Newland Formation (4543'N, 11237'W); JS-24, JS-30, and JS-36, pore space-filling, replacement, and fracture-filling
sulfides in Newland Formation (4645'N, 11105'W). Lead isotopic ratios normalized to UBC Broken Hill #1 values of
2oGpb/
204
Pb = 16.007, 207pbP04Pb = 15.397, 208Pb/204Pb = 36.675. Stacey and Kramers (1975) model ages, T
uc
' and 11
values, Iluc' are calculated relative to second-stage parameters (extrapolated in the case of ages >3700 Ma), assuming
a mineralization age of 1400 Ma.
show the signature of ancient continental crust.
Here, within distances of kilometers to tens of
kilometers of the present and, presumably, original
border of the Belt basin, occurrences of galena
display a large departure from the tighter isotopic
grouping of the 'primary' array (Figure 4). These
new analyses define a steep slope on a 207PbPo4Pb us
206PbPo4Pb diagram (Figure 4B), which is an up-
ward extension of the 'primary' array. By contrast,
resolution between the 'primary' array, including its
extension to include the new analyses, and the
elongate array is not obvious on the corresponding
208Pb/
204
Pb us 206PbPo4Pb diagram (Figure 4A). Al-
though the 'primary' and elongate arrays carry very
different age information, they do not allow a dis-
tinction in Th/u (3.4) despite the different histories
proposed for their radiogenic lead components.
Apparently, geographic position within the Belt
basin plays the dominant role in determining whe-
ther the lead will have an elevated 207PbPo4Pb.
However, except for proximity to the eastern and
southern shoreline, little can be said yet about
either horizontal or vertical variation until a sub-
stantially larger database is acquired. So far, lead
with distinctly elevated 207PbPo4Pb has been found
only in units of the Lower Belt, extending from the
LaHood to the Newland Formations. The highest
207PbPo4Pb does occur in galena hosted by the La
Hood Formation, an observation supporting the
prediction of a correlation between isotopic com-
position and content of locally derived clastic
material in the sediment.
The older literature - published prior to 1970
and reporting less precise isotopic analyses - re-
veals two additional localities from the southern
Lewis and Clark Range that may contain galena
with elevated 207PbPo4Pb (Mudge and others, 1968).
If confirmed by modern mass spectrometric tech-
niques, these data would further support the gen-
eral presence of recognizable Archean lead near
those parts of the basin abuting and/or overlapping
onto the Wyoming province.
Archean crustal lead in the Helena Embayment 705
An investigation based only on the isotopic
composition of galena leaves unanswered important
questions about the pathway and ultimate source of
the mineralizing fluids. In order to pursue such
questions further, several samples of Lower Belt
sedimentary rocks were analyzed, too. Table 2 pre-
sents D, Th, and Pb concentration and lead isotopic
composition data for five whole-rock samples of the
LaHood, Newland, and Greyson Formations collec-
ted near the southern margin of the Helena Em-
bayment. Before making a comparison of the ore
Figure 4. 208
Pb
/204
Pb
us 206pbP04
Pb
(A) and 207
Pb
/204
Pb
us 206
Pb
/204
Pb
(B)
for Middle Proterozoic (Coeur d'Alene-
type) lead, including labelled, newly
analyzed samples.
Symbols same as Figure 3, with the
addition of the new analyses repre-
sented by: solid circles, Lower Belt-
hosted ore deposits; crosses, Lower Belt
sedimentary rocks.
Also shown are the trajectory lines
calculated to correct the isotopic com-
position of sedimentary rocks for 1400
million years of in situ radioactive de-
cay.
.D
Q.

375
37.0
36.5
36.0
35.5

350L ,
160
15.6
.D
Q.
o
and rock isotopic composition, however, a substan-
tial correction is required for in situ growth of
radiogenic lead due to radioactive decay. Accord-
ingly, the isotopic ratios corrected for an estimated
age of 1400 million years (Obradovich and others,
1984) are also given (in parentheses) in Table 2.
Having no a priori knowledge of the initial iso-
topic composition of these rocks allows a consider-
able latitude in the range of acceptable values, but
certainly the following lie outside of that permis-
sible range: 208Pb/
204
Pb (= 45.13) of sample WC-48,
?
%
,..0 PCD-2. %
'6 J&-24
S, LJ&-30
J&-36"., P 2067
o P 2063
5J
I
165
206 PDI 204 Pb
P 2067.
P 2063
I
170 175
152L-__ L-__ __ __ __ -L __ -L __ __ -J ____
160 165 170 175
206 Pb/204 Pb
706 R. E. Zartman
Table 2. U, Th, and Pb concentration and lead isotopic composition of whole-rock samples of the LaHood,
Newland, and Greyson Formations, Belt Supergroup, Montana.
U Th Pb
206Pb
207
Pb
208
Pb
Sample Location
(ppm) (ppm) (ppm)
204Pb
204
Pb
204
Pb
JC-l LaHoodFm, 0.99 3.38 18.3 17.878 15.878 37.063
Jefferson Canyon, (16.778) (15.805) (36.217)
Jefferson Co
JC-2 LaHoodFm, 0.94 2.99 5.44 18.989 15.867 37.925
Jefferson Canyon, (16.318) (15.603) (35.325)
Jefferson Co
WC-48 LaHoodFm, 0.65 6.36 2.52 22.697 16.078 61.544
Highland Mtns, (17.180) (15.588) (45.130)
Silver Bow Co
WC-15 NewlandFm, 7.1 24.3 12.2 20.141 15.753 39.570
Highland Mtns., (10.732) (14.918) (29.832)
Silver Bow Co
PP-17 FreysonFm, 3.6 12.0 21.0 19.307 15.643 38.886
Highland Mtns, (16.646) (15.407) (36.145)
Silver Bow Co
Samples: JC-l, olive, well-indurated, conglomeratic, arkosic wacke of the lowermost LaHood Formation
(4550'55"N, 111 55'OO"W); JC-2, grayish-green, well-indurated, medium-grained, arkosic wacke ofthe LaHood
Formation (4550'20"N, 11154'30"W); WC-48, light gray, micaceous, conglomeratic, arkosic wacke of the
LaHood Formation (4542'OO"N, 11232'OO''W); WC-15, pinkish-red to tan, laminated, marly shale of the
Newland Formation (4543'OO"N, 11235'30"W); PP-17, bluish-gray, finely-laminated, silty argillite of the
Greyson Formation (4545'30"N, 11229'OO"W). Lead isotopic ratios normalized to UBC Broken Hill #1 values of
206PbP04Pb = 16.007, 207Pb/
204
Pb = 15.397,208Pb/204Pb = 36.675. Lead isotopic ratios in parentheses have been
corrected for 1400 million years of in situ radioactive decay.
and 206Pb/
204
Pb (= 10.73) and 208Pb/
204
Pb (= 29.83)
of sample WC-15. Although these impossible values
demand open-system behavior of the rocks and cast
doubt on the accuracy of all the calculated initial
ratios, the insensitivity of the critical 207Pb;204Pb to
such error still make the intended comparison valid.
The initial ratios of the five sedimentary rocks are
plotted on Figure 4, where they may be compared
with the galena isotopic composition. For isotopic
systematics involving only uranogenic lead (Figure
4B), erroneous corrections due to uranium and/or
lead migration in the rock will in most instances
cause displacement along a sub horizontal trajectory
line (the gain of some exotic lead being a possible
exception) and, thus, produce little change in
207Pb;204Pb.
As depicted in Figure 4, the evidence would
seem conclusive that the high 207Pb;204Pb signature
observed in the galena of the Helena Embayment is
likewise present in the coarse conglomeratic rocks of
the LaHood Formation but not in the fine-grained
calcareous Newland or shaly Greyson Formations.
Indeed, these latter two formations yield a lead that
resembles isotopically the 'normal' marine lead spe-
culated by LeCouteur (1973) to dominate through-
out most of the Belt basin. The demonstration of a
similar range in 207PbPo4Pb for both ore deposits and
sedimentary rocks supports the hypothesis that the
mineralizing fluids do originate as interstitial solu-
tions in the enclosing sediments. The ore appar-
ently can later move into a different stratigraphic
unit where its isotopic signature no longer matches
that of the new host rock, as, for example, seems to
be the case with galena (samples P 2063 and P 2067,
Table 1) hosted by the Newland Formation (re-
presented by sample WC-15) in the Highland
Mountains.
Archean crustal lead in the Helena Embayment 707
DISCUSSION
The association of lead having high 207Pb/
204
Pb
(relative to its 206Pb/
204
Pb) with residence in an old,
high P environment, and the identification of that
high p environment with Archean continental crust,
is now well established (Stacey and Zartman, 1978;
Gariepy and others, 1985). Not only is this isotopic
characteristic revealed in preserved ancient crustal
rocks, but the anomaly can survive transport and
redeposition of such old rock into younger strata. In
the latter case, of course, other contributing com-
ponents to the sediments, such as contemporaneous
volcanic rocks and recycled younger continental
crust, will compete in determining isotopic com-
position. Accordingly, a distinctly high 207Pb/
204
Pb
should occur only where old crustal rocks comprise a
significant source of the lead in sediments. Lead
from both ore deposits (samples PCD-1 and PCD-2,
Table 1) and conglomeratic rock (samples JC-1 and
JC-2, Table 2) of the LaHood Formation yield the
highest 207PbPo4Pb found in this study, but without
detailed information about the isotopic composition
and element concentration of the various competing
source materials, a quantitative evaluation cannot
be made of the proportion of each component pre-
sent.
The next question to ask of the Pb-isotope data
is just how well can one constrain the age of any
older Archean component(s). A unique property
arising from the coupled radioactive decay of 238U
(_ 206Pb) and 235U (_ 207Pb) locks into the lead
isotopes a rigid relationship with time. Thus, if the
'primary' array consisted entirely of samples that
grew to their present isotopic composition due to
variable UlPb from one originally isotopically homo-
geneous source, a best-fit line through the array
could be interpreted as a simple secondary isochron.
The 207PbPo4Pb us 206Pb/
204
Pb slope of that line can
then be used to calculate an average source age, T
1
,
of 3800 Ma, assuming a mineralization age, T
2
, of
1400 Ma. However, the entire array is not likely to
be comprised of lead derived from a single source,
and caution demands that Tl not be uncritically
accepted as a meaningful age estimate of sediment
provenance.
LeCouteur (1973) argued persuasively for at
least two components of lead to explain the overall
isotopic pattern found throughout the Belt basin. It
is doubtful that the proposed marine and continen-
tal components would bear so simple an age rela-
tionship as is implicit in a secondary isochron. If,
instead, the 'primary' array is interpreted as a mix-
ing line, an age assessment of the high 207PbPo4Pb
samples derives more from a comparison with the
general behavior of terrestrial lead than from an
intrinsic attribute of the isotopic ratio itself.
Several models to explain isotopic evolution of lead
among major crustal and mantle reservoirs have
been proposed (Russell, 1972; Stacey and Kramers,
1975; Doe and Zartman, 1979). All of these models
recognize that lead lying distinctly above some
average growth curve acquires this isotopic com-
position by residing in a high p environment early
in the Earth's history when the shorter half-lived
235U was considerably more abundant.
An extensive database now exists to uniquely
associate that high 11 environment with the upper
crust and, furthermore, to relate the size of the ano-
maly to the upper crustal age (Stacey and Zartman,
1978; Brevart and others, 1982; Gariepy and others,
1985). Using the approach of Stacey and Kramers
(1975) model ages, T
uc
' of 3560-4110 Ma and cor-
responding values of Puc of 10.46-11.15 were calcula-
ted for the high 207PbPo4Pb samples (Table 1).
Although the model ages are computed mathe-
matically not unlike secondary isochron ages, they
represent not an average source age, as discussed
above, but rather times of departure of individual
samples from the Stacey and Kramers average
growth curve. However, both approaches lead to a
similar conclusion about the great antiquity of the
high 207Pb/
2o4
Pb lead. Although the Wyoming
province records a major tectonic episode at 2900-
2700 Ma, evidence for substantially older rocks and
some component of recycled crust in the late
Archean granites has been reported (Wooden and
others, 1988; Peterman and Futa, 1987; Stevenson
and others, 1988). At this preliminary stage of
investigation, no identification of sediment or fluid
provenance that would directly relate the observed
lead isotopic ratios of the anomalous samples to a
specific source rock has been attempted. Certainly
this additional evidence of continental crust with an
age > 3500 Ma stimulates the search for its
existence adjacent to the Belt basin.
Nd-isotope studies of the Belt Supergroup by
Frost and O'Nions (1984), Frost and Winston (1987),
and Burwash and others (1988) assign average crus-
tal residence ages to each analyzed sample, which
also serve as an estimate of the proportion of older
components present in the sedimentary rocks. Per-
haps not surprisingly, their conclusions are quite
similar to those derived from the lead isotopes. A
708
R. E. Zartman
sample of conglomeratic LaHood Formation - the
stratigraphic unit with the highest 207PbFo4Pb -
also gives the oldest Nd crustal residence age (2900-
3800 Ma, depending on model assumptions; Frost
and Winston, 1987). Interestingly, quartz arenites
of the Fort Steele Formation at the base of the
Canadian Purcell Supergroup, which were sampled
just 25 km east of Kimberley, Britich Columbia, also
have Archean Nd model ages (Burwash and others,
1988). By contrast, both Pb and Nd isotopes imply
that only a small fraction of the fine-grained
sediments from throughout most of the Belt basin
can be of Archean crustal origin.
CONCLUSIONS
The successful search for Archean crustal-
derived lead in the Belt basin confirms the
usefulness of radiogenic isotopes to identify the
provenances for ore deposits and sedimentary rocks.
A number of conclusions can be drawn from the
cumulative lead isotopic database that now exists,
which serve to summarize our current knowledge of
the basin and to set the stage for future work.
1. Of the two lead-isotope populations found for
galena hosted by Middle Proterozoic Belt Super-
group rocks, the syngenetic to diagenetic Precam-
brian variety (commonly referred to as Coeur
d'Alene-type) of lead offers important insight into
the provenance(s) for Belt sedimentation.
2. Most of the Belt basin is characterized by Coeur
d'Alene-type lead of rather uniform isotopic compo-
sition. The minor observed variability in isotope
ratios can be related to: (a) the acquisition of a
radiogenic component during subsequent remobi-
lization of the original stratabound occurrences, and
(b) dispersion of the original lead isotopic com-
position along a 'primary' array. This 'primary'
array has been interpreted to be a short mixing line
between 'normal' marine and continental com-
ponents contributing to sedimentation within the
Belt basin.
3. The steep slope of the 'primary' array on a
207Pb/
204
Pb us 206Pb/
204
Pb diagram can be translated
into information on the average age of the con-
tinental component. Based on only the short mixing
line defined by published analyses, earlier workers
speculated on the presence of Archean continental
crustal, in addition to dominantly 'normal' marine,
lead within some of the Coeur d'Alene-type deposits.
Little was known, however, either about the
geographic and stratigraphic distribution of such an
ancient component in Belt rocks or about the
provenance(s) from which it was derived.
4. A search to find occurrences of lead with a more
distinctive signature of an ancient continental com-
ponent focused on the Helena Embayment - a
southeastern projection of the Belt basin having
close sedimentologic and tectonic association with
adjacent Archean crystalline rocks. Herein is re-
ported the discovery of three unequivocal occur-
rences of such ore deposits in the Highland Moun-
tains, Bridger Range, and Little Belt Mountains.
Although the host rocks for these distinctively high
207Pb/
204
Pb lead deposits include both the LaHood
and Newland Formations, only the conglomeratic
LaHood Formation has been shown to bear a similar
isotopic signature of an ancient provenance.
Samples of fine-grained Newland and Greyson
Formations yielded isotopic compositions identical
with the 'normal' marine lead found in the prepon-
derance of Coeur d'Alene-type deposits.
5. The general similarities in lead isotopic be-
havior of ore deposits and sedimentary rocks strong-
ly suggest that the mineralizing fluids originated as
interstitial solution, which interacted and isotopi-
cally equilibrated with the enclosing sediments.
Subsequent redeposition of the ore at a different
stratigraphic level could account for the present
mismatch in isotopic signature between some ore
deposits and their new host rock.
6. The model ages of Stacey and Kramers (1975)
were calculated for each of the newly analyzed
samples, resulting in upper crustal residence ages of
3570 to 4090 Ma and 11 values of 10.46 to 11.15. This
evidence of an unexpectedly old component of con-
tinental crust within certain proximal sediments of
the Belt basin invites further geochronologic inves-
tigation of the adjacent Archean Wyoming province.
ACKNOWLEDGMENTS
I am particularly thankful to the donors of the
galena and rock samples that revealed the high
207Pb/
204
Pb encountered in this study. Donald A.
Hull, formerly of Homestake Mineral Development
Company, provided drill core containing sulfides
hosted by the Newland Formation in the Highland
Mountains, Silver Bow County. Ray Womack,
Womack and Associates, Inc., provided galena from
a shear zone of the Cross Range fault cutting the
LaHood Formation in the Bridger Range, Gallatin
Archean crustal lead in the Helena Embayment 709
County. Juergen Schieber, University of Texas at
Arlington, provided galena hand-picked from drill
core of the Newland Formation in the Little Belt
Mountains, Meagher County. Rock samples of the
LaHood, Newland, and Greyson Formations from
the Highland Mountains, Silver Bow County, were
contributed by J. Michael O'Neill of the U.S.
Geological Survey; those of the LaHood Formation
from Jefferson Canyon, Jefferson County, were
collected by myself at the stop described by Lewis
(1988).
The chemistry and mass spectrometry for
making the isotopic measurements was performed
by Maryse H. Delevaux and Loretta M. Kwak.
Discussions with U.S. Geological Survey colleagues,
especially John S. Stacey, Kenneth R. Ludwig, J.
Michael O'Neill, Jack E. Harrison, and Joseph L.
Wooden, have been invaluable to my interpretation
of the lead isotope systematics and understanding of
the geology of Belt Supergroup and Archean
Wyoming province rocks. An earlier version of the
manuscript was reviewed by Richard Berg and
Ronald Burwash, whose helpful comments are
appreciated and have been adopted into this final
draft.
REFERENCES
BREVART, 0., B. DUPRE, and C. J. ALLEGRE, 1982,
Metallogenic provinces and the remobilization process
studied by lead isotopes: Lead-zinc ore deposits from the
Southern Massif Central, France: Economic Geology, v. 77,
p.564-575.
BURWASH, R. A., P. A. CAVELL, and E. J. BURWASH,
1988, Source terranes for Proterozoic sedimentary rocks in
southern British Columbia: Nd isotopic and petrographic
evidence: Canadian Journal of Earth Sciences, v. 25, p.
824-832.
DOE, B. R., and R. E. ZARTMAN, 1979, Plumbotectonics I:
The Phanerozoic; pp. 22-70 in H. L. Barnes (ed.), Geo-
chemistry of Hydrothermal Ore Deposits, 2nd Edition:
Wiley Interscience, New York, New York, USA, 798 p.
FROST, C. D., and R. K. O'NIONS, 1984, Nd evidence for
Proterozoic crustal development in the Belt-Purcell
Supergroup: Nature, v. 312, p. 53-56.
__ , and D. WINSTON, 1987, Nd isotope systematics of
coarse- and fine-grained sediments: Examples from the
Middle Proterozoic Belt-Purcell Supergroup: Journal of
Geology, v. 95, p. 309-327.
GARIEPY, C., C. J., ALLEGRE, and R. H. XU, 1985, The Pb-
isotope geochemistry of granitoids from the Himalaya-
Tibet collision zone: Implications for crustal evolution:
Earth and Planetary Science Letters, v. 74, p. 220-234.
GODWIN, C. I., J. E. GABITES, and A. ANDREW, 1988,
Leadtable: A Galena Lead Isotope Data Base for the
Canadian Cordillera: British Columbia Ministry of
Energy, Mines and Petroleum Resources, Paper 1988-4,188
p.
LECOUTEUR, P. C., 1973, A Study of Lead Isotopes
From Mineral Deposits in Southeastern British
Columbia and in the AnvU Range, Yukon Territory:
PhD dissertation, University of British Columbia, Van-
couver, British Columbia, Canada, 142 p.
LEWIS, S. E., 1988, Fieldguide to mesoscopic features in
the LaHood Formation, Jefferson River Canyon area,
southwestern Montana; pp. 155-158 in S. E. Lewis and R.
B. Berg (eds.), Precambrian and Mesozoic Plate Mar-
gins: Montana, Idaho and Wyoming: Montana Bureau
of Mines and Geology, Special Publication 96, 195 p.
MARVIN, R. F., and R. E. ZARTMAN, 1984, A Tabulation
of Lead Isotopic Ratios for Lead Minerals from Mines
and Prospects in the Belt-Purcell Basin in the United
States and Canada: U.S. Geological Survey Open File
Report 84-210, 15 p.
MUDGE, M. R., R. L. ERICKSON, and D. KLEINKOFF, 1968,
Reconnaissance Geology, Geophysics, and Geochem-
istry of the Southeastern Part of the Lewis and Clark
Range, Montana: U.S. Geological Survey Bulletin 1252-
E,35p.
OBRADOVICH, J. D., R. E. ZARTMAN, and Z. E. PETER-
MAN, 1984, Update of geochronology of the Belt Super-
group; p. 82-84 in S. W. Hobbs (ed.), The Belt: Belt Sym-
posium II (1983), Abstracts with Summaries: Montana
Bureau of Mines and Geology, Special Publication 90, 117
p.
PETERMAN, Z. E., and K. FUTA, 1987, Is the Archean
Wyoming province exotic to the Superior craton? Evidence
from Sm-Nd model ages of basement cores [abstract]:
Geological Society of America Abstracts with Programs, v.
19, p. 803.
RUSSELL, R. D., 1972, Evolutionary model for lead isotopes
in conformable ores and in oceanic volcanics: Reviews in
Geophysics and Space Physics, v. 10, p. 529-549.
SCHMIDT, C. J., and J. M. GARIHAN, 1986, Middle Pro-
terozoic and Laramide tectonic activity along the southern
margin of the Belt Basin in the Belt Supergroup; pp. 217-
235 in S. M. Roberts (ed.), Belt Supergroup: A Guide to
Proterozoic Rocks of Western Montana and Adjacent
Areas: Montana Bureau of Mines and Geology, Special
Publication 94, 311 p.
710 R. E. Zartman
STACEY, J. S., and J. D. KRAMERS, 1975, Approximation
of terrestrial lead isotope evolution by a two-stage model:
Earth and Planetary Science Letters, v. 26, p. 207-221.
__ , and R. E. ZARTMAN, 1978, Lead and strontium
isotopic study of igneous rocks and ores, Gold Hill district,
Utah: Utah Geology, v. 5, p. 1-15.
STEVENSON, R. K., P. J. PATCHETT, D. BRIDGWATER, A.
KRONER, and P. A. MUELLER, 1988, Evolution of Archean
continental crust implied by Hf isotopes in sedimentary
zircons [abstract): EOS, Transactions of the American
Geophysical Union, v. 69, p. 517.
WINSTON, D., 1986, Stratigraphic correlation and
nomenclature of the Middle Proterozoic Belt Supergroup,
Montana, Idaho and Washington; pp. 69-84 in S. M. Ro-
berts (ed.), Belt Supergroup: A Guide to Proterozoic
Rocks of Western Montana and Adjacent Areas: Mon-
tana Bureau of Mines and Geology, Special Publication 94,
311p.
WOODEN, J. L., P. A. MUELLER, and D. W. MOGK, 1988,
Review of the geochemistry and geochronology of the
Archean rocks of the northern part of the Wyoming
Promince; pp. 383-410 in W. G. Ernst (ed.), Metamor-
phism and Crustal Evolution of the Western United
States: Prentice-Hall, New York, New York, USA, 1153 p.
ZARTMAN, R. E., 1974, Lead isotopic provinces in the
Cordillera of the western United States and their geologic
significance: Economic Geology, v. 69, p. 792-805.
__ , and J. S. STACEY, 1971, Lead isotopes and minera-
lization ages in Belt Supergroup rocks, northwestern
Montana and northern Idaho: Economic Geology, v. 66, p.
849-860.
INTERNATIONAL BASEMENT TECTONICS ASSOCIATION PUBLICATION NO.8
Tectonic development of base-metal and barite-vein deposits
associated with the early Mesozoic basins of
eastern North America
GILPIN R. ROBINSON, Jr.
U.S. Geological Survey, MS 913, National Center, Reston, Virginia 22092, USA
(received December 31,1988, revision accepted July 12, 1989)
ABSTRACT
Epithermal base-metal and barite-vein deposits are common in and near the margins of rift basins in
extensional continental settings, and such veins are structurally, temporally, and almost certainly genetically
associated with these rift basins. Approximately 100 vein occurrences of this type occur near early Mesozoic
basins ofthe eastern United States. The veins appear to form as a result of transport of moderate temperature
(100-250C) brines from within the basins and adjacent basement rocks to shallow sites of mineral pre-
cipitation. Mineral deposition appears due to cooling, sulfate reduction, wall-rock reaction, and possibly fluid
mixing, with the relative importance of these processes varying in different deposits. Saline fluids of
appropriate chemistry and temperature to form vein mineralization occur in the deep portions of basins and
adjacent basement underlying the basins. Early Mesozoic basins filled with red-bed sediments may be
important as recharge areas of saline brines and leachable metals. Pre-existing fractures and faults are
important in localizing vein development, and seismic activity may induce the movement of subsurface
hydrothermal fluids. Cycling of stress and fluid pressure during faulting can move significant quantities of
hot, deep-seated fluids into the upper crust and provides a mechanism for rapid mixing of fluids that may lead
to mineral deposition. A period of fluid migration appears to occur late in the history of basin development
due to the dynamic and changing stress regime developed during the tectonic transition from continental
rifting to sea-floor spreading. Stratabound base-metal deposits that also form by transport of metal-bearing
fluids may occur in the vicinity of these vein deposits.
INTRODUCTION
Base-metal and barite-vein deposits are com-
mon in or near the margins of rift basins in con-
tinental settings worldwide. Approximately 100
such vein occurrences are located in and near the
Fundy (Nova Scotia, Canada), Hartford (New
Hampshire-Massachusetts-Connecticut) , Newark
(New Jersey-Pennsylvania), and Culpeper (Mary-
land-Virginia) early Mesozoic basins of eastern
North America (Robinson and Woodruff, 1988).
Locations of 14 districts containing most of these
vein deposits in the eastern United States are shown
in Figure 1, and a summary of vein characteristics
is given in Table 1. Base-metal and barite veins
also occur within and bordering Late Proterozoic rift
basins in the Lake Superior region of Canada
(Franklin and Mitchell, 1977, Haynes, 1988), the
Cretaceous Benue trough of Nigeria (Olade and
Morton, 1985; Maurin and Lancelot, 1987), the
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason), pp. 711-725. Kluwer Academic Publishers, Dordrecht, The
Netherlands, 1992.
711
712 G. R. Robinson
Paleozoic Midland Valley of Scotland (Jassim and
others, 1983; Patrick and others, 1983; Samson and
Banks, 1988), rift basins associated with the Rhine
Graben system of Germany (Hoffmann and Bau-
mann, 1984), and in pull-apart basins and linear
fault zones along the Grenville front in Scandinavia
(Andreasson and others, 1987) and south of the
Rhodesian craton in South Africa (Reimer, 1986).
Spatial association of base-metal and barite-vein
deposits with rift basins in extensional continental
settings implies their genetic connection, which is
discussed here as it applies to early Mesozoic rift
Figure 1. Map showing distribution of mine districts
containing base-metal and barite-vein deposits asso-
ciated with early Mesozoic basins of Massachusetts,
Connecticut, New York, New Jersey, Pennsylvania,
Maryland, and Virginia. Early Mesozoic basins are
indicated by stippled pattern. Distribution of mine
districts and deposits indicated by solid black pattern,
with identifying number:
1. Cheshire district, New Hampshire
2. Leverett district, Massachusetts
3. Whatley-Hatfield district, Massachusetts
4. Loudville district, Massachusetts
5. New Britain district (north) and Bristol
deposit (south), Connecticut
6. Cheshire-Jinny Hill district, Conneticut
7. Hopewell-Solebury district, New J ersey-
Pennsylvania
8. New Galena deposit, Pennsylvania
9. Audubon (north), and Phoenixville (south)
districts, Pennsylvania
10. Safe Harbor deposit, Pennsylvania
11. Boyds barite occurrence, Maryland
12. Cedar Run district, Virginia
13. St. Stephens deposit, Virginia
14. Gear-Kemper district, Virginia
basins of eastern North America, beginning with a
summary of vein characteristics.
CHARACTERISTICS OF BASE-METAL AND
BARITE-VEIN DEPOSITS
An inventory of"'" 100 vein occurrences associa-
ted with early Mesozoic basins of eastern North
America and a summary of vein characteristics is
given by Robinson and Woodruff (1988). Additional
reasonably detailed descriptions of the veins include


,j'"

\
\,/
/
2
3
HART'FORD
BASI
I 0
.,
1CO MUS
\
5
I
i
I
I
I
\/
0
.,
100 KtOMfT9lS

\ ",,1 "'O'?-

...
,
,../
-

/

NEWARK
,../
BASIN
"\
\<&

",,,0,,\1>'
., tJ.,f
GE TYSBUAG
BASIN
,../
Base-metal and barite-vein deposits, early Mesozoic basins of eastern North America 713
Bannerman (1941) - fluorite deposits of the Che-
shire district, New Hampshire; Emerson (1898a,b)
- lead- and barite-vein deposits in the Turners
Falls area, Leverett, Whatley, Hatfield, and Loud-
ville districts, Massachusetts; Bateman (1923) -
Bristol copper mine, Connecticut; Fritts (1962) -
barite-vein deposits of the Cheshire district, Con-
necticut; Smith (1977) - lead-, zinc- and copper-
vein deposits of New Galena, Audubon, and Phoe-
nixville districts, Pennsylvania; and Edmundson
(1938) - barite-vein deposits in the Culpeper basin,
Virginia.
Individual veins range from a few centimeters
to ten meters in width, and groups of veins extend
laterally for as much as a few kilometers. Minera-
lized intervals extend in some cases to depths of
> 500 m, such as in the Magnet Cove barite-sulfide
deposit that borders the eastern margin of the
Fundy basin (Boyle, 1962; Boyle and Jambor, 1966).
This deposit is the largest base-metal and barite-
vein deposit in the province. Through 1964, pro-
duction from this deposit totalled = 3.3 million tons
of barite and 16.7 thousand tons of sulfide concen-
trates, with a total metal content of 1.3 million
pounds of copper, 9 million pounds of lead, 4.3
million pounds of zinc, and 1.3 million ounces of
silver (Boyle and Jambor, 1966, p. 396).
Vein mineralogy typically consists of quartz,
galena, sphalerite, and (or) chalcopyrite, with minor
pyrite and tetrahedrite and local areas of barite,
fluorite, and (or) carbonate minerals and bitumen.
Sulfide minerals in the veins characteristically have
high silver contents (galena typically 250-700 ppm
Ag; bornite, 500 ppm Ag; and chalcocite, 1500 ppm
Ag), and sphalerite may contain high cadmium,
germanium, and gallium concentrations (Robinson
and Woodruff, 1988). Silver is typically a minor
component in abundant sulfide minerals such as
galena and is locally present as a constituent in
trace sulfosalts and sulfides like tetrahedrite and
acanthite. High silver contents make the veins
attractive exploration targets in places. Some veins
(New Britain district, Connecticut) contain minor
bitumen (Ryan, 1984, 1986; Gray, 1982). Quartz-
carbonate veins (New Britain area) subparallel to
metal-bearing veins contain liquid hydrocarbons as
well as bitumen (Pratt and Burruss, 1988).
Veins occur in both basin sediments and adja-
cent basement. Vein-related alteration of quartzo-
feldspathic gneiss, sandstone, siltstone, or mudstone
is not evident, but subtle modal increases in albite,
chlorite, muscovite, and (or) other layer silicates
may be present against the veins (Robinson and
Woodruff, 1988). Wallrock alteration is intense and
obvious only when veins are in basalt or diabase.
These mafic wallrocks are bleached and altered to 1
m from vein walls, although alteration haloes of 5-
10 cm are more typical (Ryan, 1986; Robinson and
Woodruff, 1988). Albite, chlorite, and (or) mont-
morillonite are the principal alteration minerals. A
slight to moderate degree of proton (hydrogen)
metasomatism is required to preserve charge
balance in altered material, and the intensity of this
proton metasomatism generally increases toward
vein contacts. These alteration patterns are consis-
tent with slightly acidic, oxidized fluids equilibrated
with sericite + feldspar and (or) clay assemblages.
Fluid inclusion studies of quartz, sphalerite,
and carbonates from veins (New Britain district and
Jinny Hill deposit, Connecticut - Ryan, 1984,
1986; and Audubon district, Phoenixville district,
and New Galena deposit - Lawler, 1981) indicate
fluid inclusion homogenization temperatures in the
range 90-210C (Figure 2). Source fluids are brines
of moderate salinity (10-16 equivalent wt% NaCI;
=2 molar NaCI solutions); fluid inclusions at one
occurrence in Connecticut have a high CO
2
content
(Ryan, 1986). A pressure correction of = 10-
NEW BRITAIN
JINNY HILL
NEW GALENA
AUDUBON
PHOENIXVILLE
MESOZOIC BASE METAL VEINS
FLUID INCLUSION HOMOGENIZATION
TEMPERATURE
f---f=====::;-.----j
I f------4
I
H
SALINITY
H
100 120 140 160 180 200 10 15
T (0C) WT. PERCENT
NaCI EQUIVALENT
Figure 2. Range of fluid-inclusion homogenization-
temperatures and fluid salinity determined by freezing
temperature from inclusions in quartz and sphalerite from
vein deposits associated with early Mesozoic basins of the
eastern United States indicated by barred lines. Data are
from three deposits in New Britain district, Connecticut;
Jinny Hill deposit, Connecticut; three dump areas from
New Galena deposit, Pennsylvania; five mines from
Audubon district, Pennsylvaina; and three mines from
Phoenixville district, Pennsylvania. Data for Connecticut
deposits from Ryan (1984, 1986) and for Pennsylvania
deposits from Lawler (1981).
T
a
b
l
e

1
.

B
a
s
e
-
m
e
t
a
l

a
n
d

b
a
r
i
t
e

v
e
i
n
s

a
s
s
o
c
i
a
t
e
d

w
i
t
h

e
a
r
l
y

M
e
s
o
z
o
i
c

b
a
s
i
n
s

o
f
t
h
e

e
a
s
t
e
r
n

U
n
i
t
e
d

S
t
a
t
e
s
,

b
y

g
e
o
g
r
a
p
h
i
c

l
o
c
a
t
i
o
n
.

D
e
p
o
s
i
t

H
o
s
t

R
o
c
k

H
o
s
t

R
o
c
k

C
o
m
m
o
d
i
t
y

D
e
p
o
s
i
t

N
a
m
e

a
n
d

L
o
c
a
t
i
o
n
,

W
i
t
h

N
u
m
b
e
r

T
y
p
e

L
i
t
h
o
l
o
g
y

A
g
e

P
r
i
n
c
i
p
a
l

A
c
c
e
s
s
o
r
y

A
.

F
u
n
d
y

b
a
s
i
n

&

v
i
c
i
n
i
t
y
,

N
o
v
i
a

S
c
o
t
i
a
,

C
a
n
a
d
a

W
a
l
t
o
n
-
C
h
e
v
e
r
e
,

N
o
v
a

S
c
o
t
i
a

d
i
s
t
r
i
c
t

M
a
g
n
e
t

C
o
v
e
,

H
a
n
t
s
,

N
S

m

s
a
n
d
s
t
o
n
e
/
s
h
a
l
e

C
a
r
b
o
n
i
f
e
r
o
u
s

B
a
,
P
b

Z
n
,
C
u
,
A
g

B
.

H
a
r
t
f
o
r
d
-
D
e
e
r
f
i
e
l
d

b
a
s
i
n

&

v
i
c
i
n
i
t
y
,

N
e
w

H
a
m
p
s
h
i
r
e
,

M
a
s
s
a
c
h
u
s
e
t
t
s
,

a
n
d

C
o
n
n
e
c
t
i
c
u
t

1
.

C
h
e
s
h
i
r
e

N
H

d
i
s
t
r
i
c
t

7
m
,
1
p
,
2
0

g
n
e
i
s
s

P
a
l
e
o
z
o
i
c

F
,
P
b

B
a
,
C
u
,
A
g

C
h
e
s
h
i
r
e

C
o
.
,

N
H

T
u
r
n
e
r
s

F
a
l
l
s
,

F
r
a
n
k
l
i
n

C
o
.
,

M
A

2
0

s
a
n
d
s
t
o
n
e

J
u
r
a
s
s
i
c

C
u

B
a

2
.

L
e
v
e
r
e
t
t

d
i
s
t
r
i
c
t

1
m
,

4
0

g
r
a
n
i
t
e

D
e
v
o
n
i
a
n

P
b

B
a
,
C
u

F
r
a
n
k
l
i
n

C
o
.
,

M
A

3
.

W
h
a
t
l
e
y
-
H
a
t
f
i
e
l
d

d
i
s
t
r
i
c
t

5
m
,
3
0

s
c
h
i
s
t

S
i
l
u
r
i
a
n

P
b

B
a

F
r
a
n
k
l
i
n

C
o
.
,

M
A

g
r
a
n
i
t
e

D
e
v
o
n
i
a
n

4
.

L
o
u
d
v
i
l
l
e

d
i
s
t
r
i
c
t

8
m

g
r
a
n
i
t
e

D
e
v
o
n
i
a
n

P
b
,
B
a

A
g
,
Z
n
,
C
u

H
a
m
p
s
h
i
r
e

C
o
.
,

M
A

5
.

N
e
w

B
r
i
t
a
i
n

d
i
s
t
r
i
c
t

3
m
,

9
0

s
i
l
t
s
t
o
n
e
/
b
a
s
a
l
t

J
u
r
a
s
s
i
c

C
U
,
Z
n

B
a
,
P
b

H
a
r
t
f
o
r
d

C
o
.
,

C
T

5
.

B
r
i
s
t
o
l

m
i
n
e
,

H
a
r
t
f
o
r
d

C
o
.
,

C
T

m

s
a
n
d
s
t
o
n
e

T
r
i
a
s
s
i
c

C
u

A
g
,
U

6
.

C
h
e
s
h
i
r
e
.
.
J
i
n
n
y

H
i
l
l

d
i
s
t
r
i
c
t

9
m
,

2
p
,

1
0

s
a
n
d
s
t
o
n
e

T
r
i
a
s
s
i
c

B
a
,
C
u

P
b
,
A
g

N
e
w

H
a
v
e
n

C
o
.
,

C
T

C
.

N
e
w
a
r
k

b
a
s
i
n

&

v
i
c
i
n
i
t
y
.

N
e
w

J
e
r
s
e
y

a
n
d

P
e
n
n
s
y
l
v
a
n
i
a

M
e
n
l
o

P
a
r
k

m
i
n
e
,

M
i
d
d
l
e
s
e
x

C
o
.
,

N
J

m

s
i
l
t
s
t
o
n
e

T
r
i
a
s
s
i
c

C
u

7
.

H
o
p
e
w
e
l
l
-
S
o
l
e
b
u
r
y

d
i
s
t
r
i
c
t

1
m
,

2
p
,

2
0

d
i
a
b
a
s
e

J
u
r
a
s
s
i
c

B
a
,
C
u

M
e
r
c
e
r

C
o
.
,

N
J
,

a
n
d

B
u
c
k
s

C
o
.
,

P
A

s
i
l
t
s
t
o
n
e

T
r
i
a
s
s
i
c

8
.

N
e
w

G
a
l
e
n
a
,

B
u
c
k
s

C
o
.
,

P
A

m

b
l
a
c
k

s
h
a
l
e

T
r
i
a
s
s
i
c

P
b
,
Z
n

A
g
,
A
u
,
C
u

S
c
h
u
y
l
k
i
l
l

F
a
l
l
s
,

P
h
i
l
a
d
e
l
p
h
i
a

C
o
.
,

P
A

0

T
r
i
a
s
s
i
c

P
b

Z
n

-
-
-
-
-
_
.
_
-
-


(
c
o
n
t
i
n
u
e
d
)

-
l

>
-
'

*
'
"

o

;
:
0

8
1

0
"

S


m

o

i
:
j

T
a
b
l
e

1

C
o
n
t
i
n
u
e
d

D
e
p
o
s
i
t

H
o
s
t

R
o
c
k

H
o
s
t

R
o
c
k

C
o
m
m
o
d
i
t
y

D
e
p
o
s
i
t

N
a
m
e

a
n
d

L
o
c
a
t
i
o
n
,

W
i
t
h

N
u
m
b
e
r

T
y
p
e

L
i
t
h
o
l
o
g
y

A
g
e

P
r
i
n
c
i
p
a
l

A
c
c
e
s
s
o
r
y

C
.

N
e
w
a
r
k

b
a
s
i
n

&

v
i
c
i
n
i
t
y
,

N
e
w

J
e
r
s
e
y

a
n
d

P
e
n
n
s
y
l
v
a
n
i
a

(
c
o
n
t
i
n
u
e
d
)

9
.

A
u
d
u
b
o
n

d
i
s
t
r
i
c
t

4
m
,

2
p
,

1
0

s
a
n
d
s
t
o
n
e

T
r
i
a
s
s
i
c

P
b
,
C
u

Z
n
,
A
g

M
o
n
t
g
o
m
e
r
y

C
o
.
,

P
A

P
h
o
e
n
i
x
v
i
l
l
e

d
i
s
t
r
i
c
t

1
0
m
,

1
p
,

1
0

g
n
e
i
s
s

P
r
e
c
a
m
b
r
i
a
n

P
b
,
Z
n

C
U
,
B
a
,
A
g

C
h
e
s
t
e
r

C
o
.
,

P
A

s
a
n
d
s
t
o
n
e

T
r
i
a
s
s
i
c

D
.

G
e
t
t
y
s
b
u
r
g

b
a
s
i
n

&

v
i
c
i
n
i
t
y
,

P
e
n
n
s
y
l
v
a
n
i
a

I

1
0
.

S
a
f
e

H
a
r
b
o
r
,

L
a
s
n
c
a
s
t
e
r

C
o
.
,

P
A

0

d
o
l
o
m
i
t
e

C
a
m
b
r
i
a
n

Z
n
,
B
a

I

I

E
.

C
u
l
p
e
p
e
r

b
a
s
i
n

&

v
i
c
i
n
i
t
y
,

M
a
r
y
l
a
n
d

a
n
d

I

V
i
r
g
i
n
i
a

I

1
1
.

B
o
y
d
s
,

M
o
n
t
g
o
m
e
r
y

C
o
.
,

M
D

0

d
i
a
b
a
s
e

J
u
r
a
s
s
i
c

B
a

1
2
.

C
e
d
a
r

R
u
n

d
i
s
t
r
i
c
t

1
m
,
1
0

s
i
l
t
s
t
o
n
e

T
r
i
a
s
s
i
c

B
a

i

P
r
i
n
c
e

W
i
l
l
i
a
m

C
o
.
,

V
A

I

1
3
.

S
t
.

S
t
e
p
h
e
n
s
,

F
a
u
q
u
i
e
r

C
o
.
,

V
A

m

s
i
l
t
s
t
o
n
e

T
r
i
a
s
s
i
c

B
a

,

1
4
.

G
e
a
r
-
K
e
m
p
e
r

d
i
s
t
r
i
c
t

2
p
,
1
0

s
i
l
t
s
t
o
n
e

T
r
i
a
s
s
i
c

B
a

F
a
u
q
u
i
e
r

C
o
.
,

V
A

F
.

O
t
h
e
r

a
r
e
a
s
,

V
i
r
g
i
n
i
a

A
l
b
e
r
m
a
r
l
e

m
i
n
e
,

A
l
b
e
r
m
a
r
l
e

C
o
.
,

V
A

m

g
n
e
i
s
s

C
a
m
b
r
i
a
n

Z
n
,
P
b

A
g

S
c
o
t
t
s
v
i
l
l
e
,

A
l
b
e
r
m
a
r
l
e

C
o
.
,

V
A

0

s
a
n
d
s
t
o
n
e

T
r
i
a
s
s
i
c

B
a

D
o
l
a
n

p
r
o
p
e
r
t
y
,

N
e
l
s
o
n

C
o
.
,

V
A

0

s
h
a
l
e

T
r
i
a
s
s
i
c

C
u

G
.

D
e
e
p

R
i
v
e
r

b
a
s
i
n

&

v
i
c
i
n
i
t
y
,

N
o
r
t
h

C
a
r
o
l
i
n
a

H
a
r
r
i
s
v
i
l
l
e
,

M
o
n
t
g
o
m
e
r
y

C
o
.
,

N
C

0

d
i
a
b
a
s
e

J
u
r
a
s
s
i
c

B
a

-
-
_
.
_
-
-
-
-
-
-
-
-
-
-
-
-
-
-
L
o
c
a
t
i
o
n

n
u
m
b
e
r

r
e
f
e
r
s

t
o

t
h
e

n
u
m
b
e
r
e
d

a
r
e
a
s

s
h
o
w
n

o
n

F
i
g
u
r
e

1
.

D
e
p
o
s
i
t
s

w
i
t
h
o
u
t

n
u
m
b
e
r
s

a
r
e

n
o
t

s
h
o
w
n

o
n

F
i
g
u
r
e

1

a
n
d

t
h
e
y

d
o

n
o
t

n
e
c
e
s
s
a
r
i
l
y

o
c
c
u
r

i
n

t
h
e

"
d
i
s
t
r
i
c
t
s
"

t
h
e
y

a
r
e

c
l
o
s
e
s
t

t
o

i
n

t
h
e

t
a
b
l
e
.

T
h
e

d
e
s
i
g
n
a
t
i
o
n
s

"
m
,
"

"
p
,
"

a
n
d

"
0
"

i
n
d
i
c
a
t
e

d
e
p
o
s
i
t

t
y
p
e

-
r
e
s
p
e
c
t
i
v
e
l
y
,

m
i
n
e
,

p
r
o
s
p
e
c
t
,

o
r

m
i
n
e
r
a
l

o
c
c
u
r
r
e
n
c
e
,

a
n
d

t
h
e

f
i
g
u
r
e

p
r
e
c
e
d
i
n
g

d
e
p
o
s
i
t

t
y
p
e

i
n
d
i
c
a
t
e
s

t
h
e

e
s
t
i
m
a
t
e
d

n
u
m
b
e
r

o
f

e
a
c
h

t
y
p
e

i
n

t
h
e

d
i
s
t
r
i
c
t
.

A
d
a
p
t
e
d

f
r
o
m

R
o
b
i
n
s
o
n

a
n
d

W
o
o
d
r
u
f
f

(
1
9
8
8
,

T
a
b
l
e

1
A
)
.

t
o

I
l:
>

g
;


e
.


p
.
.
.

C
"

I
l:
>

>
o
j
.
.
.
.
.

.
,
.
.
.

C
D

.
:
:

C
D

.
.
.
.
.

I
:
j


o

f
!
J
.

.
&
r
-
C
D

I
l:
>

:
!
.

'
<


C
D

m

l
a

8
.

(
'
)

C
"

I
l:
>

m

5
'

m

S
.

C
D

I
l:
>


C
D


Z

o


>

S

C
D

:
:
t

(
'
)

I
l:
>

"
'
l

.
.
.
.
.
.

0
1

716 G. R. Robinson
11 C/100 bars trapping pressure for fluids in this
salinity range (Potter and Brown, 1977; Haas, 1976)
must be applied to fluid-inclusion homogenization-
temperatures to determine the trapping tempera-
ture (Roedder and Bodnar, 1980). The fluid-
trapping pressure is probably somewhere between
lithostatic and hydrostatic in this vein situation
(Roedder and Bodnar, 1980) and, for shallow open-
vein development, trapping pressures are probably
close to hydrostatic pressure. Using generous un-
certainty estimates regarding fluid-inclusion trap-
ping pressure, a correction of + 10C to + 40C (cor-
responding to estimated depths of 3000-4000 m for
Pennsylvania veins and < 2500 m for Connecticut
veins, and estimated hydrostatic pressures in the
range of 100-400 bars) is applied to fluid-inclusion
homogenization-temperatures to estimate trapping
temperature. Using this correction, a fluid-inclu-
sion trapping-temperature range of 100-250C is
calculated for these vein deposits.
Limited lead isotope data on galenas from some
veins (Hartford basin - Parnell, 1986) indicate that
these galenas are moderately radiogenic, with iso-
tope ratios of 18.5-18.6 for 206Pb/
204
Pb, 15.64-15.66
for 207PbPo4Pb, and 38.4-38.6 for 208Pb/
204
Pb. Their
moderately high 207Pb/
204
Pb ratios suggest lead deri-
vation from predominantly continental sources with
an enrichment in uranium and uranium/thorium
relative to deep crustal or mantle sources. Model
lead ages fall in the range of 15020 Ma. The
easily mobilized nature of uranium and much of its
daughter lead in first-cycle red-bed sediments of
continental basins (Doe and Zartman, 1979) implies
that model lead ages of galenas in these vein
deposits may be younger than the age of deposition.
However, the vein-hosted galenas are only slightly
anomalously radiogenic (J-type lead) relative to the
lead composition of the average orogene of Doe and
Zartman (1979), implying that the model lead ages
may be similar to the formation ages.
Sulfur isotope values for suites of coexisting
minerals from twenty veins from eight deposit dis-
tricts in Massachusetts, Connecticut, Pennsylvania,
and Virginia are reported by Robinson and Wood-
ruff (1988). The results of their sulfur isotope study
(Figure 3) illustrate the range of sulfur isotope
variation of individual mineral samples from the
various deposits and districts. Sulfur isotope varia-
tion in barite samples from all deposits from
Massachusetts to Virginia lie within a relatively
small range of 0
34
S values (+ 10 to + 20%0). In depo-
sits where barite is an abundant mineral, 0
34
S
values for barite generally lie at the lower end of the
range (+ 10 to + 15%0); in deposits where barite is a
minor or paragenetically late mineral, 0
34
S values
for barite are at the higher end (+ 15 to + 20%0).
Districts in proximity (Leverett-Hatfield, Massa-
chuestts; New Britain-Jinny Hill, Connecticut; and
Audubon-Phoenixville, Pennsylvania) have similar
and overlapping 0
34
S values for barite. In hydro-
thermal solutions at temperatures> 100C, little or
no isotope fractionation occurs between aqueous
sulfate and coexisting sulfates (Ohmoto and Rye,
1979); therefore sulfur isotope values for sulfate
minerals represent aqueous sulfate values. These
results imply a relatively uniform regional fluid
source for sulfate sulfur, at least on the scale of
individual basins.
In contrast, sulfur isotope values for coexisting
sulfides from the vein deposits show a considerable
range in 0
34
S values (-17 to + 19%0). For most
deposits, the range in 0
34
S values for individual
sulfide minerals is close to 5%0. In the New Galena
deposit (Pennsylvania), the range of galena 0
34
S
values is 18%0 (-4 to +14%0). This type of large
0
34
S variation in sulfides is characteristic of systems
that have undergone sulfate reduction or have
received fluids from multiple sources or both
(Ohmoto and Lasaga, 1982).
Sulfur isotope values for sulfides in some veins
differ significantly from those of pre-existing sul-
fides in adjacent host rocks, implying that sulfide
phases in host rocks are not a dominant source of
reduced sulfur for the vein mineralization. For
example, a sample of diagenetic pyrite from black
shale at the New Galena deposit, has a 0
34
S value of
-19.29. This value differs significantly from the
range of 0
34
S values (- 4 to + 19%0) of the vein-
hosted sulfides (Figure 3). If this diagenetic pyrite
formed by bacterial reduction of sulfate-bearing
aqueous fluids, and a biogenetic isotopic fractiona-
tion of +4520%0 existed between fluid sulfate and
biogenetic sulfides (Ohmoto and others, 1988), then
the aqueous sulfate source at New Galena would
have had a value of + 25 20%0 0
34
S. This range of
aqueous sulfate values overlaps the measured range
of barite sulfate values, in general agreement with a
regional aqueous sulfate source of + 10 to + 20%0.
Calculated sulfur isotope fractionation tem-
peratures for coexisting sulfides (90-280C) in the
vein samples using equations of Ohmoto and Rye
(1979, their Tables 10-1 and 10-2) are consistent
with temperatures indicated from fluid inclusion
studies (Figure 4), with the exception of two
Base-metal and barite-vein deposits, early Mesozoic basins of eastern North America 717
MESOZOIC BASE-METAL VEINS
Q!l!!!!
LEVERETT
HATFIELD IZlZI2l rn
NEW BRITAIN I2I'N OJ
JINNY HILL
o
m
OJ]
DO
o
NEW GALENA El
AUDUBON
PHOENIXVILLE
VIRGINIA
[2J [2J [2J [2J rn
[2J fSl [2J['lli]
[2J rn2] I2EEl

o
moo
rn
I I
-20 -15 -10 -5 10 15 20
S 34 5
[Zl Galena
El
Pynte
0 Barite
iSJ Sphalerite Other sulfides
OJ Chalcopyrite
Figure 3. Sulfur isotope values for suites of coexisting
minerals from 20 vein deposits in the Leverett and
Whatley-Hatfield districts, Massachusetts; New Britain
district and Jinny Hill deposit, Connecticut; New Galena
deposit, Pennsylvania; and various barite deposits in the
Culpeper Basin in Virginia are shown with mineral types
indicated by symbol. The sulfur isotope values were
measured by L. G. Woodruff at USGS laboratories in
Reston, Virginia.
MESOZOIC BASE-METAL VEINS
SULFUR ISOTOPE FRACTIONATION
DISTRICT
LEVERETT
HATFIELD IZIZI lSI-
NEW BRITAIN f-----1 IZl rn
JINNY HILL H
NEW GALENA
AUDUBON
IZl
PHOENIXVILLE
100 200 300 400 500 600
T (0C)
[2] Galena CD Chalcopyrite 0 Barite
[SJ Sphalerite o Other sulfides 1-------1 Fluid inclusion
Th range
Figure 4. Sulfur isotope fractionation temperatures for
coexisting sulfur-bearing minerals calculated using equa-
tions of Ohmoto and Rye (1979, their Tables 10-1 and 10-
2), modified with revised aqueous sulfate-aqueous sulfide
fractionation curve of Ohmota and Lasaga (1982). Mineral
pairs are indicated by symbol. Fluid-inclusion homogeni-
zation-temperatures (Th) for corresponding samples are
indicated by barred lines.
sphalerite-galena pairs from the New Galena depo-
sit. However, fractionation temperatures calculated
for coexisting sulfide-sulfate mineral pairs using
Ohmoto and Rye's (1979) equations modified to
account for the revised aqueous sulfate-aqueous
sulfide fractionation curve of Ohmoto and Lasaga
(1982), are generally inconsistent with crystalliza-
tion temperatures indicated by fluid inclusion
studies; they typically give nonequilibrium values
of 300 to >600C (Figure 4). This type of non-
equilibrium sulfate-sulfide fractionation is common-
ly observed in low-temperature vein occurrences
(Ohmoto and Lasaga, 1982). Three important
processes that may produce this relationship are:
(1) inheritance of higher-temperature fractionation
characteristics from deeper parts of the hydro-
thermal system; (2) mixing of sulfide-rich with
sulfate-rich solutions near the deposition site; and
(or) (3) partial oxidation or reduction of the fluid
near the deposition site (Ohmota and Lasaga, 1982).
Inheritance of relict high-temperature frac-
tionation characteristics between aqueous sulfate
and aqueous sulfide from deeper and hotter portions
of the hydrothermal system for some vein deposits
may explain the discrepancy between fluid-inclu-
sion filling-temperatures of 200C or less and
calculated sulfate-sulfide isotope fractionation tem-
peratures of ""400C. The districts where inherited
fractionation characteristics may be important are
New Britain (Connecticut) and Audubon and Phoe-
nixville (Pennsylvania), where calculated fractiona-
tion temperatures are ""400C or less. The high-
temperature fractionation characteristics are
preserved by rapid cooling of the fluid, which also
results in quartz, sulfide, and (or) barite deposition
due to their prograde solubility in 2 molar NaCI
solutions with temperature (Holland and Malinin,
1979).
Aqueous sulfate reduction may be an important
sulfide deposition mechanism at the New Galena
deposit. This interpretation is supported by the
large range in 5
34
S values for galena and sphalerite,
the absence of sulfate minerals at the deposit, and
the overlapping range of 5
34
S values for galena and
sphalerite with the range of barite 5
34
S values for
the other deposits. Shales, rich in organic matter in
the mineralized area of the New Galena deposit,
may have acted as the fluid reductant. The large
range in sulfide 5
34
S ( - 8 to + 5%0) from deposits in
the Audubon and Phoenixville district, and the
variable barite sulfate values in the heavy range of
0
34
S values from these deposits suggest that sulfate
718
G. R. Robinson
reduction could have played a role in formation of
these deposits as well, even though reductants are
not evident. Hydrocarbons that may be involved as
reduction agents are reported from veins in the New
Britain district (Ryan, 1986; Pratt and Burruss,
1988).
Mixing of sulfide-rich and sulfate-rich solu-
tions, particularly where a contrast in solution
temperatures exists, is an effective mechanism for
depositing barite and (or) sulfides without much
silica gangue (Ohmoto and Rye, 1979). Sulfide- and
(or) barite-rich veins are found in the Whatley-
Hatfield district (Massachusetts), the New Britain
and Cheshire-Jinny Hill districts (Connecticut), the
Hopewell-Solebury district (New Jersey-Pennsyl-
vania), and barite deposits in the Culpeper basin
(Maryland-Virginia) (Table 1). The narrow range
of salinities reported from some of these vein
systems implies that, if fluid mixing occurred, the
mixed fluids represent partially reduced and
unreduced fluids of similar character, possibly from
the same source area.
In places, the veins cut Early Jurassic rocks,
and the veins are younger in age than both in-
trusion and some faulting and block rotation in the
early Mesozoic basins. The age of vein formation
may coincide with a 175 Ma thermal event in the
basins reported by Sutter (1988) based on 4Ar/39Ar
isotopic analysis of potassium feldspar from Early
Jurassic intrusives in the Culpeper basin and the
Newark basin. Sutter (1988) interpreted the 175
Ma apparent ages of these feldspars as due to a
moderate-temperature ( = 200C) heating event
caused by migration of hydrothermal fluids. Model
lead ages reported for galena from some of these
veins (Parnell, 1986) are compatible with this 175
Ma thermal event. This 175 Ma age is broadly
correlative with the age of Mesozoic magnetic
lineaments and of marine sedimentation during
initial opening of the Atlantic Ocean, which
resulted from the physical separation of North
America from Africa (Klitgord and Schouten, 1986,
p.364).
INTERPRETATION
Formation of vein deposits of this type is
controlled by local physical and chemical conditions
at the site of deposition. Mineral deposition mech-
anisms responsible for the vein deposits appear to be
cooling, sulfate reduction, wallrock reaction, and
possibly fluid mixing, with the relative importance
of these processes varying in different deposits. The
style, scale, and timing of vein mineralization, how-
ever, can also be interpreted in a process-oriented
framework that is infuenced by larger scale tectonic
and magmatic events which relate to evolution of
rift basins in an extensional tectonic setting. In this
context, ore deposits result from interaction of a
number of processes involving element sources,
chemical transport, and traps. A chain of coinci-
dence, where all of these processes and interactions
occur in proper sequence may result in an ore
deposit; large deposits may form when these
processes are maximized, whereas small deposits
form when one or more of the processes is inefficient.
Energy, sometimes in the form of tectonic
deformation, is needed to drive fluid transport.
These vein deposits appear to be associated with
transport of moderate-temperature (100-250C)
brines from within the basins and adjacent base-
ment rocks to shallow sites of mineral precipitation.
Therefore, the geologic processes that influence
development and transport of metal-bearing saline
fluids in this setting also influence many critical
genetic aspects of vein deposits. In the following
section, development of metal-bearing saline fluids
is discussed in the context of rift basin development
and sedimentation, and fluid transport is discussed
relative to the structural and tectonic setting of
intracontinental rift basins.
DISCUSSION
Exposed early Mesozoic basins in the eastern
United States are typically fault-bounded half-
grabens with dimensions of = 100-300 km length,
10-30 km width, and 2-8 km depth. Sedimentary fill
for individual basins generally is 3000-20,000 km
3
.
Depositional environments in a typical Triassic-
Jurassic rift basin in eastern North America
(Figure 5) vary systematically from alluvial fan to
fluvial to fluvio-playa to playa-lacustrine facies.
Most basins develop as closed hydrologic and sedi-
mentary systems during part of their history
(Smoot, 1985). Base-metal mineralization, as veins
and stratabound occurrences, appears to be associa-
ted with transport of deep-seated, metal-bearing
brines from within extensional basins and adjacent
basement to shallow sites of precipitation. Develop-
ment of metalliferous brines and their subsequent
migration seems to be a typical occurrence gene-
Base-metal and barite-vein deposits, early Mesozoic basins of eastern North America 719
Fal,llt
PthZn-Ba ve,ns
Uritnlum
Setl,n'@l'I thoSled
.... u.;IlIlo'm ')\JIf.oes
o O,IIoDue. bilul l
ConglomeUIIif! .and
_ EU...c. k .nd Qreen
S.,ldSIOI'III:. "nel stl., lo.:!!o
o Shea'eo 'OC'-S
1////1 MornlC'ls
P.lleOlt)'1: 1, ...... eSIQlir.(lQklml l e
f 1w ......
o 2 MILES
I I I I
Sh ..
o I 2
Figure 5. Depositional environments in a typical Triassic-Jurassic rift basin in eastern North America vary systematically
from 1) alluvial fan, 2) fluvial, 3) fluvio-playa, to 4) playa-lacustrine facies. Most basins develop as closed hydrologic
sedientary systems during part of their history. Base-metal mineralization, as veins and stratabound occurrences, appears to
be associated with transport of deep-seated, metal-bearing brines from within the extensional basins and adjacent basement to
shallow sites of precipitation. Development of metalliferous brines in extensional basin settings and their subsequent
migration appears to be a typical occurrence that is genetically associated with the tectonic setting and developmental history
of rift basins.
tically associated with the tectonic setting, high
heat flow, and developmental history of rift basins
(Hanor, 1979; Sverjensky, 1987; Eugster, 1985).
Thermal anomalies caused by increased heat
flow associated with crustal extension on a regional
scale generally coincide with a broad zone of maxi-
mum crustal subsidence and basin development
(Buck and others, 1988). Thermal gradients in the
rift areas may be nearly 1.5 times greater than
average continental values (Lysak, 1987). The high
heat flow values and thermal effects from igneous
activity associated with crustal extension lead to
development, at shallow crustal levels, of warm
formation waters that are capable of carrying high
solute loads. Peak heat-flow values and thermal
anomalies, however, post-date initiation of exten-
sion by "., 5-20 million years (Buck and others,
1988); hence, connate brines may evolve along com-
plex temperature, pressure, and composition paths.
Basement lithologies bordering the basins are
possible sources of metals transported to the basins
in detritus formed by erosion or in solution in
surface or groundwater. The extent of metal release
may depend on the extent of diagenesis. The dis-
solution, alteration, and recrystallization of both
framework minerals (particularly feldspars and
amphiboles) and cements (particularly oxides, hy-
droxides, carbonates, sulfides, and sulfates) may
release transition metals and control the pH and
chemistry of the fluid. Dominant aquifers are in
"red-bed" siliciclastic and quartzofeldspathic rocks,
and the pH of groundwaters is generally buffered by
sericite (illite) + clay or feldspar assemblages. The
median pH for groundwaters from sandstone aqui-
fers in the Gettysburg basin (Pennsylvania) is 6.6,
and groundwaters from most areas of sandstones
and conglomerates have a pH lower than 7 (Wood,
1980, p. 53).
The depositional setting of the basins promotes
development of closed basin hydrology (Smoot,
1985). This results in development of saline surface
and ground waters (Hardie and Eugster, 1970),
oxidized sulfate-bearing sediments, and lacustrine
units with extensive transgressive and regressive
contacts (Smoot, 1985; Olsen, 1988) providing inter-
faces between reduced black shales with organic
720
G. R. Robinson
matter and oxidized tan and red siltstones. Inter-
action of saline groundwaters with sediments of this
type apparently leads to development of deep-seated
metal-rich brines (Sverjensky, 1987). The few data,
from early Mesozoic basins of eastern United States,
on the chemistry of waters wells > 1000 ft (gen-
erally beyond the range of potable water), indicate
their saline nature. In the Gettysburg basin at York
(Pennsylvania) well waters "" 3000 ft below the land
surface contain 9800 ppm CI, 8000 ppm Na, 870 ppm
Ca, and 4800 ppm sulfate (Bain, 1972, p. 100).
Water from the lowest level (625 m) of a 640-m test
well in the Newark basin at Patterson (New Jersey)
was reported (Cook, 1885, p. 115-117) to be "strongly
saline" and contained "" 6900 mgll of NaCl. This
saline water was interpreted as residual formation
waters by Carswell and Rooney (1976, p. 23); how-
ever, most deep groundwater samples are believed
to represent the upper part of a deeper brine system
which has chemical characteristics similar to those
of oil field brines that have been diluted by mixing
with near-surface groundwaters (Bain and Brown,
1980).
The active tectonic setting of basins may be an
important influence on transport of fluids. Primary
porosity and permeability of most rocks in the
basins are low, due to their poorly sorted nature and
pervasive cementation (Marine and Siple, 1974;
Bain and Brown, 1980). Much of their primary
porosity was lost through diagenesis (cementation),
lithification, and compaction. However reactive
groundwaters that created porosity and permea-
bility are evident in places. Secondary porosity,
created by dissolution of framework feldspar
fragments, was reported from sandstones in the
Danville basin of Virginia (Thayer, 1987). In addi-
tion, the variety and complexity of cements in
apparently impermeable mudstones in the Newark,
Gettysburg, Culpeper, and Danville basins indicate
that fluids of differing composition and character
passed through these lithologies over an extended
period of time (Smoot and Horowitz, 1988).
Low hydraulic conductivity of the basins is
illustrated by the behavior of water levels within
wells sited in the basins. Slight water-level distur-
bances in water wells require many years to recover
to levels comparable to adjacent wells (Marine and
Siple, 1974). Most fluid transport occurs in narrow
secondary openings such as joints and faults that
tend to be well developed only in a few thin beds or
zones (Carswell and Rooney, 1976; Biemesderfer
and Leske, 1961, p. 40). Due to this bedding control,
the ratio of lateral to vertical permeablility is
> 100:1 in many cases (Wood, 1980, p. 17), and the
hydraulic connectivity between individual bedding
aquifers is poor. In general, secondary fracture
porosity decreases markedly with depth; however,
some high porosity and permeability of unknown
depth and lateral extent exist in deep artesian
aquifers, as evidenced by high yields in some deep
wells (Wood, 1980). Deep circulation of ground-
water appears to be largely fault controlled, based
on analogy with hydrology of the Basin and Range
Province (Cole, 1982). Intrabasin fluid circulation
may be modified greatly by faults, diabase intru-
sives, and basalt extrusives - which all act as aqui-
tards. A schematic hydrologic model, based on Cole
(1982), illustrates the pattern of groundwater move-
ment in a representative rift basin setting (Figure
6). The active tectonic setting of the basins provides
topographic relief generating hydrologic gradients
which drive groundwater movement. Extensive
dilational fracture networks in major fault systems
are important conduits for deep fluid movement.
Fluid temperatures and salinity generally increase
along the transport path of the groundwater system,
leading to development of warm saline fluids at
moderate depths. The vertical change in chemical
facies with increasing depth or flow path typically is
from Na-Ca-Mg bicarbonate to Na-Ca-Mg or Ca
sulfate to Na-chloride brines. Na-chloride brines
apparently dominate at depth in all of the basins.
Multiple hydrologic regimes typically develop (Cole,
1982) and the basins may appear hydrologically
isolated. For example, deep groundwaters from the
Durham basin (North Carolina) are markedly
different in chemistry from waters in crystalline
basement below the basin (Marine and Siple, 1974).
Many of the veins are localized in the vicinity of
fault and fracture systems, and these fault systems
may be important conduits for fluids. Most mature
fault zones show mineralogical and chemical evi-
dence of extensive hydration and chemical altera-
tion relative to their host rocks (Kerrich and others,
1984; Ratcliffe and Burton, 1988), and it is probable
that faulting occurs in the presence of an aqueous
fluid of at least hydrostatic pressure (Sibson, 1981).
Active fault and fracture systems, however, are not
simply passive fluid conduits. Fault displacement
in the upper crust occurs by seismic slip and is
accompanied by stress and fluid pressure cycles that
induces fluid flow (Sibson, 1981; this volume).
Repeated seismic slip can move significant quan-
tities of hot, deep-seated fluids as discrete pulses
Base-metal and barite-vein deposits, early Mesozoic basins of eastern North America 721
into the upper crust and provides a mechanism for
rapid mixing of fluids which may lead to mineral
deposition. Fluid transport may accompany seismic
slip in situations where the fault system acts as a
"valve" on the fluid reservoir or it may follow
seismic slip in situations where the relaxation of
stress and dilatency in the system "pumps" fluids
(Sibson, 1981; this volume).
For situations where the fault system acts as a
valve on the fluid reservoir, seismic slip is induced
by increasing fluid pressure above hydrostatic
values. Faulting creates fracture permeablility,
allowing partial draining of the fluid reservoir and
reduction of fluid pressure. Self-sealing of the frac-
ture system occurs by mineral deposition from
fluids, and the pressure cycle can be repeated
(Sibson, 1981). The "valve" fault model is consistent
with evidence that much of the crustal extension
associated with formation of early Mesozoic basins
in eastern North America is accomplished by re-
activation of existing fault discontinuities rather
than creation of new faults (Robinson, 1979; Rat-
cliffe, 1971; Ratcliffe and Burton, 1985; Hutchinson
and Klitgord, 1988). For frictional reactivation of
such faults to occur at other than an optimal orien-
tation relative to stress trajectories, in preference to
the formation of a new favorably orientated fault,
fluid pressure conditions must be high, at least
intermittantly, and fault-related hydraulic exten-
sion fractures in the form of mineral-filled veins
should occur (Sibson, 1985, 1987; Phillips, 1972).
Transient post-seismic surface discharge of sig-
nificant quantities of hydrothermal fluids following
earthquakes in consolidated rocks (Sibson, 1981)
provides evidence for a seismic "pumping" mech-
anism that is interpreted as a consequence of the
dilatancy/fluid-diffusion model for shallow earth-
quakes (Sibson and others, 1975). Prior to fault
failure, a large region around the focus of the sub-
sequent earthquake dilates in response to rising
tectonic stress with the opening of extension cracks.
The developing fracture porosity reduces fluid pres-
sure in the dilatent zone, which promotes fluid infil-
tration. Seismic faulting partially reduces stress
levels allowing the dilatent cracks to collapse, and
the fluids they contain are rapidly expelled into the
fault system (Sibson and others, 1975).
Both of these seismically induced fluid trans-
port mechanisms are consistent with mUltiple
periods of mineralization, fluid sources, and pulses
of fluid migration which are evident in many vein
systems associated with early Mesozoic basins of the
_ ______ ll..!ln.<r.'.uucllUtJ..___ Sl'\i!II1)ow 'We-II
_ ______ I"'n.""...., .. "' "-. ___ Ctilcridc/tllCortlonete
(Gradu.l)
./ Km

Figure 6. Schematic hydrologic model of a rift basin
similar to early Mesozoic basins of eastern North America.
The model is based on that developed by Cole (1982, his
Fig. 8) for the East Shore area of Utah. Deep fluid
circulation is largely controlled by faults. Surface hot
springs occur along some major fault systems, and warm
saline brines occur at shallow crustal levels. Legend:
basin fill - pattern; crystalline rocks - no pattern; faults
or fractures - heavy solid and dashed lines; normal faults
have relative displacement shown by arrows; heavy arrows
indicate direction of groundwater movement.
eastern United States. Such fluid transport mech-
anisms are also attractive because they can drive
significant quantities of hot, deep-seated fluids into
the upper crust, perhaps from depths as great as 10
km, and provide a mechanism for rapid mixing of
different fluids.
Other types of ore deposits, formed by transport
of metal-bearing fluids in similar extensional basin
settings, may occur in the vicinity of these vein
deposits, such as sediment-hosted stratiform base-
metal sulfide deposits (Gustafson and Williams,
1981; Plimer, 1986) and stratabound Revett-type
sandstone-hosted copper-silver deposits in the Belt
Supergroup, Montana-Idaho (Lange and Sherry,
1983). In fact, these base-metal and barite-vein
deposits share a similar geologic setting and geo-
chemical association of Ag with Pb, Zn, and Cu with
a number of silver vein districts, such as the Cobalt-
Gowganda and Thunder Bay districts, Canada
(Andrews, 1986) and the Coeur d'Alene district,
Montana-Idaho (Fryklund, 1964; Landis and others,
1984), and they may be considered distant cousins.
722
G. R. Robinson
In this broad class of deposits, a consistent set of
geologic features includes: (1) a tectonic environ-
ment characterized by basinal subsidence and crus-
tal extension in a continental rift setting; (2) spatial
association of the vein and many stratabound
deposits with regional faults; and (3) fluid inclusion
evidence indicating moderate salinity brines with
temperatures generally < 250C during mineraliza-
tion.
CONCLUSION
Epithermal base-metal and barite-vein deposits
are common in and near the margins of rift basins
in extensional continental settings, and such veins
are structurally, temporally, and almost certainly
genetically associated with these rift basins.
Approximately 100 veins of this type are known to
be associated with early Mesozoic basins of eastern
North America. The veins, which occur as open-
space filings in high-angle faults and fractures,
typically contain quartz, barite, and (or) lead-, zinc-,
and copper-sulfide minerals and are characterized
by high sliver contents. Multiple fluid sources and
periods of mineralization and brecciation are
evident. Fluid inclusion data indicate moderate
temperatures (100-250C) and salinites (10-16
equivalent wt% NaC!) during mineralization. The
veins formed late in the history of basin develop-
ment by the transport of brines from deep sources
within the basins and adjacent basement to shallow
sites of mineral precipitation. Stratabound base-
metal deposits that also form by the transport of
metal-bearing fluids may occur in the vicinity of
these vein deposits.
Formation of vein deposits of this type is in-
fluenced not only by chemical and physical condi-
tions of vein fluids and host rocks at the site of
deposition but also by tectonic, magmatic, and
sedimentary events that are part of the development
and evolution of rift basins in extensional tectonic
regimes. Early Mesozoic basins characterized by
first-cycle red-bed sediments, closed-basin hydro-
logy, and high geothermal gradients may be
important as recharge areas of saline brines and
leachable metals (Zielinski and others, 1983). This
tectonic setting apparently creates deep-seated,
metal-rich saline fluids. The structural and deposi-
tional setting of these basins promotes development
of closed -basin hydrology, saline surface and ground
waters, and lacustrine units with extensive trans-
gressive and regressive contacts that are redox
interfaces between oxidized sediments containing
hematite and sulfate minerals and reduced
sediments containing organic material.
The active tectonic setting of the basins may be
an important influence on the transport of fluids.
Deep circulation of groundwaters appears to be
largely fault controlled, by analogy with hydrology
of the Basin and Range province. The active tec-
tonic setting of the basins developed extensive
dilational fracture networks associated with major
fault systems. Active tectonics of rift basins pro-
vides a driving force for fluid migration by produc-
ing topographic relief that creates regional hydro-
logic gradients. Fluid temperatures and salinities
generally increase along the transport path of the
groundwater system, leading to development of
warm saline fluids at moderate depths. High geo-
thermal gradients and thermal effects of igneous
intrusions also promote development of hot reactive
fluids at shallow crustal levels. Repeated cycling of
fluid pressures in active fault systems can also move
significant quantities of hot, deep-seated fluids into
the upper crust and provides a mechanism for rapid
mixing of fluids which may lead to mineral deposi-
tion. Fluid migration may be due to tectonic pro-
cesses that were activated in Middle Jurassic time
by the changing stress regime developed during
initial opening of the Atlantic Ocean, which
followed early basin development along the proto-
Atlantic margin of eastern North America by "'" 50
million years.
REFERENCES
ANDREASSON, P. G., Z. SOLYOM, and 1. JOHANSSON,
1987, Geotectonic significance of Mn-Fe-Ba and Pb-Zn-Cu-
Ag mineralizations along the Sveconorwegian-Grenville
front in Scandinavia: Economic Geology, v. 82, p. 201-207.
ANDREWS, A. J., 1986, Silver vein deposits: Summary of
recent research: Canadian Journal of Earth Sciences, v.
23, p. 1459-1462.
BAIN, G. L., 1972, Feasibility Study of East Coast
Triassic Basins for Waste Storage: u.s. Geological
Survey, Open-file Report, 113 p.
__ , and C. E. BROWN, 1980, Evaluation of the Dur-
ham Triassic Basin of North CaroUna and Techniques
Used to Characterize Its Waste-Storage Potential:
U.S. Geological Survey, Open-file Report 80-1295,133 p.
BANNERMAN, H. M., 1941, New Hampshire Mineral
Resource Survey, Part V: The Fluorite Deposits of
Base-metal and barite-vein deposits, early Mesozoic basins of eastern North America 723
Cheshire County, NH: New Hampshire State Planning
Commission, 11 p.
BATEMAN, A. M., 1923, Primary chalcocite: Bristol copper
mine, Connecticut: Economic Geology, v. 18, no. 2, p. 122-
166.
BIEMESDERFER, G. K., and R H. LESKE, 1961, Ground
water control at Grace mine: Mining Congress Journal, v.
47, no. 10, p. 39-44.
BOYLE, R. W., 1962, Geology of the Barite, Gypsum,
Manganese, and Lead Zinc-Copper-Silver Deposits of
the WaUon-Cheverie Area, Nova Scotia: Geological
Survey of Canada, Paper 62-25, 26 p.
__ , and J. L. JAMBOR, 1966, Mineralogy, geochemistry
and origin of the Magnet Cove barite-sulphide deposit,
Walton, Nova Scotia: Canadian Institute of Mining and
Metallurgy Transactions, v. 69, p. 394-413.
BUCK, W. R, F. MARTINEZ, M. S. STECKLER, and J. R
COCHRAN, 1988, Thermal consequences of lithospheric
extension: Pure and simple: Tectonics, v. 7, no. 2, p. 213-
234.
CARSWELL, L. D., and J. G. ROONEY, 1976, Summary of
Geology and Ground-Water Resources of Passaic
County, New Jersey: U.S. Geological Survey, Water-
Resources Investigations 76- 75, 47 p.
COLE, D. R, 1982, Tracing fluid sources in the East Shore
area, Utah: Groundwater, v. 20, p. 586-593.
COOK, G. H., 1885, Annual Report of the State Geologist:
New Jersey Geological Survey, p. 115-117.
DOE, B. R, and R E. ZARTMAN, 1979, Plumbotectonics,
the Phanerozoic; p. 22-70 in H. L. Barnes (ed.), Geo-
chemistry of Hydrothermal Ore Deposits, Second
Edition: Wiley-Interscience Publication, New York, New
York,798p.
EDMUNDSON, R S., 1938, Barite Deposits of Virginia:
Virginia Division of Mineral Resources, Bulletin 53, 85 p.
EMERSON, B. K., 1898a, Holyoke Folio: U.S. Geological
Survey, Geologic Atlas No. 50, 13 p.
__ , 1898b, Geology of Old Hampshire County,
Massachusetts, Comprising Franklin, Hampshire,
and Hampden Counties: U.S. Geological Survey, Mono-
graph 29,790 p.
EUGSTER, H. P., 1985, Oil shales, evaporites, and ore
deposits: Economic Geology, v. 49, p. 619-635.
FRANKLIN, J. M., and R H. MITCHELL, 1977, Lead-zinc-
barite veins of the Dorian area, Thunder Bay district,
Ontario: Canadian Journal of Earth Sciences, v. 14, no. 9,
p. 1963-1979.
FRITTS, C. E., 1962, The Barite Mines of Cheshire: The
Cheshire Historical Society, Cheshire, Connecticut, 36 p.
FRYKLUND, V. C., Jr., 1964, Ore Deposits of the Coeur
d'Alene District, Shoshone County, Idaho: U.S. Geo-
logical Survey, Professional Paper 445, 103 p.
GRAY, N. H., 1982, Copper occurrences in the Hartford
basin of northern Connecticut; p. 195-211 in R Joesten and
S. S. Quarrier (eds.), Guidebook for Fieldtrips in Con-
necticut and South Central Massachusetts: 74th New
England Intercollegiate Geological Conference, Guidebook
No.5, Storrs, Connecticut, 482 p.
GUSTAFSON, L. B., and N. WILLIAMS, 1981, Sediment-
hosted stratiform deposits of copper, lead, and zinc: Econo-
mic Geology, 75th Anniversary Volume, p. 139-178.
HAAS, J. L., Jr., 1976, Physical Properties of the Coe-
xisting Phases and Thermodynamic Properties of the
H
2
0 Component in Boiling NaCl Solutions: U.S. Geo-
logical Survey, Bulletin 1421-A, 73 p.
HANOR, J. S., 1979, The sedimentary genesis ofhydrother-
mal fluids; p. 137-172 in H.L. Barnes (ed.), Geochemistry
of Hydrothermal Ore Deposits, Second Edition: Wiley-
Interscience Publication, N ew York, New York, 798 p.
HARDIE, L. A., and H. P. EUGSTER, 1970, The evolution of
closed-basin brines: Mineralogical Society of America,
Special Publications, no. 3, p. 273-290.
HAYNES, F. M., 1988, Fluid-inclusion evidence of basinal
brines in Archean basement, Thunder Bay Pb-Zn-Ba
district, Ontario, Canada: Canadian Journal of Earth
Sciences, v. 25, p. 1884-1894.
HOFFMANN, R, and A. BAUMANN, 1984, Preliminary
report on the Sr isotopic composition of hydrothermal vein
barites in the Federal Republic of Germany: Mineralium
Deposita, v. 19, p. 166-169.
HOLLAND, H. D., and S. D. MALININ, 1979, The solubility
and occurrence of non-ore minerals; p. 461-508 in H. L.
Barnes (ed.), Geochemistry of Hydrothermal Ore Depo-
sits, Second Edition: Wiley-Interscience Publication,
New York, New York, 798 p.
HUTCHINSON, D. R, and K. D. KLITGORD, 1988, Deep
structure of rift basins from the continental margin around
New England; p. 211-219 in A. J. Froelich and G. R Robin-
son, Jr. (eds.), Studies of the Early Mesozoic Basins of
the Eastern United States: U.S. Geological Survey,
Bulletin 1776, 423 p.
JASSIM, R Z., R A. D. PATTRICK, and M. J. RUSSELL,
1983, On the origin of the silver + copper + cobalt +
baryte mineralization of Ochil Hills, Scotland: A sulphur
isotope study: Transations of the Institute of Mining and
Metallurgy, Section B, v. 92, p. B213-B216.
724
G. R. Robinson
KERRICH, R., T. E. LA TOUR, and L. WILLMORE, 1984,
Fluid participation in deep fault zones: Evidence from
geological, geochemical and 180/160 relations: Journal of
Geophysical Research-B, v. 89, no. 6, p. 4331-4343.
KLITGORD, K. D., and H. SCHOUTEN, 1986, Plate kin-
ematics of the central Atlantic; p. 251-378 in P. R. Vogt
and B. E. Tucholke (eds.), The Geology of North Am-
erica, Volume M: The Western North Atlantic Region:
Geological Society of America, Boulder, Colorado, 720 p.
LANDIS, G. P., D. L. LEACH, and A. H. HOFSTRA, 1984,
Silver-base metal mineralization as a product of meta-
morphism - Coeur d'Alene district, Shoshone County,
Idaho: Concepts of genesis: Montana Bureau of Mines and
Geology, Special Publication 90, p. 68.
LANGE, I. M., and R. A. SHERRY, 1983, Genesis of the
sandstone (Revett) type of copper-silver occurrences in the
Belt Supergroup of northwestern Montana and north-
eastern Idaho: Geology, v. 11, p. 643-646.
LAWLER, J. P., 1981, Fluid Inclusion Evidence for Ore-
Forming Solutions: Phoenixville, Audubon and New
Galena Mine Districts, Pa.: MA Thesis, Bryn Mawr
College, Bryn Mawr, Pennsylvania, 74p.
LYSAK, S. V., 1987, Terrestrial heat flow of continental
rifts; in I. B. Ramberg, E. Milanovsky, and G. Qvale (eds.),
Continental Rifts - Principal and Regional Charac-
teristics: Tectonophysics, v. 143, p. 31-41.
MARINE, L. W., and G. E. SIPLE, 1974, Buried Triassic
basin in the central Savannah River area, South Carolina
and Georgia: Geological Society of America Bulletin, v. 85,
p.311-320.
MAURIN, J.-C., and J. R. LANCELOT, 1987, Origine des
mineralisations de Pb-Zn de la Vallee de la Benoue
(Nigeria) d'apres la composition en Pb des galenes et de
l'encaissant: Mineralium Deposita, v. 22, p. 99-108.
OHMOTO, H., and A. C. LASAGA, 1982, Kinetics of re-
actions between aqueous sulfates and sulfides in hydro-
thermal systems: Geochimica et Cosmochimica Acta, v. 46,
p.1727-1745.
__ , and R. O. RYE, 1979, Isotopes of sulfur and carbon;
p. 509-567 in H. L. Barnes (ed.), Geochemistry of
Hydrothermal Ore Deposits, Second Edition: Wiley-
Interscience Publication, New York, New York, 798 p.
__ , C. J. KAISER, and K. A. GEER, 1988, Systematics of
SUlphur isotopes in Recent marine sediments and ancient
sediment-hosted basemetal deposits; p.70-120 in H. K.
Herbert and S. E. Ho (eds.), Proceedings of the
Conference on Stable Isotopes and Fluid Processes in
Mineralization: Geology Department and Extension Ser-
vice, University of Western Australia, Publication 23, 382 p.
OLADE, M. A., and R. D. MORTON, 1985, Origin of lead-
zinc mineralization in the southern Benue Trough, Nigeria
- Fluid inclusion and trace element studies: Mineralium
Deposita, v. 20, no. 2, p. 76-80.
OLSEN, P. E., 1988, Continuity of strata in the Newark
and Hartford basins; p. 6-18 in A. J. Froelich and G. R.
Robinson, Jr. (eds.), Studies of the Early Mesozoic
Basins of the Eastern United States: U.S. Geological
Survey, Bulletin 1776, 423 p.
PARNELL, J., 1986, Hydrocarbons and metalliferous
mineralization in a lacustrine rift basin: The Hartford-
Deerfield Basin, Connecticut Valley: Neues Jahrbuch fur
Mineralogie, Abhandlungen, v. 145, no. 1, p. 93-110.
PATTRICK, R. A. D., M. L. COLEMAN, and M. J. RUSSELL,
1983, Sulphur isotopic investigation of vein lead-zinc
mineralization at Tyndrum, Scotland: Mineralium Depo-
sita, v. 18, p. 477-485.
PHILLIPS, W. J., 1972, Hydraulic fracturing and minera-
lization: Journal of the Geological Society of London, v.
128, p. 337-359.
PLIMER, I. R., 1986, Sediment-hosted exhalative Pb-Zn
deposits: Products of contrasting ansialic rifting: Trans-
actions of the Geological Society of South Africa, v. 89, no.
I, p. 57-73.
POTTER, R. W., II, and D. L. BROWN, 1977, The volumetric
properties of aqueous sodium chloride solutions from 0 to
500C at pressures up to 2000 bars based on a regression of
available data in the literature: U.S. Geological Survey
Bulletin, no. 1421-C, p. CI-C36.
PRATT, L. M., and R. C. BURRUSS, 1988, Evidence for
petroleum generation and migration in the Hartford and
Newark basins; p. 74-79 in A. J. Froelich and G. R. Robin-
son, Jr. (eds.), Studies of the Early Mesozoic Basins of
the Eastern United States: U.S. Geological Survey,
Bulletin 1776, 423 p.
RATCLIFFE, N. M., 1971, Ramapo fault system in New
York and adjacent New Jersey: A case of tectonic heredity:
Geological Society of America Bulletin, v. 82, p. 125-141.
__ , and W. C. BURTON, 1985, Fault reactivation models
for origin of the Newark basin and studies related to
eastern U.S. seismicity; p. 36-45 in G. R. Robinson, Jr. and
A. J. Froelich (eds.), Proceedings of the Second U.S.
Geological Survey Workshop on the Early Mesozoic
Basins of the Eastern United States: U.S. Geological
Survey, Circular 946,147 p.
__ , and W. C. BURTON, 1988, Structural analysis of the
Furlong fault and the relation of mineralization to faulting
and diabase intrusion, Newark pasin, Pennsylvania; p.
176-193 in A. J. Froelich and d. R. Robinson, Jr. (eds.),
Studies of the Early Mesozoic Basins of the Eastern
Base-metal and barite-vein deposits, early Mesozoic basins of eastern North America 725
United States: U.S. Geological Survey, Bulletin 1776, 423
p.
REIMER, T. 0., 1986, Phanerozoic barite deposits of South
Mrica and Zimbabwe; p. 1351-1393 in C. R Anhaeusser
and S. Maske (eds.), Mineral Deposits of Southern
Africa: Geological Society of South Africa, Johannesburg,
South Africa, 2 vols., 2335 p.
ROBINSON, G. R, Jr., 1979, Pegmatite cutting mylonite-
Evidence supporting pre-Triassic faulting along the
western border of the Danville Triassic basin, southern
Virginia [abstract]: Geological Society of America, Ab-
stracts with Programs, v. 11, no. 4, p. 208.
__ , and L. G. WOODRUFF, 1988, Characteristics of base-
metal and barite vein deposits associated with rift basins,
with examples from some early Mesozoic basins of eastern
North America; p. 377-390 in A. J. Froelich and G. R
Robinson, Jr. (eds.), Studies of the Early Mesozoic
Basins of the Eastern United States: U.S. Geological
Survey, Bulletin 1776, 423 p.
ROEDDER, E., and R J. BODNAR, 1980, Geologic pressure
determinations from fluid inclusion studies: Annual
Review of Earth and Planetary Sciences, v. 8, p. 263-301.
RYAN, S. S., 1984, Carbonate-quartz-barite veins of the
Hartford basin [abstract): Geological Society of America,
Abstracts with Programs, v. 16, no. 1, p. 61.
__ , 1986, Description and Parage netic Interpreta-
tion of Quart z-Carbon ate-Barite Veins of the Hartford
Basin: MS Thesis, The University of Connecticut, Storrs,
Connecticut, 132 p.
SAMSON,!' M., and D. A. BANKS, 1988, Epithermal base-
metal vein mineralization in the southern uplands of
Scotland: Nature and the origin of the fluids: Mineralium
Deposita, v. 23, p. 1-8.
SIBSON, R H., 1981, Fluid flow accompanying faulting:
Field evidence and models; in D. W. Simpson and P. G.
Richards (eds.), Earthquake Prediction: An Inter-
national Review: American Geophysical Union, Maurice
Ewing Series, v. 4, p. 593-603.
__ , 1985, A note on fault reactivation: Journal of
Structural Geology, v. 7, p. 751-754.
__ , 1987, Earthquake rupturing as a hydrothermal
mineralizing agent: Geology, v.15, p. 701-704.
__ , J. M MOORE, and A. H. RANKIN, 1975, Seismic
pumping - A hydrothermal fluid transport mechanism:
Journal of the Geological Society of London, v. 131, p. 653-
659.
SMITH, R C., II, 1977, Zinc and Lead Occurrences in
Pennsylvania: Pennsylvania Geological Survey, Mineral
Resource Report 72, 318 p.
SMOOT, J. P., 1985, The closed-basin hypotheSis and its use
in facies analysis of the Newark Supergroup; p. 4-10 in G.
R. Robinson, Jr. and A. J. Froelich (eds.), Proceedings of
the Second U.S. Geological Survey Workshop on the
Early Mesozoic Basins of the Eastern United States:
U.S. Geological Survey, Circular 946,147 p.
__ , and M. R. HOROWITZ, 1988, Vug-filling diagenetic
minerals in early Mesozoic lacustrine mudstones of the
Newark Supergroup [abstract]: Geological Society of
America, Abstracts with Programs, v. 20, no. 7, p. A52.
SUTTER, J. F., 1988, Innovative approaches to the dating of
igneous events in the early Mesozoic basins of eastern
North America; p. 194-200 in A. J. Froelich and G. R
Robinson, Jr. (eds.), Studies of the Early Mesozoic
Basins of the Eastern United States: U.S. Geological
Survey, Bulletin 1776, 423 p.
SVERJENSKY, D. A., 1987, The role of migrating oil field
brines in the formation of sediment-hosted Cu-rich
deposits: Economic Geology, v. 82, p. 1130-1141.
THAYER, P. A., 1987, Petrology, diagenesis, and petro-
physical characteristics of Upper Triassic Dan River Group
sandstones, NC and V A [abstract]: Society of Economic
Paleontologists and Mineralogists, Abstracts, v. 4, p. 83-84.
WOOD, C. R, 1980, Groundwater Resources of the
Gettysburg and Hammer Creek Formations, South-
eastern Pennsylvania: Pennsylvania Geological Survey,
Water Resources Report 49 (Fourth Series), 57 p.
ZIELINSKI, R A., S. BLACK, and T. R WALKER, 1983, The
mobility and distribution of heavy metals during the
formation of first cycle red beds: Economic Geology, v. 78,
p.1574-1589.
CHAPTER 7
ABSTRACTS OF OTHER
CONFERENCE PRESENTATIONS
[REPRINTED FROM THE PROGRAM WITH ABSTRACTS VOLUME,1988]
728
Abstracts - From Program with Abstracts Volume, 1988
Features of volcanic rocks and tectonic settings
ALFRED T. ANDERSON, Jr.
Department of Geophysics Science, Uniuersity of Chicago,
5734 South Ellis Auenue, Chicago, illinois 60637, USA
Many physical (structural and textural) and composi-
tional (chemical and mineralogical) attributes of volcanic
rocks are correlated with tectonic setting. However, only a
few causal relations are known. Modern subduction-
related volcanic rocks are persistently highly porphyritic
and dominantly fragmental, especially if subaerial. These
features are expected for magmas rich in H20, because
ascending magmas become supercooled by effervescence
leading to crystallization and explosive eruption. The high
H
2
0 is plausibly related to subduction of cool lithosphere
rich in hydrous minerals. Has subducting lithosphere
always been cool and hydrous? Deep submarine extrusions
of identical magmas would be less porphyritic, less vesi-
cular and not so explosive. Variable vesicularity and asso-
ciated sulfide deposits could, however, help reveal original
depths of extrusion. Magnesian olivine is stabilized by
H
2
0 in siliceous melts; consequently, olivine in blocky
and/or fragmental (highly viscous) volcanic rocks is sug-
gestive of a hydrous magma. Nearly all tholeiitic basalts
crystallize cpx before opx at low pressures. Anomalous
early opx in them seems best explained by siliceous con-
tamination. But how do we account for Mauna Loa?
Granitic contamination may deplete basalt in Na20 with
consequent reversal of plagioclase zoning. Little altered
oceanic ridge basalts have diagnostic chemical composi-
tions, but most old rocks are strongly altered. Reclosure of
the Atlantic would likely preserve parts of Iceland and
Azores. How would we interpret them?
The role of water in the formation of granulite and
amphibolite facies rocks
LINDA M. ANGELONI
Department of Geology, Uniuersity of Montana, Missoula,
MT 59812, USA
Dry mafic rocks in an amphibolite facies environment
develop granulite facies mineral assemblages. Field evi-
dence from Archean rocks in southwestern Montana
supports the theoretical model proposed by Powell (1983)
for the generation of granulite facies assemblages. Pow-
ell's model uses a metabasic system to illustrate internal
buffering of melting reactions which produces a low and
variable activity of water thereby permitting the forma-
tion of granulite facies assemblages. This variable activity
of water produces a mixed pattern of metamorphic facies at
the outcrop scale; uneven patches of granulites are mixed
with amphibolites.
In the Tobacco Root Mountains granulite facies as-
semblages are restricted to mafic lithologies and include
hyp-plag-hbl and diop-gar-plag-qtz. Surrounding pelitic,
quartzofeldspathic, and calcic lithologies have upper am-
phibolite facies assemblages.
Magmatic underplating in a collisional island-arc set-
ting may account for the widespread occurrence of mafic
sills in the Tobacco Root Mountains. Mafic sills intruding
the lower crust are probably drier than the surrounding
metamorphic environment. The inherent dryness of these
sills appears to foster the development of granulite facies
assemblages restricted to mafic lithologies in an otherwise
upper amphibolite facies terrane. A possible alternative is
that the mafic sills supplied additional heat to produce
slightly higher temperature granulite facies assemblages.
Comparison olthe late Precambrian-early Paleozoic
and Mesozoic continental margins of the northeast
United States
P.J. BAROSH
35 Potter Street, Concord, Massachusetts 01742, USA
The Appalachian orogeny occurred between two
colliding continental margins as the Paleo-African Plate
moved relatively west-southwest against North America.
The orogeny began on the east near the subduction zone
and the orogen crumpled inward towards the North
American Plate. The northwest-dipping subduction zone,
now expressed by the Nashoba Thrust Belt, crosses
southern and eastern Connecticut and eastern Massa-
chusetts, where it is locally 22 km wide. It continues
northeastward offshore of the Maine coast. The major
tectonic episode was in the Late Precambrian (600-620
m.y.) with diminishing episodes in the Early Paleozoic.
The ocean closed in the Devonian and the remainder of the
Paleozoic is dominated by strike-slip faulting, uplift and
formation of extensional terrestrial basins.
The position of the Mesozoic continental breakup is
controlled by and parallels the earlier collision border. The
western edge of Triassic and Early Jurassic grabens lies a
little northwest of the subduction zone. The extensional
faults follow ancient ones. The later break forming the
Atlantic continental margin is offshore southeast of the
subduction zone that lies near the present coastline. A
deviation from parallelism at a bend in the zone left a
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
Abstracts - From Program with Abstracts Volume, 1988 729
larger fragment of Paleo-Africa attached to form Rhode
Island and eastern Massachusetts.
The position of the major northwest-trending trans-
form fracture zones along the early Atlantic continental
margin was controlled by older transverse basement frac-
ture zones on the continent.
The structural pattern of the Mesozoic margin and
adjacent ocean basin thus mimics that on the continent.
The pattern of basement fractures in the crust therefore
can be extended in addition to being inherited.
Tectonic setting of Tertiary hydrothermal systems in
the Great Basin region
BYRON R. BERGER
U.S. Geological Survey, Office of Mineral Resources,Denver,
Colorado 80225, USA
The tectonic evolution of the Great Basin region since
the Mesozoic has been related to processes along the
western margin of the North American plate. Until the
late Mesozoic, a magmatic arc had been continuous along
the length of the plate margin with northwesterly and
north-southerly faulting in the back-arc region. More
cryptic east-westerly trends are also evident. From the
Eocene to the early Miocene, the arc was not continuous
along the plate margin, and consequently back-arc
faulting was not uniformly distributed throughout the
region. During this period, back-arc extensional faulting
and related magmatic activity swept from northwest to
southeast across the region. Oligocene polymetallic vein
deposits, gold-bearing skarns, sedimentary-rock hosted
gold deposits, and porphyry-type deposits in the Great
Basin occur within these northwesterly trending regions
and showing east-westerly alignments that may reflect
back-arc extensional faulting that is bounded by deep-
seated, east-westerly transform-related zones inherited
from pre-Tertiary plate-tectonic processes. Tertiary hydro-
thermal activity appears to be restricted to the region west
of the pre-Tertiary miogeoclinal slope-shelf hinge line
except along east-west trends. A continuous volcanic arc
was established briefly in the early Miocene followed by
subsequent rapid retreat of the arc to the north. WNW-
ESE basin and range extensional faulting east of the San
Andreas transform has been the predominant tectonic
style since the middle Miocene. The basin and range fault-
ing has been an important control on mineralization
occurring primarily in regions coincident with pre-basin
and range deep-seated, through-going faulting. The Great
Basin physiography commenced development in the Plio-
cene.
Precise geochronology and isotope systematics as
tools in tectonic and petrologic analysis
M. E. BICKFORD
Department of Geology, University of Kansas, Lawrence,
Kansas 66045, USA
Modern tectonic analysis requires high resolution of
time relationships among igneous, sedimentary, struc-
tural, and metamorphic events. This is accomplished, par-
ticularly in non-fossiliferous rocks, by precise radiometric
geochronology. The most commonly used method for ig-
neous rocks is U -Pb analysis of zircons. The accuracy and
precision of this method depends upon the characteristics
and history of the zircons, but in favorable cases ages can
be determined to within O.5% by application of careful
handpicking and abrasion techniques to improve concor-
dance. Metamorphic chronology can also be learned from
high-grade rocks in which zircons have completely re-
crystallized or developed rims of metamorphic zircon. In
lower grade rocks, U -Pb analysis of sphene or monazite, as
well as Ar-Ar studies of micas and amphiboles, has been
effective in determining tectono-thermal histories because
the temperature response of the U-Pb and K-Ar isotopic
systems in these minerals is reasonably well understood.
Rb-Sr analysis of whole-rocks and minerals is less reliable,
both for determination of primary crystallization ages and
studies of metamorphic history, because the systematics of
radiogenic Sr loss in rocks and minerals is still not well
understood.
Isotopic systematics is a powerful tool for determining
crustal histories of rocks. In particular, the Sm-Nd, Rb-Sr,
and U-Pb systems provide information about the source
regions of igneous rocks and, to some extent, the protoliths
of metamorphic rocks. The Sm-Nd system is now widely
applied, despite the very long half-life of 147Sm, because
both parent and daughter elements are rare-earths and
therefore not easily fractionated during crustal processes
including high-grade metamorphism and even melting.
Thus, Sm-Nd model ages are commonly interpreted as
mantle separation events because mantle melting, in
which Sm and Nd are fractionated as Sm is preferentially
held by residual garnet, is considered to be the last event
that establishes the SmlNd ratio. Although subject to
continuing effects of crustal processes, the Rb-Sr and U-Pb
systems also can provide information about magma sour-
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
730 Abstracts - From Program with Abstracts Volume, 1988
ces, metamorphic protoliths, and crustal residence time.
Commonly, all of these isotopic systems are studied simul-
taneously to obtain maximum information about rock-
forming events and the greatest resolution of their his-
tories.
Regional structural setting of sediment-hosted gold
deposits in Nevada
HAROLD F. BONHAM, Jr.
Uniuersity of Neuada-Reno, Neuada Bureau of Mines and
Geology, Reno, Neuada 89557-0088, USA
Sediment-hosted gold deposits of Carlin type in
Nevada occur along regional trends that clearly reflect
deep crustal flaws. The Carlin trend (N35W) with present
reserves in excess of 40 million ounces of gold, is the most
important alignment of sediment-hosted gold deposits.
Other important trends are the Getchell trend (NE), the
Eureka-Battle Mountain trend (N30'400W), and the
Independence Range trend (N100E). Madrid (1988) has
suggested that these deposits are localized in the crestal
areas of antiforms developed during the Mesozoic Era.
Precambrian continental margins
KEVIN BURKE
Lunar & Planetary Institute, 3303 Nasa Road 1, Houston,
Texas,and Department of Geosciences, U niuersity of
Houston, 4800 Calhoun Rd.,Houston, Texas 77004, USA
All known main types of modern continental margin
can be discerned in the Precambrian geologic record and
although it would be interesting to be able to identify
subtle differences from the modern environment, extreme
tectonism and very limited stratigraphic resolution (Pre-
cambrian rocks do not have diagnostic faunas) have so far
prevented success in the search for differences.
During Archean times, when the record of the early
stages of continental assembly was particularly complete,
Andean margins appear to have been rather widely repre-
sented, although generally extremely tectonized. Strike-
slip dominated sections of Atlantic-type margins have not
been recognized in the Precambrian, but even in young
oceans these are not always easily distinguished from
rifted margins.
One obvious environment to seek in the tectonized
collisional record is the giant delta complex comparable to
those of the Niger, Mississippi and Ganges.
The western Canada convergent margin: Structure
and tectonic history
R M.CLOWES
Lithoprobe, 6339 Stores Road, Uniuersity of British
Columbia, Vancouuer, V6T 2B4F, Canada
As part of the Lithoprobe project, multichannel re-
flection profiles, coupled with a wide range of other geo-
physical and geological studies, have permitted detailed
delineation of the structure and tectonic history of the
convergent margin of western Canada. The modern con-
tinental margin was built against a pre-Tertiary continent
consisting of an amalgamation of older terranes, the oldest
and westernmost being Wrangellia. Eocene subduction
was interrupted by a seaward jump in the trench axis in
latest Eocene, trapping a section of marine volcanics
(Crescent Terrane), together with sections of Mesozoic
marine sedimentary rocks (Pacific Rim Terrane). These
were placed against and under the margin on steeply
dipping thrust faults. Seaward and underlying the Cres-
cent Terrane is the modern accretionary prism. Develop-
ment of the deformation front at the toe of the continental
slope is clearly shown by landward and seaward verging
faults in the sediments. From the deep ocean to Vancouver
Island, the top of the downgoing oceanic plate is imaged
well at depths which are consistent with those determined
for the top of the plate from refraction and earthquake
data. Above the subducting slab, two broad bands of high
reflectivity dip beneath Vancouver Island. The deeper
band coincides with an isothermal dipping layer of high
conductivity as interpreted from magnetotelluric and heat
flow data. One tectonic model for the origin of the re-
flective bands suggests that they are thick sections of
underplated volcanic and sedimentary material, the lower
one of which would be saturated with hot saline fluids. An
alternative interpretation suggests that the lower band is
caused by fluids that are driven off the downgoing oceanic
plate and are trapped below an impermeable horizon
formed at a metamorphic front.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
Abstracts - From Program with Abstracts Volume, 1988 731
Two types offoreland basin in Sichuan micro craton
margin, China
DONG HUOGENl, LU HUAFUl, DENG XIYANGl, and WU
BAOQING2
lDepartment of Earth Sciences, Nanjing University,
Nanjing, China; 2Sichuan Petroleum Administration,
Sichuan, China
Owing to the Indosinian and Yenshanian tectonic
events the early Paleozoic turbidities and Paleozoic-
Mesozoic carbonate platform sequence became a fold-
thrust belt; and a retroarc foreland basin consisted of a
Jura type fold belt in southeastern margin of Sichuan
microcraton, respectively. Consequently, A-type subduc-
tion (Bally) and a shallow, broad, long wavelength basin
with lower potential are recognized. In contrast, the
northwestern Sichuan basin collided with a volcanic arc
and became a peripheral foreland basin composed of the
step thrust nappes and fold thrust nappes due to the Indo-
sinian event and the gravity sliding nappes in consequence
of Tibet plateau uplifting in Yenshanian and Hima-
layanian time. Thus, an A-type subduction (Weber) and a
short wavelength basin with greater potential were dis-
tinguished.
The formation and emplacement oflarge thrust
sheets on the eastern margin of North America
MARK A. EVANS
Department of Geology and Planetary Science, University of
Pittsburgh, Pittsburgh, Pennsylvania 15260, USA
The eastern margin of North America has been sig-
nificantly shortened during Paleozoic collisional events. In
the central Appalachians three large thrust sheets domin-
ate the structural geometry. The Inner Piedmont thrust
sheet (lPTS) is the most easterly of these and the first to
have formed. It is a composite crystalline thrust sheet
composed of smaller imbricate thrust sheets. The IPTS
overrides the central and much larger Blue Ridge thrust
sheet (BRTS).
The BRTS is also a composite crystalline thrust sheet,
and contains smaller thrust sheets that were emplaced
during the Taconic to earliest Alleghanian. Seismic data
and extensive ductile deformation zones throughout the
Blue Ridge provide evidence for these thrusts. Continued
Alleghanian thrusting resulted in the detachment of the
Precambrian crystalline rocks and formation of a ramp
from the crystalline rocks to the Ordovician Martinsburg
Formation. These crystalline rocks were subsequently
transported as the BRTS over lower Paleozoic sedimentary
rocks. The leading edge of the BRTS was deformed into a
large basement-cored fault-bend fold that underwent
intense ductile shear and later brittle thrusting.
After emplacement of the BRTS the lower Paleozoic
rocks became detached along the Cambrian Waynesboro
Formation. Transport along this detachment resulted in
the folding and imbrication of the Cambro-Ordovician
carbonates and the formation of a ramp to the Ordovician
Martinsburg Formation. Continued thrusting moved the
previously deformed carbonates over the ramp as the
North Mountain thrust sheet (NMTS). The NMTS and the
attached BRTS were ultimately transported over 60 km
toward the foreland above a similar section of carbonates.
This lower section of carbonates was later deformed into a
duplex.
Sedimentology and organic geochemistry of the
Jurassic Newark Basin, northern New Jersey
MICHAEL S. FEDOSHl, JOSEPH P. SMOOT2, and RAMA K.
KOTRA3
1 U.S. Army Corps of Engineers, 26 Federal Plaza, New
York, New York 10278, USA, 2Mail Stop 926,U.S.
Geological Survey, Reston, Virginia 22092, USA; 3Mail
Stop 923, U.S. Geological Survey, Reston, Virginia 22092,
USA
The U.S. Army Corps of Engineers has drilled core
holes along the alignment of a proposed flood diversion
tunnel in northern New Jersey. The tunnel alignment is
in the Mesozoic Newark Basin and trends across the
regional northeast strike. About 1650 meters of the ap-
proximately 5000 meter thick Basin strata have been
recovered in stratigraphically overlapping rock core.
Although 430 meters of core have been recovered from
Late Triassic strata, examinations have concentrated on
the almost completely recovered Early Jurassic strata.
The Early Jurassic section of the Newark Basin
consists of three sedimentary formations separated by
three tholeiitic basalt formations. Each sedimentary for-
mation consists of fluvially deposited red sandstone and
siltstone with cyclical deposits of interbedded black
lacustrine shales and gray deltaic siltstones. Differences
among the sedimentary formations reflect an overall in-
crease in standing water, a decrease in depositional
gradient, and a trend to a drier climate during Early Jur-
assic times.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
732 Abstracts - From Program with Abstracts Volume, 1988
Organic geochemical studies are being performed to
determine the nature and amount of carbon compounds in
the cored Jurassic strata. Results of elemental, pyrolytic,
and extract analyses will be presented.
Tectonic controls on gold mineralization in the
southem Appalachian Piedmont
P. GEOFFREY FEISS
Department of Geology, Uniuersity of North Carolina,
Chapel Hill, North Carolina 27599-3315, USA
The southern Appalachian Piedmont is again about to
become, after a hiatus of over 100 years, a major producer
of gold. By 1989, annual production will be close to
175,000 oz from three properties in South Carolina.
In recent years, the dominant emphasis on the control
of gold mineralization in the Carolina volcanic slate belt of
the southern Appalachian piedmont has been on strata-
bound, epithermal to exhalative models. Regional to local
stratigraphic considerations, the mineralogy of ore and of
alteration suites, lead and oxygen isotopes, structural
studies, and analogies with modern volcanogenic systems
support such concepts. This would imply that tectonic
controls on primary gold mineralization are secondary;
major structures, particularly those controlled by the
basement, would serve only to localize volcanic centers
and/or geothermal systems which drive mineralization.
Such models would suggest that those deposits that show
strong tectonite fabrics, including the gold-rich massive
sulfide occurrences of the Gold Hill-Silver Hill fault zone,
vein deposits of the same area and of the Charlotte belt,
pyrophyllite-associated deposits of the Glendon/Robbins
area in North Carolina, and the 50 m.t. Ridgeway deposit
in South Carolina, are instances of structurally remobi-
lized syngenetic mineralization.
Epigeneticists have, in recent years, suggested that
the Haile and Ridgeway deposits, in particular, are meta-
morphogenic mineralized systems wherein gold minera-
lization and associated silicification and argillic alteration
are genetically linked to Taconic (?) brittle to ductile
deformation. Such models place basement control of gold
mineralization in the fore-front -- assuming we knew the
location and the nature of the Carolina slate belt basement
at that time.
Activation of the central crystalline root zone of the
Himalaya
L. N. GUPTA and N. CHAUDHRI
Centre of Aduanced Study in Geology,Panjab Uniuersity,
Chandigarh 160014, India
The central crystalline zone (CCZ) of the Himalaya
has been recognized as the root zone of the Lesser Hima-
layan crystalline nappes. The Badrinath-Kedarnath crys-
talline complex constituting the basement rocks of the
Garhwal nappe, comprises a thick Precambrian sequence
of pelitic, psammitic and calcareous metasediments which
include pelitic schists and gneisses (garnet, kyanite,
sillimanite, cordierite, staurolite, anthophyllite, spinel,
etc.), quartzites, skarn rocks, amphibolites and meta-
pyroxenites.
Imprints of polyphase deformation and metamor-
phism can be recognized in the CCZ. The regional fold
pattern is that of a refolded anticlinal dome in which the
cauliflower-shaped feature is clearly seen in the Badrinath
area. The rocks have been metamorphosed up to horn-
blende-hornfels granulite subfacies. The core of the CCZ
exhibits all the characteristics of mobilization and ana-
texis, the culmination of which resulted in the formation of
granitic magmas which in turn intruded the rocks of the
CCZ during the Miocene period.
At present, the CCZ of the Badrinath-Kedarnath
region is sandwiched between the Main Central Thrust
(MCT) and Malari Thrust. The uplift of this zone was ini-
tiated during the Lower Palaeozoic orogenic movements
and has been exposed as a result of uplift and southward
movement along the MeT.
Recognition of ancient continental margins in the
Canadian Shield
HERW ART H. HELMSTAEDT
Department of Geological Sciences, Queen's
Uniuersity,Kingston, Ontario, K7L 3N6, Canada
Early Proterozoic orogenic belts joining the Archean
nuclei of the Canadian Shield resemble Phanerozoic
mountain belts thought to have originated at converging
plate margins. As indicated by remnants of initial rift
deposits, passive margin sequences, and diabase dike
swarms, many of the Proterozoic belts evolved at the rifted
margins of Archean continental fragments. The develop-
ment of foredeeps, fold and thrust belts, and calc-alkaline
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
Abstracts - From Program with Abstracts Volume, 1988 733
magmatic arcs suggest an origin by collision with other
Archean cratons or juvenile arc terranes. Until recently,
the apparent lack of ophiolites was used as argument that
the Proterozoic rifts did not evolve into true ocean basins.
However, the discovery of a well preserved ophiolite suite,
including sheeted dykes, in the Cape Smith belt of north-
ern Quebec confirms that seafloor spreading did indeed
occur.
Recognition of continental margins within the Ar-
chean nuclei of the shield is not as straightforward as in
the Proterozoic, though it is a myth that the Archean
lithologic record is not compatible with plate tectonics. If
the 2.8 Ga komatiite-quartz arenite association of the
Prince Albert and Woodburn Lake groups in the north-
western Churchill Province was deposited in a continuous
rift structure, Archean micro-continents measuring more
than 800 km in diameter must have existed. As indicated
by the remnants of true oceanic crust in the Yellowknife
greenstone belt, rifting of such microcontinents did proceed
to the seafloor-spreading stage. Archean plate conver-
gence is suggested by the recognition of accretionary
prisms along subprovince boundaries in the Superior
Province. Plate convergence is also held responsible for
large-scale horizontal imbrications, obduction of green-
stone belts and, in the case of regional granulite-facies
terranes, for the doubling of continental crust.
U sing paleomagnetism and aeromagnetism to
delineate ancient continental margins
JOHNW. HILLHOUSE
U.S. Geological Survey, Menlo Park, California 94025,
USA
The interaction, past and present, between the
Earth's magnetic field and ferromagnetic minerals in the
crust provides clues to the tectonic history of the con-
tinents. As interpreted from paleomagnetic poles of
Mesozoic age and geologic mapping, the general tectonic
style of western North America involves accretion of
volcanic arcs and nearshore sedimentary wedges, slivering
of newly-added crust by strike-slip faults, and finally major
translation of these slivers (e.g., Wrangellia and Salinia
terranes). During Paleozoic time, similar tectonic pro-
cesses probably operated along the active margins between
Laurasia and Gondwana. Inferences vary concerning the
Proterozoic record, as some paleomagnetists see evidence
for mobile cratons and others a stable Proterozoic super-
continent. Validity of conclusions drawn from paleomag-
netism rests on the: 1) the time-averaged geomagnetic
field has always maintained the form of a geocentric axial
dipole keeping with the record of the last 20 million years,
2) the magnetic remanence accurately recorded the ancient
field and is not severely contaminated by post-depositional
components. Aeromagnetic surveys, which reveal con-
trasts in magnetic susceptibility and remanent mag-
netization in the upper crust, are useful for tracing ancient
suture zones, rifts, and strike-slip faults from areas of good
exposure into regions having young sedimentary cover.
Midcontinental magnetic features reflect the underlying
geologic structures; however, subsurface models derived
from aeromagnetic data are non-unique. Model para-
meters, such as the source depths, variability of mag-
netization, and the relative contributions of induced and
remanent magnetizations, must be constrained by all
available geological and geophysical information.
The knowledge offaults and lithostratigraphy for a
reevaluation of the Abitibi and Pontiac subprovinces
MICHEL HOCQ
Service Geologique de Quebec, Ministere de l'Energie et des
Ressources, 1620, boul. del'Entente, Quebec, G1S 4N6,
Canada
The Abitibi and Pontiac subprovinces belong to the
Superior Province. They are located between the Kapu-
skasing zone (W), the Opatica Subprovince (N), and the
Grenville Province (E). The first geological compilation for
both subprovinces, which accompanies the GAC Special
Paper 28, was provided by the OGS and MER (Quebec) in
1983. The Quebec part of the map is presently revaluated
from the lithologic, magmatic and tectonic point of view.
The presence of longitudinal faults and deformation
corridors has been identified in the field and deduced
indirectly. Most of the faults are underlined by EM
anomalies. They are located at lithological interfaces, i.e.,
between volcanics and sediments; others are situated in
volcanics and some are cross-cutting supracrustal rocks
and large intrusives as well. All these faults or fault zones
are subvertical or steeply dipping and have not yet been
studied along their whole length in terms of their tectonic
evidences and significance.
These faults delimit specific lozenge-shaped blocks
characterized by their own lithology, stratigraphy and
granitoid intrusions.
Weare presently contemplating the possibility that
these structures represent remanents of horizontal tec-
tonics during the Kenoran orogeny. Most of these blocks
could in fact be considered as allochthonous terranes or
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
734 Abstracts - From Program with Abstracts Volume, 1988
thrust sheets limited by listric faults. Several qualitative
cross-sections are drawn in order to illustrate and propose
a different interpretation from what is now accepted. In
that sense, there is a good agreement, from the Cadillac
fault up to 60 km north in the vicinity of Rouyn, with the
nonmigrated recordings from the deep seismic profile of
LITHOPROBE.
Tectonics of gold mineralization in the Precambrian
of South India
D.JAYAKUMAR
Department of Geology, Madras University Post Graduate
Extension Centre, Government College of Engineering
Campus, Salem, Tamil Nadu, 636 all, India
The shield area of South India may be modelled as (1)
Dharwar craton, (2) mobile-belt of Eastern Ghat and (3)
mobile-belt south of Cauvery. The interrelation among the
three tectonic provinces is being attempted by many
workers. Most of the supra-crystal belts occurring in the
Dharwar craton are comparable to the early Proterozoic
geosynclines of Canada and Australia and some are of
typical greenstone belts. They are coeval (2600Ma).
In the Archean belt of Dharwar craton, Kolar schist
belt with profuse gold mineralization as gold-quartz-
sulfide lodes and gold-quartz calcite veins occur. The N-S
lodes are associated with banded iron-formation, inter-
banded with komatiitic and tholeiitic amphibolites. The
gold concentration varies from 1 to 6 ppm in gold quartz-
sulfides and 1 to 10 ppm in gold-quartz-calcite veins. The
lodes have depth persistence even to a depth of 3500 m.
The source for gold could be iron-rich tholeiite derived from
komatiitic melt components of metasomatized mantle
sources. Based on the tectonic model it is suggested that
gold mineralization may extend southwards in the tran-
sition zone parallel to N -S tensional thermal axis.
Metallotectonic analysis ofthe Canadian
Appalachians
J. DUNCAN KEPPlE
Department of Mines and Energy, P. O. Box 1087, Halifax,
Nova Scotia, B3J 2Xl, Canada
The northern Appalachians may be subdivided into
the following terranes (from northwest to southeast)
associated with characteristic mineral deposits: (i) North
American terrane consisting on its eastern side of Gren-
villian basement overlain by a Late Precambrian-Ordo-
vician miogeocline containing Nb in rift-related alkaline
intrusives, remobilized sedimentary Cu Mo Ag in
Late Precambrian-Cambrian rift sediments, and Missis-
sippi Valley type Zn deposits in Lower Ordovician plat-
formal limestone; in turn overlain by an Ordovician fore-
deep; (ii) Hydrothermal Cu-Zn-Ag veins associated with
tholeiitic volcanics in the continental rise terranes that
have been thrust over the North American terrane; (iii)
Late Cambrian-Early Ordovician ophiolitic terranes typi-
cally contain chromite and asbestos in the ultramafic
rocks, Cu Zn ( Ag Pb Au) veins and massive
sulfide deposits in the overlying tholeiitic volcanic rocks,
and metamorphic/hydrothermal Sb Mo Au Cu-Ph-
Zn quartz veins produced during regional metamorphism;
(iv) Ordovician volcanic arc and rifted arc terranes are host
to Cu-Zn-Pb Au Ag Sb massive sulfide deposits;
(v) Precambrian Avlon composite terrane contains Pb-Zn-
Cu-W skarns in Helikian metasediments, Zn-Cu-Pb Au
Ag massive sulfides in Late Precambrian rifted volcanic
arc rocks and is associated with polymetallic Cu-Mo por-
phyries.
These terranes are overstepped by Silurian-Devon-
ian-Carboniferous sequences, the latter containing Cu-Zn-
Ba mineralization in Visean limestones. Devonian (-Car-
boniferous) plutonism is also widespread across the entire
Appalachians, and is associated with polymetallic Sn-W-
U Mo Sb Bi F Be Li Au Ag Cu
Pb Zn mineralization.
From passive to active to transform: Structural
history of the European-Apulian margins during the
Alpine cycle
B.LAMMERER
Institut f. Allgermeine U. Angewandte Geologie,
Universitat Munchen, Luisenstrasse 37, D-800 Munchen 2,
Germany
In the European Alps, three subsequent stages of
structural history of the continental margins of the Paleo-
European and Apulian plates may be observed: (1) crustal
stretching with subsidence, rifting, elevated heat flow and
magmatism led finally to continental fragmentation with
the development of passive continental margins and
oceanization in between (Triassic to Lower Cretaceous); (2)
subduction, obduction and collision, characterized by
thrusting, decollement and folding, high pressure meta-
morphism and plutonism led to a complex fold-and-thrust
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
Abstracts - From Program with Abstracts Volume, 1988 735
belt and a thickened crust (Upper Cretaceous-Lower
Tertiary); and (3) dextral transform (transpressive) move-
ments between Central Europe and the Adria-realm
caused superposed folding and thrusting, horizontal
stretching of large crustal volumes and connecting Alpine
to Apenninic tectonics (Upper Tertiary to Recent).
Tampere Schist Belt: Structural style within a
convergent plate margin in the Early Proterozoic of
southern Finland
MIKKO NIRONEN
University of Helsinki, Department of Geology, P. O. Box
115, SF-OOl71 Helsinki, Finland
The early Proterozoic Svecofennian terrain consists of
mantle-derived continental crust that was formed mainly
during a short time interval 1.90-1.86 Ga ago adjacent to
the late Archean (about 2.7 Ga) craton to the NE. The
Tampere Schist Belt (TSB) which is part of the Sveco-
fennian terrain, is a narrow, E-W-trending belt about 200
km long and 20 km wide at maximum. It is comprised of
low to medium grade metamorphosed turbiditic sediments
overlain by pyroclastic rocks of predominantly inter-
mediate composition and calc-alkaline affinities.
The F 1 folds with subvertical axial surfaces and
originally subhorizontal fold axes dominate the structure
of the TSB. Well preserved primary structures indicate a
major E-W trending syncline in the northern part of the
belt with minor upright, open to tight synclines and
anticlines in the limb areas. F 1 fold axes in the limbs have
been rotated toward the steeply plunging stretching
lineation. Dextral F 2 folds with subvertical, E-W to NE
striking S2 crenulation cleavage and steeply plunging fold
axes overprint penetrative Sl schistosity. D3 kinking and
fracturing are restricted to narrow zones. The 200 m wide,
steeply dipping shear zone between the TSB and the mig-
mati tic rocks to the south is a thrust fault with vergence to
the N. Another major fault with vertical displacement is
postulated in the northern part of the TSB. Both faults
were formed at a late stage of progressive D1 deformation.
On the basis of these structures, and published sedi-
mentological, geochemical and isotopic data, the TSB is
interpreted as an upper part of a thrust sheet within a
forearc of an ancient convergent margin. The vergence of
structures suggest subduction toward the present S. Thus
a more complex model is required than the formerly
presented (accretionary) island arc models which assume
subductions in the Svecofennian terrain toward the N. A
microplate system with slabs subducting in variable
directions would better explain the intensive and extensive
generation of new crust in the early Proterozoic of southern
Finland.
Archean volcanic-sedimentary (greenstone) belts
RICHARD W. OJAKANGAS
University of Minnesota at Duluth, Duluth, Minnesota
55812, USA, and U.S. Geological Survey
The elongated belts are of two ages, 3.5 to 3.4 Ga and,
dominantly, 2.7 to 2.6 Ga. They consist of thicknesses of
several km of mafic to felsic volcanic rocks (commonly
bimodal) and associated sedimentary rocks. Komatiitic
rocks are present in variable quantities.
The mafic basal sequences, rarely on discernible base-
ment, are commonly composed of pillowed lavas and con-
sanguinous intrusions, both of tholeiitic (andlor calc-
alkaline) composition. Lensoid Algoma-type iron forma-
tions (siliceous, fine-grained, laminated) commonly occur
in the mafics and at the interface with overlying felsics.
The calc-alkaline felsic upper portions of the sequen-
ces are largely volcaniclastic. Dacite is common as dikes
cross-cutting the mafics and as thick pyroclastic accumula-
tions of variable texture. Rhyolite domes, debris flows, hot
ash flows, and submarine and airfall tuffs are locally pre-
sent. massive sulfides, when present, generally occur with
the felsics.
The felsic volcaniclastics pass outward from volcanic
center into sediments. Thick turbidite sequences of con-
glomerate graywacke and mudstone dominate and con-
stitute a resedimented facies association deposited on
submarine fans or ramps flanking volcanic edifices. Also
important is the alluvial fan-braided stream association of
planar- to cross-bedded lithfeldspathic sandstones. Both
facies contain abundant felsic volcanic and coeval plutonic
detritus. Shallow-marine facies between the above two
facies are rare, although some orthoquartzite and carbon-
ate units (some stromatolitic) occur.
The belts are structurally complex; most have under-
gone two or more deformations. Bedding is commonly
vertical to overturned and lineations are steep. Faults
(some mineralized) are abundant. The metamorphic grade
is generally greenschist. The belts are surrounded by and
intruded by K-poor (older) and K-rich felsic plutons.
Various tectonic models, including rifting, sagduction and
plate tectonic models, have been proposed.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
736
Abstracts - From Program with Abstracts Volume, 1988
Use of granitic type in tectonic analysis: Mid-
Paleozoic magmatism, eastern Massachusetts
MATTHEW L. PAIGE and RUDOLPH HON
Department of Geology and Geophysics, Boston College,
Chestnut Hill, Massachusetts 02167, USA
Granitic magmatism of different affinities (A, I and S-
type) could be successfully applied to unravel paleotectonic
environments in areas where there exists a lack of ade-
quate stratigraphic or structural controls. Eastern Massa-
chusetts is one example. This region has been divided into
four different tectonic blocks by semi-parallel fault
systems. The Boston Avalonian Block, the most eastern
block abuts the Nashoba Block and the Newbury Inlier to
the west. Farther to the west lies the Merrimack Trough.
Each block is characterized by different metamorphic
grade, metasediment composition, structure, and Mid-
Paleozoic magmatism. Attempts to correlate these tectonic
blocks by structure and stratigraphy were not successful.
The geological feature that each block has in common
is the presence of mid-Paleozoic magmatic suites. The
Merrimack Trough and the Nashoba block are charac-
terized by both S and I-type plutonic magmatism. The
Newbury Inlier is characterized by I-type volcanism and
the Boston Avalonian block is intruded by a series of A-
type plutons. The tectonic setting for these granitic types
has been examined on a block by block basis instead of
examining the presence of Mid-Paleozoic magmatism
within all of the blocks.
Detailed geochemical study of the Mid-Paleozoic Cape
Ann Complex (late Ordovician-early Devonian) within the
Boston Avalonian Block shows that this type of
magmatism is consistent with formation in the behind-arc
environments. If contrasted with the characteristic
magmatism in the other blocks, we can suggest a tectonic
model of an easterly dipping subduction zone and with the
plate boundary lying to the west of the Merrimack Trough.
In this model the fault systems are not considered to be
plate sutures, but rather internal fracturing of a single
plate.
Thermal history of the Bakersville metagabbro and
metadiabase dikes in the Blue Ridge, northwestern
North Carolina and adjacent Tennessee
LAURA D. RAINEY, STEVEN A. GOLDBERG, and J.
ROBERT BUTLER
Department of Geology, University of North Carolina,
Chapel Hill, NC 27599-3315, USA
Bakersville metagabbro sills and metadiabase dikes
are tholeiitic intrusions associated with Late Proterozoic
rifting and formation of the Iapetus Ocean. The Bakers-
ville suite intrudes Grenville age and older (1.0 to 1.8 Ga)
basement within the western Blue Ridge thrust complex
(BR) and can be used to distinguish Precambrian from
Paleozoic metamorphism. The BR has been divided into
four major thrust sheets, each with distinct lithologic,
deformational and metamorphic characteristics. A NW -SE
traverse across the BR from Pardee Point, TN, to Bakers-
ville, NC, shows aphanitic to fine-grained amygdaloidal
porphryitic metadiabase abundant and commonly intrudes
metagabbro. In the NW, the Paleozoic metamorphic grade
is lower greenschist facies increasing to upper amphibolite
facies in the SE. The high grade metamorphic assem-
blages of the metabasites represent a single peak meta-
morphic event. At any given locality, metamorphic tex-
tures vary within the suite and can be correlated with the
relict igneous textures. Locally, ductile deformation
imposed on these metabasites produces poikiloblastic
garnet bearing amphibolites. A petrogenetic grid indicates
peak metamorphic conditions of 750-800C at 7-9 kb and
is correlated with the Taconic orogeny. Applied Fe-Mg ex-
change geothermometry on the coarse-grained garnet
amphibolites shows close agreement, with a decrease in
temperature to the northwest. Lower temperatures ob-
tained from fine-grained coronas consisting of the peak
metamorphic assemblage are consistent and probably
reflect the closure temperature during very slow cooling.
A comparison of the systematic change of closure tempera-
tures with respect to the various metamorphic textures of
the metabasites will be discussed in evaluating the cooling
rates of the higher grade thrust sheets within the BR.
The Midland fish-bed: An Early Jurassic
transgressive-regressive lacustrine sequence in the
Culpeper Basin, Virginia
SUSAN C. ROBERTS and PAMELA J. W. GORE
Department of Geology, Emory University, Atlanta, Georgia
30322, USA
A series of meter-square sample blocks through a 65
cm transgressive-regressive lacustrine shale sequence was
examined to determine temporal changes in an ancient
lake. The Early Jurassic Midland fish-bed, Culpeper
Basin, Virginia, preserves a sequence resulting from rapid
deepening and gradual shallowing of a rift valley lake.
Preserved below the fish-bed is mudcracked shale con-
taining plant fragments and representing a vegetated
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
Abstracts - From Program with Abstracts Volume, 1988 737
mudflat environment. The lower 20 cm of the fish-bed
represent a transgressive sequence and is very organic rich
(3.03 to 7.32% TOC) and contains abundant, well preserved
fossil fish, fish coprolites, ostracodes, and conchostracans.
This portion of the sequence was deposited in anoxic water
in a deep stratified lake. The upper 40 cm of the fish-bed is
a regressive sequence and contains less organic material
(1.01 to 3.07% TOC), and poorly preserved fossil fish and
fish scales. Conchostracans, ostracodes and coprolites
disappear gradually up through the sequence and fish
become less abundant and more poorly preserved upward.
The rapid deepening of the lake is attributed to a
combination of continuous subsidence, high precipitation,
and low sedimentation rates. Later, gradual shallowing of
the lake is the result of increased sedimentation rates, and
decreased precipitation. The presence of fluvial units over-
lying the fish bed suggests that increased sedimentation
rates could be the result of delta progradation slowly filling
the lake. The Midland fish-bed is an organic rich lacus-
trine sequence recording the rapid deepening and gradual
shallowing of a stratified rift valley lake with anoxic
bottom waters.
Varieties of Proterozoic orogeny
JOHN J. W. ROGERS
Department of Geology, CB#3315, University of North
Carolina, Chapel Hill, North Carolina 27599-3315, USA
Compression and magmatism associated with the
closure of an ocean basin can be demonstrated in some Pro-
terozoic terrains. An excellent example of this "Phanero-
zoic-style" orogeny is in the central Nubian-Arabian
shield, where very late Proterozoic accretion of island arcs
formed a crust that is currently more mafic than older
shields. Other examples are the Singhbhum orogenic belt
ofIndia (1,500 m.y.) and the Penokean belt of the northern
U.S. (1,850 m.y.). This type of orogenic belt is charac-
terized by ensimatic and subduction-zone sedimentary and
magmatic suites and by widespread development of primi-
tive, post-orogenic granites immediately following the
accretionary phase.
A more common type of Proterozic orogenic belt can-
not clearly be related to oceanic closure. These belts lack
ophiolites, oceanic sediments, calcalkaline magmatic
suites, and compositionally simple post-orogenic granites.
In India, for example, most of the 1,500 m.y. thrust zones
simply juxtapose rocks of granulite grade against lower-
grade suites without evidence of subduction or remnants of
oceanic suites; the prevalence of mid-Proterozoic plat-
formal sediments in India supports the concept of
intracontinental thrusting. Similarly, most of the latest
Proterozoic orogeny in Africa (Pan African) seems merely
to rework older sialic rocks, with formation of new
continental crust restricted to the northeastern part of the
continent. Other examples of Proterozoic ensialic orogeny
have been proposed in parts of the Australian and Cana-
dian shields. At present, it is not clear whether this type of
orogeny is restricted to the Proterozoic or whether it has
occurred in the Phanerozoic and simply is not exposed.
Metamorphism and tectonics
J. W.SEARS
Geology Department, University of Montana, Missoula,
Montana 59812, USA
Typical geothermal gradients place the deeper parts
of most continents under steady-state metamorphic con-
ditions in which quartz-rich rocks have virtually no plastic
strength. However, most exposed metamorphic complexes
crystallized during distinct tectonic episodes in linear belts
along ancient plate boundaries. Metamorphic/tectonic
episodes correspond to times of crustal thickening along
convergent plate boundaries, and crustal thinning along
divergent boundaries, when changes in lateral support
allow the plastic rocks to spread.
In western Montana, the thickened Cordilleran
orogenic welt spread laterally during a late Cretaceous
metamorphic/tectonic episode when the rocks became too
soft to support the load of the mountains because of heat
brought in by widespread igneous intrusions. A second
metamorphic episode occurred when changing plate boun-
dary conditions permitted crustal extension to draw out
metamorphic core complexes in early Tertiary time.
The Red Sea line - A Late Proterozoic transcurrent
fault
ARYEH E. SHIMRON
Geological Survey of Israel, 30 Malkhe Israel, Jerusalem,
Israel
A mismatch of considerable dimensions between the
African and Asian segments of the Arabian-Nubian Shield
is indicated by recent tectonic, lithostratigraphic and
geochronological data. This is manifested in the left-
lateral, Late Proterozoic displacement of about 350 km,
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
738
Abstracts - From Program with Abstracts Volume, 1988
along the present Red Sea axis, of the corridor of the NW-
trending Najd Fault System. A reconstruction made by
shifting Arabia and Sinai southward, and towards the -500
m contour, until the best match is obtained on the Najd
corridor, also results in good continuity of accreted iso-
chronous microplates and major suture-ophiolite com-
plexes on both sides of the Red Sea. In the northern shield
the matching mega-units comprise the northern margins
of the ca. 800-700 Ma South Eastern Desert, Sinai and
Midyan magmatic arc terranes and certain anomalous (ca.
>800 Ma) high grade, passive continental margin clastic-
carbonate sequences. In consequence of such major trans-
current faulting, the North Eastern Desert and Sinai were
in an "extensional basin-dome" relationship during the
Late Proterozoic. This final tectonic evolutionary episode
in the Arabian-Nubian Shield took place in two, possibly
related and overlapping phases: (a) deformation along the
Najd Fault System (ca. 640-600 Ma); and (b) Red Sea
transcurrent faulting (ca. 610-580 Ma).
The geological definition of seismogenic continental
crust
RICHARD H. SmSON
Department of Geological Sciences, University of
California, SantaBarbara, California 93106, USA
In deforming continental crust with moderate to high
heat flow (e.g., San Andreas fault system, Basin and
Range), microearthquake activity is largely restricted to
the top 10-15 km of the crust. Larger shocks (M>5.5) tend
to nucleate towards the base of this seismogenic regime,
rupturing upwards and laterally so that their aftershocks
are mostly restricted to the same zone. Maximum depth of
activity deepens in colder cratonic crust, but shallows
locally to as little as 4-5 km in regions of active geothermal
fluid upwelling.
Examination of the style and rock products of shear-
ing in ancient fault zones exhumed from different crustal
levels provides a geological basis for interpreting these
variations in the depth of seismic activity. For quartzo-
feldspathic crust, a reconstructed depth sequence of
dominant fault rock types within localized shear zones is:
lonite
Gneiss
The cataclasite-mylonite transition represents a
major change in the mechanical response of the crust, with
crystal plastic flow of quartz in mylonites becoming
significant at temperatures in excess of 300 -350C corres-
ponding to the onset of greenschist facies metamorphic
conditions. The base of the seismogenic zone thus re-
presents the transition from unstable frictional (FR)
behavior, where shear resistance is dominantly pressure-
dependent and increases with depth, to quasi-plastic (QP)
aseismic shearing in mylonite belts where flow shear
resistance is temperature-dependent and decreases with
depth. Peak shear resistance is expected in the vicinity of
the FRlQP transition, the zone of large rupture nucleation,
where mixed continuous and discontinuous deformation
must occur over a large range of strain rates.
It is apparent that much epithermal fault-hosted
mineralization develops in the boiling zone occupying the
top 1-2 km of seismogenic crust, while some mesothermal
lode gold deposits developed near the base of the seismo-
genic regime in the vicinity of the FRlQP transition.
Abrupt reductions in fluid pressure may be induced by
earthquake rupturing in both these environments, trig-
gering episodes of hydrothermal precipitation.
Late Precambrian and Paleozoic accretionary
history of the central and southern Appalachians:
Characterization through magmatic and
deformational events
A. K. SINHA, R. L. BADGER, E. HUND, J. P. HOGAN, and
R. E. GUY
Department of Geological Sciences, Virginia Polytechnic
Institute and State University, Blacksburg, Virginia
24061, USA
The relationship between generation of thermal
anomalies (magmatism), and terrane accretion or
continental breakup can be identified through spatial and
temporal distribution of igneous rocks and episodes of
faulting. Late Precambrian extension of the eastern
margin of North America is comprised of two distinct
episodes, an early phase marked by alkaline plutonism and
volcanism (-700 m.y.) and a later event (-570 m.y.)
marked by the outpouring of the voluminous flood basalts
of the Catoctin Formation. Subsequent to a transition
period, a magmatic arc (-500 m.y.) is indicative of
subduction tectonics. Eventual collapse of the arc terrane
(-460-440 m.y.) provided the thermal instability for
generation of epidote-bearing granites. Ages of faulting
(Brevard, Conowingo) are consistent with thrust tectonics
during this time. The Siluro-Devonian (-425-380 m.y.) is
marked by parallel belts of plutonism. The plutons in the
western overthrust terrane are a result of delayed
decompressional melting and may not be the result of
subduction processes. Igneous activity in the Charlotte
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
Abstracts - From Program with Abstracts Volume, 1988 739
belt (gabbro-syenite-granite) is probably related to strike
slip accretion and transtensional environment in the
'Avalon' terrane. The Permo-Carboniferous (-325-270
m.y.) igneous activity is related to oblique subduction and
suturing of Africa.
The thermal evolution of the central and southern
Appalachians is consistent with the suturing of four
distinct plates over a 200 m.y. interval where three distinct
tectonic processes (convergent, tensional, overthrust) have
occurred.
Crustal extensional processes in Cordilleran
metamorphic core complexes
ARTHUR W. SNOKE
Department of Geology and Geophysics, University of
Wyoming, Laramie, Wyoming 82071, USA
Cordilleran metamorphic core complexes provide a
unique view of the processes involved in continental
extension at contrasting crustal levels. The upper plates of
core complexes consist of a mosaic of high-angle and low-
angle normal faults truncated by a detachment fault that
commonly serves as the structural boundary between a
plastically-deformed footwall and a brittlely-deformed
hanging wall. The plastic deformation of the footwall can
be of any age but clearly indicates that upper crustal rocks
have been structurally juxtaposed onto middle crustal
rocks. Detachment fault, as used in this context, therefore
implies important crustal omission. Detailed geochrono-
logical data from some core complexes indicate that
Tertiary plutonic rocks were involved in the plastic defor-
mational history thereby demonstrating the importance of
high-temperature Tertiary strain during core complex
evolution. Pressure-temperature-time paths for several
well-studied core complexes indicate fault slip rates > 1
cm/year, suggesting high strain rates during Tertiary
crustal extension. Palinspastic restorations of upper plate
rocks associated with many core complexes indicate a large
percent extension (i.e., > 100%). The ubiquitous presence
of two-mica granitoids, commonly characterized by very
high initial 87Sr/86Sr ratios, indicates the role of partial
melting and high-grade metamorphism in the deep-crust
beneath core complexes. The seismic reflectivity of core
complex crust also suggests the importance of plastic
deformation and magmatic underplating in the lower crust
beneath these highly extended terranes. In summary, core
complexes provide much insight into an important stage of
crustal extension intermediate between continental rifts
and oceanic basins.
Fluids in the crust: Use of stable isotopes to
distinguish magmatic/hydrothermal events in
fragments of oceanic and arc crust
DEBRA S. STAKES
Department of Geological Sciences, University of South
Carolina, Columbia, South Carolina 29208, USA
The construction of oceanic and island arc crust is
accompanied by pervasive seawater hydrothermal altera-
tion which imparts a characteristic 180 versus depth pro-
file. Studies in ophiolite, and assumed for normal oceanic
crust, suggest that gabbroic rocks are depleted in 180 to
values near 5 per mil. Shallow dike and extrusive rocks
are enriched in
18
0 to values over 8 per mil. This profile is
complicated by multiple magmatic and hydrothermal
events of back arc or incipient arc formation observed in
the northern Semail ophiolite. Here, plagiogranite-diorite
intrusions result from off-axis magmatism. These are
surrounded by metagabbroic rocks that are highly depleted
in
18
0, and cut by quartz-sulfide-epidote-jasper veins that
reflect temperatures above 300C. Mafic dikes predate, or
are contemporaneous with, or postdate plagiogranite in-
trusion and presumably fuel the shallow hydrothermal
systems. Dikes and hydrothermally altered country rock
are present as xenoliths at various stages of assimilation.
Mafic dikes and plagiogranites are intruded along fault
intersections which also acted as conduits for hydro-
thermal fluids.
The only sample of modern oceanic crust is ODP Site
735 from the intersection of the southwest Indian Ridge
and the Atlantis II fracture zone. Transform-enhanced
cracking has produced hornblende-filled vertical veins of
invariably deformed gabbroic rocks. These cracks allow
fluids to hydrate the lower crust and upper mantle at
temperatures above 500C. Assemblages that include
diopside-sphene-rutile-clinozoisite-plagioclase-phlogopite-
anthophyllite may result from fluids from the hydrated
mantle. All gabbroic rocks are depleted in
18
0 to values
near 5 per mil. Such fragments of hydrothermally altered
crust are potentially incorporated into active continental
margins. In addition, the processes observed in seafloor
systems are analogous to the early stages of continental
arc development with meteoric hydrothermal systems.
The transition from marine to meteoric fluids is an
important factor in the magmatic/tectonic history of a
region.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (BuUe, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
740 Abstracts - From Program with Abstracts Volume, 1988
Large-scale hydrothermal systems and gold
deposition along major fault zones in collisional
margins: Characteristics of the Mother Lode,
California, and similar systems in B.C. and their
comparison to other types of geothermal systems
BRUCE E. TAYLOR
Geological Survey of Canada, 601 Booth Street, Ottawa,
Ontario, KIA OEB, Canada
Characteristics of hydrothermal systems and gold
deposition associated with major fault zones in collisional
margins (e.g., Mother Lode and northern Sierra districts,
California) include: discontinuous, narrow alteration
zones many km in length (e.g., >200 km; Mother Lode),
fluids oflow salinity, moderate CO
2
, relatively
low S, high 6
18
0, variable 6D, generally restricted range of
6
13
C, and temperatures generally <350C. Hydrothermal
activity postdated formation of the major fault zone by ca.
15-45 Ma and was not spatially associated with synchron-
ous magmatism.
Regional/district patterns of 6
18
0 of quartz veins
indicate lithologic control (i.e., low w/r ratios) of fluid che-
mistry. High 6
18
0 of vein quartz (ca. + 13-20) and altera-
tion minerals imply fluids with high
18
0/
16
0, in contrast to
low 6
18
0 waters and large 6
18
0 gradients in fossil and
modern magmatically-heated geothermal systems. Data
from B.C. suggest highly evolved meteoric waters, whereas
Mother Lode data imply mixing of several waters. De-
watering (plus metamorphism) of thickened crust would
provide evolved fluids (ultimately meteoric or sea water) of
complex histories. Deep penetration km?) of surface
waters could have been associated with extensional defor-
mation during uplift. S in the Mother Lode is ultimately of
mantle origin by degassing and/or remobilization. CO
2
in
the fluids could reflect CO
2
dominantly degassed from
mantle-derived magmas which also provided the heat for
felsic magma generation.
Structural analysis ofthe central Columbia Plateau
utilizing radar, Landsat, digital topography, and
magnetic data bases
RICHARD L. THIESSEN 1, JAY R. ELIASONl, CRAIG W.
BROUGHER 1, KENT R. JOHNSONl, GREG B. MOHLl, M.
FOLEy2, and D. E. BEAVER2
1 Geology Department, Washington State University,
Pullman, Washington 99164, USA?Battelle PNL, Battelle
Blvd, Richland, Washington 99352, USA
Interest in the Hanford site (Washington) as a nuclear
production, power, and waste disposal site has led to a vast
quantity of geophysical and remote sensing data sets of the
central Columbia Plateau. To date these various studies,
including 13 independent magnetic linear and image
lineament studies, have not been adequately correlated,
but provide a unique opportunity to compare and contrast
the viability of the different geophysical and remote
sensing techniques.
The geology of the central Columbia Plateau is char-
acterized by subdued topography and limited outcrop, with
most of the exposure concentrated in localized folded!
faulted mountains (the Yakima folds) and along river
canyons.
Magnetic anomalies are being produced by structures
at depth that mayor may not extend to the surface. Con-
versely, the images and DEMs show surface structures
that may possibly continue to depth. Features that are
exhibited by both data types will be the best constrained.
Due to their topographic expression and associated defor-
mation, the Yakima folds are detected by nearly all of the
studies. Elsewhere, alignments of windblown surficial
sediments can be easily distinguished due to their strong
image expression and lack of correlation to magnetic
features. Similarly, magnetic linears due to gently tilted
basalts or dikes at depth had no correspondence in the
images and DEMs. One of the few mapped faults not asso-
ciated with the Yakima folds that is presently defined to be
8 km long, could be extended for more than 110 km based
upon eight of the above studies. Other features that do not
correspond to known structures are detected in a number of
the above data types and so are likely to have a strong
structural control.
Late Cretaceous tectonics of the Blythe-Quartzite
region, SE California and SW Arizona: Deformation
within an oblique-slip orogen
RICHARD M. TOSDAL
U.S.Geological Survey, 345 Middlefield Road, Menlo Park,
California 94025, USA
Oblique-slip orogens are characterized by (1) lateral,
vertical, and rotational movements, (2) lateral variations
in sedimentary facies and thicknesses, (3) oblique com-
pressional and extensional structures, and (4) shifting
deformation styles along the belt. A Late Cretaceous oro-
genic belt in the Blythe-Quartzite region, SE California
and SW Arizona has many of these features, and it is here
proposed to be an oblique-slip orogen.
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
Abstracts - From Program with Abstracts Volume, 1988 741
The approximately 200-km long, easterly-trending
Blythe-Quartzite belt lies at a slight angle to the north-
westerly Cretaceous batholithic trend of California, and,
by inference, to the plate margin at that time. The three
structural domains of the belt, from north to south, are the
(1) Jurassic (?) and Cretaceous Maria fold and thrust belt
(MFTB), (2) "McCoy Basin", where the entire Jurassic (?)
and Cretaceous McCoy Mountains Formation (MCM)
currently crops out, and (3) Late Cretaceous Mule Moun-
tains thrust system (MMT).
Deformation in the belt began in the Arizona MFTB
where southward directed folding and thrusting affected
rocks as young as the lower MCM; an 85 Ma granite
intrudes one thrust. The lower MCM in the "McCoy Basin"
southwest of the MFTB was uplifted and eroded at this
time. Erosion was followed by deposition of the upper
MCM, beginning prior to 78 Ma; presumably the uplifted
MFTB was the source region. In California, MCM depo-
sition was apparently uninterrupted by thrusting as in
Arizona. In the California MFTB, regional metamor-
phism, deformation, and pegmatite intrusion occurred at
mid-crustal depths. At higher crustal levels, the deforma-
tion and metamorphism affected the entire MCM. Prior to
70 Ma, the metamorphosed MCM was buried beneath a
cool crystalline block along the southward-dipping MMT;
movement was towards the northeast. Granite intruded
parts of the belt at about 70 Ma. Uplift and cooling of the
belt followed.
The obliquity of the Blythe-Quartzite belt to the
inferred continental margin and the rapidly shifting
tectonic environments are consistent with deformation in
an oblique-slip orogen. This interpretation is supported by
the inferred Late Cretaceous oblique subduction and ac-
companying dextral shear along the continental margin,
as evidenced by plate tectonic reconstructions for this time
and the paleomagnetic data for large scale translations.
Mesoscopic analysis of brittle deformation zones and
fault rocks: Mesozoic extensional reactivation of the
late Paleozoic Hylas fault zone, Virginia
R. VENKATAKRISHNAN andA. WATKINS
Department of Geological Sciences, Old Dominion
University, Norfolk, Virginia 23529, USA
Brittle deformation zones (BDZ) occur as bands (inter-
leaved?) 0.5-1 km in width along the eastern portions of the
Hylas Fault Zone (HFZ) in eastern Virginia. The HFZ
(N300E/42S) is a highly mylonitized right-lateral trans-
pressional thrust that was reactivated during Mesozoic
extensional events. The Richmond Triassic Basin forms an
east-facing half graben on the HFZ. The BDZs pervasively
transect mylonite fabric and kinematic indicators on the
BDZ display components of SE-directed extension over-
printed by oblique right- and left-handed strike slip com-
ponents. Extensional fractures are commonly coated by an
assortment of calcite and zeolite antitaxial vein minerals.
These are often overprinted by mUltiple sets of shear-fibre
overgrowths displaying well-developed planar fabric.
Fractures operating in extension display pseudotachylite
(?) at restraining steps, whereas implosion or crackle
breccia forms at releasing bends. Tabular or pipe-like (2-
3m wide) zones of chloritic breccia transect the entire
width of BDZ. Reworking of breccia clasts by large fluid
fluxes "pumped" through these zones is evident; the clasts
are often exploded, rotated, and may show signs of pluck-
ing and injection of late stage veins and the chloritic
matrix into the wall rocks. The contact of breccia zones
with wall rock is generally knife-sharp. The entire BDZ is
cut by Jurassic-age diabase dikes, that are themselves cut
across, but not tilted, by late stage anastomizing veins.
The upper plate Triassic sedimentary rocks display
detachment-style deformation. The interplay of BDZ and
hanging-wall structural style indicates that the marginal
faults to Triassic Basins may in fact be shallower in dip
than the Paleozoic thrust faults that exerted preexisting
controls to basin formation. The overall geometry of de-
formation described above is not unlike the upper level
brittle deformation observed in core-complex detachment
zones in the Western United States.
Ph isotopic characteristics of Mesozoic intrusive
magmatism along the craton margin in the western
USA
J.L. WOODEN,J. S. STACEY,A.C. ROBINSON,R. W.
KISTLER, R. M. TOSDAL, and M. J. WHITEHOUSE
U.S. Geological Survey, MS 937, Menlo Park, California
94025, USA
Pb isotopic compositions for Mesozoic intrusive rocks
associated with subduction magmatism in the western
U.S. establish that different environments with respect to
sources and/or processes were present. In the eastern
Mojave Desert, where the magmatic arcs cut across the
Proterozoic craton, the data for the intrusive rocks show
large spreads in 206/204 Pb ratios and define Pb-Pb iso-
chrons with ages similar to those of the exposed basement
rocks. This pattern holds in spite of major differences in
composition and age in these plutons suggesting that a
In: Basement Tectonics 8: Characterization and Comparison of Ancient and Mesozoic Continental Margins -
Proceedings of the 8th International Conference on Basement Tectonics (Butte, Montana, 1988) (edited by M. J.
Bartholomew, D. W. Hyndman, D. W. Mogk, and R. Mason). Kluwer Academic Publishers, Dordrecht, The Netherlands,
1992.
742 Abstracts - From Program with Abstracts Volume, 1988
combined crust and ancient lithospheric mantle system
dominates all source materials. Other suites of plutons
that intruded basement rocks at the craton edge in SE
California and SW Arizona have more restricted ranges of
Pb isotopic compositions. Some of these compositions are
consistent with continental arc sources, others with upper
or lower crustal sources; and intrusive events with
different ages show distinct data ranges. Magmatism in
the northern Sierra Nevada and in central Nevada was
involved with a very different craton margin. Here a thick
sequence of Late pC and early Paleozoic sediments overlap
the crystalline craton. Pb isotopic data show a single
linear trend, best described as an extended mixing line
(206/204 Pb == 18.6-19.6). In eastern Nevada, 208/204 Pb
values increase rapidly where Zartman (1974) located the
crystalline craton edge. In this area, Sr vs Pb isotopic data
have correlations with pluton age. We conclude that Pb
isotopic data are sensitive indicators of the variations in
magmatic sources along a craton margin, and that these
variations are related to the complexity and previous
history of the crust-lithospheric mantle system exposed at
that margin.
AUTHOR INDEX
744
Author Index
Agrawal, A. ................................... 583
Alt,D. ........................................ 385
Ameen, M. S. .................................. 559
Anderson, A. T., Jr. ............................. 728
Anderson, J. L. ................................. 205
Angeloni, L. M. ................................ 728
Arav, S. ....................................... 425
Badger, R. L. .................................. 738
Barosh, P. J. ................................... 728
Barth, A. P. .................................... 205
Bartholomew, M. J. ............................ 443
Beaver, D. E. .................................. 740
Bender, E. E. .................................. 205
Berger, B. R. ................................... 729
Bickford, M. E. ................................. 729
Bilodeau, W. L. ................................ 241
Bonham, H. F., Jr. .............................. 730
Bowes, D. R. ................................... 283
Brew, D. A. .................................... 169
Brougher, C. W. ................................ 740
Burke, K. ..................................... 730
Butler, J. R. ................................... 736
Card, K. D. .................................... 615
Chaudhri, N. .................................. 732
Clowes, R. M. .................................. 730
Cummings, M. L. ............................... 323
Davis, M. J. .................................... 205
Deng,X. ...................................... 731
Dilek, Y. ...................................... 179
Dong,H. ...................................... 731
Drury, M. J. ..................................... 27
Dudas, F. O. ..................................... 93
Edelman, S. H. ................................. 197
Eliason, J. R. .................................. 740
Ernst, W. G. ................................... 145
Erslev, E. A. ................................... 313
Evans, M. A. ................................... 731
Farber, D. L. ................................... 205
Fedosh, M. S. .................................. 731
Feiss, P. G. .................................... 732
Foley, M. ...................................... 740
Fullagar, P. D. ................................... 37
Gallagher,J.J.,Jr. ............................. 591
Garihan, J. M. ................................. 539
Goldberg, S. A. ................................. 736
Gore, P. J. W. .................................. 736
Guilbert, J. M. ................................. 129
Gupta, L. N. ................................... 732
Guthrie, G. E. .................................. 341
Guy, R. E. ..................................... 738
Hamilton, W. .................................... 3
Hansen, V. L. ................................... 51
Hayes, E. M. ................................... 205
Helmstaedt, H. H. .......................... 625, 732
Hillhouse, J. W. ................................ 733
Hocq, M. ....................................... 733
Hogan,J.P. .................................... 738
Hon,R. ........................................ 736
Hozik, M. J. .................................... 503
Hund,E. ....................................... 738
Hunter, D. R. .................................. 265
Jayakumar, D.
Johnson, K. A.
Johnson, K. R.
734
205
231,740
Keppie, J. D. ................................... 734
Kish, S. A. ..................................... 481
Kistler, R. W. .................................. 741
Kotra, R. K. .................................... 731
Lammerer, B. .................................. 734
LU,H. ......................................... 731
McBride, B. C. ................................. 341
McCarthy, T. S. ................................ 689
McCulloch, W. R. ............................... 323
McMillan, W. J. ................................ 633
Mason, R. .................................. 625, 679
Mawer, C. K. .................................... 67
Meen, J. K. .................................... 299
Minzoni, N. .................................... 579
Mogk, D. W. ................................... 283
Mohl, G. B. ................................ 231,740
Moores, E. M. .................................. 179
Mueller, P. A. .................................. 283
Myers, R. E, .................................... 689
Nironen, M. .................................... 735
Ojakangas, R. W. ............................... 735
Paige, M. L. .................................... 736
Poulsen, K. H. .................................. 615
Panteleyev, A. ................................. 633
Prucha, J. J. .................................... 83
Rainey, L. D. ................................... 736
Ranson, W. A. .................................. 539
Rast,N. ....................................... 395
Robert, F. ...................................... 615
Roberts, S. C. .................................. 736
Robinson, A. C. ................................. 741
Robinson, G. R., Jr. ............................. 711
Rogers, J. J. W. ............................. 583,737
Author Index 745
Root, S. 1. ...................................... 469 Thiessen, R. L. ............................. 231,740
Roots, C. F. ..................................... 359 Thomas, M. D. ................................... 5
Ryberg, P. T. ................................... 527 Thompson, R. 1. ................................ 359
Timmins, E. A. ................................. 373
Sears, J. W. ................................ 385,737 Tollo, R. P. ..................................... 425
Schmidt, C. J. .................................. 341 Tosdal, R. M. ............................... 740,741
Sheedlo, M. K. .................................. 341
Shimron, A. E. ................................. 737 Ugarkar, A. G. ................................. 599
Sibson, R. H. ............................... 603, 738
Sillitoe, R. H. ................................... 665 Venkatakrishnan, R. ........................... 741
Sinha, A. K. .................................... 738
Smoot, J. P. .................................... 731 Watkins,A. .................................... 741
Snoke, A. W. ................................... 739 Wehr, F. ....................................... 407
Stacey,J.S ..................................... 741 Whitehouse,M.J. .............................. 741
Stahl, S. D. ..................................... 249 Wooden, J. L. .............................. 283,741
Stakes, D. S. ................................... 739 Woodward, L. A. ................................ 653
Stanistreet, 1. G. ................................ 689 Wu,B. ......................................... 731
Steltenpohl, M. G. .............................. 491
Symons, D. T. A. ................................ 373 Young,E.D .................................... 205
Taylor, B. E. ................................... 740 Zartman, R. E. ................................. 699
Proceedings of the International Conferences on Basement Tectonics
1. R. Mason (ed.): Basement Tectonics 7. Proceedings of the Seventh Interna-
tional Conference on Basement Tectonics, held in Kingston, Ontario, Canada,
August 1987. 1992 ISBN 0-7923-1582-0
2. J. Bartholomew, D. W. Hyndman, D. W. Mogk and R. Mason (eds.):
Basement Tectonics 8. - Characterization and Comparison of Ancient and
Mesozoic Continental Margins. Proceedings of the Eighth International
Conference on Basement Tectonics, held in Butte, Montana, USA, August
1988. 1992 ISBN 0-7923-2088-3
3. M. J. Rickard, H. J. Harrington and P. R. Williams (eds.): Basement Tec-
tonics 9 - Australia and Other Regions. Proceedings of the Ninth Interna-
tional Conference on Basement Tectonics, held in Canberra, Australia, July
1990. 1992 ISBN 0-7923-1559-6
KLUWER ACADEMIC PUBLISHERS - DORDRECHT / BOSTON / LONDON

Você também pode gostar