Você está na página 1de 30

0361-0128/01/3439/887-30 $6.

00 887
The Lithologic, Stratigraphic, and Structural Setting of the
Giant Antamina Copper-Zinc Skarn Deposit, Ancash, Peru
DAVID A. LOVE,

ALAN H. CLARK,
Department of Geological Sciences and Geological Engineering, Queens University, Kingston, Ontario, Canada K7L 3N6
AND J. KEITH GLOVER*
Glover Consulting Ltd., 146 Simcoe St., Victoria, B.C., Canada V8V 1K4
Abstract
Antamina, located at latitude 9 32' S and longitude 77 03' W in the Ancash Department of north-central
Peru, is the largest known Cu-Zn skarn ore deposit. It incorporates a mineral reserve of 561 Mt, which has an
average grade of 1.24 percent Cu, 1.03 percent Zn, 13.71 g/t Ag and 0.029 percent Mo, calculated at a 0.7 per-
cent Cu equiv cutoff grade. The grandite-dominated calcic skarn formed in and around an upper Miocene por-
phyritic monzogranite stock emplaced into Upper Cretaceous carbonate strata that had experienced thin-
skinned, northeast-verging thrusting and folding in the late Eocene Incaic orogeny. The exoskarn Cu-Zn ore is
discordant to the strata of the Jumasha and overlying Celendn Formations, which comprise, respectively, mas-
sive to thick-bedded, relatively pure limestones and thin-bedded, predominantly marly limestones. The Ju-
masha Formation, the upper contact of which is locally defined as the top of the uppermost thick-bedded lime-
stone or marble unit, hosts approximately three-quarters of the known exoskarn. Approximately the same
fraction of the contiguous endoskarn Cu ore occurs adjacent to this formation. The overlying Celendn For-
mation is less extensively mineralized but, because it is widely metamorphosed to hornfels and locally con-
verted to diopsidic skarnoid, may have inhibited the upward and outward migration of hydrothermal fluids,
thereby promoting the development of the unusually large endoskarn ore zone. Ore also occurs in late hy-
drothermal breccias emplaced during the formation of mineralized endoskarn.
The preskarn thermal metamorphic aureole around the ore deposit is expressed differently in the two host
formations. Jumasha Formation limestone is coarsened and bleached to banded gray marble and locally to
white marble peripheral to the intrusion and skarn. Minor scapolite occurs in dark gray bands in marble, con-
centrated in a discontinuous halo tens of meters wide and commonly separated from the skarn by tens of me-
ters. Three facies of calc-hornfels are recognized in the marl beds of the Celendn Formation adjacent to the
intrusion extending hundreds of meters beyond sulfide-bearing skarn: a peripheral, very fine grained, light
brown phlogopitic facies; an intermediate, fine-grained, gray tremolitic facies; and a proximal, medium-grained,
light green diopsidic facies. At an X
CO
2
of 0.1 to 0.9 and P = 100 MPa, these zones reflect temperatures in-
creasing to circa 495C adjacent to the intrusion. In addition, in nodular beds of the Celendn Formation that
have been metamorphosed to hornfels, diagenetic calcite nodules are selectively replaced by diopside for dis-
tances of tens of meters beyond the skarn front. Such calc-silicate formation through both metamorphism and
metasomatism, together with a 9 km
2
cluster of Pb-Zn-Ag vein deposits, provides district-scale vectors to ore.
The Antamina deposit lies on a newly recognized cross-strike structural discontinuity in the segmented In-
caic Maran thrust and fold belt, the northeast-trending Querococha arch. Southeast of the arch, Incaic folds
and thrust faults strike north-northwest, but northwest of the arch they strike northerly. The plunge of fold axes
concomitantly changes from south-southeast to north. Stratigraphic relationships indicate that the arch was a
paleohigh, at least in the Jurassic and possibly throughout the late Paleozoic-early Mesozoic interval. The mid-
dle Miocene Carhuish pluton is exposed on the arch 30 km southwest of Antamina, whereas coeval Calipuy Su-
pergroup volcanic units lie at similar altitudes to the north and south. Only scattered hydrothermal centers of
late Miocene age are known in the Cordillera Negra, but an apparent swarm of intrusions, including the Anta-
mina stock, occurs along the Querococha arch.
Antamina is situated where the locus of changes in the strike of folds and faults and the plunge of folds steps
left along the arch. At Antamina, a pair of fault-bend folds above frontal thrust ramps show approximately 500
m of dextral apparent offset across the deposit and are inferred to have been separated by a northeast-strik-
ing transfer fault or lateral ramp, itself localized by a left-stepping jog in the Valley fault, an underlying, sim-
ilarly oriented transverse structure. The jog in the Valley fault is inferred to have also controlled intrusion and
skarn development. This local-scale jog in the Valley fault mimics the regional step along the arch. The arch
may reflect a transform segment of the originally jagged, rifted continental margin, which persisted as a trans-
verse basement weakness. Northeast-striking, originally sinistral, basement structures affected regional-scale
sedimentation and structural patterns, including articulation of the thrust and fold belt. At a local scale, they
2004 by Economic Geology
Vol. 99, pp. 887916

Corresponding author: e-mail, love@geol.queensu.ca
*In Memoriam. On July 17, 2001, Keith Glover died unexpectedly but peacefully in his sleep, immediately after returning from fieldwork. He had an in-
fectious passion for rocks and great patience in teaching about them. He was a skilled structural geologist and a professional ore deposit specialist with acute
perception and a love for walking the rocks. Keith had been consulting internationally for 14 years and was respected for his intellect and breadth of geolog-
ical understanding. He also managed admirably to balance his love for geology with that for his family. Keith was well liked for his generous spirit and hu-
manity, and he is greatly missed.
Introduction
DESPITE their potential ore genetic and metallogenic impor-
tance, the lithologic, stratigraphic, and structural settings of
skarn mineralization have rarely been comprehensively docu-
mented. It is therefore difficult to assess their influence on
the localization of skarn-generating hydrothermal systems
and, in particular, to envisage the specific environments in
which exceptional deposits have developed. In this paper, we
describe the host rocks and structural relationships of the An-
tamina Cu-Zn(-Ag-Mo) deposit, north-central Peru, and pro-
pose a model for the stratigraphic and tectonic environment
in which this largest known Cu-Zn skarn orebody formed.
These aspects are controversial, in part because of the poorly
defined local stratigraphic succession and because of the de-
formation, metamorphism, and metasomatism imposed on
the ore-hosting strata. Following a brief summary of the geol-
ogy of the deposit, we document the regional-scale (ca. 5,000
km
2
) geologic and geodynamic setting of the mineralization
before focusing on the district scale (ca. 120 km
2
).
Copper mineralization was known at Antamina (anta: cop-
per in Quechua) in pre-Colonial times, but only modest
amounts of Pb and Ag are known to have been produced in
the district prior to 2001 (Redwood, 1999). The skarn con-
tains proven and probable reserves of 561 Mt with an average
grade of 1.24 percent Cu, 1.03 percent Zn, 13.71 g/t Ag, and
0.029 percent Mo (calculated at a 0.7% Cu equiv cutoff
grade). Compaa Minera Antamina S.A., which operates the
Antamina open-pit mine, is owned by BHP Billiton (33.75%),
Noranda (33.75%), Teck Cominco (22.5%), and Mitsubishi
(10%). Production of copper-silver and zinc concentrates, as
well as lead, molybdenum, and bismuth byproducts, began in
July 2001 (Zuzunaga, 2003).
The Antamina mine is located at approximately 9 32' S and
77 03' W, 270 km north of Lima and 130 km from the Pacific
coast, in Ancash Department in north-central Peru (Fig. 1). It
lies in the eastern part of the Cordillera Occidental, east of
the Cordillera Blanca and west of the Ro Maran valley.
The skarn is exposed between approximately 4,200 and 4,800
m a.s.l., at the head of a southwest-draining glacial valley, but
prior to mining much of the orebody was covered by Lago
Antamina, a glacial tarn (Fig. 2). The history of exploration at
Antamina and the general geology of the deposit are summa-
rized by Redwood (1998, 1999). OConnor (2000) reviewed
the geologic and geophysical approaches that delineated the
ore and outlined the development of the mine, metallurgical
testing, and resource calculations. Both authors generally
supported previous geologic descriptions and interpretations
(e.g., Petersen, 1965).
The Antamina deposit formed at 9.86 to 10.18 Ma
(
40
Ar/
39
Ar step-heating data of Love et al., 2003) around a
small monzogranitic porphyry intrusion. It is hosted by Upper
Cretaceous carbonate strata within the Maran thrust and
fold belt, formed by the late Eocene Incaic orogeny (Fig. 1;
Noble et al., 1979; Mgard, 1984). The western Andes of
Peru were the site of episodic arc magmatism from the Late
Triassic to the late Miocene (Cobbing et al., 1981). However,
from latitudes 2 S to 15 S they are now underlain by a flat
subduction zone widely ascribed to underthrusting of the
Nazca Ridge (Barazangi and Isacks, 1976; Pilger, 1981; Ham-
pel, 2002) and, in the northern part, the postulated Inca
Plateau (Gutscher et al., 1999). This major flat-slab domain
separates the Northern and Central Volcanic zones of the
Andes and has been apparently amagmatic since emplace-
ment of the last phase of the Cordillera Blanca batholith at
6.3 to 8.2 Ma (Mukasa, 1984; McNulty et al., 1998). Intrusion
and mineralization at Antamina took place shortly before this
terminal magmatism. Although the carbonate rocks that host
the deposit have long been recognized as Upper Cretaceous,
they have been assigned to various formations. The strati-
graphic relationships in the mine area are herein clarified
through examination of the carbonate rocks around the skarn
and their comparison with well-described measured sections
elsewhere in north-central Peru. This analysis permits both
elucidation of the structure of the area and characterization of
the types of rocks replaced by the skarn.
The Antamina deposit shares numerous common features
with other large porphyry-related Cu skarns (Einaudi, 1982a,
b), but it differs from most in the exceptional development of
mineralized endoskarn and the association of ore-grade Cu,
Zn, and Mo in contiguous zones. An additional unusual fea-
ture is the widespread development of chalcopyrite-rich hy-
drothermal breccias, the extent and above-average Cu con-
tent of which were unrecognized prior to our research. Many
of the observations on the deposit- and district-scale geology
recorded herein, and their interpretation, were introduced in
unpublished reports prepared by the authors (D.A. Love and
A.H. Clark, 1998a, b, 2000, unpublished reports to Compana
Minera Antamina S.A., Lima; J.K. Glover, 1998a, 1998b,
1998c, unpublished reports to Compana Minera Antamina
S.A., Lima).
The Antamina deposit
The Antamina deposit comprises endoskarn and exoskarn,
with subordinate breccia bodies that cut both skarn and in-
trusion within the perimeter of skarn. The mineralized skarn
is dominated by grandite garnet, which grades from brown to
green nearer the host limestone (Petersen, 1965). Thus the
deposit conforms to the oxidized calcic clan of Einaudi et al.
(1981). The geology of the deposit and its immediate sur-
roundings is summarized in Figure 3a on three northwest-
southeast drill-hole cross sections, spaced 50 m apart,
through the middle of the orebody. The simplified surface ge-
ology map in Figure 3b was constructed by projecting the
lithologic boundaries defined on these and thirty other cross
sections. The area delimited by the mineralized skarn front is
approximately 1.18 km
2
overall. The skarn zone straddles the
original intrusive contact and surrounds a circa 0.24 km
2
core
of porphyry that forms a crudely parallelogram-shaped prism
with a vertical axis (Fig. 3b). The outer boundary of the min-
eralized skarn is more elongated northeast-southwest than
this core because it expands around faults and dikes extending
to the east, northeast, and southwest. It is therefore roughly
888 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 888
influenced lateral ramp formation and related fracture development in the overlying thrust sheets. In the pro-
posed model, they also localized later uplift and the rapid transit of small volumes of productive melt into a
shallow crustal setting, conditions favorable for formation of a giant magmatic-hydrothermal ore deposit.
elliptical in plan and has a northwest-southeast width of up to
1,000 m, and a northeast-southwest length of more than 2,500
m. The long axis parallels the Antamina valley and is perpen-
dicular to the regional structural grain of the deformed car-
bonate host rocks. As recognized by Petersen (1965) the
outer limit of skarn is generally subvertical. The skarn nar-
rows with depth as the core of porphyry widens (Fig. 3a), but
there is no significant change in the Cu and Zn grades to
depths of at least 400 m below the original valley floor. Except
at high elevations in the eastern part of the deposit, the ex-
oskarn is almost everywhere mineralized.
In detail, the individual skarn facies and breccia bodies are
complexly shaped and discontinuous, but the deposit can be
simplified as comprising an inner shell of endoskarn, stock-
work, breccia, and brown garnet exoskarn that contains the
copper molybdenum ore and an outer shell of green garnet
exoskarn comprising the copper-zinc ore. Molybdenite is dis-
seminated in irregular zones within and at the margin of the
intrusion. Chalcopyrite is the dominant copper mineral ex-
cept at shallow depths in the southwestern part of the deposit,
where bornite predominates in wollastonitic exoskarn, which
forms an enclave in green garnet Cu-Zn exoskarn.
Hydrothermal breccia is common at or near the endoskarn-
exoskarn contacts along the northwest and southeast sides of
the deposit. Breccia also cuts the porphyry core as anasto-
mosing sheets and pipes that are commonly enveloped by
fine-grained maroon garnet endoskarn (Fig. 3). The breccia
zones contain minor chlorite and were originally described by
Petersen (1965) as chlorite skarn. The breccias are generally
poorly sorted and comprise angular to rounded fragments of
the skarn and metallic minerals supported in a sand-sized
matrix of similar composition. As in skarn, sphalerite and
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 889
0361-0128/98/000/000-00 $6.00 889
FIG. 1. Location map of the Antamina deposit and general geology of part of Ancash and La Libertad Departments, Peru.
Compiled and modified after Egeler and De Booy (1956), Cosso (1964), Wilson and Reyes (1964), Cosso and Jan (1967),
Wilson et al. (1967, 1995), Myers (1976, 1980), Reyes (1980), Snchez (1995), Allende (1996), Cobbing et al. (1996), Jacay
(1996), Snchez et al. (1998), INGEMMET (1999), and Strusievicz et al. (2000). The Tapacocha axis delineates the western
edge of the Cretaceous shelf (Myers, 1974, 1975). Other deposits (Pierina and Pasto Bueno) and prospects (Magistral) men-
tioned in the text are shown, as are the areas illustrated in Figs. 5 and 6. MTFB = Maran thrust and fold belt.
molybdenite do not commonly occur together in the breccias;
sphalerite is found in breccias in or near exoskarn, but molyb-
denite occurs in breccias in endoskarn. The common metallic
minerals in the breccias are pyrite, chalcopyrite, and mag-
netite, which are mostly comminuted but also occur as veins,
massive bodies, and large fragments. This mineral assemblage
is rare in exoskarn but forms widespread stockworks and
sheeted vein swarms in fine-grained maroon garnet en-
doskarn, in many places grading into crackle, mosaic, and ma-
trix-dominated breccia. We estimate that approximately one-
third of the ore at Antamina may have formed during this late
brecciation, veining, and endoskarn-forming stage.
Widely spaced, late calcite-tetrahedrite sphalerite
galena veinlets are common and cut all skarn and breccia
types, although they are also locally dismembered in breccia,
probably because of settling. Scarce realgar veinlets occur in
calc-hornfels above and peripheral to the skarn.
Regional Geologic Setting
The Upper Cretaceous strata enclosing the Antamina de-
posit are part of a metallogenically important Albian and
Upper Cretaceous package of carbonate rocks, the Machay
Group, that hosts many ore deposits in the polymetallic skarn
and carbonate-replacement belt of central Peru (Soler et al.,
1986). These rocks formed during the later of two Permian to
Paleocene episodes of basin development that deposited a
succession of alternating siliciclastic and carbonate facies in
western South America (Sempere et al., 2002). This Late
Jurassic to Paleogene subsidence formed the West Peruvian
trough, which separated the magmatic arcs to the west from
the eastern geanticline now represented by the Maran
metamorphic complex (Benavides, 1956; Wilson, 1963;
Atherton et al., 1983; Mgard, 1987). The Cretaceous sedi-
mentary rocks that crop out in the Antamina area accumu-
lated in the shallow-water portion of this trough, the Yauli
shelf (Szekely, 1967). The Tapacocha axis, now a north-north-
westtrending high strain zone, separates the western,
deeper-water portion of the trough from the shelf sedimen-
tary rocks (Fig. 1; Myers, 1974, 1975).
Stratigraphic relationships
The host Machay Group (Figs. 1 and 4) includes all Albian
to mid-Campanian carbonate rocks south of 9 S and east of
the Tapacocha axis in north-central Peru. Szekely (1967) in-
troduced the Machay Group in central Peru and Samam-
Boggio (1980) applied it throughout Peru. These rocks also
have been referred to informally as the middle Cretaceous
limestone series (Harrison, 1940), the upper Cretaceous
and Albian carbonate series (Mgard, 1984), and the upper
carbonate sequence (Manrique, 1998). The group is under-
lain by a predominantly siliciclastic sequence comprising
Upper Jurassic marine black shales of the Chicama Group
and Lower Cretaceous (Berriasian to Aptian) continental to
shelf sandstones, shales, and minor limestones of the Goyllar-
isquisga Group (Fig. 4). It is widely overlain, conformably or
slightly unconformably, by red beds (Wilson, 1963), which are
mainly Campanian to Paleocene but as old as Santonian in
central Peru (Jaillard, 1987). Not preserved in the immediate
Antamina area, these nonmarine, coarse clastic rocks have
been variously described as the Pocabamba Formation, 25
km southeast of Antamina at La Unin (Wilson, 1963, after
McLaughlin, 1924), the Chota Formation, 20 km north of An-
tamina (Benavides, 1956: after Broggi, 1942), and the Cas-
apalca Group, 25 km southwest of Antamina in the Cordillera
Huayhuash (Coney, 1971, after McLaughlin, 1924).
The Machay Group contains two transgressive sequences
separated by a disconformity ascribed to late-middle Albian
uplift and erosion related to the Mochica orogeny (Mgard,
1984). In the lower part of the group, the successive Pari-
ahuanca, Chulec, and Pariatambo Formations (Fig. 4) record
a transition from near-shore, calcareous sandstone and mas-
sive, shelly limestone, through thin-bedded limestone and
marl, to deep-water, thin-bedded, bituminous, dark gray marl
and limestone (Benavides, 1956, 1999; Wilson, 1963; Jaillard,
1987). Following the late-middle Albian hiatus, carbonate
sedimentation on the platform resumed with the deposition
of the shallow-water, upper Albian to upper Turonian Ju-
masha Formation (Jaillard, 1987), originally defined by
McLaughlin (1924) in central Peru. This formation is overlain
by the muddier, deeper-water Celendn Formation (Bena-
vides, 1956), largely Coniacian to Santonian in age (Jaillard,
1987) but attaining the mid-Campanian in northern Peru
(Mourier et al., 1988). The upper part of the group, compris-
ing the Jumasha and Celendn Formations (Fig. 4), thus rep-
resents a second major transgressive sequence. The lower to
middle Albian carbonate strata (the first transgressive sequence)
are similar in lithology and thickness in both northern and
890 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 890
FIG. 2. Premine physiography of the Antamina area, with place names re-
ferred to in the text, illustrating the clustering of Ag-bearing Pb-Zn vein de-
posits around Lago Antamina documented by Bodenlos and Ericksen (1955).
Also indicated are the Contonga Pb-Zn-Cu-Ag mine 5 km to the north-north-
west of Lago Antamina and veins about 500 m northeast of Contonga. B =
Barrn, C = Casualidad, Cc = Condorcoccha, F = Fortuna, JE = Julia Eloisa,
P = Poderosa, Pp = Putapuquio, R = Recompensa, RdO = Rosita de Oro, SF
= San Francisco, SR = Santa Rosa, UP = Usu Pallares. The viewpoints for
photographs in Figures 7, 8, 12, and 15 are shown as eyes. (Contour inter-
val is 200 m.)
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 891
0361-0128/98/000/000-00 $6.00 891
FIG. 3. Simplified cross sections (a) and projected surface geology map (b) of the Antamina deposit, illustrating the
crudely elliptical, vertical zones of endoskarn and exoskarn developed between a core of largely skarn-free porphyry and the
limestone host rocks. Simplified surface geology map is based on drill-hole geology projected to surface and on surface map-
ping by D.A.L. and J.K.G., combined with that by L. Hathaway (Inmet) and M. Wunder (Noranda). Prominent northwest-
striking Incaic folds and the left-stepping, transverse Valley fault (VF) are indicated. The location of the proposed Valley lat-
eral ramp (VLR), inferred to be responsible for the apparent dextral offset of the Antamina anticline (AA), is also shown.
central Peru (Benavides, 1956; Jaillard, 1987). However,
north of approximately 9 S, the overlying upper Albian to
mid-Campanian carbonate rocks are much thicker, more fos-
siliferous, and lithologically more variable than in central
Peru, and the Jumasha Formation interval is divided into five
formations (Benavides, 1956). Jaillard (1987) provides corre-
lations between these stratigraphic sections in northern and
central Peru.
Tectonic relationships
Published descriptions of the structural setting of the Anta-
mina deposit (Bodenlos and Ericksen, 1955; Terrones, 1958;
Petersen, 1965; Redwood, 1999) have focused on the
Maran thrust and fold belt, which developed circa 30 m.y.
before mineralization, with scant consideration of the re-
gional tectonic environment that existed in the mid-Miocene.
We argue, however, that Antamina lies athwart a large-scale,
cross-strike (northeast-southwest) structural discontinuity
(Wheeler, 1978) that was tectonically active at the time of
skarn formation and hence has metallogenic significance.
Love et al. (2001) termed the structure the Querococha arch
because its southwestern limit at the margin of the Callejon
de Huaylas lies close to Laguna Querococha (Figs. 5 and 6).
Its influence on the abundance of Neogene intrusions, the re-
gional strikes in the Maran thrust and fold belt, and the
stratigraphic relationships of underlying Mississippian to
Lower Jurassic strata (Fig. 4) is described below.
The overall strike of the Maran thrust and fold belt
changes, and the common plunge directions reverse, across
the proposed northeast-trending cross-strike structural dis-
continuity (Fig. 5). These thin-skinned Eocene structures, at-
tributed to the Incaic orogeny, are the dominant tectonic ele-
ments in the region, although two regional unconformities in
the Cretaceous succession in central Peru, one at the base of
the Jumasha Formation and the other below the Casapalca
Group and its equivalents, represent earlier emergence dur-
ing, respectively, the Mochica and Peruvian tectonic phases
(Noble et al., 1979; Mgard, 1984). The Maran thrust and
fold belt extends from 5 S to 12 30' S, generally striking
north-northwestsouth-southeast, parallel to the present
892 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 892
Uchupata Type-Sections
Celendn Formation (163 m, from Section 19)
Marl: light gray, nodular, soft, white weathering
Shale: calcareous, slightly silty, yellowish, with
a few interbeds of dark-brown limestone
Marl: light gray, nodular, white weathering, with
sparse interbeds of limestone
Marl: light gray, nodular, soft, white weathering
Marl: light gray to tan, nodular, with few interbeds
of massive light-gray limestone
Jumasha Formation (820 m, from Section 20)

Limestone: medium gray, thick bedded,
weathering dark dove gray,
Foraminifera-bearing


Limestone: argillaceous
Limestone: medium gray, massive, thick
bedded, weathering dark-brownish gray
Dolostone: thin bedded, brown
Dolostone: light gray to orange-brown,
massive, thick-bedded, karstic, weathering
dark orange-brown

Dolostone: silty, medium gray, somewhat nodular
Antamina Mine Area
Upper Sequence
Thin-bedded shaly limestone: dark gray to black, weathers
medium to light gray, thin- to thick-bedded, generally very fine
grained, dominated by biomicrite to microsparite, ranges in
carbonate content from muddy calcisiltite with 50 - 75 %
calcite to calcareous siltstone with < 50 % calcite, variably
nodular, 10-90% nodules 1-15 cm diameter, nodules are
rounded to multilobate and have sharp to indistinct contacts
with the matrix
Lower Sequence
Impure limestone interbedded near the top:
rare, medium to thick beds (20-50 cm), very dark gray to
black, silty limestone, weathers medium gray
Predominantly thick-bedded, relatively pure limestone:
dark gray, medium- to very thick-bedded, ranges in grain size
from mudstone to wackestone, weathers light to pale gray
Locally fossiliferous bioclastic wackestone:
with broken pelecypods and gastropods, but no apparent
diagnostic fauna
S
K
A
R
N
Campanian
- ? - - - ? - - - ? -
Santonian
- ? - - - ? - - - ? -
Coniacian

Turonian
- ? - - - ? - - - ? -
Cenomanian
- ? - - - ? - - - ? -
Upper Albian
0
100
200 m
Outline of Regional Stratigraphy

M Eocene - M Miocene Calipuy Supergroup

Campanian - Paleocene Casapalca Group

Albian - U Cretaceous Machay Group
Coniacian - Santonian Celendn Formation
U Albian - Turonian Jumasha Formation
M Albian Pariatambo Formation
L - M Albian Chulec Formation
L Albian Pariahuanca Formation

L Cretaceous Goyllarisquisga Group
(Berriasian - Aptian)
U Jurassic Chicama Group

U Triassic - L Jurassic Pucar Group
L Permian - L Triassic Mitu Group

Mississippian Ambo Group
FIG. 4. Inferred stratigraphic column in the mine area compared with that for the Jumasha and Celendn Formations,
compiled from observations (Benavides, 1956) on the measured sections in the Ro Puchca valley, approximately 20 km north
of Antamina. The contact between the Jumasha and Celendn Formations coincides with the boundary between the Turon-
ian and Coniacian stages and is indicated with a solid line; unknown stage boundary locations have been approximated and
indicated with dashed lines and question marks. The stratigraphic interval interpreted to host the skarn is shown. Inset shows
an outline of the regional stratigraphic section.
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 893
0361-0128/98/000/000-00 $6.00 893
FIG. 5. Structural geology of the Maran thrust and fold belt in the vicinity of Antamina, illustrating the marked change
in the orientations of thrust faults and fold axes across a northeast-trending zone through Antamina (after Egeler and de
Booy, 1956; Wilson et al., 1967, 1995; Cobbing et al., 1996; Jacay, 1996; and Strusievicz et al., 2000). The northeast-south-
westtrending loci of these changes in structural attitude is indicated by the heavy dashed line, which is offset in a left-step-
ping sense in the vicinity of Antamina. This area is the location of the proposed cross-strike structural discontinuity discussed
in the text. Inset shows the location of the map area relative to the Huari (19-i), Singa (19-j), Requay (20-i) and La Unin
(20-j) quadrangles. No attempt has been made to establish continuity between the map units of Wilson et al. (1967, 1995)
north of 9 30' S and those of Cobbing et al. (1996) to the south. The undated volcanic rocks of Pampa Junn have been as-
signed to the Calipuy Supergroup (see text for discussion).
894 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 894
FIG. 6. Simplified geology of approximately 8,000 km
2
of the Maran thrust and fold belt in the region around Antam-
ina; the dominantly upper Miocene Cordillera Blanca batholith is removed (cf. Figs. 1 and 5), reflecting the geology at the
time of intrusion and mineralization in the late Miocene. The northeast-southwest trending cross-strike structural disconti-
nuity that passes through Antamina is delimited by heavy dashed lines. Mississippian to Lower Jurassic sedimentary rocks
are absent beneath the Cretaceous Goyllarisquisga Group northeast of Antamina along the cross-strike structural disconti-
nuity but are present north and southeast of the cross-strike structural discontinuity, except where cut out by faulting. An un-
usual abundance of igneous bodies intrudes the Maran Belt along the cross-strike structural discontinuity, compared to
transects to the north and south. After Egeler and de Booy (1956), Wilson et al. (1967, 1995), Cobbing et al. (1996), Jacay
(1996), and Strusievicz et al. (2000).
plate boundary, and comprises structures that predomi-
nantly verge northeast (Mgard, 1984; Fig. 5). However,
southeast of the cross-strike structural discontinuity, folds
and thrust faults strike north-northwest, whereas to the
northwest of it, they strike northerly (Figs. 1, 5, and 6).
Moreover, fold plunges are reversed across this zone: to the
southeast, most major anticlines and synclines plunge to the
south-southeast, whereas to the northwest, they plunge
north (Fig. 5). The locus of changes in strike and plunge ex-
tends northeast from Laguna Querococha, but about 5 km
southwest of Antamina it steps 8 km to the north before con-
tinuing northeastward (Fig. 5). Faults with the same overall
northeast strike as the cross-strike structural discontinuity
control some present-day drainages, such as the northeast-
trending Ro Puchca valley, 20 km north of Antamina, which
is discordant to the overall north to north-northwest grain of
the terrain (Fig. 5).
The deflection in the Antamina area is one of several that
articulate the Maran thrust and fold belt. Northerly strikes
continue to Llamalln, 50 km to the north of Antamina in the
eastern part of the belt (Fig. 1). Still farther north, the over-
all north-northwest regional strike of the Maran thrust and
fold belt resumes. A comparable sharp deflection to north-
south strikes occurs at the northern end of the Cordillera
Blanca (Fig. 1), 175 km to the north-northwest of Antamina,
where the Casma-Pasto Bueno fault zone intersects the re-
gional north-northwest strikes (Rivera, 1996). Benavides
(1999) identified many segments in the fabric of the Maran
thrust and fold belt, including the two described above, al-
though he proffered a different mechanism for their forma-
tion, as discussed below.
Stratigraphic variations across the cross-strike structural
discontinuity: The contact relationships of the upper Paleo-
zoic and Mesozoic strata (Fig. 4) to the lower Paleozoic
Maran metamorphic complex vary in accordance with the
segmentation of the thrust and fold belt. Figure 6 shows the
relationships along the western margin of the Maran com-
plex throughout an area more extensive than that shown in
Figure 5. Mississippian to Lower Jurassic strata that normally
separate the pre-Ordovician metamorphic rocks from the
Cretaceous sedimentary rocks are absent near the proposed
cross-strike structural discontinuity. In the north-striking seg-
ment of the fold belt immediately north of the Antamina area,
the eastern limit of the Mississippian to Cretaceous succes-
sion is a subhorizontal unconformity that strikes north overall
but has an irregular surface trace owing to the incised topog-
raphy (Figs. 5 and 6). In the north-northweststriking seg-
ments of the belt, however, the eastern extent of the Meso-
zoic rocks is generally delimited by north-northweststriking
faults (e.g., southeast of Antamina, and northwest of Llamal-
ln; Fig. 6). Southeast of Antamina, the southwest margin of
the Maran metamorphic complex is largely defined by a se-
ries of major northeast-verging reverse faults that involved
basement, but the Mississippian to Lower Jurassic strata are
preserved between the pre-Ordovician metamorphic rocks
and the Cretaceous sedimentary rocks (Fig. 6). However, in
the north-northweststriking segment northwest of Llamal-
ln, the Maran complex is backthrust over the Cretaceous
rocks and the thickness of the Mississippian to Lower Juras-
sic strata is only locally apparent (Fig. 6).
Northeast of Antamina, Mississippian to Lower Jurassic
strata (Fig. 4) are absent along the proposed cross-strike
structural discontinuity, and the clastic rocks of the Lower
Cretaceous Goyllarisquisga Group lie unconformably on the
pre-Ordovician Maran metamorphic complex (Figs. 5 and
6). In contrast, north and southeast of the cross-strike struc-
tural discontinuity, a relatively thick sequence of Mississip-
pian to Lower Jurassic strata separates these units (Fig. 6).
Sandstones and shales of the Mississippian Ambo Group lo-
cally unconformably overlie the Maran complex near its
western limit. Similarly, continental sedimentary rocks and al-
kaline to subalkaline volcanic rocks of the Lower Permian
Mitu Group commonly overlie the Ambo Group and also lo-
cally lie unconformably on the Maran complex along its
western edge but are absent northeast of Antamina. Both
north and south-southeast of the cross-strike structural dis-
continuity, these siliciclastic sedimentary rocks are overlain by
Upper Triassic to Lower Jurassic carbonate rocks and shales
(Pucar Group). The absence of these three groups directly
northeast of Antamina records either a sub-Cretaceous ero-
sional unconformity or nondeposition from Mississippian to
Late Jurassic times. The cross-strike structural discontinuity
therefore also coincided with a topographic high or arch, at
least in the Middle to Late Jurassic, but possibly persisting
throughout the Mississippian to Jurassic interval.
The cross-strike structural discontinuities probably also in-
fluenced the distribution of the uppermost Cretaceous to Pa-
leocene red beds that unconformably overlie the Cretaceous
strata (Fig. 4). These are absent near the proposed cross-
strike structural discontinuity through the Antamina area, and
they also thin significantly near the more northerly Casma-
Pasto Bueno deflection, but they attain a considerable thick-
ness between these two transverse zones as well as to the
south of the arch (Fig. 1).
Igneous activity along the cross-strike structural disconti-
nuity: Between the Cordillera Blanca and the Maran meta-
morphic complex, the 1:100,000 quadrangle maps of Cobbing
et al. (1996) and Wilson et al. (1967, 1995) record a greater
abundance of Tertiary hypabyssal and extrusive rocks along
the proposed cross-strike structural discontinuity than to the
northwest or southeast (Fig. 6). In addition, west of the thrust
and fold belt, volcanic rocks of the Calipuy Supergroup
(Strusievicz et al., 2000) are abundant northwest and south-
east of this cross-strike structural discontinuity (e.g., in the
Nevado Huantsn area and the Cordillera Huayhuash; Fig.
6). These rocks are interpreted to be lower-middle Miocene
because hypabyssal intrusions associated with similar volcanic
rocks elsewhere in the region (Huaraz Group of the Calipuy
Supergroup; Fig. 5) have been shown by
40
Ar/
39
Ar step-heat-
ing geochronology to have persisted to 14.2 Ma (Strusievicz et
al., 2000; Love et al., 2001). However, the 115 km
2
middle
Miocene granodioritic Carhuish pluton crops out on the axis
of the proposed cross-strike structural discontinuity (Fig. 5),
representing deeper-seated rocks of broadly equivalent age
(13.7 Ma U/Pb zircon date on the main phase, Mukasa, 1984;
16.5 Ma K/Ar date on the marginal phase, Cobbing et al.,
1981). The volcanic rocks of Pampa Junn (Egeler and de
Booy, 1956) northeast of the Carhuish pluton (Fig. 5) have
not been dated, so their significance with regard to the trans-
verse structure is uncertain. However, we propose that the
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 895
0361-0128/98/000/000-00 $6.00 895
cross-strike structural discontinuity focused the distribution
of igneous activity as it diminished through the mid-Miocene,
prior to intrusion and mineralization at Antamina. In Figure
6 the arch is depicted with a width of approximately 20 km to
incorporate the left-stepping locus of changes in Incaic strikes
and plunges (Fig. 5), the absence of Mississippian to Lower
Jurassic rocks beneath the sub-Cretaceous unconformity at
the northeast end of the arch, the diameter of the Carhuish
pluton at its southwestern margin, and the array of Miocene
intrusions.
Local Geologic Setting
Sedimentary host rocks
The contact of the Jumasha and Celendn Formations
within the Machay Group has not previously been defined in
the Antamina mine area, owing to the intense skarn and horn-
fels development and to the paucity of biostratigraphic mark-
ers. The cliff-forming strata surrounding the Antamina deposit
were mapped initially by Bodenlos and Ericksen (1955) as ma-
rine limestones of the Jumasha Formation. J.J. Wilson (un-
published report to Cerro de Pasco Corp., Lima, 1959, in Pe-
tersen, 1965) observed that the beds on the northwest slopes
of Quebrada Antamina were stratigraphically higher than
those at similar elevations on the southeast side, but he did
not locate the interformational contact. Petersen (1965) de-
scribed the carbonate units as the Machay Formation, and
they were subsequently reassigned to the Jumasha Formation
by Cobbing et al. (1996).
The skarn-hosting limestone strata at Antamina are herein
subdivided into two sequences, the upper markedly more
shaly than the lower, and assigned to the Celendn and Ju-
masha Formations, respectively. These sequences are de-
scribed and compared with a section of the relevant strati-
graphic interval constructed from two sections measured by
Benavides (1956) at Uchupata in the Ro Puchca valley, 20 km
north of Antamina (Fig. 4). A measured section has not been
established for the Antamina minesite because of the struc-
tural complexity of the area (described below) and the ab-
sence of marker beds in the exposed host rocks. The inferred
stratigraphic interval occupied by the skarn is also indicated
in Figure 4. The upper sequence of thin-bedded rocks that
prior to mining underlay the ridge crests flanking Quebrada
Antamina (Fig. 2) is widely altered to hornfels adjacent to the
Antamina intrusive center and grades into shaly limestone
with increasing distance from it (Fig. 7). The lower sequence
896 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 896
FIG. 7. The two distinct host rock types of the Antamina skarn system: an upper sequence of thin-bedded, silty limestones,
here largely converted to calc-hornfelses, assigned to the Celendn Formation (Ce), and a lower sequence of thick-bedded,
relatively pure limestones that form marbles, interpreted as the Jumasha Formation (J). Locations from which the pho-
tographs were taken are indicated in Fig. 2. Photographs taken in 1997 through 1999; this ridge has now been largely re-
moved by mine development. (a) The southeast side of Quebrada Antamina, looking southeast. The contact (long black
dashes) between the Jumasha and Celendn Formations is placed at the top of the uppermost massive thick-bedded lime-
stone (pale gray band) deformed by the Antamina anticline. Note the irregular upper, southeastern contact of the skarn (long
white dashes), here largely confined to the Jumasha Formation. (b) Looking northeast at the head of Quebrada Antamina
showing the distinct control of bedding in the Celendn Formation on skarn development.
of calcitic marble contains only subordinate intercalated diop-
side-rich units and grades outward into predominantly thick-
bedded, relatively pure limestone (Fig. 7).
The lower sequence (Figs. 4, 7a, and 8a) generally contains
more than 75 percent calcite and comprises calcitic, variably
bioclastic limestones that range in grain size from mudstone
to wackestone. This sequence is generally thick-bedded and
massive and displays karstic weathering (Fig. 8b). Toward the
top of this sequence, medium to thick beds (2050 cm) of im-
pure, silty limestone are interbedded with the pure limestone
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 897
0361-0128/98/000/000-00 $6.00 897
FIG. 8. The Jumasha and Celendn Formations. (a) Cliff-forming thick beds of massive limestone on the southeast flank
of Quebrada Callapo (see Fig. 2 for location). (b) Massive limestone of the Jumasha Formation, showing fluted weathering
(south of Yanacancha area, Fig. 2). (c) Upper: massive (i.e., unlaminated and with no preferred orientation of fossils) pele-
cypod carbonate wackestone of the Jumasha Formation (DDH CMA-039, 155 m). Lower: sheared fossiliferous carbonate
wackestone, equivalent to upper piece, not marmorized, but with a strong preferred orientation (DDH CMA-039, 160 m).
(d) More recessive weathering, vegetated, thin-bedded limestones and marls of the Celendn Formation cropping out on the
southeast slopes of Quebrada Ayash near its intersection with Quebrada Tucush (see Fig. 2 for location). Photograph taken
in 1997; the lower half of this area is now obscured by construction of the tailings dam. (e) Nodular limestone of the Ce-
lendn Formation, consisting of light gray, coalescing, carbonate-rich nodules separated by wisps of dark gray calcareous silt-
stone (Fortuna Mine area, see Fig. 2; hammer for scale). (f) Upper: Celendn Formation nodular limestone (DDH CMA-
C3, 221 m). Lower: lenticular bedding interpreted as sheared nodular limestone of the Celendn Formation (DDH
CMA-C5, 87.5 m).
(Figs. 4 and 7a). This sequence hosts ore at the surface and at
depth in the southwest part of the deposit, but only in the
subsurface in the northeast. Few whole fossils are preserved
(Fig. 8c) and no ammonites were observed in the mine area,
precluding biostratigraphic correlations. However, we assign
this sequence to the Jumasha Formation on lithologic
grounds. Sheared fossiliferous carbonate wackestone occurs
locally in this sequence (Fig. 8c). In north-central Peru, Be-
navides (1956) described the Jumasha Formation as domi-
nated by massive, thick-bedded, light orange-brown to yel-
lowish-brown and gray, fossil-poor dolostones and limestones
that weather dark yellowish-brown to brownish-gray. The for-
mation has been further described as comprising topographi-
cally prominent, cliff-forming, light-gray limestones and yel-
lowish dolostones that are characteristically bioclastic
(Wilson, 1963). In the Uchupata section (Fig. 4), the upper
437 m is limestone, and the lower 353 m is dolostone. Unlike
the overlying Celendn Formation, this formation only rarely
contains calcareous siltstones, although it incorporates marly
limestone beds near its top (Jaillard, 1987). The upper limit of
the Jumasha Formation was defined lithostratigraphically
where medium- or thick-bedded limestones pass upward into
thin-bedded marls and limestones (Wilson, 1963).
The upper sequence in the mine area (Figs. 4, 7, and 8d)
comprises thin- to thick-bedded impure limestone-marl that
varies in carbonate/silicate ratio from relatively calcite-rich
muddy limestone (generally 5075% carbonate) to calcareous
siltstone (less than 50% carbonate). Many units contain light-
gray, calcitic nodules, composing 10 to 90 percent of the rock,
enclosed by dark, silty calcareous mudstone (Fig. 8e, f). No
limestone beds with siliceous nodules have been observed.
The nodular texture is interpreted to be diagenetic. Sheared
limestone with lenticular bedding exposed at the northeast
head of Quebrada Antamina represents deformed nodular
limestone in which the nodules have been flattened into
lenses during folding and faulting (Fig. 8f). The upper se-
quence lacks identifiable fossils but is assigned to the Ce-
lendn Formation on lithologic grounds. In the Uchupata sec-
tion, 20 km north of the mine, the Celendn Formation
comprises very soft, friable, fossil-poor, light greenish-gray,
nodular, moderately silty marls and calcareous shales (Fig. 4;
Benavides, 1956). This formation is generally medium-bed-
ded (0.30.8 m) and variably dolomitic, and it ranges from
fine-grained to pelletal (Wilson, 1963).
The contact between the Jumasha and Celendn Forma-
tions in the mine area is conformable and generally grada-
tional throughout several meters and is marked by upward-in-
creasing siltiness and decreasing bed thickness. This contact
is interpreted to be the top of the uppermost thick-bedded
limestone or marble unit on the basis of lithology and bedding
characteristics. This stratigraphic position is illustrated in Fig-
ure 7a for outcrops on the southeast wall of the Antamina val-
ley. The contact is folded by the anticlines exposed on the val-
ley walls adjacent to the southwestern part of the deposit, and
it dips northeast in the subsurface in the northeast part of the
deposit, as do the overlying strata exposed in that area. The
thin-bedded calc-hornfelses that predominate on the ridge
crests flanking Quebrada Antamina grade laterally away from
skarn into the marly, variably nodular sequence assigned to
the Celendn Formation, but they do not contain the shales
characteristic of its upper part. The relatively pure calcitic
marble of the Jumasha Formation that hosts ore at surface in
the southwest part of the deposit and at depth in the north-
east sector does not contain the thick dolostones characteris-
tic of the lower part of that formation, and no magnesian
skarn has been found. On this basis, we conclude that the
skarn developed in the upper part of the Jumasha Formation
and the lower part of the Celendn Formation (Fig. 4). How-
ever, we estimate that approximately three-quarters of the
mineralized exoskarn, and all of the wollastonite exoskarn, de-
veloped in the Jumasha Formation, and that a similar propor-
tion of the contiguous and genetically related endoskarn was
formed in the stock adjacent to that formation. It is not
known whether magnesian skarn developed in dolostone
below the limit of exploration and development drilling.
Alteration of sedimentary rocks adjacent to the skarn
The outer contact of coarse-grained mineralized exoskarn is
abrupt, convoluted, and uninfluenced by bedding in the mar-
bles formed from the limestones of the Jumasha Formation
(Fig. 7a), but it is more gradational and stratigraphically con-
trolled in the fine-grained calc-silicate rocks developed in the
interbedded limestones and marls of the overlying Celendn
Formation (Fig. 7b). In the Jumasha Formation, the mineral-
ized skarn is juxtaposed with marbles containing local hori-
zons rich in calc-silicate minerals that may be ascribed to
thermal metamorphism. In contrast, the calc-silicate-bearing
rocks developed in the marly Celendn Formation include
rock types that are interpreted as the products of either meta-
morphism or metasomatism. The latter do not host sulfide
minerals and differ radically in texture and mineralogy from
the exoskarn ore. Their wide distribution around the upper
part of the orebody constitutes a significant exploration tar-
get.
Jumasha Formation: On approaching skarn, dark gray Ju-
masha limestone is converted to coarsely crystalline, gray to
white calcitic marble (Table 1, Fig. 9), forming an aureole
ranging from tens to hundreds of meters wide. Within the au-
reole, local development of sparse diopside, wollastonite, or
scapolite porphyroblasts (Fig. 9a) or slight differential erosion
and color variation (Fig. 9b) in calcite marble record minor
compositional variations in the limestone. Rare, fine-grained,
light green, diopsidic layers in calcite marble (Fig. 9c), the
porcellanite of Terrones (1958), are interpreted as calc-horn-
fels, representing thermally metamorphosed, medium to
thick beds of dolomitic, muddy, fine-grained limestone near
the top of the formation.
Within the marble aureole, medium-grained, mottled, or
banded gray marble predominates, locally grading inward to
medium- to coarse-grained, pure white marble, reflecting ei-
ther degraphitization or metamorphism of organic matter to
clear vitrinite. Rare, medium- to coarse-grained, thin, buff
garnet layers in both white and mottled facies of marble
mimic the forms of folded silicate-rich laminae (Fig. 9d) and
appear to have replaced them. This type of garnet develop-
ment is interpreted as a bimetasomatic reaction skarn. In gray
marble, scapolite is locally developed in some darker-gray
bands (Fig. 9e) but does not persist into white marble or
skarn (Table 1). Scapolite therefore forms a discontinuous
halo up to tens of meters wide and commonly separated from
898 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 898
the skarn front by tens of meters. White marble develops in
patches (Fig. 9d and f) and is concentrated along stylolites
and fractures (Fig. 9f). The latter probably represent fluid
pathways, suggesting that development of white marble in-
volved the local channeling of fluids. Such pathways were not
consistently used by later skarn-related fluids and are locally
cut by fractures with wollastonite-rich selvages (Fig. 9f).
Immediately southwest of the skarn, the Jumasha Forma-
tion has a distinct planar banded fabric, locally a spaced cleav-
age, that dips northeast at a high angle to bedding (Fig. 9a).
To the northeast, this fabric steepens and becomes vertical
adjacent to the skarn and along the strike of a major fold axis.
The consistent strike, systematically changing dip, and rela-
tionship to folds visible in the adjacent limestones indicate
that this is an upward-fanning cleavage associated with local
folding. In the adjacent skarn, alternating andradite-rich and
sphalerite-rich bands, generally 5 to 50 cm wide, are subpar-
allel to this fabric, suggesting that at least some preexisting
structures influenced mineralization. This banding may cor-
respond with the coarse-grained garnet-defined fabric ob-
served by Terrones (1958).
Celendn Formation: In the Celendn Formation, on the
ridges around Lago Antamina, the skarn has a halo of miner-
alogically diverse fine-grained calc-silicate rocks with subcon-
choidal to conchoidal fractures. With increasing distance
from the skarn, these rocks grade into interbedded limestone
and marl that reacted differentially to thermal metamorphism
around the porphyry intrusion and orebody, thus highlighting
the bedding (Fig. 7b).
Dark gray massive units change in color, mineralogy, and
grain size proximal to monzogranitic dikes. Although we have
not mapped the various facies of these rocks in detail, we rec-
ognize the development of three distinct zones: a distal very
fine grained, light-brown phlogopitic facies; an intermediate
fine-grained, light-gray, tremolitic facies; and a proximal
medium-grained, light-green diopsidic facies (Table 1, Fig.
10a, b). The boundaries between these facies are commonly
sharp and smooth but are locally irregular where controlled
by stockwork fractures (Fig. 10a, b). The outer limit of the
distal brown facies is typically hundreds of meters from the
boundary of sulfide-bearing skarn, and the progression from
brown through gray to green facies occurs over distances
ranging from tens of meters adjacent to the main intrusion to
tens of centimeters (Fig 10b) adjacent to dikes extending be-
yond the main intrusion.
Although local fluid channeling and probably metasoma-
tism occurred at the boundaries of the facies, the systematic
mineralogic zonation (Table 1) most likely reflects increasing
temperature. The first appearance of light brown calc-horn-
fels corresponds with the phlogopite-in reaction, Dol + Kfs =
Phl + Cal, which occurs in the range 350 to 465C under ge-
ologically reasonable conditions (i.e., P = 1,000 bars and X
CO
2
= 0.1 to 0.9: Tracy and Frost, 1991). The gray calc-hornfels
contains tremolite with or without phlogopite, whereas brown
calc-hornfels is tremolite free (Table 1). The transition from
brown to gray calc-hornfels therefore is interpreted to repre-
sent a phlogopite-out reaction that coincides with the forma-
tion of tremolite, probably through the reaction Phl + Cal +
Qtz = Tr + Kfs (Tracy and Frost, 1991; Table 1). The light-
green calc-hornfels records the first appearance of massive
diopside (Table 1), although veinlets and small blebs of this
mineral locally occur farther from the intrusive contacts. This
zone probably formed through the reaction Tr + Qtz + Cal =
Di, which under similar pressure and X
CO
2
occurs at temper-
atures of between 405 and 495C (Tracy and Frost, 1991).
The local narrowness of the gray calc-hornfels facies and the
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 899
0361-0128/98/000/000-00 $6.00 899
TABLE 1. Mineral Assemblages in Calc-Hornfels in the Celendn Formation, and in Marbles of the Jumasha Formation,
Adjacent to Ore-Bearing Garnet Skarn, Antamina, Peru
1
Formation, rock type,
Facies
and mineralogy Distal Intermediate Proximal
Celendin Formation
2
Calc-hornfels Brown, very fine grained Gray, fine grained Green, fine to medium grained
Common minerals Calcite, anorthite, calcite, Calcite, quartz, tremolite Diopside, calcite, quartz, microcline
quartz, phlogopite
Minor minerals Diopside Diopside, Vesuvianite,
wollastonite, grossular,
phlogopite, hedenbergite
orthoclase, ferrosilite
Jumasha Formation
Marble Gray; fine to medium grained, Gray; medium grained, banded Gray; medium to coarse grained, massive
banded
Common minerals Calcite Calcite Calcite
Minor minerals Vitrinite or other organic matter Scapolite, vitrinite, graphite, Colorless vitrinite,
or other organic matter wollastonite,
diopside
1
Determined by X-ray powder diffraction using a Philips PANalytical XPert PRO diffraction system with XPert Plus and XPert HighScore software for
peak matching and phase identification
2
Distal, intermediate, and proximal facies zones are broader in the Celendn Formation than in the Jumasha Formation
900 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 900
FIG. 9. Marmorized Jumasha Formation. (a) Subtle bedding, defined by porphyroblast abundance (outlined) in calcitic
marble, dips moderately southwest and is overprinted by a northeast-dipping planar fabric (looking northwest, portal of bulk
sample adit, 450 m southwest of west shore of Lago Antamina; hammer for scale). (b) Subtle bedding, defined by slight color
difference and differential erosion (outlined), reflecting cryptic difference in porphyroblast abundance in calcitic marble
(looking northwest, 350 m southwest of west shore of Lago Antamina; 20 cm notebook in foreground for scale). (c) Medium-
bedded layer of light green, fine-grained, diopsidic calc-hornfels defines bedding within coarse-grained calcitic marble near
the top of the formation (looking northeast, 300 m southwest of the western shore of Lago Antamina; 15 cm ruler for scale).
(d) Upper: mottled gray calcitic Jumasha Formation marble with tightly folded, boudinaged, dark-gray silty laminae. Lower:
mottled white and gray marble with tightly folded buff garnet-rich layer, similar in shape to the silty layer in the upper core
(DDH CMA-086, 218 m). (e) Banded gray Jumasha Formation marble showing development of approximately 10 percent
black scapolite crystals within darker layers (DDH CMA-136, 232 m). (f) Upper: white marble in gray Jumasha Formation
marble both as patches throughout and locally around veinlets and stylolites (DDH CMA-136, 228 m). Lower: veinlet-con-
trolled white marble in gray Jumasha Formation marble cut by irregular veinlets with chalky-white wollastonitic selvages
(DDH CMA-236, 441 m). Scale bars in centimeters.
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 901
0361-0128/98/000/000-00 $6.00 901
FIG. 10. Hornfels, skarnoid, and wrigglite developed in the Celendn Formation. (a) Irregular stockwork of sealed frac-
tures controlling development of light-gray tremolitic calc-hornfels in light-brown phlogopitic calc-hornfels (on the southeast
ridge crest, center of Fig. 12c; 15 cm pencil for scale). (b) Calc-hornfels facies with distal, light-brown, very fine grained phl-
ogopitic calc-hornfels at upper left, light-gray tremolitic in the center, and proximal, light-green medium-grained diopsidic
calc-hornfels at lower right. These facies are unusually closely spaced because they are on the margin of a narrow dike, ap-
proximately 350 m southeast of the main intrusion (same location as [a]; 55 mm lens cap for scale on the far left). (c) Sparse,
large, medium-green diopsidic nodules in pale gray calc-hornfels (base of cliffs, northeast of Lago Antamina; 10 cm knife for
scale). (d) Very abundant coalescing diopsidic nodules with calcite-bearing calc-hornfels matrix (base of cliffs, east of Lago
Antamina; 15 cm pencil for scale). (e) Irregular concentric banding, interpreted by Bodenlos and Ericksen (1955) to be of
algal origin, but reinterpreted as wrigglite texture of metasomatic replacement origin (base of cliffs, east of Lago Antamina;
15 cm pencil for scale). (f) Banding developed in calc-hornfels and apparently controlled by northeast striking fractures (on
the ridge crest southeast of Antamina valley, just east of the east end of Fig. 12c; 55 mm lens cap for scale).
absence of tremolite in the phlogopite facies suggest that the
metamorphism took place at high X
CO
2
. Under such condi-
tions the diopside-in and phlogopite-in reactions are sepa-
rated by only circa 10C. Also, the development of tremolite
through the elimination of phlogopite is promoted under
these conditions, but its formation within the field of phlogo-
pite stability through the reaction Dol + Qtz = Tr + Cal is in-
hibited.
In contrast to these calc-hornfelses that formed under es-
sentially thermal metamorphic conditions, calc-silicate devel-
opment through metasomatism is revealed by development of
skarnoid in some beds of the Celendn Formation, extending
tens of meters beyond the mineralized skarn front. Skarnoid
is a descriptive term for calc-silicate rocks that are relatively
fine grained, calcium rich, and iron poor, and that reflect, at
least in part, an aluminosilicate component in the protolith
(Zharikov, 1970). It is genetically intermediate between a
purely metamorphic hornfels and a purely metasomatic,
fluid-controlled skarn, which is typically coarser grained and
does not as closely reflect the composition or texture of the
immediately surrounding rocks (Einaudi, 2000). In some
beds in the Celendn Formation, diopsidic nodules stand out
in relief against the calcite-bearing calc-silicate matrix (Fig.
10c, d). The nodules, locally concentrically banded, range
from >10 cm to <1 cm in diameter and from sparse to abun-
dant, independent of their size (Fig. 10c, d). They are inter-
preted as products of the preferential metasomatism of cal-
citic nodules (cf. Fig. 8e, f).
Both close to the skarn front and tens to hundreds of me-
ters from it, sulfide-free concentrically layered structures
occur in some massive (nonnodular) beds of altered Celendn
limestone (Fig. 10e). These were interpreted as algal struc-
tures by Bodenlos and Erickson (1955). However, they do not
have the requisite three-dimensional geometry, and no un-
ambiguous algal structures have been seen in unaltered lime-
stones in the area. Also, at the transition from brown to gray
facies in calc-hornfels, rhythmic banding is developed adja-
cent to through-going, northeast-striking fractures (Fig. 10f).
Similar laminated features occur in ore-bearing wollastonite
skarn. Because of their similarity to banded features de-
scribed in other skarns (e.g., Knopf, 1908; Eskola, 1951), we
interpret the layering as metasomatic wrigglite (Kwak and
Askins, 1981). From these relationships, we conclude that
minor metasomatism occurred beyond the extent of metallic
mineralization in the Celendn Formation.
Structural geology of the mine area
The Upper Cretaceous strata of the Machay Group that
host the skarn were intensely deformed in the Maran
thrust and fold belt. Southwest of the mine, the Jumasha For-
mation is thrust over the younger Celendn Formation, and
the Lower Cretaceous formations of the Goyllarisquisga
Group are thrust both over the Celendn Formation and over
each other in reverse stratigraphic order (Fig. 11).
Local folds and faults: The Antamina deposit is hosted by a
relatively flat-lying section of the Jumasha and Celendn For-
mations, is deformed by open folds, and is cut by numerous
small thrust-fault ramps and bedding-parallel thrust faults
(Fig. 11), all interpreted to have developed during the late
Eocene Incaic orogeny. Prior to open-pit development, the
steep slopes around the Antamina valley provided clear,
three-dimensional images of the local structural relationships
(Fig. 12), summarized in an isometric block diagram in Fig-
ure 13. The vergence of the thrust faults are interpreted from
cutoff angles, and it is assumed that the strata are all upright,
unless obviously overturned. The local thrusting and related
folding are described below in sequence from southwest to
northeast.
Southwest of Antamina, a northeast-verging thrust, herein
named the Yaquirsh-Buque Punta thrust (YBPT in Figs. 11,
12, and 13), has caused structural repetition of the Jumasha
Formation. A few hundred meters to the east of this fault, the
Jumasha Formation is thrust over the younger Celendn For-
mation on the northeast-verging Antamina thrust fault (mine
terminology; AT in Figs. 11, 12, and 13). The imbricate thrust
stack is developed in massive limestones of the Jumasha For-
mation (Fig. 8a) in the next transcurrent valley to the north-
west of Antamina (Quebrada Callapo in Fig. 2). This is part of
a duplex in which the Antamina thrust is the sole and one of
the unnamed thrusts west of the Yaquirsh-Buque Punta
thrust is the roof, although the latter has been extensively
eroded. These imbricate slices do not all continue south from
Quebrada Callapo into Quebrada Antamina because some
fault surfaces merge.
On the northwest slope of the Antamina valley, only the Ju-
masha Formation occurs between the Antamina thrust and
the Yaquirsh-Buque Punta thrust, whereas the Jumasha and
Celendn Formations in conformable stratigraphic sequence
separate these two thrusts on the opposing southeast slope
(Figs. 11, 12, and 13). The immediately underlying thrust
slice, carried on the northeast-verging Fortuna thrust (mine
terminology: FT), occurs only on the northwest slope of the
valley, east of and below the Antamina thrust. In this area, the
Fortuna thrust repeats the Celendn Formation and has a
small recumbent anticline in its hanging wall (Figs. 11, 12a,
and 13). However, neither this thrust nor its hanging-wall an-
ticline can be traced across the valley to its southeast side
(Fig. 12b, c), either because the thrust surfaces merged or be-
cause the branch line of the two thrusts plunges north and
therefore climbs above the erosion level to the south. The An-
tamina thrust cuts the shallowly dipping Fortuna thrust (Figs.
11, 12a, 13) as well as other minor, bedding-parallel thrusts in
the Celendn Formation.
Most of the folds and thrust faults at Antamina are north-
east verging, as is characteristic of the Maran thrust and
fold belt (Mgard, 1984), but two southwest-verging thrusts
occur southeast of Lago Antamina (Figs. 11, 12c, and 13).
These thrusts are interpreted as backthrusts related to the un-
derlying northeast-verging Oscarina thrust (mine terminol-
ogy: OT), which is the lowest in the immediate mine area.
This thrust can be traced along the crest of the southwest
slope of the Tucush valley (Fig. 2), east of Antamina. It also
repeats the Celendn Formation and deforms it into a hang-
ing-wall anticline (Figs. 11; 12a, b, d; and 13). Many other
minor thrust faults are recorded by ramps with bedding cut-
offs for distances of tens of meters, especially in the Celendn
Formation, but are not illustrated here.
Prominent northwest-striking, open, upright anticlines are
exposed above the deposit on the northwest and southeast
walls of the valley (Figs. 7a; 12a, c; and 13). They become
902 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 902
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 903
0361-0128/98/000/000-00 $6.00 903
FIG. 11. Local geology of the Antamina area (modified after Cobbing et al.,1996; Glover, 1997, unpublished report to
Compana Minera Antamina S.A., Lima; Palomino, 1997, unpublished report to Compana Minera Antamina S.A., Lima; and
Love and Clark, 1998, unpublished report to Compana Minera Antamina S.A., Lima). The physiography of this area is
shown in Fig. 2. The areas depicted in Fig. 13 are indicated by the three diagonal rectangles. AT = Antamina thrust, FT =
Fortuna thrust, VF = left-stepping Valley fault, YBPT = Yaquirsh-Buque Punta thrust.
904 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 904
FIG. 12. Photomosaics illustrating the structure and stratigraphy of the Antamina mine area. Locations from which the
photographs were taken are indicated in Fig. 2. The views in (a) and (b) look northwest, whereas those in (c) and (d) look
south. Photographs taken in 1997 through 1999; much of (a), (b), and (c) has now been removed by mine development, and
the skyline of (d) has been modified. (a) The northwest side of the Antamina valley; the irregular upper or northwestern con-
tact of the skarn (long white dashes) cuts off the stratigraphic contact (long black dashes) of the Jumasha (J) and Celendn
(Ce) Formations. (b) The southeast flank of the ridge extending northeast of Cerro Buque Punta (i.e., the opposite side of
the ridge illustrated in [c]), viewed from the Yanacancha area. (c) The southeast side of the Antamina valley; the irregular
upper, southeastern contact of the skarn (long white dashes) cuts off the stratigraphic contact (long black dashes) of the Ju-
masha (J) and Celendn (Ce) Formations. (d) The northwest flank of the ridge extending northeast of Cerro Yaquirsh (i.e.,
the opposite side of the ridge illustrated in [a]). Faults and fold axes are shown by solid black lines, the stratigraphic contact
between the Jumasha (J) and Celendn (Ce) Formations by long black dashes, the contact of the skarn by long white dashes,
intrusive contacts by solid white lines, intrusions by crosses. AT = Antamina thrust, Ce = Celendn Formation, FT = Fortuna
thrust, J = Jumasha Formation, OT = Oscarina thrust, VF = surface expression of the northeastern segment of the Valley
fault, YBPT = Yaquirsh-Buque Punta thrust.
broader and more open upward but the intervening synclines
are tight. Such concentric folds are common elsewhere in the
Maran thrust and fold belt (e.g., Coney, 1971). The largest
of the parallel folds visible on the flanks of the Antamina val-
ley has been widely referred to as the Antamina anticline, the
axial plane of which has an apparent dextral offset of approx-
imately 500 m in a northeast-southwest direction across the
valley. This offset has been variously ascribed to local bends
in the strike of the folds (Bodenlos and Ericksen, 1955), dex-
tral offset on a postulated northeast-striking Valley fault (Ter-
rones, 1958), transverse normal faulting (Petersen, 1965,) and
cross-folding (McKee et al., 1979). However, in this study we
show that the offset is best explained by a lateral ramp model.
In this model (Fig. 14), the anticlines represent fault-bend
folds related to hanging-wall cutoffs folded over the top of
northwest-striking frontal ramps that are offset because a
northeast-striking, 500-m-long transfer fault or lateral ramp
separates the frontal ramps in an apparent dextral sense. This
postulated structure, the Valley lateral ramp, would account
for the absence of corresponding offsets of other structures to
the southwest and northeast (Fig. 11).
Minor diking and skarn alteration extend beyond the main
intrusion and orebody to the southwest down Quebrada An-
tamina and to the northeast and east at the head of the valley.
The vertical dikes that extend east from the main porphyry
mass to the Rosita de Oro area (Figs. 11 and 12c) offset two
minor forethrusts and backthrusts, indicating that there was
minor, apparently vertical, faulting between the time of thrust
faulting and folding and that of intrusion and mineralization.
The orientation of these dikes changes considerably along
strike; near the main stock they are nearly vertical, strike east,
and cut strata (Fig 12c), whereas farther east beyond the
ridge crest they become sills and follow either strata or bed-
ding-parallel thrust flats, strike south, and dip moderately to
the west (Fig. 12b). The dikes become progressively more al-
tered to endoskarn toward the main intrusion, and although
they transect zoned exoskarn they do not appear to have in-
truded it. A similarly variable orientation is shown by a dike
that extends northeast from the northern shore of Lago Anta-
mina; it locally crosses strata at a steep angle but also follows
flat and moderately dipping strata (Fig. 12a). The parallel fea-
tures at opposite ends of the main intrusion and orebody sug-
gest an offset, left-stepping, northeast-southwest fracture
zone that is longer than the 500 m lateral ramp, the postu-
lated Valley fault (Figs. 3b and 11).
A well developed northeast-striking, widely spaced, nearly
vertical fracture set occurs in the sedimentary host rocks
throughout the area (e.g., on the southwest flank of Cerro
Racpe; Fig. 15a), and we interpret it as regional a-c jointing
related to folding. A similarly oriented fracture set is in-
tensely developed near the deposit, especially along the
southeast side of the Antamina valley (Fig. 15b, c). In many
places, calc-hornfels is more intensely developed and wide-
spread where these fractures are closely spaced. All of the
northeast-striking fractures may have been related to Eocene
folding and thrust faulting, but they were better developed in
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 905
0361-0128/98/000/000-00 $6.00 905
C
e
J
J
C
e
C
e
J
C
e
C
e
O
T
A
T
Y
B
P
T
O
T
N
E
S
W
C
e
J
C
e
C
e
C
e
J
0
5
0
0
1
0
0
0
m
C
e
J
J
J
J
C
e
C
e
J
J
C
e
A
T
A
A
A
A
C
e
J
J
V
F
V
F
V
L
R
C
e
C
e
O
T
J
J
O
T
C
e
C
e
Y
B
P
T
A
T
F
T
O
T
A
A
C
e
J
J
C
e
C
e
C
e
C
e
C
e
C
e
C
e
N
E
F
T
A
T
O
T
Y
B
P
T
C
e
S
W
J
J
C
e
C
e
J
C
e
C
e
FIG. 13. Schematic, isometric block diagram of the pre-Miocene geology of the Antamina area, looking north, summa-
rizing the major Eocene Incaic folding and thrust faulting in the host Cretaceous strata. In the late Miocene, intrusion and
formation of the Antamina deposit occurred in the central block on this diagram, localized by the Valley lateral ramp (VLR)
and the step in the Valley fault (VF). The locations of the three blocks are indicated by rectangles in Fig. 11. The Jumasha
(J) and Celendn (Ce) Formations are indicated. Faults and fold axes are shown by solid black lines, bedding by black dashes.
Other features abbreviated as in Figures 11 and 12.
the mine area because of flexure and tearing of strata above
a transfer fault or lateral ramp (discussed below). Following
intrusion, they provided access for fluids to the hornfels-
bearing strata.
In the Antamina district, therefore, the Jumasha and Ce-
lendn Formations have been thrust-faulted, folded, and jux-
taposed into a thick, complex thrust stack. The total thickness
of rock that overlay the site of mineralization in the late
Miocene cannot be estimated because the local thickness of
overlying uppermost Cretaceous and Paleocene red beds is
unknown and because it is unclear if the subaerial volcanic
rocks and associated sedimentary rocks of the Eocene to mid-
dle Miocene Calipuy Supergroup (Strusievicz et al., 2000) ex-
tended into this area.
Structural control of mineralization: As in most intrusion-
related skarn deposits, the most obvious feature controlling
mineralization at Antamina is the contact between the main
stock and the host limestones (Terrones, 1958; Petersen,
1965; Redwood, 1999). However, the southeast and northwest
intrusive contacts are themselves parallel to, and along strike
from, the two segments of the proposed northeast-striking
Valley fault. The hydrothermal breccia sheets that are com-
mon at or near the endoskarn-exoskarn contact also have this
orientation. The irregular zones of breccia and endoskarn
within the intrusion are interpreted to strike predominantly
north-south (Fig. 3b) and may have been controlled by cross-
cutting structures. The western end of the main body of the
intrusion also generally strikes north-south, as does the skarn
front in that area (Fig. 3b). The parallelogram shape, in plan,
of the main body of the intrusion and the surrounding skarn
is complicated by the network of anastomosing dikes with en-
velopes of fine-grained garnet skarn extending to higher ele-
vations to the east of Lago Antamina (Fig. 12c). At least one
of these dikes intrudes a normal fault on which movement oc-
curred between the time of thrust faulting and folding and
that of intrusion and mineralization (Fig. 12c). Several other
dikes with skarn envelopes also extend beyond the main mass
of porphyry and most are controlled by Incaic structures.
Along the southern edge of the porphyry and skarn, minor pe-
ripheral rhyodacitic dikes and sills and their skarn envelopes
locally follow bedding and thrust faults. Several minor Pb-Zn-
Ag veins (Fig. 2), some of which have been mined on a small
906 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 906
Foot-wall Block
Flat
b
Frontal Ramp
L
a
te
r
a
l R
a
m
p
transport
direction
Anticline
c d
L
a
te
ra
l A
n
tic
lin
e
Fault bend fold
Anticline
Hangingwall Cutoff
Anticline
Future
Hanging-wall
Block
Future
Foot-wall
Block
Trace of Future Fault
a
FIG. 14. Schematic diagram illustrating the lateral ramp model for the structural setting of the Antamina deposit. Look-
ing south so that the face of the lateral ramp is exposed, showing offset anticlines produced in the hanging wall of a thrust
fault by two frontal ramps separated by a lateral ramp. (a) Geometry of the fault prior to movement. (b) Geometry of the
footwall (hanging-wall block removed), showing the southwest-dipping frontal ramps linked by a northeast-striking lateral
ramp or transfer fault. (c) After minor thrust movement, offset ramp-cutoff anticlines produced in the hanging wall of the
thrust fault, with more intense northeast-striking fracturing above the lateral ramp, represented by dashed lines. (d) Hang-
ing-wall cutoff anticlines separated from fault-bend anticlines by additional thrust movement. Extensive fracturing developed
in the thrust sheet where it flexed over the lateral ramp.
scale, occur at or near the contacts between these dikes and
the host limestone, describing a circa 3 3 km area of dis-
persed hydrothermal activity centered on the Antamina skarn
(Bodenlos and Ericksen, 1955).
Discussion
Regional stratigraphic and tectonic relationships
The postulated Valley lateral ramp at Antamina may have
been controlled by underlying, northeast-striking basement
faults. Lateral ramps are thought to be largely the result of
the interaction of thrust sheets and old basement fracture sys-
tems (Pohn, 2000). Both Mgard (1987) and Benavides
(1999) proposed that the segmentation and articulation of the
Maran thrust and fold belt probably reflect control by base-
ment structures, although they differ in their interpretations
of the mechanism. Benavides (1999) attributed the segmen-
tation of the Maran thrust and fold belt to major northeast-
striking, dextral basement faults, without specifying when the
faults were active. Further, Bussell and Pitcher (1985) sug-
gested that a well developed set of northeast-striking, en ech-
elon, dextral faults in the Cretaceous-Paleogene Coastal
batholith may have controlled some of the contacts of early
Paleocene intrusive ring complexes (Fig. 16). However, no
significant tear faulting parallel to the northeast-trending de-
flections is apparent in the deformed Mesozoic cover strata,
suggesting that the basement faults have not been active since
the Jurassic, and that any dextral offsets are apparent, not
real. Mgard (1987), in contrast, proposed that en echelon
sinistral growth faults in the basement were the common
boundary of the miogeosyncline and Yauli shelf in the eastern
part of the West Peruvian trough, and controlled the attitude
of the present thrust and fold belt.
The effects of the cross-strike structural discontinuities on
sedimentation have varied through time and with rock type.
A compilation of documented stratigraphic columns (Fig.
16b) shows that from Huancayo in the south to 50 km north
of Antamina the Cretaceous strata vary in aggregate thickness
from approximately 500 to 1,500 m and thicken overall to the
north. Generally, only the easternmost sections have been
used in this compilation, minimizing the effect of thickness
variations perpendicular to the margin, and thus emphasizing
along-strike variations that may be related to segmentation of
the margin. It is evident that the Cretaceous strata exhibit no
clear relationship between thickness and proximity to a cross-
strike structural discontinuity.
A model for the geologic history of the Antamina region
The geologic history of the Antamina region from the for-
mation of the West Peruvian trough in the Jurassic to the
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 907
0361-0128/98/000/000-00 $6.00 907
FIG. 15. Northeast-striking, nearly vertical fracturing in the eastern and
southeastern part of the Antamina mine area. Locations from which the pho-
tographs were taken are indicated in Fig. 2. Photographs taken in 1998; much
of (b) and (c) has now been removed by mine development. (a) Looking east-
northeast at the southwest flank of Cerro Racpe where all bedding dips
southwest, toward the point of view. Widely spaced northeast-striking frac-
tures are expressed as lineaments that converge in the distance and are per-
pendicular to the strike of the Celendn (Ce) and upper Jumasha (J) Forma-
tions on the lower slope, but are more obvious as vertical fractures in the
lower Jumasha Formation on the steep upper slope. Long black dashes indi-
cate the stratigraphic contact between the Celendn and Jumasha Forma-
tions. (b) Looking north-northeast along the ridge crest on the southeast side
of the Antamina valley in the immediate vicinity of the mine. The more in-
tensely developed northeast-striking fracture set in the Celendn Formation
can be seen as closely spaced vertical fractures in the cliff face. (c) Looking
northeast along the ridge bounding the southeast side of the Antamina valley,
near its head, in the immediate mine area. The closely spaced northeast-
striking fractures in the Celendn Formation strike away from the point of
view across the steep slope in the lower-right foreground. The fractures are
less strongly developed along strike to the northeast on Cerro Aparina (upper
right).
Miocene intrusion and formation of the orebody is summa-
rized schematically in Figure 17. In this model, the margin-
parallel West Peruvian trough formed during the Middle or
Late Jurassic by extension on en echelon, northwest-striking
normal faults separated by northeast-striking transform faults
(Fig. 17a, b; Mgard, 1987). The normal faults on the western
margin of the basement are thus right-stepping, but the seg-
ments experienced no relative displacement and each north-
east-striking transform fault experienced only sinistral move-
ment. This distribution of growth faults in the Jurassic would
result in promontories and reentrants in the margin of the
West Peruvian trough, similar to those of the larger-scale
early Paleozoic eastern margin of North America (Thomas,
1977).
The Querococha arch coincided with an intermittent topo-
graphic high that developed at least in the Middle Jurassic, or
even throughout the Mississippian to Middle Jurassic inter-
val, and which also influenced the distribution of Cretaceous
carbonate rocks, Paleocene red beds, and Miocene igneous
rocks. It is apparent that the development of the northeast-
striking basement structures predated Jurassic-Cretaceous
sedimentation. The Querococha arch apparently influenced
the distribution of Mississippian to Lower Jurassic rocks
northeast of Antamina, resulting in either local nondeposition
in the Mississippian to Early Jurassic or a Middle to Late
Jurassic erosional unconformity (Fig. 17b). Additional minor
extension at any angle discordant to the northeast-trending
faults could have resulted in reactivation of the originally
908 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 908
0 100 200 kms
0 100 200 miles
YANACOCHA
Lima
Huaraz
Huarmey
Trujillo
Chimbote
Hunuco
Huancayo
La Oroya
Huancavelica
Cerro de Pasco
7
9

W
7
8

W
7
7

W
7
6

W
7
5

W
Maran
Coastal Batholith
Cordillera
6
9
11
14
17
20
22
D
B-14,15
B-16
B-19
Cajamarca
Rio
ANTAMINA
B
la
n
c
a
1
2

S
1
0

S
8

S
Casma
PASTO
BUENO
KEY
Section
Intrusion
Anticline
Fault
Segment boundaries
discussed in text
Other segment boundaries
of Benavides (1999)
7
PIERINA
1
1

S
9

S
N
Pacific Ocean
Querococha
Arch
Casma - Pasto
Bueno zone
a
b
14
Jatunhuasi
Basin,
S of La Oroya
& SW of
Huancayo
C
enom
anian
Albian (97.0 2)
clastic
sequence
C
re
ta
c
e
o
u
s
J
u
ra
s
s
ic
(1
4
4
.8

3
.1
)
calcareous
sequence
Campanian and
younger (< 83 4 Ma)
Ks-Ce
558m
Ks-Ca
800m
B-16
KsP-Ch
(equivalent to
top 1/5 of Ks-J )
>920m
Ks-R (equivalent to
lowest 1/5 of Ks-J)
B-14, 15
Ki-Cr
613m
Ki-G
>516m
135m
Paleogene
~~~~~~~~~~
Cretaceous
Upper
~~~~~~~~~~
Lower
163m
750m
B-19
190m
>413m
KsP-Ch
Ki-Cr
>180m
Ks-J
22
44m
280m
851m
337m
KsP-C
>250m
20
230m
167m
59m
135m
>200m
17
115m
232m
138m
185m
>200m
52m
11
44m
266m
>250m
617m
>500m
9
122m
600m
70m
174m
25m
>150m
6
KsP-C
Ki-Pt
Ki-Cl
138m
>485m
25m
433m
131m
>200m
D
Pucar Gp.
~700m
Carcapuquio Fm.
Chunumayo Fm.
Ki-G
~850m
Ki-Cl ~280m
Ki-Pt ~50m
Ks-J ~340m
KsP-P ~250m
>1000m
443m
39m
>280m
107m
25m
ANTAMINA
NORTH
(north of approx. 930' S)
Chota Fm. (KsP-Ch),
Celendn Fm.(Ks-Ce)
Cajamarca Fm. (Ks-Ca)
Coor Fm.
Romirn Fm.
Mujarrn Fm.
Rosa Fm. (Ks-R)
Crisnejas Fm. (Ki-Cr)
Inca Fm.
Goyllarisquisga Gp. (Ki-G)
NORTH-CENTRAL
(south of approx. 930' S)
Casalpalca Gp. (KsP-C)
Celendn Fm.(Ks-Ce)
Jumasha Fm. (Ks-J)
Pariatambo Fm. (Ki-Pt)
Chulec Fm. (Ki-Cl)
Pariahuanca Fm.
Goyllarisquisga Gp. (Ki-G)
Otuzco Gp.
Quilquian Gp.
Pulluicana Gp.
CENTRAL
Pocobamba Fm. (KsP-P)
Machay Group
(north-central and central)
FIG. 16. Segmentation and articulation of the Maran thrust and fold belt and its effect on Cretaceous sedimentation.
(a) Simplified geology of the western Andes of central and north-central Peru, showing the major anticlinal axes, faults and
segment boundaries in the Maran thrust and fold belt (Benavides, 1999), the major plutonic centers of the Coastal
batholith and Cordillera Blanca batholith (Pitcher et al., 1985), and the numbered locations of the sections used in (b). The
major producing mines, Yanacocha, Pierina, Antamina and Cerro de Pasco are also shown. (b) Longitudinal fence diagram
of the Cretaceous stratigraphic section of central and north-central Peru along the eastern margin of the Maran belt. Sec-
tions B-14, B-15, B-16, and B-19 from Benavides (1956), sections 6, 9, 11, 14, 17, 20, and 22 from Wilson (1963), and sec-
tion D from Manrique (1998). Antamina is located between sections 6 and B-19.
sinistral faults as north-side-down normal faults and hence
erosion of preexisting sedimentary rocks from the transverse
highs (Fig. 17b). The effect of the Querococha arch in the
Late Jurassic cannot be deduced because the Upper Jurassic
Chicama Group is not exposed in the eastern part of the
Maran thrust and fold belt.
In the Cretaceous, the Querococha arch and other trans-
verse structures that segment the Maran thrust and fold
belt had little apparent effect on the thickness of clastic
(Goyllarisquisga Group) and carbonate (Machay Group) sed-
imentary rocks on the Yauli shelf (Figs. 16b and 17c), but it
appears to have affected the distribution of the carbonate
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 909
0361-0128/98/000/000-00 $6.00 909
a b
c
b
a
s
e
m
e
n
t
M
id
d
le
J
u
ra
s
s
ic
m
id
d
le
-L
a
te
C
re
ta
c
e
o
u
s
N
Q
A
CPB
f
M
io
c
e
n
e
e
E
o
c
e
n
e
d
P
a
le
o
c
e
n
e
Miocene
intrusions
Paleocene
Casapalca Gp.
Albian - Upper Cretaceous
Machay Gp.
Upper Jurassic - Lower Cretaceous
Chicama & Goyllarisquizga Gps.
Upper Triassic - Lower Jurassic
Pucar Gp.
Mississippian - Lower Triassic
Ambo & Mitu Gps.
pre-Ordovician
Maran metamorphic complex
Faults
transform

normal

thrust
Fold axis, with regional plunge

Regional tectonic forces

Q
CPB
FIG. 17. Regional-scale schematic diagrams, looking east, illustrating the proposed structural evolution of the Antamina
area. (a) In the Middle or early Late Jurassic, the West Peruvian trough started to develop through formation of an en ech-
elon pattern of left-lateral growth faults on the western edge of the Maran metamorphic complex. The Mississippian to
Lower Jurassic sedimentary rocks that overlay the metamorphic complex are removed from the diagram for clarity (after M-
gard, 1987). (b) Also in the Middle or Late Jurassic, minor extension at an angle to the transform faults segmented the con-
tinental basement, producing on the promontories structural highs parallel to the original offsets, such as the Querococha
arch, along which the Mississippian to Lower Jurassic sedimentary rocks were eroded. In the Late Jurassic the Chicama
Group was deposited in the western, deeper-water part of the basin, not illustrated. (c) Cretaceous clastic and carbonate
(Goyllarisquisga and Machay Groups) sedimentation on the Yauli shelf was not strongly controlled by the segmentation al-
though carbonate facies may have extended farther seaward adjacent to promontories. (d) Latest Cretaceous to Paleocene
deposition of red beds (checkerboard patterned) was controlled by the segmentation, with depocenters in the reentrants. (e)
In the Eocene Incaic orogeny, major strike deflections and en echelon fold patterns that conform to the segmentation of the
basement developed in the Maran thrust and fold belt. Reentrants became structural salients, and promontories became
structural recesses. (f) In the Miocene, additional uplift of the Querococha arch reactivated basement structures, which al-
lowed intrusions to extend farther inland along the arch and to reach shallow levels. A = Antamina mine, CPB = Casma-
Pasto Bueno zone, Q = Querococha arch.
rocks. The Machay Group does not extend as far west to the
north of the arch as it does on the arch and to the south of it.
From the arch north the westernmost limit of outcrop of the
Machay Group therefore crosses the belt in a northeast-
southwest direction (Fig. 1). Thomas (1977) noted a similar
relationship between the extent of carbonate facies and the
locations of structural salients and recesses in the Appalachi-
ans. The distribution of latest Cretaceous and Paleocene red
beds has also been influenced by the transverse structures
(Fig. 17d). The red beds pinch out toward both the Quero-
cocha arch and the Casma-Pasto Bueno zone, and they at-
tain their greatest thickness between these arches in a struc-
tural salient, which would have been a reentrant and
depocenter in the margin of the West Peruvian trough at the
time of sedimentation.
Shortening and variations in the regional attitudes of folds
and thrust faults generated in the Eocene Incaic orogeny are
represented in Figure 17e. The folds and thrusts in the An-
tamina region constitute an articulated structural recess in
the margin of the craton, which probably formed through de-
formation around a basement promontory, the northwestern
edge of which is now delineated by the Querococha arch.
Because the orientation of folds and thrust faults in thin-
skinned tectonic belts generally reflects the underlying
ramps rather than the translation direction of deformation
(Pohn, 2000), the strike and articulation of the Maran
thrust and fold belt mimic the geometry of the basement, de-
spite variations in the Cenozoic plate convergence direction
and rate (Pardo-Casas and Molnar, 1987; Somoza, 1998;
Norabuena et al., 1999). Old transform faults in the margin
of the West Peruvian trough did not experience extensive
later strike-slip movement because the maximum shortening
direction was not parallel to the orientation of movement
during rifting. We propose that the sinuous configuration of
the mountain belt generally reproduces the original zig-zag
margin, although the articulation is pronounced on the east-
ern side of the belt, and shortening in the interior of the belt
had a smoothing effect on the segmentation. Thus, the re-
gional strike at the western extent of the Cretaceous carbon-
ate strata gradually changes throughout a distance of approx-
imately 175 km along strike, from northerly near Antamina
to north-northwesterly at the northwest end of the Cordillera
Blanca (Fig. 1). As Thomas (1977) concluded in the context
of eastern North America, the Maran thrust and fold belt
has formed a best-fit curve around old promontories and
reentrants.
The persistence of Miocene igneous rocks farther from the
main axis of the magmatic arc into the foreland thrust and
fold belt along a proposed basement transverse structure
(Fig. 17f) is consistent with observations in other belts. Ig-
neous intrusions have been documented directly over, and
elongated parallel to, three of the four lateral ramps in the
Appalachians analyzed in detail by Pohn (2000). Further-
more, the difference in exposure level of the Miocene ig-
neous rocks on and adjacent to the southwest end of the arch
implies that post-Eocene uplift enhanced the plunge rever-
sals of the Eocene folds across the arch (Love et al., 2001).
Thus, faults parallel to the arch may have accommodated its
uplift and furnished the structural anisotropies that provided
conduits for magma ascent in the late Miocene.
Regional effects on local structure and mineralization
At Antamina, the northeast-trending fracture set, the Valley
fault and the Valley lateral ramp are parallel to the regional
cross-strike structural discontinuity, the Querococha arch,
and may have been controlled by similarly oriented underly-
ing basement structures. At the local scale, the Valley lateral
ramp and the Antamina intrusion have been localized by the
left-stepping jog in the Valley fault. Regionally, about 5 km
southwest of Antamina, the northeast-trending locus of
changes in strike and plunge in the Maran thrust and fold
belt steps left by approximately 8 km (Fig. 5). Thus the step
in the Valley fault developed within, and mimics, a similar,
larger-scale, left-stepping jog within the Querococha arch.
The local-scale structural evolution of the Antamina area,
including intrusion and formation of the orebody in the late
Miocene, is summarized schematically in Figure 18. Whereas
other models, such as en echelon folding, could explain the
apparent dextral offset of the Antamina anticline, the lateral
ramp model is preferred here because it provides a locus for
later, northeast-elongated intrusion and hydrothermal activ-
ity. The northeast-striking fracture set in the host rocks pe-
ripheral to the ore at Antamina (Figs. 15 and 18a) was, we
contend, formed by deformation associated with thrust trans-
lation along the underlying and similarly oriented transfer
fault or lateral ramp (Fig. 14). The overall form of a thrust
sheet does not record its passage over a lateral ramp beyond
the limit of that ramp but, in overlying thrust sheets, longitu-
dinal fractures would form in the lateral anticline over the
ramp. These fractures would have sheared vertically in a
north-sidedown sense or opened owing to flexure above the
lateral ramp or transfer fault and would persist beyond it (Fig.
15d), providing evidence that a thrust sheet had traversed
such a ramp. At Antamina, the inferred Valley lateral ramp ex-
tends only 500 m from one offset anticlinal axis to the other,
but strong, northeast-striking, nearly vertical fracturing, in-
terpreted herein as evidence of tearing in the overriding
thrust sheet, extends farther northeast (Figs. 15 and 18a) and
is interpreted as trace evidence of the Valley lateral ramp.
The upward-fanning axial planar cleavage related to the An-
tamina anticline also formed at this time (Fig. 18a).
In this model, the left-stepping jog in the transcurrent Val-
ley fault localized the Eocene development of the lateral
ramp that resulted in formation of the offset anticlines and
also focused the Miocene igneous activity responsible for the
Antamina stock and skarn. Prior to intrusion of the main por-
phyry mass, the early east-striking dikes east of the main in-
trusion formed in association with east-west shortening local-
ized at the northeast end of the southwestern segment of the
Valley fault (Fig. 18b). Many of the peripheral dikes and min-
eralized veins beyond the main porphyry body intrude, or
branch from, the same transcurrent structures. The dike
branches were controlled by other preexisting structures, pre-
dominantly bedding and Incaic thrusts, and they may have in-
truded at this time. In the Jumasha Formation the develop-
ment of white marble was localized by preexisting axial
cleavages, stylolites, and fractures, providing early fluid path-
ways. Also, in the Celendn Formation calc-hornfels forma-
tion was localized in many places where the northeast-strik-
ing, nearly vertical fracture set was strongly developed.
910 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 910
The primary control on the skarn and some of the breccia
zones was the margin of the main intrusion, which we argue
was itself localized by the postulated step in the Valley fault.
The northwest and southeast margins of the main intrusion
appear to be defined by the northeast-trending Valley fault,
and the main mass of the intrusion occupies the left-stepping
gap between the two offset segments of this structure (Figs.
3b and 11). In such a system, any relaxation in east-west com-
pression could allow sinistral movement on oblique north-
east-trending structures such as the Valley fault and forma-
tion of extensional north-south-oriented structures within the
left-stepping jog (Fig. 18c), thereby providing a locus for
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 911
0361-0128/98/000/000-00 $6.00 911
e
d
V
a
l
l
e
y

F
a
u
l
t
V
a
l
l
e
y

F
a
u
l
t
A
ntam
ina A
nticline fault bend fold
A
ntam
ina A
nticline
fault bend fold
frontal ram
p
frontal ram
p
Valley Lateral R
am
p
upw
ards-
fanning
cleavage
f
r
a
c
t
u
r
e
s

p
a
r
a
l
l
e
l

t
o
f
a
u
l
t

a
n
d

l
a
t
e
r
a
l

r
a
m
p
a
N
b
c
Strike and dip of cleavage
Anticline
Thrust fault
Normal fault
Fractures
Dike
Intrusion
Regional east-west shortening
Regional east-west extension
FIG. 18. Schematic diagrams, looking east, illustrating the proposed local-scale structural evolution of the Antamina area.
(a) Structural elements in the Antamina area prior to intrusion and mineralization: frontal thrust ramps offset by the Valley
lateral ramp, which was localized by a jog in the transcurrent Valley fault; fault-bend-fold anticlines overlying the tops of the
offset frontal ramps; strongly developed northeast-striking fracturing parallel to and overlying the lateral ramp; and upward-
fanning axial cleavage in the overlying anticlines. (b) East-west compression oblique to the Valley fault induced north-south
extension and allowed intrusion of dikes in wing cracks localized at the end of one segment of the Valley fault. (c) Relaxation
of east-west compression allowed east-west extension manifested as north-southstriking normal faulting in the jog between
the two segments of the Valley fault. (d) Monzogranitic magma intruded the parallelogram-shaped extensional regime be-
tween the two segments of the Valley fault, and ultimately generated skarn ore. (e) Renewed east-west compression formed
east-weststriking veins in skarn.
intrusion (Fig. 18d). Development of northeast-striking brec-
cia zones within skarn on the northwest and southeast mar-
gins of the intrusion and north-striking breccia and endoskarn
zones within the intrusion represent, in this model, structural
reactivation of the major transcurrent faults and the minor
north-south extensional faults, respectively.
Although the recrystallization of the host rocks during skarn
development obscured evidence of controls by preexisting
fractures or bedding, we deduce that the intersecting struc-
tures that locally control ore grades in skarn may have origi-
nated at substantially different times. Terrones (1958) re-
ported higher exoskarn ore grades where a set of 100
mineralized sheeted veins intersects structures extending
from a northwest-striking anticlinal axis in marble, and which
we interpret as upward-fanning axial planar cleavages. The in-
tersecting sheeted vein set differs in orientation from the
northeast-striking fracture system related to the lateral ramp
described above. We propose that these veins, which cut
skarn and were therefore late, formed in tension fractures as-
sociated with renewed east-west shortening (Fig. 18e) after
the brief episode of relaxation. Thus an Incaic axial planar
cleavage was intersected by Miocene tension veins and devel-
oped a permeable path for hydrothermal fluids.
Implications for local exploration and ore genesis
The surrounding zones of marble in the Jumasha Forma-
tion, and particularly of hornfels and skarnoid in the Celendn
Formation, provide larger exploration targets than the skarn
itself. An isometric block diagram of the simplified geology of
the deposit (Fig. 19) shows the structure and alteration in the
adjacent host rocks. Gray marble in the Jumasha Formation
generally extends tens of meters beyond skarn, although lo-
cally it extends beyond 100 m, but the outer limit of the dis-
tal, brown, phlogopitic facies of calc-hornfels in the Celendn
Formation typically extends several hundreds of meters from
the boundary of sulfide-bearing skarn. The most extensive
halo around the skarn is represented by the swarm of Ag-
bearing Pb-Zn vein deposits (Fig. 2). These occur within and
beyond marble and calc-hornfels in a 9 km
2
area surrounding
Antamina and up to a kilometer from the skarn front.
The development of strong exoskarn mineralization may
have been promoted by the relatively pure calcitic limestone
of the Jumasha Formation, which experienced intense calcite
destruction and Ca mobilization. In contrast, the overlying
Celendn Formation, having been metamorphosed to horn-
fels, may have acted as a cap on this hydrothermal system,
912 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 912
FIG. 19. Schematic, exploded isometric block diagram of the Antamina deposit. Looking north, showing the major folds
and thrust faults, the Valley fault (VF), the extent of skarn development within and adjacent to the intrusion, the exten-
sive calc-hornfels and skarnoid development in the Celendn Formation, and the restricted marmorization of the Jumasha
Formation. The offset Antamina anticline (AA) is indicated, but the inferred Valley lateral ramp is hidden from view in
this perspective.
suppressing the upward and outward escape of fluids and
thereby fostering development of extensive endoskarn.
The same lithologic succession of pure carbonate overlain
by muddy carbonate to shale that could contain and concen-
trate a developing hydrothermal system, also occurs in the
lower transgressive sequence of the Machay Group where the
shelly Pariahuanca Formation limestone is overlain by the
marl of the Chulec Formation (Figs. 4 and 16). In addition, in
northern Peru, similar prospective successions may occur at
other levels within the upper transgressive sequence of the
Machay Group because at least six shallowing-upward marl to
limestone sequences (Jaillard, 1987) are recognized in the
five formations correlative with the Jumasha Formation.
Metallogenic and geotectonic implications
Several other ore deposits, in addition to Antamina, coin-
cide with deflections in the strike of the Maran thrust and
fold belt. At the northern extremity of the Cordillera Blanca
(Fig. 1), 175 km to the north-northwest of Antamina, signifi-
cant middle to late Miocene intrusion-related mineral de-
posits, such as the Magistral Cu-Mo skarn prospect and the
formerly productive Pasto Bueno W-Cu-Ag vein system (Fig.
1), are associated with the Casma-Pasto Bueno zone. Farther
north, the Yanacocha Au district is localized in the northeast-
trending trans-Andean Chicama-Yanacocha structural corri-
dor (Teal et al., 2002). Such large-scale structural controls on
the location of ore deposits in Peru have been suggested by
Vidal and Noble (1994), Petersen and Vidal (1996), and
Rivera (1996), but not at Antamina. Similarly, in the Ap-
palachians, two of the four lateral ramps studied by Pohn
(2000) are associated with an unusual abundance of mineral
deposits, and many minor mineral occurrences are associated
with another (Coleman, 1988a, b).
The Machay Group has long been recognized as a metallo-
genically important stratigraphic interval in central Peru (Pe-
tersen, 1965). It hosts many skarn and carbonate replace-
ment-type deposits such as, from northwest to southeast,
Magistral, Antamina, Tuco-Chira, Pachapaqui, Raura, Uchuc-
chacua, Chungar, Santander, Yauricocha, and Pursima Con-
cepcin. Equivalent Albian to Turonian carbonate rocks of
the Arcurquina and Ferrobamba Formations in south-central
Peru host Oligocene skarn deposits in the Andahuaylas-Yauri
belt (Noble et al., 1984; Soler et al., 1986).
The Eocene to Miocene Calipuy Supergroup resulted
from suprasubduction zone magmatism (Noble et al., 1999).
The terminal event of this Supergroup is represented by the
middle Miocene Carhuish pluton, dated at 13.7 Ma (U/Pb
zircon date, Mukasa, 1984), and coincided with the forma-
tion of many hydrothermal centers in the Cordillera Negra
such as the Pierina high-sulfidation epithermal Au-Ag de-
posit (Figs. 1 and 16; Strusievicz et al., 2000). Subsequently,
an approximately 5.5 m.y. hiatus in major igneous activity
preceded the late Miocene intrusion of the main phase of
the Cordillera Blanca batholith, the Cohup leucogranodior-
ite, at 8.2 0.2 Ma (McNulty et al., 1998). During this pe-
riod of relative magmatic quiescence, only minor volumes of
a tonalitic to dioritic marginal phase of the Cordillera Blanca
batholith were intruded (Beckinsale et al., 1985), and only a
few scattered hydrothermal centers developed in the
Cordillera Negra (Strusievicz et al., 2000). However, along
the Querococha arch, an apparent swarm of intrusive bodies
was emplaced during this interval, one of which generated
the Antamina hydrothermal system (Love et al., 2001).
Models for the genesis of giant porphyry Cu systems in the
central Andean orogen (Zentilli et al., 1988; Clark, 1993; Zen-
tilli and Maksaev, 1995; Richards, 2000) postulate that rapid
ascent of magma is an important contributing factor in ore
formation because it would minimize modification through
assimilation-fractional crystallization processes, which are en-
visaged to decrease the Cu content of melts. Further, an un-
restrictive structure may be necessary to allow small bodies of
felsic magma access to the upper crust, because their low ef-
fective buoyancy would normally result in slow ascent, cool-
ing, and hence deeper crystallization. Thus, during the
magmatic lull, the basement structures controlling the Que-
rococha arch may have provided the conduit necessary for
rapid emplacement of small volumes of fertile melt into the
upper crust, allowing the Antamina porphyry to crystallize in
a suitably shallow environment favorable for mineralization.
Therefore, we conclude that the confluence of fertile
Miocene magmatism and reactive carbonate strata, essential
for ore genesis at Antamina, was afforded by the Querococha
arch cross-strike structural discontinuity.
Conclusions
The world-class late Miocene Antamina skarn deposit is
hosted by deformed Upper Cretaceous carbonate strata of
the Machay Group. The relatively pure marbles and minor in-
tercalated calc-hornfels that host the ore deposit at surface in
the southwest and at depth in its northeast sector represent
the upper part of the Jumasha Formation, whereas the strata
that form the ridges around Antamina and host the upper-
northeast part of the deposit are assigned to the lower part of
the overlying, originally muddier, Celendn Formation (Fig.
18), here uncharacteristically resistant to erosion owing to
hornfels formation in proximity to the Antamina intrusive
center. Marble is developed for tens of meters adjacent to the
skarn in the Jumasha Formation, but distinctive calc-horn-
felses and skarnoids persist for hundreds of meters from the
skarn front in the overlying Celendn Formation (Fig. 18).
Around the Antamina deposit, both this and the 9 km
2
swarm
of Pb-Zn-Ag vein deposits provide much larger exploration
targets than the skarn itself. Moreover, the systematic miner-
alogic zonation from peripheral phlogopitic through
tremolitic to proximal diopsidic facies in the calc-hornfelses
provides a vector toward the intrusion and, by extension, the
associated mineralization.
The sedimentary succession that hosts the Antamina de-
posit was folded, thrust-faulted, and thickened into a complex
thrust stack during the late Eocene Incaic orogeny, which
generated the orogen-parallel Maran thrust and fold belt.
Within this stack, a left-stepping jog in the transverse Valley
fault apparently controlled the Eocene formation of the Val-
ley lateral ramp, and together they localized the Miocene in-
trusion and skarn. East-west diking east of the main intrusion
and the northeast end of one segment of the Valley fault is in-
terpreted as early and associated with north-south extension
related to east-west shortening. Relaxation of this compres-
sion allowed east-west extension, which would have been fo-
cused in the left-stepping jog in the Valley fault, forming the
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 913
0361-0128/98/000/000-00 $6.00 913
locus for the main mass of intrusion and the associated skarn
ore. Renewed east-west shortening could have again induced
north-south extension and resulted in the late east-west vein
system. The original intrusive contact unambiguously con-
trolled the location of the skarn, yet was itself controlled by
larger-scale structures.
The Antamina hydrothermal activity occurred in a regional-
scale, northeast-trending, cross-strike structural discontinuity,
the Querococha arch, which articulates the Maran thrust
and fold belt. About 5 km southwest of Antamina, the locus of
this articulation steps left along strike, a feature mimicked on
a smaller scale by the Valley fault. The arch is inferred to have
affected sedimentary and magmatic processes at least from
the Jurassic to the Miocene and to have been controlled by a
basement structure, perhaps a transform segment of the orig-
inal margin of the West Peruvian trough. The arch was the
northwestern edge of a promontory on which the Cretaceous
Machay Group carbonate rocks were deposited farther west
than elsewhere along the belt. It also allowed Miocene mag-
matism to extend toward the foreland and intrude the
Machay Group.
We envisage that the carbonate rocks of the Machay Group
provided both chemical and physical traps for ore-forming
fluids. Intense exoskarn developed in relatively pure Jumasha
Formation limestone, whereas the Celendn Formation horn-
felses capped this system, promoting recirculation of hy-
drothermal fluids and extensive endoskarn development. The
Querococha arch provided a suitable structure for the intru-
sion to reach the Machay Group at the hypabyssal depths re-
quired for fertile fluid release.
Acknowledgments
We thank Inmet Mining Corporation and Rio Algom Ltd.,
and especially Frank Balint and Kelly OConnor, for initial
support of this project, a contribution to the Queens Uni-
versity Central Andean Metallogenetic Project (QCAMP),
and for repeatedly employing J.K.G. to examine various
structural aspects of Antamina during 1997 and 1998. A post-
doctoral fellowship at Queens University for the senior au-
thor was funded in 1997 and 1998 by Inmet Mining Corpo-
ration and Rio Algom Ltd., and in 1998 through 2000 by Rio
Algom. James Macdonald of BHP Billiton, Bill Mercer of
Noranda, and John Thompson of Teck Cominco subse-
quently provided support and encouragement. This research
was also funded by Natural Science and Engineering Re-
search Council of Canada Discovery Grants to A.H.C. We
also thank Ca. Minera Antamina S.A. for unstinting logistic
support and are grateful to the many geologists involved in
the mine development, including Leo Hathaway, Stewart
Redwood, Manuel Pacheco, Jose Sales, Richard Ct, Diane
Nicolson, Matt Wunder, Rick Schwarz, Jeff Hussey, Scott
Smith and, especially, Eric Lipten, for stimulating discus-
sions of mine geology and ore genesis. We would also like to
thank the Economic Geology reviewers Gerry Ray, Greg
Dipple, Larry Meinert, and Andreas Mueller, associate edi-
tors Dave Cooke and Steve Garwin, and especially editor
Mark Hannington for their large contributions to the prepa-
ration of the final manuscript.
September 10, 2002; March 22, 2004
REFERENCES
Allende, T., 1996, Geologa del cuadrngulo de San Pedro de Chonta (18-j),
Boletn 68, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero
y Metalrgico, Sector Energa y Minas, Per, 218 p., 1 map, 1:100,000
scale.
Atherton, M.P., Pitcher, W.S., and Warden, V., 1983, The Mesozoic marginal
basin of central Peru: Nature, v. 305, no. 5932, p. 303306.
Barazangi, M., and Isacks, B., 1976, Spatial distribution of earthquakes and
subduction of the Nazca plate beneath South America: Geology, v. 4, p.
686692.
Beckinsale, R.D., Snchez-Fernndez, A.W., Brook, M., Cobbing, E.J., Tay-
lor, W.P., and Moore, N.D., 1985, Rb-Sr whole-rock isochron and K-Ar age
determinations for the Coastal batholith of Peru, in Pitcher, W.S., Atherton,
M.P., Cobbing, E.J., and Beckinsale, R.D., eds., Magmatism at a plate
edge: The Peruvian Andes: Glasgow, Blackie and Son Ltd., p. 177202.
Benavides, V., 1956, Cretaceous System in northern Peru: Bulletin of the
American Museum of Natural History, v. 108, p. 353494.
1999, Orogenic evolution of the Peruvian Andes: The Andean cycle, in
Skinner, B.J., ed., Geology and ore deposits of the central Andes: Society of
Economic Geologists Special Publication 7, p. 61107.
Bodenlos, A.J., and Ericksen, G.E., 1955, Lead-zinc deposits of Cordillera
Blanca and northern Cordillera Huayhuash, Peru: United States Geologi-
cal Survey Bulletin 1017, 102 p.
Broggi, J.A., 1942, Geologa del embalse del Ro Chotano en Lajas: Boletn
Sociedad Geolgica del Per, v. 12, p. 123.
Bussell, M.A., and Pitcher, W.S., 1985, The structural controls of batholith
emplacement, in Pitcher, W.S., Atherton, M.P., Cobbing, E.J., and Beckin-
sale, R.D., eds., Magmatism at a plate edge: The Peruvian Andes: Glasgow,
Blackie and Son Ltd., p. 167176.
Clark, A.H., 1993, Are outsize porphyry copper deposits either anatomically
or environmentally distinctive?, in Whiting, B.H., Hodgson, C.J., and
Mason, R., eds., Giant ore deposits: Society of Economic Geologists Spe-
cial Publication 2, p. 213283.
Cobbing, E.J., Pitcher, W.S., Wilson, J.J., Baldock, J.W., Taylor, W.P., Mc-
Court, W.J., and Snelling, N.J., 1981, The geology of the western Cordillera
of northern Peru: Institute of Geological Sciences (London) Overseas
Memoir 5, 143 p.
Cobbing, E.J., Snchez, A., Martinez, W., and Zrate, H., 1996, Geologa de
los cuadrngulos de Huaraz (20-h), Recuay (20-i), La Unin (20-j),
Chiquin (21-i), Yanahuanca (21-j), Boletn 76, Serie A: Carta Geolgica
Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y
Minas, Per, 281 p., 5 maps, 1:100,000 scale.
Coleman, J.L., Jr., 1988a, CSDs of the eastern United States, in Coleman,
J.L., Jr., Groshong, R.H., Jr., Rheams, K.F., Neathery, T.L., and Rheams,
L.J., eds., Structure of the Wills Valley anticline-Lookout Mountain syn-
cline between the Rising Fawn and Anniston CSDs, northeast Alabama:
Alabama Geological Society Annual Field Trip Guidebook 25, p. 4951.
1988b, Geology of the Anniston CSD, in Coleman, J.L., Jr., Groshong,
R.H., Jr., Rheams, K.F., Neathery, T.L., and Rheams, L.J., eds., Structure
of the Wills Valley anticline-Lookout Mountain syncline between the Ris-
ing Fawn and Anniston CSDs, northeast Alabama: Alabama Geological So-
ciety Annual Field Trip Guidebook 25, p. 4143.
Coney, P.J., 1971, Structural evolution of the Cordillera Huayhuash, Andes of
Peru: Geological Society of America Bulletin, v. 82, p. 18631883.
Cosso, A., 1964, Geologa del cuadrngulos de Santiago de Chuco (17-g) y
Santa Rosa (18-g), Boletn 08, Serie A: Carta Geolgica Nacional: Instituto
Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, 69 p., 2
maps, 1:100,000 scale.
Cosso, A., and Jan, H., 1967, Geologa de los cuadrngulos de Pumape
(16-d), Chocope (16-e), Otuzco (16-f), Trujillo (17-e), Salaverry (17-f) y
Santa (18-f), 1967, Boletn 17, Serie A: Carta Geolgica Nacional: Instituto
Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, 141 p., 4
maps, 1:100,000 scale.
Egeler, C.G., and De Booy, T., 1956, Geology and petrology of part of the
southern Cordillera Blanca, Peru: Verhandelingen van het Koninklijk Ned-
erlands Geologisch Mijnbouwkundig Genootschap, Geologische Serie, pt.
17, 86 p.
Einaudi, M.T., 1982a, Description of skarns associated with porphyry copper
plutons: Southwestern North America, in Titley, S.R., ed., Advances in ge-
ology of the porphyry copper deposits, Southwestern North America: Tus-
con, Arizona, University of Arizona Press, p. 139184.
1982b, General features and origins of skarns associated with porphyry
copper plutons: Southwestern North America, in Titley, S.R., ed., Advances
914 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 914
in geology of the porphyry copper deposits: Southwestern North America:
Tuscon, Arizona, University of Arizona Press, p. 185210.
Einaudi, M.T., 2000, Skarns of the Yerington district, Nevada: A triplog and
commentary, in Dilles, J.H., Barton, M.D., Johnson, D.A., Proffett, J.M.,
and Einaudi, M.T., eds., Contrasting styles of intrusion-associated hy-
drothermal systems: Society of Economic Geologists Guidebook Series, v.
32, pt. 1, p. 101125.
Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits:
ECONOMIC GEOLOGY SEVENTY-FIFTH ANNIVERSARY VOLUME, p. 317391.
Eskola, P., 1951, Around Pitkranta: Suomalaisen Tiedeakatemian Toimituk-
sia Sarja A (Annales Academi Scientiarum Fennic Series A), v. III, n. 27,
p. 196.
Gutscher, M.-A., Olivet, J.-L., Aslanian, D., Eissen, J.-P., and Maury, R.,
1999, The lost Inca Plateau: Cause of flat subduction beneath Peru?:
Earth and Planetary Science Letters, v. 171, p. 335341.
Hampel, A., 2002, The migration history of the Nazca Ridge along the Peru-
vian active margin: A re-evaluation: Earth and Planetary Science Letters, v.
203, p. 665679.
Harrison, J.V., 1940, The geology of the central Andes in part of the Province
of Junn, Peru: Quarterly Journal of the Geological Society, London, v. 99,
p. 136.
INGEMMET, 1999, Mapa geolgico del Per: Instituto Geolgico, Minero
y Metalrgico, Sector Energa y Minas, Per 1:1,000,000, Digital Version
CD-ROM.
Jacay, J., 1996, Geologa del cuadrngulo de Singa (19-j), Boletn 67, Serie A:
Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sec-
tor Energa y Minas, Per, 215 p., 1 map, 1:100,000 scale.
Jaillard, E., 1987, Sedimentary evolution of an active margin during middle
and upper Cretaceous times: The north Peruvian margin from late Aptian
up to Senonian: Geologische Rundschau, v. 76, p. 677697.
Knopf, A., 1908, Geology of the Seward Peninsula tin deposits, Alaska:
United States Geological Survey Bulletin 358, 71 p.
Kwak, T.A.P., and Askins, P.W., 1981, The nomenclature of carbonate re-
placement deposits, with emphasis on Sn-F(-Be-Zn) wrigglite skarns:
Journal of the Geological Society of Australia, v. 28, p. 123136.
Love, D.A., Clark, A.H., and Schwarz, F.P., 2000, The Antamina deposit, An-
cash, Peru: Anatomy and petrology of a giant copper skarn [abs.]: Geologi-
cal Society of America Abstracts with Programs, v. 32, no. 7, p. A137.
Love, D.A., Clark, A.H., Strusievicz, O.R., and Lee, J.K.W., 2001, The re-
gional tectonic setting of the giant Antamina Cu-Zn skarn deposit, north-
central Peru [abs.]: Geological Society of America Abstracts with Programs,
v. 33, no. 6, p. A358.
Love, D.A., Clark, A.H., Ullrich, T.D., Archibald, D.A., and Lee, J.K.W.,
2003,
40
Ar-
39
Ar evidence for the age and duration of magmatic-hydrother-
mal activity in the giant Antamina Cu-Zn skarn deposit, Ancash, north-cen-
tral Peru [abs.]: Geological Association of Canada/Mineralogical Associa-
tion of Canada/Society of Economic Geologists Abstracts, v. 28, Abstract
no. 396 CD-ROM.
Manrique, A.I., 1998, Promocin de la minera del carbn en El Peru: Ge-
ologa economica de las cuencas de Alto Chicama, Santa, Oyn y Jatun-
huasi: PROCARBON, Instituto Geolgico, Minero y Metalrgico, Sector
Energa y Minas, Per, 66 p.
McKee, E.H., Noble, D.C., Scherkenbach, D.A., Drexler, J.W., Mendoza, J.,
and Eyzaguirre, V.R., 1979, Age of porphyry intrusion, potassic alteration,
and related Cu-Zn skarn mineralization, Antamina district, northern Peru:
ECONOMIC GEOLOGY, v. 74, p. 928930.
McLaughlin, D.H., 1924, Geology and physiography of the Peruvian
cordillera, Departments of Junn and Lima: Geological Society of America
Bulletin, v. 35, p. 591632.
McNulty, B.A., Farber, D.L., Wallace, G.S., Lopez, R., and Palacios, O.,
1998, Role of plate kinematics and plate-slip-vector partitioning in conti-
nental magmatic arcs: Evidence from the Cordillera Blanca, Peru: Geology,
v. 26, p. 827830.
Mgard, F., 1984, The Andean orogenic period and its major structures in
central and northern Peru: Journal of the Geological Society, London, v.
141, p. 893900.
1987, Structure and evolution of the Peruvian Andes, in Schaer, J.-P.,
and Rodgers, J., eds., The anatomy of mountain ranges: Princeton, New
Jersey, Princeton University Press, p. 179210.
Mourier, T., Bengtson, P., Bonhomme, M., Buge, E., Cappetta, H., Crochet,
J.-Y., Feist, M., Hirsch, K.F., Jaillard, E., Laubacher, G., Lefranc, J.P., Moul-
lade, M., Noblet, C., Pons, D., Rey, J., Sig, B., Tambareau, Y., and Taquet,
P., 1988, The Upper Cretaceous-lower Tertiary marine to continental
transition in the Bagua basin, northern Peru: Paleontology, biostratigraphy,
radiometry, correlations: Newsletters on Stratigraphy, v. 19, p. 143177.
Mukasa, S.B., 1984, Comparative Pb isotope systematics and zircon U-Pb
geochronology for the Coastal, San Nicols and Cordillera Blanca
batholiths, Peru: Unpublished Ph.D. dissertation, University of California,
Santa Barbara, 362 p.
Myers, J.S., 1974, Cretaceous stratigraphy and structure, western Andes of
Peru between latitudes 10 and 10 30': American Association of Petroleum
Geologists Bulletin, v. 58, p. 474487.
1975, Vertical crustal movements of the Andes in Peru: Nature, v. 254,
p. 672674.
1976, Erosion surfaces and ignimbrite eruptions, measures of Andean
uplift in northern Peru: Geological Journal, v. 11, p. 2944.
1980, Geologa de los cuadrngulos de Huarmey (21-g) y Huayllapampa
(21-h), Boletn 33, Serie A: Carta Geolgica Nacional: Instituto Geolgico,
Minero y Metalrgico, Sector Energa y Minas, Per, 153 p., 1 map,
1:100,000 scale.
Noble, D.C., McKee, E.H., and Mgard, F., 1979, Early Tertiary Incaic
tectonism, uplift and volcanic activity, Andes of central Peru: Geological
Society of America Bulletin, v. 90, p. 903907.
Noble, D.C., McKee, E.H., Eyzaguirre, V.R., and Marocco, R., 1984, Age
and regional tectonic and metallogenetic implications of igneous activity
and mineralization in the Andahuaylas-Yauri belt of southern Peru: ECO-
NOMIC GEOLOGY, v. 79, 172176.
Noble, D.C., Wise, J.M., Vidal, C.E., 1999, Episodes of Cenozoic extension
in the Andean orogen of Peru and their relation to compression, magmatic
activity and mineralization, in Machar, J. Benavides, V., and Rosas, S.,
eds., 75 Anniversario de la Sociedad Geologica del Peru, Julio, 1999: So-
ciedad Geologica del Peru, Volumen Jubilar 5, p. 4566.
Norabuena, E., Dixon, T., Stein, S., and Harrison, C.G.A., 1999, Decelerat-
ing Nazca-South America and Nazca-Pacific motions: Geophysical Re-
search Letters, v. 26, p. 34053408.
OConnor, K., 2000, Yacimiento polimetlico Antamina: historia, exploracin
y geologa, in primer volumen de monografas de yacimientos minerales
PeruanosHistoria, exploracin y geologa, volumen Luis Hochschild
Plaut: Lima, Peru, Instituto de Ingenieros de Minas del Peru, p. 231244.
Pardo-Casas, F., and Molnar, P., 1987, Relative motion of the Nazca (Faral-
lon) and South American plates since Late Cretaceous time: Tectonics, v. 6,
p. 233248.
Petersen, U., 1965, Regional geology and major ore deposits of central Peru:
ECONOMIC GEOLOGY, v. 60, p. 407476.
Petersen, U., and Vidal C., C.E., 1996, Magmatic and tectonic controls on
the nature and distribution of copper deposits in Peru, in Camus, F., Silli-
toe, R.S., and Peterson, R., eds., Andean copper deposits: New discoveries,
mineralization, styles and metallogeny: Society of Economic Geologists
Special Publication 5, p. 118.
Pilger, R.H., 1981, Plate reconstructions, aseismic ridges, and low-angle sub-
duction beneath the Andes: Geological Society of America Bulletin, v. 92,
p. 448456.
Pitcher, W.S., Atherton, M.P., Cobbing, E.J., and Beckinsale, R.D., eds.,
1985, Magmatism at a plate edge: The Peruvian Andes: Glasgow, Blackie
and Son Ltd., 328 p.
Pohn, H.A., 2000, Lateral ramps in the folded Appalachians and in overthrust
belts worldwideA fundamental element of thrust-belt architecture:
United States Geological Survey Bulletin 2163, 71 p.
Redwood, S.D., 1998, The Antamina copper-zinc skarn deposit, northern
Peru [abs.]: Geological Association of Canada/Mineralogical Association
of Canada/Canadian Geophysical Union Program with Abstracts, v. 23, p.
153.
1999, The geology of the Antamina copper-zinc skarn deposit, Peru: The
Gangue, Newsletter of the Mineral Deposits Division, Geological Associa-
tion of Canada, issue 60, p. 17.
Reyes, L., 1980, Geologa de los cuadrngulos de Cajamarca (15-f), San Mar-
cos (15-g) y Cajabamba (16-g), Boletn 31, Serie A: Carta Geolgica Na-
cional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas,
Per, 70 p., 3 maps, 1:100,000 scale.
Richards, J.P., 2000, Lineaments revisited: Society of Economic Geologists
Newsletter, 42, p. 1 and 1420.
Rivera, J.N., 1996, El megafracturamiento pre-Mendaa de Casma-Pasto
Bueno y su influencia en la metalognia Andina: Instituto de Ingenieros de
Minas del Per Revista Minera, v. 240, p. 612.
Samam-Boggio, M., 1980, ed., El Per Minero: Instituto Geolgico, Minero
y Metalrgico, Sector Energa y Minas, Lima, Per, 18 volumes.
LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 915
0361-0128/98/000/000-00 $6.00 915
Snchez, A., 1995, Geologa de los cuadrngulos de Culebras (20-g), Casma
(19-g), Chimbote (19-f), Boletn 59, Serie A: Carta Geolgica Nacional: In-
stituto Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per,
263 p., 3 maps, 1:100,000 scale.
Snchez, J., lvarez, D., and Lagos, A., 1998, Geologa de los cuadrngulos
de Juscusbamba (16-i) y Plvora (16-j), Boletn 119, Serie A: Carta Ge-
olgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector En-
erga y Minas, Per, 262 p., 2 maps, 1:100,000 scale.
Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe,
O., Neeraudeau, D., Crdenas, J., Rosas, S., and Jimnez, N., 2002, Late
Permian-Middle Jurassic lithospheric thinning in Peru and Bolivia, and its
bearing on Andean-age tectonics: Tectonophysics, v. 345, p. 153181.
Soler, P., Grandin, G., and Fornari, M., 1986, Essai de synthse sur la mtal-
logenie du Prou: Godynamique, v. 1, p. 3368.
Somoza, R., 1998, Updated Nazca (Farallon)-South America relative motions
during the last 40 m.y.: Implications for mountain building in the central
Andean region: Journal of South American Earth Sciences, v. 11, p.
211215.
Strusievicz, O.R., Clark, A.H., Lee, J.K.W., Farrar, E., Slauenwhite, M., and
Hodgson, C.J., 2000, Metallogenetic relationships of the Huaraz, Ancash,
segment of the precious-base metal subprovince of northern Peru [abs.]:
Geological Society of America Abstracts with Programs, v. 32, no. 7, p.
A504.
Szekely, T.S., 1967, Geology near Huallacocha Lakes, central high Andes,
Peru: American Association of Petroleum Geologists Bulletin, v. 51, p.
13461353.
Teal, L., Harvey, B., Williams, C., and Goldie, M., 2002, Geology of the Yana-
cocha gold deposits, northern Peru [extended abs.], in Marsh, E.E., Gold-
farb, R.J., and Day, W.C., eds., Global exploration 2002: Integrated meth-
ods for discovery: Denver, Colorado, Society of Economic Geologists,
Abstracts, p. 4344.
Terrones, A.J., 1958, Structural control of contact metasomatic deposits in
the Peruvian cordillera: Mining Engineering, Transactions of the American
Institute of Mining Engineers, v. 11, p. 365372.
Thomas, W.A., 1977, Evolution of Appalachian-Ouachita salients and re-
cesses from reentrants and promontories in the continental margin: Amer-
ican Journal of Science, v. 277, p. 12331278.
Tracy, R.J., and Frost, B.R., 1991, Phase equilibria and thermobarometry of
calcareous, ultramafic and mafic rocks, and iron formations, in Kerrick,
D.M., ed., Contact metamorphism: Reviews in Mineralogy, v. 26, p.
207289.
Vidal, C.E., and Noble, D.C., 1994, Yacimientos hidrotermales controlados
por estructura y magmatismo en la region central del Per: Instituto de In-
genieros de Minas del Per Revista Minera, v. 230, p. 1619.
Wheeler, R.L., 1978, Cross-strike structural discontinuities, possible explo-
ration tool in detached forelands [abs.]: Geological Society of America,
Southeastern Section Abstracts with Programs, v. 10, no. 4, p. 201.
Wilson, J.J., 1963, Cretaceous stratigraphy of central Andes of Peru: Ameri-
can Association of Petroleum Geologists Bulletin, v. 47, p. 133.
Wilson, J.J., and Reyes, L., 1964, Geologa del cuadrngulo de Pataz, Boletn
09, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y Met-
alrgico, Sector Energa y Minas, Per, 91 p., 1 map, 1:100,000 scale.
Wilson, J.J., Reyes, L., and Garayar, J., 1967, Geologa de los cuadrngulos de
Pallasca (17-h), Tayabamba (17-i), Corongo (18-h), Pomabamba (18-i),
Carhuaz (19-h) y Huari (19-i), Boletn 16, Serie A: Carta Geolgica Na-
cional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas,
Per, 95 p., 6 maps, 1:100,000 scale.
1995, Geologa de los cuadrngulos de Pallasca (17-h), Tayabamba (17-
i), Corongo (18-h), Pomabamba (18-i), Carhuaz (19-h) y Huari (19-i), Bo-
letn 60, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y
Metalrgico, Sector Energa y Minas, Per, 63 p., 6 maps, 1:100,000 scale.
Zentilli, M., and Maksaev, V., 1995, Metallogenetic model for the late
Eocene-early Oligocene supergiant porphyry event, northern Chile, in
Clark, A.H., ed., Giant ore deposits II: Controls on the scale of orogenic
magmatic-hydrothermal mineralization, April 1995, Queens University,
Kingston, Ont., Giant Ore Deposits Workshop, 2nd, Proceedingssecond
(corrected) printing, 1996, p. 152165.
Zentilli, M., Doe, B.R., Hedge, C.E., Alverez, O., Tidy, E., and Daroca, J.A.,
1988, Istopos de plomo en yacimientos de tipo prfido cuprfero com-
parados con otros depsitos metalferos en los Andes del norte de Chile y
Argentina (English abstract): Actas, V Congreso Geolgico Chileno, Santi-
ago, v. 1, p. B331369.
Zharikov, V.A., 1970, Skarns: International Geology Review, v. 12, p. 541559,
619647, 760775.
Zuzunaga, A., 2003, Closing the loopUsing actual concentrator perfor-
mance to determine the true value of ore sources [abs.]: Canadian Institute
of Mining, Metallurgy and Petroleum, Annual Meeting, 105th, Montreal,
Quebec, May 57, 2003, Program with Abstracts, p. 170.
916 LOVE ET AL.
0361-0128/98/000/000-00 $6.00 916

Você também pode gostar