Você está na página 1de 21

Miner Deposita (2013) 48:653673

DOI 10.1007/s00126-013-0455-6

ARTICLE

Lithologic controls on mineralization at the Lagunas Norte


high-sulfidation epithermal gold deposit, northern Peru
Luis M. Cerpa & Thomas Bissig & Kurt Kyser &
Craig McEwan & Arturo Macassi & Hugo W. Rios

Received: 26 August 2011 / Accepted: 15 January 2013 / Published online: 5 February 2013
# Springer-Verlag Berlin Heidelberg 2013

Abstract The 13.1-Moz high-sulfidation epithermal gold


deposit of Lagunas Norte, Alto Chicama District, northern
Peru, is hosted in weakly metamorphosed quartzites of the
Upper Jurassic to Lower Cretaceous Chim Formation and
in overlying Miocene volcanic rocks of dacitic to rhyolitic
composition. The Dafne and Josefa diatremes crosscut the
quartzites and are interpreted to be sources of the pyroclastic
volcanic rocks. Hydrothermal activity was centered on the
diatremes and four hydrothermal stages have been defined,
three of which introduced Au Ag mineralization. The first
hydrothermal stage is restricted to the quartzites of the
Chim Formation and is characterized by silice parda, a
tan-colored aggregate of quartz-auriferous pyriterutile
digenite infilling fractures and faults, partially replacing
silty beds and forming cement of small hydraulic breccia
bodies. The 34S values for pyrite (1.72.2) and digenite
Editorial handling: F. Tornos
Electronic supplementary material The online version of this article
(doi:10.1007/s00126-013-0455-6) contains supplementary material,
which is available to authorized users.
L. M. Cerpa (*)
Departamento de Ciencias Geolgicas, Universidad Catlica del
Norte, Av. Angamos 0610, Antofagasta, Chile
e-mail: lcerpa@gmail.com
T. Bissig
Mineral Deposit Research Unit, University of British Columbia,
Vancouver, Canada
K. Kyser
Queens University, Kingston, Ontario, Canada
C. McEwan : A. Macassi : H. W. Rios
Minera Barrick Misquichilca, Av. Victor Andrs Belaunde,
Lima, Peru

(2.1 ) indicate a magmatic source for the sulfur. The


second hydrothermal stage resulted in the emplacement of
diatremes and the related volcanic rocks. The Dafne diatreme features a relatively impermeable core dominated by
milled slate from the Chicama Formation, whereas the
Josefa diatreme only contains Chim Formation quartzite
clasts. The third hydrothermal stage introduced the bulk of
the mineralization and affected the volcanic rocks, the diatremes, and the Chim Formation. In the volcanic rocks,
classic high-sulfidation epithermal alteration zonation
exhibiting vuggy quartz surrounded by a quartzalunite
and a quartzalunitekaolinite zone is observed. Company
data suggest that gold is present in solid solution or micro
inclusions in pyrite. In the quartzite, the alteration is subtle
and is manifested by the presence of pyrophyllite or kaolinite in the silty beds, the former resulting from relatively high
silica activities in the fluid. In the quartzite, gold mineralization is hosted in a fracture network filled with coarse
alunite, auriferous pyrite, and enargite. Alteration and mineralization in the breccias were controlled by permeability,
which depends on the type and composition of the matrix,
cement, and clast abundance. Coarse alunite from the main
mineralization stage in textural equilibrium with pyrite and
enargite has 34S values of 24.829.4 and d 18 OSO4 values
of 6.813.9, consistent with H2S as the dominant sulfur
species in the mostly magmatic fluid and constraining the
fluid composition to low pH (02) and logfO2 of 28 to 30.
Alunitepyrite sulfur isotope thermometry records temperatures of 190260 C; the highest temperatures
corresponding to samples from near the diatremes. Alunite
of the third hydrothermal stage has been dated by 40Ar/39Ar
at 17.00.22 Ma. The fourth hydrothermal stage introduced
only modest amounts of gold and is characterized by the
presence of massive alunitepyrite in fractures, whereas

654

barite, drusy quartz, and native sulfur were deposited in the


volcanic rocks. The d 18 OSO4 values of stage IV alunite vary
between 11.5 and 11.7 and indicate that the fluid was
magmatic, an interpretation also supported by the isotopic
composition of barite (34S=27.1 to 33.8 and d 18 OSO4 =
8.1 to 12.7). The 34Spyalu isotope thermometry records
temperatures of 210 to 280 C with the highest values
concentrated around the Josefa diatreme. The Lagunas
Norte deposit was oxidized to a depth of about 80 m below
the current surface making exploitation by heap leach methods viable.

Miner Deposita (2013) 48:653673

sediment samples (Dunin-Borkowski 2000). Minera Barrick


Misquichilca acquired the property in 2001 and announced
the discovery of the deposit in April 2002 with an initial
resource of 3.5 Moz of Au with an average grade of 1.95 g/t
(Araneda et al. 2003). Gold production commenced in late
2005 and attained annual production of 1.2 Moz at a cash
cost of $125/oz in 2008. The total production and current
reserves is 13.1 Moz Au (Barrick Gold Corp. Annual
Report 2011).

Regional geologic setting


Keywords Diatreme . Breccia . High sulfidation .
Epithermal . Central Andes . Miocene . Landscape evolution

Introduction
Lagunas Norte (756 S, 7815 E) is one of the most recent
discoveries of world class epithermal gold deposits in northern Peru and is, in contrast to other important epithermal
deposits of the region (e.g., Pierina: Rainbow 2009;
Yanacocha: Longo et al. 2010), not only hosted in volcanic
rocks but also in Upper Jurassic to Lower Cretaceous
quartzites. The limited reactivity of the quartzites resulted
in important challenges in mapping of the alteration and
consequently the relatively recent discovery of the deposit
(Araneda et al. 2003), despite the fact that it is well exposed
at surface.
In this article, we present a genetic model of the Lagunas
Norte deposit on the basis of the paragenetic evolution,
mineralization, and alteration and its relationship with the
rocks that host the mineralization. We also present stable
isotope data for pyrite, barite, and three types of alunite,
commonly in textural equilibrium with auriferous pyrite,
which constrain the origin of the mineralizing fluids and
allow documentation of the evolution of the magmatichydrothermal system of Lagunas Norte.
Exploration history and reserves
Prior to the discovery of the Lagunas Norte AuAg deposit,
significant coal mining had been carried out in the Alto
Chicama area since the end of the nineteenth century.
Between 1880 and 1931, Compaia Minera Northern
exploited the Callacuyan coal deposit, located 5 km to the
NW of Lagunas Norte (Escudero 1979). Only small-scale
coal mining for local domestic use took place after 1931, as
larger scale operations were not profitable due to the high
sulfur content of the coal (Manrique 1986).
In 1999, Centromin-Per carried out preliminary studies
to evaluate the metallic mineral potential of the area, which
led to the identification of elevated gold values in stream

The Mesozoic units of northern Peru (Fig. 1) consist of


sedimentary rocks deposited during the Andean cycle
(Mgard 1987). Starting in the Tithonian, the western
Peruvian continental margin was dominated by the subsiding Chicama basin (Jaillard and Jacay 1989), where locally
as much as 2,500 m shale, intercalated with subordinate thin
sandstone beds (Chicama Formation: Jaillard and Soler
1996) was deposited in a dominantly deep marine sedimentary environment. In the BerriasianValanginian, a gradual
transition from deep marine to shallower siliciclastic sedimentation occurred, resulting in a succession of quartz sandstones derived from the Guyana and Brasilia cratons
(Moulin 1989). These sandstones were probably deposited
in a fluvio-deltaic environment and in the study area are
represented by the Chim Formation (Benavides-Cceres
1956; Jaillard and Jacay 1989). The transition from a pelagic
to shallow continental margin environment marks a change
in paleogeography due to a drastic change in subduction
geometry at the northern Peruvian margin at that time
(Jaillard et al. 2000).
The Valanginian was dominated by marine transgressions
and regressions giving rise to the SantaCarhuaz Formation
(Benavides-Cceres 1956) which consists of alternating
sandy and shaly beds. Carbonate and black shales of the
Chulec and Pariatambo Formations, respectively, overlie the
SantaCarhuaz Formation and indicate a progressively
deepening depositional environment. In the Aptian, the
western border of the Chicama basin was the site of intense
volcanic arc activity (Casma Group; Atherton et al. 1985;
Soler 1991). Marine sedimentation and volcanism ended in
the Albian during dextral transpressive deformation in the
arc (Soler and Bonhomme 1990). Volcanic activity was
succeeded by voluminous intrusive activity leading to the
emplacement of the 10055 Ma Coastal Batholith (Cobbing
et al. 1981; Soler 1991).
In the late Cretaceous, the Mariana-type subduction was
replaced by the present-day Andean-type subduction
(Benavides-Cceres 1999). This resulted in a tectonic inversion and intense compressive deformation which gave rise to
the Maraon Fold and Thrust Belt (Benavides-Cceres 1999).

Miner Deposita (2013) 48:653673

655

79

78
80

70
0

0
C o lo m b ia

Ecuador

PERU
T ru jillo

B ra s il

10

10
L im a
C usco

80

Yanacocha

70

CAJAMARCA

La Virgen
PA C IF IC
O CEAN
8
Quaternary

Lagunas
Norte

Miocene
Oligocene-Miocene

Callacuyan

TRUJILLO
Calipuy Gp.
volcanic rocks

Quiruvilca

Cordillera Blanca Batholith


Coastal Batholith
Late Cretaceous, siliciclastic
and carbonaceous rocks
Lower Cretaceous
siliciclastic rocks
Upper Jurassic, shales

Chim Fm.
Chicama Fm.

Lower Jurassic, siliciclastic


rocks and shales
Permo-Triassic limestones
Paleozoic, metamorphic rocks

79

78

Fig. 1 Simplified geological map of northwestern Peru and locations of Lagunas Norte and other deposits of the Miocene metallogenetic belt of
Peru (modified from INGEMMET (1999) and Noble and McKee (1999))

The folded Mesozoic rocks are unconformably overlain by the


Eocene to Miocene Calipuy Group volcanic and volcaniclastic
rocks (Cosso and Jan 1967; Wilson 1975; Rivera et al. 2005;
Montgomery 2012). Volcanism ceased in the middle to late
Miocene along most of northern and central Peru. This cessation has been attributed to the onset of flat subduction along the

Peruvian margin due to the subduction of the aseismic Nazca


ridge and oceanic Inca plateau (Gutscher et al. 1999; Hampel
2002). The emplacement of many ore deposits in Peru may be
directly related to these changes in subduction geometry
(Rosenbaum et al. 2005; Bissig et al. 2008; Bissig and
Tosdal 2009).

656

Deposit geology
Mesozoic basement
The basement at Lagunas Norte is dominated by Mesozoic
pelitic and siliciclastic rocks belonging to the Chicama and
Chim Formations, respectively (Reyes 1980). These
Mesozoic rocks are thrusted and folded into NW striking
east-verging folds (the Maraon Fold and Thrust Belt:
Benavides-Cceres 1999) and are weakly metamorphosed to
slate and quartzite. Lower Miocene volcanic rocks assigned to
the Calipuy Group were deposited unconformably over the
folded Mesozoic strata. Gold mineralization is hosted by the
siliciclastic Chim Formation and the overlying volcanic strata. The deposit stratigraphy is described below in detail
(Figs. 2 and 3).
The Jurassic Chicama Formation (Stappenbeck 1929;
Cosso and Jan 1967) crops out to the west and north of the
mining operations (Fig. 2). Its thickness is unknown in the
study area but has been estimated to be up to 1,500 m thick
40 km S of Lagunas Norte (Cosso and Jan 1967; Jaillard and
Jacay 1989). It consists of a succession of dark carbonaceous
shale and siltstone with occasional thin beds of fine-grained
sandstone (see electronic supplementary data). The unit has
been weakly metamorphosed to slate and features an intense
cleavage subparallel to the bedding. The transition to the
overlying Chim Formation (see below) is gradual and characterized by increasing abundance of quartzite intercalations.
The Chicama Formation does not crop out at the deposit, but
its presence below the mineralized zone is evident from clasts
of slate in the Dafne breccia (see below) crosscutting the
Mesozoic strata.
The Upper Jurassic to Lower Cretaceous Chim Formation
(Benavides-Cceres 1956) is the principal ore host (Figs. 2
and 3). It consists of compositionally mature quartz sandstone
(typically 95 % SiO2; see electronic supplementary data) but
contains occasional coal beds, which historically have been
exploited, as well as scarce siltstone and shale intercalations.
The sandstone has undergone weak metamorphism which
resulted in some recrystallization and cementation of quartz
grains to form quartzite. The thickness of the Chim
Formation is estimated to about 450600 m in the Lagunas
Norte area (Benavides-Cceres 1956).
Volcanic rocks of the Calipuy Group
A sequence of volcanic and volcaniclastic rocks assigned to
the Calipuy Group (Cosso and Jan 1967; Rivera et al. 2005;
Montgomery 2012) overlies the Mesozoic strata in an angular
unconformity. At Lagunas Norte, four subunits from oldest to
youngest, the Quesquenda, Dafne, Josefa, and Shulcahuanga
units, can be distinguished. The Josefa and Dafne units are
closely related to magmatic-hydrothermal breccia bodies

Miner Deposita (2013) 48:653673

interpreted as diatremes from which the eruptive products


probably originated.
The Quesquenda unit is named after an eruptive center
4 km to the north of the deposit (Rivera et al. 2005) and crops
out in the easternmost portions of the deposit. There it consists
of more than 150 m pyroclastic and volcaniclastic rocks of
andesitic composition with interstratified lithic-rich tuffaceous
deposits containing carbonized wood (see electronic supplementary data). The Quesquenda unit is interpreted a product
of pyroclastic eruptions interbedded with lahar deposits.
The Dafne unit consists of a series of breccias in the
western part of the mineralized area. The breccia body
overall has a subvertical inverted cone shape and in plan
view a NW elongated ellipsoid shape measuring up to
1 km along its long axis. The breccia body cuts the
Mesozoic basement (Chicama and Chim Formations)
and is thought to be a diatreme. Four principal lithofacies
associations have been identified and their characteristics, spatial relationships, and alteration are described
below, following the classification scheme proposed by
Davies et al. (2000, 2008) and Gifkins et al. (2005). The
four breccia lithofacies are the diatreme margin, main
body, crater, and apron lithofacies and are summarized
in Table 1 and illustrated in Figs. 4 and 5.
The diatreme margin lithofacies generally consists of
clast-supported monomict and polymict breccias in a
rock flour or juvenile volcanic matrix. Hydrothermal
cement is ubiquitous (Table 1). The breccias show coarse
stratification parallel to the diatreme margin. No clasts of
breccia within breccia have been observed. Three
domains of monomictic breccias (Table 1) contain, respectively, quartzite, siltstone, or tan-colored hydrothermal quartz clasts (locally termed silice parda, see
below). The clasts are angular to subangular and the
breccias have jigsaw-fit to slightly rotated textures. The
polymictic breccias contain subangular-rotated clasts of
quartzite and siltstone. They have a rock flour matrix and
cement of hydrothermal quartz and, at depth, pyrite. The
diatreme margin lithofacies has a gradual transition to the
wall rock with jigsaw-fit textures, angular clasts, and
locally monomictic breccia chimneys (Fig. 5b). Hydrothermal
cement of quartz and pyrite dominates over matrix.
The main body lithofacies association occupies the central
portions of the diatreme and is volumetrically the most important. It is largely composed of polymictic, matrixsupported breccias (Fig 5c, d and Table 1) featuring quartzite,
siltstone, slate, and juvenile volcanic clasts (fiamme-like
whispy clast shapes) as well as occasional silice parda hydrothermal quartz fragments. The clasts are subangular to subrounded and the breccias have no apparent internal
organization or clast sorting. The matrix generally is rock
flour derived from siltstone and shale and fine-grained
reworked volcanic material.

Miner Deposita (2013) 48:653673

657

803000

803500

804000

804500

ALEXA
9122000

9122000

JOSEFA

9121500

9121500
9121000

9121000

DAFNE
C Shulcahuanga

9120500

9120500

Tectonic Breccias
Shulcahuanga Unit
Cenozoic

Josefa Unit
Dafne Unit

Chim Fm.
Mesozoic
Chicama Fm.

803000

Cross Section

Main Faults
Pit Limits
Grade/Thickness contour
(Au g/t x meters)
> 1000 g/T * m Au

9120000

9120000

Quesquenda Unit

250 g/T * m Au
L i 125 g/T * m Au

803500

804000

804500

Fig. 2 Geological map of the Lagunas Norte deposit based on Barricks regional and local mapping. The three principal ore zones are labelled

In the upper central part of the diatreme, polymictic


unstratified and massive breccias are assigned to the crater
lithofacies. Their distinguishing characteristic is the

presence of large (up to 1.7 m in diameter) rounded to


subrounded quartzite and andesite blocks featuring striae
on the clast surfaces (Fig. 5e and Table 1). Smaller clasts

658

Miner Deposita (2013) 48:653673

Shulcahuanga Unit

CENOZOIC

Josefa Diatreme
Dafne Diatreme

CALIPUY GROUP

QPF Unit

Quesquenda Unit

MESOZOIC

Chim Fm.

Chicama Fm.

Fig. 3 Generalized stratigraphic column showing the principal lithologic units of Lagunas Norte deposit and cross-cutting relationships

include quartzite and siltstone, and the matrix consists of


rock flour and juvenile volcanic material.
The apron lithofacies located in upper peripheral parts of
the diatreme is characterized by an intercalation of polymictic and monomictic clast-supported breccias with rock flour
and volcanic matrix (Fig. 5f and Table 1). This lithofacies
shows coarse bedding (also known as tephra stratification:
Lorenz 2003). Quartzite and siltstone clasts are subrounded
to rounded and locally tabular. Clast imbrication is common

and fiamme are widely observed in the most peripheral


polymictic parts of the breccias (Fig. 5g). These polymictic
breccias are recognized up to 1 km north of the diatreme.
Clast size in this unit decreases with increasing distance
from the diatreme.
The Josefa unit is subdivided into two subunits. The first
unit is a breccia body interpreted as a diatreme and the
second consists of a volcanic and volcano-sedimentary succession probably related to the same diatreme, although
other source(s) are possible. The diatreme was emplaced in
the eastern part of the deposit in the Josefa area (Fig. 2),
whereas the eruptive products have also been preserved in
the Alexa area, 1 km north of Josefa, and overlying the
peripheral parts of the Dafne breccia, 0.5 km to the west
(Fig. 2).
The Josefa diatreme is approximately 45 by 30 m in plan
view and was only recognized in the open pit after mining
had started. It has an inverted cone shape (Fig. 6a), but in
contrast to the larger Dafne diatreme, no siltstone, shale, and
carbonaceous material is present. Quartz crystals up to
5 mm in diameter occur in the largely juvenile volcanic
matrix.
The breccias at Josefa have been classified using the
same scheme as for the Dafne diatreme (Table 1). The
breccia margin lithofacies consist of monolithic clastsupported breccias with angular quartzite clasts and quartz
crystals in a volcanic matrix (Fig. 6b). This breccia also
contains quartzalunite cement. The main body of the

Table 1 Summary of lithofacies and distribution in the Dafne and Josefa diatremes
Characteristics
Dafne
Margin
Layered monomictic and
lithofacies polymictic breccias,
subrounded to subangular
clasts, crude stratification
parallel to breccia margin;
matrix and cement support;
advanced argillic alteration
Main
Polymictic breccias, matrix
body
supported, not stratified, chaotic
lithofacies distribution, subrounded to
rounded clasts; argillic
alteration
Crater
Massive body, polymictic, and
lithofacies matrix supported; rounded clast;
chaotic to crude stratification at
border; contains large andesite
blocks with striae on surfaces;
argillic alteration
Apron
Gently dipping tephra
lithofacies stratification; stratified,
polymictic, and clast, matrix,
and cement supported, rounded
to subrounded clast; advanced
argillic alteration

Characteristics
Josefa

Distribution

Interpretation

Layered polymictic breccias,


subangular to subrounded
clasts, crude stratification
parallel to breccias margin;
cement support; advanced
argillic alteration

At the border of both


Successive phreatic and
diatremes, in contact with phreatomagmatic explosions
the bedrock; contact
generate crude stratification;
steeply dipping towards
later overprinted by
the center of diatremes
hydrothermal activity

Polymictic breccias, matrix and In the central part of both Mainly phreatomagmatic
cement supported not stratified, diatremes
explosions which reworked
chaotic distribution,
matrix and clasts
subrounded to rounded clasts;
advanced argillic alteration
Massive body, polymictic,
In upper central part
Succession of violent
matrix, and cement supported;
of both diatremes
phreatomagmatic explosions,
rounded clast; crude
capable of ejecting large
stratification at border; contains
bedrock blocks
large quartzite blocks;
advanced argillic alteration
Gently dipping tephra
Located in the northwest
Succession of phreatic and
stratification; stratified,
part of the Dafne
phreatomagmatic events
polymictic and clast supported, diatreme; also similar
resulting in bedded succession;
rounded to subrounded clast;
facies in the southern part each bed representing an
advanced argillic alteration
of the Josefa diatreme
explosive event and airfall
deposition

Miner Deposita (2013) 48:653673

659

Shulcahuanga dome

Milled and reworked


diatreme facies
crater and main body lithofacies
argilic alteration

Chim Fm. quarzite

Fault
Shulcahuanga
andesite flows

Fault

Fault

Chim Fm. quarzite


Stratified and silicified diatreme facies
apron lithofacies

Fig. 4 Panoramic view of the Dafne diatreme. Photograph shows the pit exposure in 2007, looking from the northeast

breccia is polymictic and matrix supported. Clasts are subangular to subrounded and juvenile clasts and quartz crystals are present in a tuffaceous matrix (Fig. 6c). As in the
marginal facies, quartzalunite cement is present.
The crater lithofacies in the upper part of the Josefa
diatreme is characterized by large quartzite blocks up to
80 cm in diameter (Fig. 6d) in a tuffaceous matrix with
abundant quartz crystals and juvenile volcanic clasts.
The apron lithofacies is only partly preserved at the
southern margin of the diatreme where it consists of a
series of crudely stratified beds (Fig. 6e). These deposits
are overlain by pyroclastic flow deposits which are
inferred to be related to the eruptive activity at Josefa,
on the basis of lithologic similarities of the juvenile
components in the diatreme. Two principal units have
been recognized: a quartz feldspar phyric unit (QFP
unit) and an overlying dacitic unit; the latter characterized by the absence of quartz phenocrysts. These volcanic units crop out at Josefa and Alexa as well as at
Dafne where they overlie the apron lithofacies breccias
(Figs. 2 and 3) and are generally affected by advanced
argillic alteration.
The QFP unit is characterized by monomictic breccias
containing quartzite clasts. The clast sizes increase towards
the Josefa diatreme (locally termed paleosurface breccia,
Fig. 7a). Overlying this breccia is a pyroclastic flow deposit
with small (<2 cm) altered pumice fragments and quartz
crystals up to 5 mm (Fig. 7b). This pyroclastic deposit is
overlain by lithic lapilli tuff containing small quartz crystals
and rare accretionary lapilli. The upper part of this tuff unit
shows planar stratification.

Overlying the QFP unit, the Dacitic units are a series of


pyroclastic and volcaniclastic deposits with pumice and
lithic fragments but no quartz crystals (Fig. 7c). The lower
part of this unit locally shows planar stratification (Fig. 7d)
and small paleochannels filled with lithic clasts up to 2 cm
in size.
Overlying the pumice- and lithic clast-bearing dacitic
deposits there is an ash tuff with only scarce lithic
fragments. Within those strata between Dafne and
Josefa, abundant fossilized leaves and tree trunks that
are still in vertical position have locally been found
(Fig. 7e). The pinnate leaves (Fig. 7f) are of camptodrome shape, which indicates a humid and tropical flora
(Alvarez-Ramis 1999) comparable to that present now at
lower elevations in eastern Peru. The stratigraphically
highest rock type consists of pyroclastic deposits with
scarce quartz crystals and clasts of pumice and quartzite. The alteration of the dacitic unit is dominated by
advanced argillic assemblages (see below).
Shulcahuanga unit
The Shulcahuanga unit consists of porphyritic andesite lavas
and andesitic to dacitic domes that crop out to the west and
south of the deposit around Cerro Shulcahuanga. Andesitic
lavas assigned to this unit overlie the Dafne unit (Figs. 2 and
3) and have been affected by only weak chloritesmectite/illite alteration. No conclusive stratigraphic relationships with the Josefa unit have been observed, but due
to the weaker alteration, the Shulcahuanga unit is interpreted
to be younger.

660

Shulcahuanga Unit
Tectonic breccias
Volcanic facies
Apron
Diatreme
Lithofacies

Fig. 5 Lithofacies associations


of the Dafne Diatreme complex.
a Cross section (see Fig. 1 for
location of cross section),
looking northwest, through the
Dafne diatreme, showing
lithofacies distribution. b
Monomictic clast-supported
breccia showing pyrite cement
with angular clasts of quartzite
(Q) and siltstone (L). c
Polymictic breccia with
carbonaceous matrix-supported
juvenile clast (CJ), quartzite (Q),
silice parda (SP), and siltstone
(L) from central diatreme body. d
Polymictic clast-supported
breccia with clasts of siltstone
(L), juvenile clasts (CJ), and
quartzite (Q). e Polymictic
matrix-supported breccia with
large andesite boulders (A) up to
1.70 m in diameter; corresponds
to the crater lithofacies. f Closeup photograph of a coarsely
stratified polymictic breccia with
andesite blocks (A); corresponds
to the diatreme apron facies. g
Detail of clast-supported
polymictic breccia, showing
siltstone (L), quartzite (Q), and
juvenile clast (CJ) from the distal
apron breccia lithofacies

Miner Deposita (2013) 48:653673

Crater
Main Body
Margin

Chim Fm.

Chicama Fm.

CJ
SP

E
L
CJ
A

G
L

CJ
A

Q
0

Two lithologies are recognized. Firstly, andesite, locally


known as Andesitas Azules due to the pale bluegreen hue
imposed by clay alteration, is characterized by a finegrained aphanitic groundmass with hornblende phenocrysts
and occurs as dykes that crosscut the southern margin of the

1cm

Dafne diatreme. The second lithology forms the Shulcahuanga


dome (Figs. 2 and 3) and adjacent lava flows to the east and
exhibits a porphyritic texture with plagioclase, biotite, and
hornblende phenocrysts. These rocks are characterized by
prominent flow banding (Macassi 2005). An age range of

Miner Deposita (2013) 48:653673


Fig. 6 Field photographs of the
Josefa breccia. a The Josefa
breccia body in pit exposure
(4,060 to 4,090 m.a.s.l. bench).
Letters B to E refer to close-up
photographs. b Volcanic and rock
flour matrix crackle breccias with
quartzalunite cement, here
lacking juvenile fragments. c
Detail of polymictic breccia with
quartzite and juvenile pumice
fragments. d Phreatomagmatic
breccia with a quartzite block
from the crater lithofacies. e
Close-up of the apron lithofacies
showing coarse bedding (marked
by red dashed lines)

661

A
4090
E

D
4080
4070

4060

16.8 to 17.3 Ma for the Shulcahuanga unit has been established


on the basis of 40Ar/39Ar data on biotite and hornblende
(Montgomery 2012).

Analytical methods
The alteration paragenesis defined by previous workers
(Guerra 2001; Araneda et al. 2003; Macassi 2005; Ros
2005) has been refined on the basis of field observations
and detailed petrography. Mineral assemblages have been
identified by standard optical microscopy and, where appropriate, by scanning electron microscopy (SEM) and X-ray
diffraction at the Universidad Catlica del Norte. These
analyses have been complemented by infrared spectroscopy
using a Portable Infrared Mineral Analyzer (PIMA) and

using the SPECMIN database (Thompson et al. 1999) and


local databases for interpretation of infrared spectra.
Stable O, S, and H isotopic analyses were carried out at
the Queens University Facility for Isotope Research,
Kingston, Ontario, Canada. The 34S and d 18 OSO4 values
for alunite were determined using a method modified from
Wasserman et al. (1992) and Arehart et al. (1992). Sulfur
was extracted online with continuous flow technology, using a Finnigan MAT 252 isotope ratio mass spectrometer.
Sulfate oxygen was extracted using the BrF5 technique of
Clayton and Mayeda (1963). Hydrogen isotopic compositions were measured using a Thermal Finnigan TCEA coupled to a Thermo Finnigan Delta+ XP mass spectrometer
and continuous flow technology (ConFlo III) as described in
Rainbow et al. (2005). All values are reported in units of per
mil () relative to Vienna Standard Mean Ocean Water for

662
Fig. 7 Volcanic stratigraphy of
the Josefa volcanic unit. a
Monomictic breccia with
hydrothermal cement (by mine
geologists also referred to as
Paleosurface Breccia)
representing the basal portion. b
Detail of pyroclastic flow with
pumice fragments and tiny
quartz crystals, affected by
advanced argillic alteration. c
Pumice and quartzite clastbearing pyroclastic flow deposit
affected by advanced argillic
alteration. d Fine-grained
laminated ash fall deposit of the
upper volcanic member of the
Josefa volcanic unit, affected by
pervasive advanced argillic
alteration. e Remnant of a
carbonized tree (yellow arrow)
in upright position in the ash
fall deposit of the upper Josefa
volcanic unit. f Detail of
fossilized leaf present in an ash
fall deposit in the upper Josefa
volcanic unit

Miner Deposita (2013) 48:653673

O and H and Caon Diablo Troilite for S isotopic compositions. Accuracy was monitored using standards calibrated to
NIST 8556 and 8557 for sulfur and oxygen and NIST 8538
biotite for hydrogen. Analytical precision for both 34S
a n d d 18 OSO4 v a l u e s i s 0 . 3 , f o r D 3 .
Paleotemperature for coexisting alunitepyrite pairs is
calculated using the following fractionation factors: 103
ln apyH2 S 0:40  106 T 2 (Ohmoto and Rye 1979) and
103 ln aalunSO4 H2 S 6:463  106 T 2 0:56 (Ohmoto and
Lasaga 1982).
One sample has been dated by the 40Ar/39Ar method at
the Pacific Centre for Isotopic and Geochemical Research.
Alunite was handpicked and analyzed as described in Bissig
et al. (2008). The data are included as digital appendix
(ESM).

50cm

zonation pattern typical for high-sulfidation systems


(e.g., Simmons et al. 2005) with a nucleus of vuggy
quartz, surrounded by quartzalunite and dickitekaolinite alunite zones which indicate acidic fluids that
became progressively neutralized during reaction with
the host rock. Contrasting the volcanic units, alteration
affecting the quartzite is subtle and difficult to detect
(see electronic supplementary data) but kaolinite and, in
more silty strata, pyrophyllite have been detected by
PIMA.
Four hydrothermal stages have been defined at Lagunas
Norte (Fig. 8) and are described below. Gold was introduced
during stages 1 and 3; the latter being the principal mineralization stage. Minor additional gold was also introduced
during stage 4. Supergene oxidation to depths of up to 80 m
below the current surface made the ore amenable to heap
leaching methods.

Hydrothermal evolution and mineralization


Stage I: early hydrothermal activity
The hydrothermal alteration at Lagunas Norte manifests
itself in very distinctive ways, depending on the host
rock compositions and textures. In the upper volcanichosted part of the deposit, the alteration developed a

The first hydrothermal event at Lagunas Norte is characterized by fine-grained yellowish to tan-colored aggregates of
quartz, pyrite, and minor rutile, which is referred to as silice

Miner Deposita (2013) 48:653673


STAGE I

663
STAGE II

Chim Fm.

STAGE III
Chim Fm.

STAGE IV
Volcanic units

Chim Fm. /Volcanic units

STAGE V
ChimFm. /
Volcanic units

Gold
Silice Parda

Alunite
Enargite
Pyrophylite
Stibnite
Arsenopyrite
Diaspore
Barite
Drusy Quartz
Sulfur

Dafne and Josefa Diatremes


emplacement

Pyrite
Digenite
Chalcopyrite
Rutile

Coarse

Disseminated

Massive

Carbonaceous layers

Jarosite
Scorodite
Hematite
Goethite

Fig. 8 Paragenetic sequence from Lagunas Norte deposit; thickness of lines shows the relative abundance of minerals

parda by mine geologists, a term also used herein. This


assemblage is restricted to the Chim Formation where it
was generally emplaced along a network of preexisting
fractures and is best developed in silty layers, but also forms
the cement of small fault controlled monomictic breccia
bodies (Fig. 9a, b).
In the area between Josefa and Dafne and in the southern
part of the Josefa zone, silice parda is accompanied by
chalcopyrite and digenite (Fig. 9c, d). Gold is not visible
by SEM or optical microscopy, but company internal mineralogical studies show that gold is associated with pyrite,
and we assume that gold is present in solid solution or as
nanoparticles in the pyrite, as in other Andean highsulfidation epithermal deposits (e.g., Pascua; Chouinard et
al. 2005a). As indicated by the presence of silice parda
clasts in the Dafne diatreme, the first mineralization stage
preceded the diatreme emplacement. Absolute age constraints for silice parda were not determined due to a lack
of dateable minerals in this assemblage. However,
Montgomery (2012) reports an age of paragenetically early
alunite hosted in the Chim Formation of 17.360.14 Ma,
which may be considered a minimum age for this stage.
Stage II: phreatic and phreatomagmatic activity
The breccia lithofacies present in the Dafne and Josefa
diatremes suggest that they formed by phreatic and phreatomagmatic activity which here is defined as the second
hydrothermal stage. This stage was important as ground
preparation for subsequent mineralization by fracturing the
adjacent rock and as host of a portion of the ore.
Mineralization in the diatremes is controlled by the permeability, which in turn is controlled by matrix type and

abundance, type, shape, and size of clasts. A minimum age


for the brecciation events is given by the oldest age of
alunite within the overlying volcanic sequence of 17.05
0.12 Ma (Montgomery 2012).
Stage III: main mineralization stage
Most of the gold was introduced during this stage and is
contained within the pyrite but not visible optically. The
main mineralization and alteration stage is difficult to detect
in the quartzite. However, fracture infill of coarse alunite
(Fig. 10a) associated with pyrite and enargite (Fig. 10b), at
depths below 80 m from the present surface, is observed. In
the quartzite, disseminated kaolinite has been detected by
PIMA. In the more silty beds of the Chim Formation in the
core of Lagunas Norte, pyrophyllite is present, whereas
kaolinite occurs in the periphery of the deposit. In beds
where coal is present, a sulfide assemblage containing pyrite, stibnite, and arsenopyrite is observed locally.
In both diatremes alteration patterns are lithologically controlled. The margin of the Dafne breccia is intensely silicified
with minor alunite, whereas in the main body, dickitekaolinite alteration affected juvenile fragments, and fracture controlled silicification is present locally. The crater facies shows
a weak dickitekaolinite alteration restricted to matrix and
juvenile fragments. In the apron lithofacies, the matrix composition determines the alteration intensity and assemblages.
Where the matrix is predominantly carbonaceous, the juvenile
fragments are preferentially altered to alunitedickitekaolinite, whereas in beds with volcanic matrix, quartzalunite is the
dominant alteration assemblage. The Josefa breccias are pervasively altered to quartzalunite and juvenile fragments have
commonly been replaced by pyrite and alunite.

664
Fig. 9 Photographs showing
the principal characteristics of
the first stage of mineralization.
a Monomictic breccias with
tan-colored quartz cement
(silice parda, see text for
details). b Replacement of silty
layers of the Chim Formation
by silice parda. c Polished
section photograph of first stage
of mineralization showing the
granular texture of silice parda
and interspersed small pyrite
crystals (Py). d Detail of
polished section photograph of
pyrite crystals (Py) in digenite
(Di) present in quartzite (Qz)

Miner Deposita (2013) 48:653673

D
Qz
Py

Di

Py

In the volcano-sedimentary levels at Dafne, Josefa, and


Alexa as well as in the Josefa marginal facies and the Dafne
breccia, an alteration zoning pattern typical for highsulfidation epithermal deposits is observed. The distribution
of vuggy quartz zones is controlled by small E-oriented
faults and the permeability of volcanic or breccia facies.
Within the volcanic package, vuggy quartz is best developed
in pumice- and crystal-rich pyroclastic flow deposits where
pumice fragments and feldspar phenocrysts were leached
and the volcanic matrix was completely replaced by residual
quartz (Fig. 10c, d). Surrounding the vuggy quartz zone, the
assemblage quartzalunitepyrite altered the rocks. Alunite
has replaced feldspars and pumice clasts (Fig. 10e, f) and the
groundmass has been replaced by fine-grained quartz and
pyrite. Pyrite has generally been oxidized but is preserved
together with alunite in some silicified strata at Alexa
(Fig. 11a), whereas at Josefa, it occurs together with rutile.
The matrix of the breccia at the base of the volcanic pile has
been affected by pervasive quartzalunite kaolinitedickite alteration. Disseminated alunite from the volcanic Josefa
unit representing this main mineralization stage gives an

0,1mm

0,01mm

40

Ar/39Ar plateau age of 17.00.22 Ma (Fig. 12), which is


consistent with the age range of 16.7 to 17.1 Ma inferred for
the main hydrothermal activity (Montgomery 2012). The
andesitic volcanic rocks surrounding the deposit have been
affected by weak to moderate argillic alteration where illite
partly replaces hornblende and quartzchlorite veinlets have
been observed (Fig. 11b).
Stage IV: late-stage alteration
Late-stage alteration is characterized by massive alunite
forming the cement of local fault breccias and filling thin
fractures in Chim quartzite. This alunite is white to
yellowish in color (Fig. 11c), forming fine-grained and
massive aggregates. Pyrite in textural equilibrium with
this type of alunite is observed (Fig. 11d). Traces of
kaolinite and, at Alexa, diaspore are generally present
as well. The massive alunite has cut pyrophyllite-altered
siltstone beds and overgrown coarse-grained alunite. The
last manifestation of hydrothermal activity is barite, rutile, and drusy quartz as well as late native sulfur filling

Miner Deposita (2013) 48:653673


Fig. 10 Photographs showing
characteristics of hydrothermal
stage III. a Coarse alunite
crystals in fracture of Chim
quartzite. b Polished section
microphotograph of enargite
(En) and pyrite (Py) in quartz
(Qz) gangue, indicative of
alunitepyriteenargite related
fluids (Deyell et al. 2005). c
Vuggy quartz texture in
volcanic unit at Alexa. d
Disseminated alunite (Al)
replacing feldspar, Josefa
volcanic unit. e Alunite (Al)
replacing feldspar crystals and
pumice fragments in a tuff
deposit from the Josefa volcanic
unit. f Thin-section
microphotograph of tabular
alunite crystals (Al) replacing
plagioclase phenocrysts

665

Qz

En

Py

1cm

0,25mm

Al

Al
Al

open spaces in the volcanic rocks. Barite-filled fractures


occur in the quartzite at Alexa and Josefa. Only modest
gold mineralization is associated with this late-stage alteration. The youngest 40Ar/39Ar age on alunite reported
by Montgomery (2012) is 16.450.28 Ma and is herein
interpreted to constrain hydrothermal stage IV.
Stage V: supergene stage
Lagunas Norte has been affected by extensive supergene
oxidation (up to 80 m below the current surface) which has
produced hematite, goethite, and locally jarosite and scorodite. Iron oxides occur mainly as cement of tectonic and
hydrothermal breccias as well as on fracture surfaces where
they commonly present iridescent colors. Oxidation was
crucial for liberating the gold which made exploitation of
the deposit economically viable using cyanide leaching
methods.

1mm

Stable isotopes
Stable isotopic compositions were obtained for each paragenetic stage. 34S values (Fig. 13) were obtained for
sulfides from all paragenetic stages, whereas 34S, D,
and d 18 OSO4 values were obtained for alunite from stages
III and IV (Fig. 14). The 34S values and 34Salupy
precipitation temperatures were calculated for alunite and
pyrite occurring in textural equilibrium at different locations in the deposit (Table 2 and Fig. 15).
Stage I
Three 34S values for sulfides were obtained. Pyrite has
values of 1.7 and 2.2, and coexisting digenite has a value
of 2.1. A maximum fluid temperature of 360 C is given by
34S thermometry on the digenitepyrite pair (Hubberten
1980).

666

Miner Deposita (2013) 48:653673

Fig. 11 Photographs showing


characteristics of hydrothermal
stages III and IV. a Monomictic
breccia with dark quartzpyrite
cement. b Detail of weakly
propylitically altered andesite
of Shulcahuanga unit showing
quartz veinlet with Fe oxide
halo. c Massive alunite (Al)
filling open spaces in
monomictic breccia. d Thinsection microphotograph
showing a monomictic breccia
with quartzite clasts (Qzt) in
fine-grained clastic matrix, with
alunite (Al) cement

D
Al
Al

Py

Qzt

Stage III
Both coarse and disseminated alunite from stage III was
analyzed. Coarse alunite is translucent to pale pink in color,
with a tabular crystal habit. Eight samples were analyzed
and show a range of 34S values between 24.8 and 29.4.
Six of these alunite samples are in textural equilibrium with
pyrite 34S values of 4 to 0.5 and locally with enargite
(34S=1.2). The temperatures calculated for alunite
pyrite pairs for this hydrothermal stage are between 190
and 270 C (Table 2 and Fig. 15); the highest temperatures
have been recorded near the diatremes at 200 m depth
Sample Lc5014: Stage III Alunite
30

below the current surface, while the lowest values come


from samples near the surface. Fluid inclusion microthermometry was not possible because of a lack of suitable
material. The D vs. d 18 OSO4 values of the alunite and the
calculated isotopic compositions of their apparent fluids are
consistent with magmatic vapors and a predominantly magmatic origin for the fluid (Fig. 14).
Disseminated alunite from stage III is usually white, but
locally pale pink. Eight samples were analyzed and gave 34S
values between 21.7 and 28 and a range of d 18 OSO4 values
from 6.8 to 13.9. These values are typical for alunite
precipitated from a dominantly magmatic fluid (Fig. 14; Rye
et al. 1992; Rye 2005). However, one of the samples located
near the eastern margin of the deposit has values of d 18 OSO4 =
6.8 and D=40.2, which suggest influence of a meteoric fluid component. This sample has a 34S value of 27.2.
Stage IV

Plateau age = 17.00+/- 0.22 Ma


Age (Ma)

1mm

20

10
MSWD = 1.4, probability=0.23
Includes 91.2% of the 39Ar

0
0

20

40
60
80
Cummulative 39Ar percent

100

Fig. 12 40Ar/39Ar age spectrum for alunite sample of stage III from
Lagunas Norte. Errors are given at the 2 level

The massive alunite of this stage is white and has an earthy


texture. Seven samples were analyzed and have 34S values
between 19.1 and 29.2 (Fig. 14) and d 18 OSO4 values
between 11.5 and 11.6 (Fig. 14). Pyrite is commonly
present in textural equilibrium with stage IV alunite and
has 3 4 S values between 1.4 and 1.4 (n = 5).
Temperatures calculated from the 34Salupy pair (Table 2)
range between 210 and 280 C, with higher temperatures
near the diatremes and lower temperatures near the surface
(Fig. 15). For this stage, barite has 34S values from 27.1 to
33.8 and d 18 OSO4 values from 8.1 to 12.7.

Miner Deposita (2013) 48:653673

667

Frequency

STAGE I
Pyrite

STAGE III

Digenite

10

IV
III
IV
III
IV
III
IV
III
III
III IV IV
III
III IV III IV III
IV III
IV III III IV III
III III III
IV III III III I
III I I III IV III III I I
-5

STAGE IV

III Pyrite

IV Pyrite

III Enargite

IV Alunite (massive)

III Alunite (coarse)

IV Barite

III Alunite (disseminated)

IV Sulfur (native)

IV
10

15

34

IV
IV IV
III
IV
III
IV IV
III
III IV IV IV
IV III IV III IV IV IV
III IV IV III III III IV III IV IV
III III III III III III III III III III IV III IV

20

25

30

35

S ()

Fig. 13 Histogram of 34S values of sulfides and sulfates in the Lagunas Norte deposit. Alunite samples are colored according to paragenetic stages

Supergene stage
Two samples of supergene goethite were analyzed; they
have D values of 187 and 183 and 18O compositions
of 5.9 and 5.4. These values likely reflect the isotopic
composition of local meteoric water in equilibrium with
goethite well after hydrothermal processes ended.

Discussion
Most high-sulfidation epithermal deposits are related to
magmatic-hydrothermal activity affecting volcanic or igneous rocks (e.g., Cooke and Simmons 2000), but Lagunas
Norte differs because part of the mineralization is hosted in
unreactive quartzites. Four different hydrothermal stages
have been defined.
0

VV
-20

MW
L
-40

INI

TE
L

Alunite fluids
(200to 280C)
-60

OL

()

INE

FMW

KA

Fig. 14 D and d18 OSO4 diagram


showing data for stage III and IV
alunite. The fluid data were
calculated according to Stoffregen
et al. (1994) at 200280 C (based
on temperature range obtained
from 34Salupy). The lines and
fields are: MWL = meteoric water
line (Craig 1961), kaolinite line
(Savin and Epstein 1970), FMW =
felsic magmatic water (Taylor
1988), VV = volcanic vapor, i.e.,
range of fumarole water
(Giggenbach 1992)

The first stage is restricted to the Mesozoic rocks and is


characterized by silice parda (quartz pyrite rutile),
which was precipitated in fractures and preexisting faults
and also replaced siltstone beds of the Chim Formation.
Silice parda locally cemented monomictic breccias, which
suggests that the magmatic-hydrothermal fluids of this stage
had enough pressure to hydraulically fracture the quartzite.
Locally and 40 m below the surface, pyrite, chalcopyrite,
and digenite form the sulfide assemblage. The presence of
digenite suggests a high sulfidation state for the fluid
(Einaudi et al. 2003; Rye 2005). Sulfur isotopic compositions (34S 1.7 to 2.2 and 2.1 for pyrite and digenite,
respectively) are consistent with a magmatic source of sulfur. The fluid temperature of 360 C estimated on the basis
of sulfur isotope fractionation between sulfide (pyrite
digenite pair) species likely overestimates the true paleotemperature which is unreasonably high for epithermal

Fluids

-80
STAGE III
STAGE IV

Coarse Alunite
Disseminated Alunite
Massive Alunite

-100
-15

-5

15

OSO

18

()

25

668
Table 2 Summary of isotopic
data for coexisting alunitepyrite
pairs. T is calculated using the
following fractionation factors: 1
03 ln apyH2 S 0:40  106 T 2
(Ohmoto and Rye 1979); 103 ln
aalunSO4 H2 S 6:463  106
T 2 0:56 (Ohmoto and
Lasaga 1982)

Miner Deposita (2013) 48:653673

Sample name

Type

Alunite 34S

Pyrite 34S

34Salupy

T (C)

LD6-006
LD6-196
LD6-104
LD6-070
LD6-176
LD7-178
LD6-111
LC5-014
LD7-002
LD6-177
LD6-101
LD6-112
LD6-068
LD7-168

Coarse
Coarse
Coarse
Coarse
Coarse
Coarse
Disseminated
Disseminated
Disseminated
Massive
Massive
Massive
Massive
Massive

28.93
25.7
24.82
27.39
25.65
23.39
25.25
26.82
22.7
25.48
23.88
24.64
19.13
27.19

0.76
0.52
2.61
0.33
0.59
4.43
0.73
2.01
4.34
1.39
1.43
1.77
1.37
2.02

28.17
26.22
22.21
27.06
26.24
27.82
25.98
24.81
27.04
26.87
22.45
26.41
20.5
25.17

195
213
256
205
213
198
215
227
205
207
253
211
278
223

systems and inconsistent with the sulfide and alteration


paragenesis and the near-surface geomorphologic setting
of the deposit where mineralization directly underlies the
2526 Ma subplanar Pampa la Julia erosional surface
(Montgomery 2012). This indicates that the calculated fluid
temperature probably has no geological meaning and that
the two sulfide species are not in isotopic equilibrium.
The second hydrothermal stage reflects the emplacement
of diatreme breccias. Diatremes have widely been documented in epithermal deposits (e.g., Sillitoe 1985; Kelian,
Davies et al. 2008; Pascua; Chouinard et al. 2005b). At
Lagunas Norte, the emplacement of the diatremes was instrumental for ground preparation for the mineralization
introduced during stage III. The Dafne diatreme intersects
shale of the Chicama Formation as well as quartzite of the
Chim Formation. The involvement of shale resulted in
some milled breccia lithofacies being relatively impermeable to later fluid circulation which is reflected by the
limited alteration of the central parts of the diatreme.
Similar relatively impermeable breccias have been documented from Kelian, Indonesia (Davies et al. 2008). The
Josefa Diatreme, in contrast, only cuts quartzite which
resulted in a more permeable breccia body that was
cemented throughout by quartzalunite. The smaller diameter of the clasts in the center of both diatremes together with
the rounding of clasts and crude stratification near the margin
of the breccias suggest multiple explosive brecciation events
(Lorenz and Kurszlaukis 2007; Walters 2006), but the absence
of breccia-in-breccia clasts indicates that the breccias have not
been consolidated or cemented hydrothermally between the
individual explosions and that brecciation probably occurred
over a short time interval. The violent emplacement of the
diatremes not only generated permeability along their borders,
but probably also improved the secondary permeability in the
surrounding host rocks by means of fracturing.

The main stage hydrothermal alteration and mineralization (stage III) affected the volcanic rocks as well as the
underlying Mesozoic basement. The alteration mineralogy
and sulfide assemblages are generally as expected for highsulfidation epithermal deposits (e.g., Simmons et al. 2005).
The distribution of gold and the alteration zonation are
mainly controlled by the permeability of the host rock.
However, a number of deviations from the norm exist and
can be related to host rock characteristics. For example,
locally in silty beds where organic carbon is present, the
assemblage pyritearsenopyrite and stibnite is present. This
sulfide assemblage would be expected in a low-sulfidation
environment (e.g., Cooke and Simmons 2000; Einaudi et al.
2003), but at Lagunas Norte, it can readily be explained by
the locally strong reducing conditions.
The d 18 OSO4 values for disseminated alunite from the
volcanic levels reflect a large component of magmatic fluid
which is common for Andean high-sulfidation systems (Rye
2005; Deyell et al. 2005; Rainbow et al. 2005). However, near
the periphery of the deposit, the hydrothermal fluid had a
meteoric component. The coarse alunite samples from depths
of more than 80 m below the current surface have 34S of 24.8
to 29.4, values consistent with a H2S-dominated fluid,
which again is typical for Andean high-sulfidation systems
(Baumgartner et al. 2009; Rainbow 2009; Rye 1993, 2005).
The observed sulfide and alteration assemblage indicates

Fig. 15 a Distribution map of principal alteration zones and fluid


temperatures obtained using the 34Salupy thermometer at Lagunas
Norte. Red numbers correspond to stage III and blue numbers to stage
IV. Note that the highest temperatures were obtained near the Dafne
and Josefa diatremes. b Schematic cross section (see A for section line)
showing the lithology and calculated fluid temperatures from samples
projected onto the section plane. Red letters correspond to temperatures
from stage III and blue letters to stage IV

Miner Deposita (2013) 48:653673

803000

804000
Vuggy quartz
Quartz+Alunite

SampleLocation
(Projected)

Alunite+Dickite+Kaolinite

Cross Section

Fm. Chim (dk+kao+po)

Pit Limit

WeaklyAltered (argillic-propylitic)

Fault

9122000

9122000

669

ALEXA

JOSEFA

9121000

9121000

DAFNE

803000

804000

A
4200

4100

Sample Location
Projected Sample
Location
Open Pit limit
Oxide/Sulfides limit
Main Faults

4000

Silice parda
Shulcahuanga Unit
Josefa Unit

256

3900

Dafne Unit
Quesquenda Unit
Santa-Carhuaz Fm.

Temp. Al. (IV)

253
?

Chim Fm.
Chicama Fm.

670

Miner Deposita (2013) 48:653673

relatively acidic and moderately oxidizing conditions at 230 C


(pH=02 and logfO2 =28 to 30: Fig. 16).
The last hydrothermal event (stage IV) is characterized
by massive alunitepyrite in fractures in the quartzites,
whereas in the volcanic rocks, late barite and quartz as
well as late native sulfur precipitated in open spaces.
Isotopic data indicate that hydrothermal stages III and
IV had overall similar characteristics and a magmatic
origin for the fluids.
The apparent fluid temperatures calculated for stages
III and IV from the sulfur isotopic composition of
alunitepyrite pairs range from 190 to 280 C and are
highest near the diatremes (Fig. 15) and lowest near the
present surface some distance away from the diatremes.
Based on the calculated fluid temperatures, hydrothermal activity was probably focused around the Dafne
diatreme for stage III, whereas stage IV was centered
at Josefa. This is also consistent with the observation
that volcanic deposits likely originating from the Josefa
diatreme post-date those sourced from Dafne. Coarse
alunitepyrite pairs (34Salupy) consistently have higher
fluid temperatures than disseminated alunitepyrite
pairs. This is in agreement with the interpretation that
coarse-bladed alunite habits may indicate fluid boiling,
whereas fine-grained disseminated alunite was precipitated from cooler, non-boiling fluids, although no supporting fluid inclusion evidence is available.

-24

Al Kao

HSO

T=230 oC
S=0.01 m
K=0.01 m

-26
-28
SO4

-30

2-

Log

O2

-32

Hem

-34

En

-36

Tn
Py

-38
-40

Mag

-42

Po

-44
H S0
2

-46
-2

HS

10

12

pH
Fig. 16 Log fO2pH diagram at 230 C and saturated vapor pressure,
showing the stability fields of alunite (Al, gray box), kaolinite (Kao),
enargite (En), tennantite (Tn), hematite (Hem), magnetite (Mag), pyrrhotite (Po), pyrite (Py), and sulfur species from Lagunas Norte deposit. The
probable fluid composition for stage III is indicated by the gray box

Boiling could have been enhanced by water-table lowering


due to erosion near the hydrothermal system. The deposit
directly underlies the 2526-Ma subplanar Pampa la Julia
erosional surface (Montgomery 2012), but the deposit lies
immediately southeast of the Rio Chicama valley pediment
which likely incised concurrently with mineralization
(Montgomery 2012) and erosion may have enhanced mineralizing processes at the steep back scarp of the valley, much like
suggested for the El Indio belt in Chile (Bissig et al. 2002).
The calculated fluid temperatures for stage III are in apparent disagreement with the presence of pyrophyllite in silty
beds of the Chim Formation, since pyrophyllite normally
forms above 300 C (Hemley et al. 1980). However, in a fluid
with high silica activity, which is a reasonable assumption for
Lagunas Norte given the abundance of quartzite, pyrophyllite
may be metastable to significantly lower temperatures
(Hemley et al. 1980; Mojares et al. 2001), such as those of
280 C from 34Salupy thermometry.
The isotopic compositions of goethite, particularly the
strongly negative D, are indicative of high elevations of
3,0004,000 m (Poage and Chamberlain 2001). Similar
isotopic signatures for goethite have also been documented
by Montgomery (2012) who suggested that the goethite
formed in response to rapid late Miocene uplift. The paleobotanic evidence, on the other hand, suggests significantly
lower elevations or warmer climate at the time of volcanism
and hypogene mineralization than at present, indicating
surface uplift of 2,0003,000 m after 17 Ma (cf., Garzione
et al. 2008; Montgomery 2012).
The hydrothermal activity at Lagunas Norte pre-dates the
intermediate sulfidation polymetallic vein system of
Quiruvilca, 10 km to the west, for which an age of 15.2 to
15.7 has been determined by 40Ar/39Ar on muscovite/illite
from the selvage of a base metal quartzcarbonate vein
(Montgomery 2012). However, all other important highsulfidation epithermal deposits in northern Peru, including
Pierina (14 Ma; Rainbow 2009) and Yanacocha (138 Ma;
Longo et al. 2010) were emplaced significantly later than
Lagunas Norte and also post-date Quiruvilca. Further south,
in Central Peru, the majority of magmatic-hydrothermal
deposits have been emplaced during the middle to late
Miocene (e.g., Bissig et al. 2008; Baumgartner et al. 2009;
Kouzmanov et al. 2008). The flattening of the subduction
angle between about 14 and 10 Ma that generally coincided
with the onset of Nazca ridge subduction (Hampel 2002) is
thought to have played a role in generating favorable conditions for mineralization in central and northern Peru
(Rosenbaum et al. 2005; Bissig et al. 2008; Bissig and
Tosdal 2009). The slightly older age of Lagunas Norte
cannot confidently be assigned to initiation of ridge subduction, but the leading edge of the Nazca ridge may have
collided with the trench at that time (cf., Rosenbaum et al.
2005).

Miner Deposita (2013) 48:653673

Conclusions
The Lagunas Norte deposit is a high-sulfidation epithermal
system that is hosted in both normally unreactive rocks (quartzite of the Chim Formation) and dacitic to rhyolitic volcanic
rocks (Miocene). It is at 17 Ma likely the oldest high-sulfidation
deposits in the middle Miocene metallogenic belt of Peru. The
magmatic and hydrothermal evolution was controlled by at
least two diatremes (Dafne and Josefa), which cut the basement
composed of quartzite of the Chim Formation and in the case
of Dafne also slate of the Chicama Formation.
At Lagunas Norte, four hydrothermal stages are recognized, and most of the gold and silver were introduced
during stages I and III. Stage I is restricted to quartzite,
where the gold is associated to pyritedigenite chalcopyrite in a quartz rutile (silice parda) gangue assemblage
mainly in fractures and faults, as well as replacing some
siltstone levels. Isotopic data are consistent with a magmatic
origin of the sulfur. The second stage is the emplacement of
the Dafne and Josefa diatremes, in addition to the volcanic
rocks and their products. This phreatic and phreatomagmatic
activity was instrumental for enhancing fracture-controlled
permeability of the otherwise impermeable quartzitic host
rock.
Stage III contains the bulk of the alteration and mineralization. In rocks of the Chim Formation, coarse alunite
enargitepyrite precipitated in fractures, but alteration is
restricted to traces of kaolinite and pyrophyllite in some
beds within the quartzite. Locally, where coal is present,
stibnite and arsenopyrite are observed. In contrast, the breccias and Miocene volcano-sedimentary units overlying the
Cretaceous rocks have been affected by alteration assemblages typically described for high-sulfidation epithermal
systems. The alteration and mineralization is largely controlled by permeability. The central portion of the Dafne
diatreme is relatively impermeable due to the matrix being
largely composed of milled slate of the Chicama Formation,
whereas quartzite clasts are dominant at Josefa which
resulted in better permeability for the fluids.
Isotopic compositions of alunitepyriteenargite and alunitepyrite from stages III and IV, respectively, indicate
that these alteration minerals, and by inference the gold,
precipitated from fluids that were acidic, H2S dominant,
and largely magmatic in origin. Fluid temperatures based
on 34Salupy thermometry range between 190 and 280 C,
with the highest values near the diatremes which are interpreted to be the focus of hydrothermal activity.
Acknowledgments This research is part of the PhD research of Luis
Cerpa. Funds for this study were provided by Minera Barrick
Misquichilca S.A., with additional support of the Hugh E. McKinstry
Students Research Fund from the Society of Economic Geologists.
This research would not have been possible without the experience and
knowledge of Lagunas Norte staff, particularly Nick Teasdale, Jose

671
Nizama, and several geologists from Servicios Tcnicos of Lagunas
NorteBarrick. Sulfur and oxygen analysis was carried out in collaboration with Kerry Klassen and QFIR Lab in Queens University,
which is supported by NSERC Discovery, CFI, and OIT grants to
Kurt Kyser. Teresa Velardes continuous assistance in spectrometric
data is very much appreciated. Vctor Carlotto, Luis Miguel Muoz,
GR-13 team, and my co-workers of Regional Geology from Geological
Survey of Per (INGEMMET) have greatly helped with this project,
and their continuing enthusiasm and support are appreciated.
Discussions with Allan Montgomery, Amelia Rainbow, Huayong
Chen, and Fernando Tornos have greatly helped with this project and
reviewers Regina Baumgartner, Noel White, and David Cooke as well
as Editor Bernd Lehmann are thanked for their constructive reviews.

References
Alvarez-Ramis (1999) Tcnicas y mtodos fundamentales en
Paleobotnica. UNSAAC Fac. Biologia. Cusco. 40 p
Araneda R, Guerra R, Gaboury F, McEwan C, Soto R, Davidson AJ,
Hodgson J (2003) Proyecto del Alto Chicama, Distrito de Quiruvilca,
Departamento de La Libertad, Per. III Congreso Internacional de
Prospectores y Exploradores Proceedings CD-ROM
Arehart GB, Kesler SE, ONeil JR, Foland KA (1992) Evidence for the
supergene origin of alunite in sediment-hosted micron gold
deposits, Nevada. Econ Geol 87:263270
Atherton MP, Warden V, Sanderson LM (1985) The mesozoic marginal
basin of Central Peru: a geochemical study of within plate-edge
volcanism. In: Pitcher MP, Cobbing EJ, Beckinsale RD (eds)
Magmatism at a plate edge: the Peruvian Andes. Blackie
Halsted, London, pp 4758
Barrick Gold Corp. Annual Repport (2011) http://www.barrick.com/
investors/annual-report/default.aspx. Accessed 04 August 2012
Baumgartner R, Fontbot L, Spikings RA, Ovtcharova M, Schaltegger
U, Schneider J, Page L, Gutjahr M (2009) Bracketing the age of
magmatic-hydrothermal activity at the Cerro de Pasco epithermal
polymetallic deposit, Central Peru: a U-Pb and 40Ar/39Ar study.
Econ Geol 104:479504
Benavides-Cceres V (1956) Cretaceous system in Northern Peru. B
Am Mus Nat Hist 108:352494
Benavides-Cceres V (1999) Orogenic evolution of the Peruvian
Andes: the Andean cycle. Spec Publ 7:61107
Bissig T, Tosdal RM (2009) Petrogenetic and metallogenetic relationship in
the Eastern Cordillera Occidental of Central Peru. J Geol 117:499518
Bissig T, Clark A, Lee J, Hodgson C (2002) Miocene landscape
evolution and geomorphologic controls on epithermal processes
in the El Indio-Pascua Au-Ag-Cu belt, Chile and Argentina. Econ
Geol 97:971996
Bissig T, Ullrich TD, Tosdal RM, Friedman R, Ebert S (2008) The
time-space distribution of Eocene to Miocene magmatism in the
central Peruvian polymetallic province and its metallogenic implications. J S Am Earth Sci 26(1):1635
Chouinard A, Paquette J, Williams-Jones E (2005a) Crystallographic
controls on trace-element incorporation in auriferous pyrite from
the Pascua epithermal high-sulfidation deposit, Chile-Argentina.
Can Mineral 43(3):951963
Chouinard A, Williams-Jones A, Leonardson R, Hodgson C, Silva P,
Tllez C, Vega J, Rojas F (2005b) Geology and genesis of the
multistage high-sulfidation epithermal Pascua Au-Ag-Cu Deposit,
Chile and Argentina. Econ Geol 100:463490
Clayton RN, Mayeda TK (1963) The use of bromine pentafluoride in
the extraction of oxygen from oxides and silicates for oxygen
analysis. Geochim Cosmochim Acta 27:4352
Cobbing EJ, Pitcher WS, Wilson JJ, Baldock JW, Taylor WP, McCourt
W, Snelling NJ (1981) The geology of the Western Cordillera of

672
Northern Peru. Institute of Geological Sciences, London, Overseas
Memoir, 5, 143 p.
Cooke DR, Simmons SF (2000) Characteristics and genesis of epithermal gold deposits. Rev Econ Geol 13:221244
Cosso A, Jan H (1967) Geologa de los cuadrngulos de Pumape,
Chocope, Otuzco, Trujillo, Salaverry y Santa. Bol. 17. Servicio
Geologa y Minera. 141 p. Lima. Per
Craig H (1961) Isotopic variations in meteoric waters. Science
133:17021703
Davies AGS, Cooke DR, Gemmel JB (2000) Breccias associated with
epithermal and porphyry systemstowards a systematic approach to their description and interpretation. In: Bucci LA,
Mair JL (eds) Gold in 2000, Lake Tahoe, Nevada, pp 98103
Davies AGS, Cooke DR, Gemmel JB, Simpson KA (2008) Diatreme
breccias at the Kelian gold mine, Kalimantan, Indonesia: precursors to epithermal gold mineralization. Econ Geol 103:689716
Deyell C, Leonardson R, Rye R, Thompson J, Bissig T, Cooke D
(2005) Alunite in the Pascua-Lama high-sulfidation deposit: constraints on alteration and ore deposition using stable isotope
geochemistry. Econ Geol 100(1):131148
Dunin-Borkowski E (2000) El Carbn de Alto Chicama. X. Congreso
Geol. Per. CD Vol. Ext
Einaudi MT, Hedenquist JW, Inan E (2003) Sulfidation state of fluids in
active and extinct hydrothermal systems: transitions from porphyry
to epithermal environments. Soc Econ Geol Spec Publ 10:285314
Escudero J (1979) El carbn del Alto Chicama. INGEMMET. Serie B:
Geologa Econmica. 77 p. Lima Per
Garzione CN, Hoke GD, Libarkin JC, Whiters S, MacFadden B, Eiler
J, Ghosh P, Mulch A (2008) Rise of the Andes. Science,
320:13041307
Gifkins C, Herrmann W, Large R (2005) Altered volcanic rocks: a
guide to description and interpretation. University of Tasmania.
Centre for Ore Deposit Research CODES. Australia
Giggenbach WF (1992) Magma degassing and mineral deposition in
hydrothermal systems along convergent plate boundaries. Econ
Geol 87:19271944
Guerra MR (2001) Alto Chicama: Un ambiente de alteracin y
mineralizacin high sulphidation, Per. Memoria de Ttulo,
Santiago de Chile, Chile, Universidad de Chile, 150 p
Gutscher MA, Olivet JL, Aslanian D, Eissen JP, Maury R (1999) The
lost Inca Plateau: cause of flat subduction beneath Peru? Earth
Planet Sci Lett 171:335341
Hampel A (2002) The migration history of the Nazca Ridge along the
Peruvian active margin: a re-evaluation. Earth Planet Sci Lett
203:665679
Hemley JJ, Montoya JW, Marinenko JW, Luce RW (1980) Equilibria
in the system Al2O3-SiO2-H2O and some general implications for
alteration/mineralization processes. Econ Geol 75:210228
Hubberten HW (1980) Sulfur isotope fractionation in the Pb-S, Cu-S
and Ag-As systems. Geochem J 14:177184
INGEMMET (1999) Mapa Geolgico del Per, escala 1:1 000 000
Jaillard E, Jacay J (1989) Les couches Chicama du Nord du Prou:
colmatage dun basin n dune collision oblique au Tithonique,
Comptes Rendus delAcadmie des Sciences, Pars, v. 308. II, pp
14591465
Jaillard E, Soler P (1996) The Cretaceous to Early Paleogene tectonic
evolution of the northern Central Andes and its relations to geodynamics. Tectonophysics 259:4153
Jaillard E, Hrail G, Monfret T, Daz-Martnez E, Baby P, Lavenu A,
Dumont JF (2000) Tectonic evolution of the Andes of Ecuador,
Peru, Bolivia and Northernmost Chile. In: Cordani UG, Milani
EJ, Thomaz Filho A, Campos DA (eds) Tectonic Evolution of
South America, 31st Int. Geol. Congress. Brazil, pp 481559
Kouzmanov K, Bendez A, Catchpole H, Ageneau M, Prez J (2008)
The Miocene Morococha District, Central Perularge epithermal
polymetallic overprint on multiple intrusion-centred porphyry

Miner Deposita (2013) 48:653673


system. PACRIM Congress 2008. Queensland, Australia,
Extended Abstracts 117121
Longo AA, Dilles JH, Grunder AL, Duncan R (2010) Evolution of
calc-alkaline volcanism and associated hydrothermal gold deposits at Yanacocha, Peru. Econ Geol 105:11911241
Lorenz V (2003) Maar-diatreme volcanoes, their formation, and
their setting in hard-rock or soft-rock environments. Geolines
15:7283
Lorenz V, Kurszlaukis S (2007) Root zone processes in the phreatomagmatic pipe emplacement model and consequences for the evolution
of maar-diatreme volcanoes. J Volcanol Geotherm Res 159:432
Macassi A (2005) Facies de brecha en el area de Dafne, Lagunas
NorteAlto Chicama. Distrito de Quiruvilca, Provincia de
Santiago de Chuco, Departamento de La Libertad. Engineer
Thesis. Universidad Nacional de Ingeniera. 123 p
Manrique AI (1986) Geologa econmica de la Cuenca de Alto
Chicama, Santa, Oyn y Jatunhuasi. Promocin de la minera
del Carbn en el Per. Reporte Cia. Minera San Ignacio de
Morococha, Lima, Per, 86 p.
Mgard F (1987) Cordilleran and marginal Andes: a review of Andean
geology north of the Arica elbow (18S). In: Monger JWH,
Francheteau J (eds.) Circum-Pacific orogenic and evolution of
the Pacific Ocean basin. American Geophysical Union,
Geodynamic series 18:7195
Mojares EM, Tomita K, Kawano M (2001) Characterization of 2M
pyrophyllite associated with argillic alteration in steam-heated
environment, Solo, Mabini, Philippines. J Mineral Petrol Sci
96:109119
Montgomery AT (2012) Metallogenetic controls on Miocene high
sulphidation epithermal gold mineralization, Alto Chicama district, La Libertad, Northern Peru. Department of Geological
Sciences and Geological Engineering. Kingston, Queens
University. PhD: 382
Moulin N (1989) Facies et sequences de depot de la plate-forme du
Jurassique moyen lAlbien, et une coupe structurale des Andes
du Prou central, Unpublished PhD. Thesis. Paris, France.
Universit Paris XI, 275 p
Noble D, McKee E (1999) The Miocene metallogenic belt of central
and northern Peru. In: Skinner BJ (ed) Geology and ore deposits
of the central Andes. Society of Economic Geologists Special
Publication No 7, College Station, Texas, pp 155193
Ohmoto HA, Lasaga AC (1982) Kinetics and reactions between aqueous sulfates and sulfides in hydrothermal systems. Geochim
Cosmochim Acta 46:17271745
Ohmoto H, Rye RO (1979) Isotopes of sulfur and carbn. In: Barnes
HL (ed) Geochemistry of hydrothermal ore deposits, 2nd edn.
Wiley, New York, pp 509567
Poage MA, Chamberlain P (2001) Empirical relationships between
elevation and the stable isotope composition of precipitation and
surface waters: considerations for studies of paleoelevation
change. Am J Sci 301:115
Rainbow A (2009) Genesis and evolution of the Pierina high-sulphidation
epithermal Au-Ag Deposit, Ancash, Per. Department of Geological
Sciences and Geological Engineering. Kingston, Queens
University. PhD: 292
Rainbow A, Clark AH, Kyser TK, Gaboury F, Hodgson CJ (2005) The
Pierina epithermal Au-Ag deposit, Ancash, Peru: paragenetic
relationships, alunite textures, and stable-isotope geochemistry.
Chem Geol 215:235252
Reyes L (1980) Geologa de los Cuadrngulos de Cajamarca, San
Marcos y Cajabamba (Hojas 15f, 15g y 15h). Boletin del
Instituto Geolgico Minero y Metalrgico. Lima, Per
Ros H (2005) El Yacimiento epitermal de Oro de Alta Sulfuracin de
Alto Chicama, controles de mineralizacin y modelo gentico
preliminar. Tesis de Ingeniero Geologo. Universidad Nacional
de Ingenieria. 95 p

Miner Deposita (2013) 48:653673


Rivera M, Monge R, Navarro P (2005) Nuevos datos sobre el volcanismo
cenozoico (Grupo Calipuy) en el Norte del Per: Departamentos de
la Libertad y Ancash. Bol Soc Geol Per 99:721
Rosenbaum G, Giles D, Saxon M, Betts PG, Weinberg RF, Duboz C (2005)
Subduction of the Nazca Ridge and the Inca Plateau: insights into the
formation of ore deposits in Peru. Earth Planet Sci Lett 239:1832
Rye RO (1993) The evolution of magmatic fluids in the epithermal
environment: the stable isotope perspective. Econ Geol 88:733753
Rye RO (2005) A review of the stable-isotope geochemistry of sulfate
minerals in selected igneous environments and related hydrothermal systems. Chem Geol 215:536
Rye RO, Bethke PM, Wasserman MD (1992) The stable isotope
geochemistry of acid-sulfate alteration. Econ Geol 87:225262
Savin SM, Epstein S (1970) The oxygen and hydrogen isotope geochemistry of clays minerals. Geochim Cosmochim Acta 34:2442
Sillitoe R (1985) Ore-related breccias in volcanoplutonic arcs. Econ
Geol 80:14671514
Simmons SF, White NC, John DA (2005) Geological characteristics of
epithermal precious and base metal deposits. In: Hedenquist JW,
Thompson JFH, Goldfarb RJ, Richards PJ (eds) Economic Geology:
one hundredth anniversary volume 19052005, Society of
Economic Geologist, Littleton, Colorado, pp 455522
Soler P (1991) Contribution ltude du magmatisme associ aux zones
de subduction. Ptrographie, gchimie et gochimie isotopique des
roches intrusive sur un transect des Andes du Prou Central.
Implications godynamiques et mtallogniques: Unpublished
Ph.D. thesis, Paris, France, Universit Paris VI, 950 p

673
Soler P, Bonhomme MG (1990) Relation of magmatic activity to plate
dynamics in central Peru from Late Cretaceous to present. In: Kay
SM, Rapla CW (eds) Plutonism from Antartica to Alaska:
Boulder Colorado, Geological Society of America Special Paper
241:173192
Stappenbeck R (1929) Geologie des Chicamatales in Nordperu und
seiner Anthrazitlagenstaetten. Geol. und Paleont. Abhandl. N. F.
16, H. 14
Stoffregen RE, Rye RO, Wasserman MD (1994) Experimental studies
of alunite: I. 18O-16O and D-H fractionation factors between
alunite and water at 250450C. Geochim Cosmochim Acta
58:903916
Taylor BE (1988) Degassing of rhyolitic magmas: hydrogen isotope
evidence and implications for magmatic-hydrothermal ore deposits. Can Inst Min Mineral Spec Vol 39:3349
Thompson AJB, Hauff PL, Robitaille AJ (1999) Alteration mapping in
exploration: application of short-wave infrared (SWIR) spectroscopy. Soc Econ Geol Newsl 39:13 (Oct.)
Walters A (2006) Dynamical constraints on Kimberlite volcanism. J
Volcanol Geotherm Res 155:1848
Wasserman MD, Rye RO, Bethke PM, Arribas Jr A (1992) Methods
for separation and total stables isotope analysis of alunite. U.S.
Geol. Surv. Open-File Rep. 92-9
Wilson PA (1975) Potassium-argon age studies in Peru with special
reference to the emplacement of the Coastal Batholith.
Unpublished Ph.D. thesis. Liverpool, England, University of
Liverpool. 104 p

Você também pode gostar