You are on page 1of 32

Economic Geology

Vol. 98, 2003, pp. 15751605

Porphyry-Style Alteration and Mineralization of the Middle Eocene to


Early Oligocene Andahuaylas-Yauri Belt, Cuzco Region, Peru
JOS PERELL,
Antofagasta Minerals S.A. Ahumada 11, Oficina 602, Santiago, Chile

VCTOR CARLOTTO,
Departamento de Geologa, Universidad Nacional San Antonio Abad del Cuzco. Avenida de la Cultura, Cuzco, Per

ALBERTO ZRATE, PEDRO RAMOS, HCTOR POSSO, CARLOS NEYRA, ALBERTO CABALLERO,
Minera Anaconda Per S.A. Avenida Paseo de la Repblica 3245, Piso 3, San Isidro, Lima 27, Per

NICOLS FUSTER, AND RICARDO MUHR


Antofagasta Minerals S.A. Ahumada 11, Oficina 602, Santiago, Chile

Abstract
Originally known for its Fe-Cu skarn mineralization, the Andahuaylas-Yauri belt of southeastern Peru is
rapidly emerging as an important porphyry copper province. Field work by the authors confirms that mineralization in the belt is spatially and temporally associated with the middle Eocene to early Oligocene (~4832
Ma), calc-alkaline Andahuaylas-Yauri batholith, a composite body with an areal extent of ~300 130 km emplaced into clastic and carbonate strata (e.g., Yura Group and Ferrobamba Formation) of Jurassic to Cretaceous age. Batholith emplacement included early-stage, mafic, cumulate gabbro and diorite between ~48 and
43 Ma, followed by pulses of granodiorite and quartz monzodiorite at ~40 to 32 Ma. Coeval volcanic rocks
make up the middle Eocene to early Oligocene Anta Formation, a sequence of >1,000 m of andesite lava flows
and dacite pyroclastic flows with interbedded volcaniclastic conglomerate. Sedimentary rocks include the red
beds of the Eocene to early Oligocene San Jernimo Group and the postmineralization late Oligocene to
Miocene Punacancha and Paruro formations. Eocene and Oligocene volcanic and sedimentary rocks are interpreted to have accumulated largely in both transtensional and contractional synorogenic basins. New and
previously published K-Ar and Re-Os ages show that much of the porphyry-style alteration and mineralization
along the belt took place during the middle Eocene to early Oligocene (~4230 Ma). Thus, batholithic magma
emplacement, volcanism, and sedimentation are inferred to have accompanied a period of intense deformation, crustal shortening, and regional surface uplift broadly synchronous with the Incaic orogeny. Supergene
mineralization is inferred to have been active since the Pliocene on the basis of geomorphologic evidence and
a single K-Ar determination (3.3 0.2 Ma) on supergene alunite.
The belt is defined by 31 systems with porphyry-style alteration and mineralization, including 19 systems
grouped in 5 main clusters plus 12 separate centers, and by hundreds of occurrences of magnetite-rich, skarntype Fe-Cu mineralization. Porphyry copper stocks are dominated by calc-alkaline, biotite- and amphibolebearing intrusions of granodioritic composition, but monzogranitic, monzonitic, quartz-monzonitic, and monzodioritic stocks occur locally. Hydrothermal alteration includes sericite-clay-chlorite, and potassic,
quartz-sericitic, and propylitic assemblages. Calcic-potassic and advanced argillic alteration associations are locally represented, and calc-silicate assemblages with skarn-type mineralization occur where carbonate country
rocks predominate.
Porphyry copper deposits and prospects of the belt range from gold-rich, molybdenum-poor examples
(Cotabambas), through deposits carrying both gold and molybdenum (Tintaya, Los Chancas), to relatively
molybdenum-rich, gold-poor end members (Lahuani). Gold-only porphyry systems are also represented (Morosayhuas). Gold-rich porphyry copper systems are rich in hydrothermal magnetite and display a positive correlation between Cu and Au in potassic alteration. The bulk of the hypogene Cu (-Au, -Mo) mineralization occurs in the form of chalcopyrite and bornite, in intimate association with early-stage potassic alteration which,
in many deposits and prospects, is variably overprinted by copper-depleting sericite-clay-chlorite alteration.
Most porphyry copper systems of the belt lack economically significant zones of supergene chalcocite enrichment. This is due primarily to their relatively low pyrite contents, the restricted development of quartzsericitic alteration, and the high neutralization capacities of both potassic alteration zones and carbonate country rocks as well as geomorphologic factors. Leached cappings are irregular, typically goethitic, and contain
copper oxide minerals developed by in situ oxidation of low-pyrite, chalcopyrite (-bornite) mineralization. Porphyry copper-bearing stocks emplaced in the clastic strata of the Yura Group and certain phases of the Andahuaylas-Yauri batholith may develop appreciable supergene chalcocite enrichment in structurally and lithologically favorable zones.

Corresponding author: e-mail, jperello@aminerals.cl

0361-0128/01/3394/1575-31 $6.00

1575

1576

PERELL ET AL.

A model for the region suggests that the calc-alkaline magmas of the Andahuaylas-Yauri batholith and subsequent porphyry-style mineralization were generated during an event of subduction flattening which triggered
the crustal shortening, tectonism, and uplift assigned to the Incaic orogeny. Shortening of the upper crust
would have impeded rapid magma ascent favoring storage of fluid in large chambers which, at the appropriate
depth in the uppermost crust, would have promoted large-scale porphyry copper emplacement. Geodynamic
reconstructions of the late Eocene to early Oligocene period of flat subduction in the central Andes suggest
that emplacement of the Andahuaylas-Yauri batholith took place at an inflection corridor in the subduction
zone broadly coincident with the position of the present-day Abancay deflection. Similarly, evidence from
southeastern Peru suggests that the Andahuaylas-Yauri belt may be continuous with the late Eocene to early
Oligocene porphyry copper belt of northern Chile and that the process of subduction flattening in southern
Peru also may have taken place in northern Chile between ~45 and 35 Ma.

Introduction
THE ANDAHUAYLAS-YAURI belt (Bellido et al., 1972; Santa
Cruz et al., 1979; Noble et al., 1984) covers an area of approximately 25,000 km2 in southern Peru and extends for
about 300 km between the localities of Andahuaylas in the
northwest and Yauri in the southeast (Fig. 1a). Until the late
1980s, the Andahuaylas-Yauri belt had received only limited
geologic scrutiny and was mainly known for its copper-bearing, magnetite skarn deposits (Terrones, 1958; Bellido et al.,
1972; Sillitoe, 1976, 1990; Santa Cruz et al., 1979; Einaudi et
al., 1981; Aizawa and Tomizawa, 1986), best exemplified by
Tintaya, Atalaya, Las Bambas, Katanga, and Quechua. For
most researchers, these occurrences were considered to be
copper skarns associated with barren intrusions (e.g., Einaudi
et al., 1981; Noble et al., 1984), although potassic alteration in
host porphyritic stocks had been described and characterized
as such (Yoshikawa et al., 1976; MMAJ, 1983; Noble et al.,
1984). During the late 1980s, regional work complemented
by detailed geologic studies at Tintaya and Katanga (Carlier et
al., 1989), followed by grass-roots exploration in the region
during the 1990s, confirmed the presence of porphyry-style
alteration and mineralization (e.g., Fierro et al., 1997) and resulted in the discovery of additional, potentially economic
porphyry copper deposits (Table 1) at Antapaccay (Jones et
al., 2000), Los Chancas (Corrales, 2001), and Cotabambas
(Perell et al., 2002), as well as porphyry-skarn mineralization
at Coroccohuayco (BHP Company Limited, 1999). Zinc-rich,
Mississippi Valley-type mineralization was also discovered in
the region (Carman et al., 2000) adding to the metallogenic
diversity of the belt.
This paper describes the salient geologic features of a
number of porphyry Cu (-Au, -Mo) deposits and prospects of
the Andahuaylas-Yauri belt that help to define this region as a
new porphyry copper province. It also provides new
geochronologic data to constrain the age of the porphyry-style
alteration and mineralization in the belt and establishes regional correlations and comparisons with nearby porphyry
copper provinces. However, the paper is not designed to
cover in full the complex geology of this still poorly understood region. Detailed geologic descriptions can be found in
Marocco (1978) and Carlotto (1998) for the area under study,
and in Clark et al. (1990) and Sandeman et al. (1995) for
nearby southeastern Peru transects. The paper focuses on
systems for which the bulk of the mineralization is of porphyry type and excludes those deposits in which skarn-type
mineralization is the dominant style. Descriptions of the latter can be found elsewhere (Terrones, 1958; Santa Cruz et al.,
1979; Aizawa and Tomizawa, 1986; Fierro et al., 1997; Zweng
0361-0128/98/000/000-00 $6.00

et al., 1997). Following a short review of the regional geologic


setting of the Andahuaylas-Yauri belt, the main geologic features of several deposits and prospects are described. The
paper concludes with a section in which regional metallogenic
aspects are reviewed.
Methods
Except for those deposits and prospects with published descriptions (e.g., Tintaya, Antapaccay, Los Chancas), much of
the work represents the product of more than three years of
exploration by the authors, including both regional (1:25,000
scale) and detailed (1:5,000 scale) mapping. Field work was
complemented by thin section petrographic studies to characterize rock types, alteration assemblages, and dominant vein
styles at each prospect. Rock names for the main batholith intrusions and porphyry copper-bearing stocks follow the
nomenclature of Streckeisen (1976, 1978) and are based on
point counts (1,500 points) for modal proportions of key silicates. Unless otherwise stated, the K-Ar ages reported here
were determined at the geochronology laboratory of the Geological Survey of Chile, Santiago and followed standard procedures and techniques (e.g., Dalrymple and Lamphere, 1966;
Steiger and Jaeger, 1977; Baksi, 1982). All ages are referred to
the geological time table of Haq and van Eysinga (1987).
Regional Setting
The Andahuaylas-Yauri belt is located at a distance of ~250
to 300 km inland from the present-day Peru-Chile trench
(Fig. 1). The region is underlain by thick sialic crust (50 to 60
km; James, 1971), and straddles the transition zone between
the southern, normal subduction regime of southern Peru
and northern Chile and the northern, flat subduction zone of
central and northern Peru (Cahill and Isacks, 1992). It is located immediately southeast of the Abancay Deflection
(Marocco, 1978). The region encompasses parts of the intermontane depressions between the Eastern and Western
Cordilleras and the northern extremity of the Altiplano (Fig.
1b; Carlier et al., 1996; Chvez et al., 1996). The western part
of the belt is characterized by a rugged, mountainous topography where ranges and snow-capped peaks above 4,500 m
are incised by deep (>2,000 m), steep-sided canyons. These
canyons constitute the main drainage system of the region
and include the Santo Toms, Urubamba, Apurmac, Vilcabamba, Mollebamba, and Antabamba rivers, all of which
drain toward the Amazon basin. The eastern and southern
parts of the region are characterized by the gently undulating
topography of the ~4,000 m-high plateaus that extends into
the Altiplano of Bolivia (Fig. 1b).

1576

1577

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

Lima 100
75
50

125 150

75

65

70

BR
A

PERU

80
10

70

60

50

40

ZIL
0

M Cuzco
A AC

10

600

BOLIVIA

15

PE

Puno

RU

e
dg

La Paz

N
az
ca

Ri

40

CH

50

IL

E
Arica

R idge

NCH

Oligocene
Crust

CHILE

TRE

ne st
ce Cru
o
E e
rly en
Ea leoc
a
P

m/yr
8.5 c

30

Arequipa

Middle-Late
Eocene Crust

20

20

Lake Titicaca

da
di

LEGEND

Calama

r
Pe

125

Benioff Zone Contours (km)

175 200

Elevation > 3 km
Study Area
75

65

70

A:

Lima

AC: Abancay

AMAZON CRATON
AC
A

Andahuaylas

M Cuzco

Y:

Yauri

M : Machu Pichu
Y

15

50

Puno
Arequipa

La Paz

Precambrian and
Paleozoic terranes with
Grenville basement

60
PACIFIC OCEAN

Contour on Moho (km)

Arica

Altiplano

70

Western Cordillera
50

Calama

PA
TE MP
RR EA
AN N
E

20

Eastern Cordillera
0

250

FIG. 1. Sketch maps showing the location of the study area in the context of main geologic, geophysical, topographic, and
physiographic features of the Central Andes. a. Area with average elevation >3,000 m and depth contours of the subducted
slab after Cahill and Isacks (1992). Oceanic features from Jaillard et al. (2000). b. The study area relative to main regional
physiographic provinces (Jaillard et al. 2000), contours of crustal thickness (James, 1971), and main Precambrian basement
units (Ramos and Aleman, 2000).
0361-0128/98/000/000-00 $6.00

1577

500 km

1578

PERELL ET AL.
TABLE 1. Geologic Resources for Main Deposits of the Andahuaylas-Yauri Belt
Tonnage ( 106)

Cu (%)

Au (g/t)

Mo (%)

Tintaya district
Antapaccay
Coroccohuayco
Ccatun Pucara
Quechua
Tintaya

383
155
24
300
139

0.89
1.57
1.44
0.68
1.39

0.16
0.16
n.a.
n.a.
0.23

n.a.
n.a.
n.a.
n.a.
n.a.

Cotabambas Area
Azulccacca
Ccalla

24
112

0.42
0.62

0.39
0.36

<0.01
<0.01

Los Chancas

200

1.00

0.12

0.08

Deposit

1BHP

Main reference
Jones et al. (2000); Fierro et al. (2002)
BHP (1999)
BHP (1999)
E. Tejada (pers. commun., 2003)
BHP Billiton (2003)1
Perell et al. (2002)
Perell et al. (2002)
Corrales (2001)

Billiton corporate website: <www. BHPBilliton.com>

Precambrian and Paleozoic basement


Precambrian gneisses at Ro Pichari, ~130 km northwest of
Cuzco (Carlotto, 1998) are probable extensions of the
Maran massif exposed farther north and are interpreted by
Ramos and Aleman (2000) to constitute remnants of
perigondwanan terranes attached to the Amazonian Craton in
the Early Paleozoic (Fig. 1b). Paleozoic rocks in the region include >10,000 m of volcanosedimentary, marine, and continental rocks of Cambrian(?) to Early Permian age (Marocco,
1978; Carlotto et al., 1996a; Carlotto et al., 1997). The upper
part of the pre-Andean basement is dominated by >1,000 m
of volcanic and clastic rocks of the Mitu Group (Permian to
Early Triassic; Fig. 2).
Mesozoic and Cenozoic stratigraphy
The Mesozoic and Cenozoic stratigraphy of the region is
chiefly made up of Jurassic and Cretaceous sedimentary sequences deposited in a paleogeographic setting dominated by
two main basins (Western and Eastern Peruvian basins) separated by the Cuzco-Puno basement high (Fig. 3; Carlotto et
al., 1993; Jaillard and Soler, 1996). The Western basin, also
known as the Arequipa basin (Vicente et al., 1982), corresponds to the present-day Western Cordillera. It contains a
sedimentary pile (Middle Jurassic to Late Cretaceous) in excess of 4,500 m thick with a lower part dominated by turbidites, a middle part with quartz arenite, and an upper part
with abundant limestone (Vicente et al., 1982; Jaillard and
Santander, 1992). The northeastern edge of this basin, coincident with the Andahuaylas-Yauri region, includes the Lagunillas and Yura groups (Marocco, 1978), made up of Early
Jurassic limestone and Middle to Late Jurassic quartz arenite
and shale, with a total thickness of approximately 800 m (Fig.
3). The top of the sequence contains the massive micritic
limestone, black shale, and nodular chert of the Ferrobamba
Formation (Marocco, 1978; Pecho, 1981). The Cuzco-Puno
high includes ~900 m of terrigenous red beds interbedded
with shale, limestone, and gypsum (Carlotto et al., 1993; Jaillard et al., 1994). The age of these rocks is Late Jurassic to Paleocene (Fig. 3). The Eastern basin, also known as Putina
basin (Jaillard, 1994), is made up of several sequences of Late
Cretaceous marine clastic and carbonate rocks, with a total
thickness of ~ 2,600 m (Jaillard et al., 1993; Jaillard, 1994;
Crdenas et al., 1997).
0361-0128/98/000/000-00 $6.00

Eocene to early Oligocene stratigraphy


Two main units characterize the Eocene to early Oligocene
stratigraphy of the region, including the sedimentary San
Jernimo Group and the dominantly volcanic Anta Formation
(Figs. 2 and 4). These units unconformably overlie the Mesozoic and early Cenozoic sequences described above. The San
Jernimo Group (Eocene to early Oligocene) consists of two
main formations (Kayra and Soncco; Fig. 4), with a total
thickness of ~4,500 m, made up of red bed terrigenous (sandstone, shale, pelitic sandstone, and volcanic microconglomerate) strata interbedded with tuffaceous horizons near the top.
The age of the San Jernimo Group is constrained by stratigraphic relations (it unconformably overlies strata with plant
fossils of Paleocene to early Eocene age) and on K-Ar and ArAr ages of 29.9 1.4 Ma and 30.84 0.83 Ma, respectively,
from the upper tuffaceous horizons of the Soncco Formation
(Fig. 4; Carlotto, 1998; Fornari et al., 2002). Sedimentation is
interpreted to have taken place initially in a fluvial environment that progressed into structurally controlled, pull-apart
basins (Crdova, 1986; Noblet et al., 1987; Marocco and Noblet, 1990; Chvez et al., 1996). Between Cuzco and Sicuani,
basal sandstone of the Soncco Formation includes horizons of
stratiform copper mineralization, up to several meters thick,
with hypogene chalcocite and bornite, and supergene copper
oxides (Crdenas et al., 1999), which have similarities to the
red bed deposits from the Bolivian Altiplano (e.g., Corocoro;
Sillitoe, 1989) and northern Chile (San Bartolo; Travisany,
1979). The San Jernimo Group is equivalent to the Puno
Group of the Peruvian Altiplano southeast of the study region
(Fig. 4), where it is overlain by the volcanic horizons of the
Tacaza Group (Klinck et al., 1986; Clark et al., 1990; Jaillard
and Santander, 1992). Farther south, sedimentary, conglomeratic, red bed sequences are known in the Altiplano of Bolivia (e.g., the lower horizons of the Tiwanaku Formation and
the Berenguela and Turco formations; Hrail et al., 1993), in
the Puna of northwestern Argentina (Geste and Quioa formations; Alonso, 1992; Kraemer et al., 1999; Coutand et al.,
2001), and in the Salar de Atacama area of northern Chile
(upper Purilactis Group; Mpodozis et al., 1999).
The Anta Formation is a >1,000 m sequence characterized by a lower member with andesite lava flows and dacite
pyroclastic flows locally interbedded with alluvial conglomerate, and an upper member of fluvial conglomerate with

1578

0361-0128/98/000/000-00 $6.00

1579

Figure.-2

Morosayhuas
Cotabambas
Chaccaro
Ferrobamba
Chalcobamba

11
12
13
14
15

Tintaya
Los Chancas
Pea Alta
Leonor
Panchita

16 Lahuani
17 Trapiche

7230

F ault

5 4

Cuzco

a chay

ult

Fa

25

Velille

50 km

7200

Santo Toms

ba
s

rco
sF
au

Pa

oF
a

ul t

10

Livitaca

Mesozoic to early Cenozoic marine sedimentary


sequences (Yura and Lagunillas Gps; Ferrobamba Fm.)

Yauri

7130

S ic
ua

Fault

Ay
av
ir

iF

au
lt

Putina
Basin

Sicuani

ul t

11

Fa

Reverse Fault

Porphyry Cu cluster/deposit

7100

Syncline

Anticline

Early Paleozoic basement (undifferentiated)

Late Paleozoic to Early Triassic basement


a: Mainy granitoids,
b b: Mainy volcanic rocks (Mitu Gp.)

Pomacanchis

rur

Urcos

Accha

lt

ul

Paruro

co-U

lt

Cu z

Fa
u

ul t

Yau ri Fault

Fa

FIG. 2. Geologic map of the study area, modified and greatly simplified after Carlotto (1998), with additions after Pecho (1981) and this study.

6 Alicia
7 Cristo de los Andes
8 Katanga
9 Portada
10 Winicocha

lt

17

Cotabambas

R ecord

bo
m

m
ba
ta
o
C

7300

14

Curahuasi

Ta
m

Fa

16

F ault

13

a
A ban ca y F

Abancay

15
Mo
lleb
amba

lt

Lima ta mbo F a ult

is

1
2
3
4
5

12

Fau

Eocene to early Oligocene Andahuaylas-Yauri Batholith


Eocene to early Oligocene continental rocks
a: Volcanic and sedimentary sequences (Anta Fm.)
b b: Red bed sequences (San Jernimo and Puno Gps.)

Oligocene to Miocene subaerial volcanic rocks


(Tacaza and Sillapaca Gps. and equivalent units)

Pliocene shoshonitic volcanic rocks


Oligocene to Miocene continental sedimentary rocks
(Paruro, Punacancha Fms. and equivalent units)

Miocene to Pliocene

h
a nc
om a c

7330

Chalhuanca

ul

Fa

ac

a
mb

ni

1430

Andahuaylas

Machu Pichu

lt

1400

1330

pu
au

i ca
C us

alhuanc
a

1300

A
h
cc
ua n
a -H

eF
uit
oq
P

Ch

Lakes

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

1579

1580
LATE CRETACEOUS
(MAASTRICHTIAN)

PERELL ET AL.

NE BORDER OF THE
WESTERN (AREQUI PA) BASIN

EASTERN (PUTINA)
BASIN

CUZCO-PUNO
HIGH

E
PU

A
H

EARLY-LATE (TURONIAN)
CRETACEOUS

HU
HU

CH
PALEOZOI C
BASEMENT

MIDDLE (BAJOCIAN) -LATE


(TITHONIAN) JURASSIC

CH

?
YURA GP.

FERROBAMBA

A AYAVACAS

M MARA

H HUANCANE

S SORAYA

HU HUAMBUTIO

G GRAMADAL

AA ANTA-ANTA

P PISTE

PU PUQUIN

CH CHUQUIBAMBILLA

50

L LAGUNILLAS GP.

100 km

FIG. 3. Schematic paleogeographic reconstruction of the backarc basin of southern Peru during the Mesozoic and the earliest Cenozoic. Main stratigraphic units and correlations after Vicente et al. (1982), Jaillard (1994), Jaillard et al. (1994, 2000),
and Carlotto (1998). See text and Figure 8 for dominant rock types of each sequence.

interbedded andesite and basaltic andesite flows (Fig. 4). Its


age is constrained to middle Eocene to early Oligocene by
stratigraphic relations and K-Ar geochronology (Carlier et al.,
1996; Carlotto, 1998). Southwest of Cuzco, two biotite-rich
dacitic flows from the middle part of the formation have returned K-Ar ages of 38.4 1.5 and 37.9 1.4 Ma, and a
basaltic horizon from the upper part of the unit yields a K-Ar
whole rock age of 29.9 1.1 Ma (Carlotto, 1998). The Anta
Formation andesites and conglomerates are interpreted to be
stratigraphic equivalents of the San Jernimo Group red beds
(Fig. 4), with the erosion products of the Anta Formation
feeding the San Jernimo basin located to the northeast. The
coarsening-upward characteristics of the sequence, with alluvial and fluvial conglomerates dominated by volcanic and plutonic clasts at the top, are interpreted to reflect topographic
rejuvenation of the source regions in response to increasing
regional tectonic uplift, with sedimentation in a piggy-back
style basin environment (Carlotto, 1998).
Late Oligocene to Miocene stratigraphy
The late Oligocene to Miocene sedimentary deposits of the
region include the Punacancha (1,5005,000 m thick) and
Paruro (>1,100 m-thick) formations (Fig. 4). They are dominated by coarsening-upward red shale and sandstone, with
0361-0128/98/000/000-00 $6.00

gypsum and conglomerate being characteristic in the upper


parts of the sequences. Sedimentation is interpreted to have
taken place in a fluvial environment with braided rivers, flood
plains and alluvial fans in structurally controlled basins (Carlotto et al., 1996a, 1997; Jaimes et al., 1997; Romero et al.,
1997). The age of these sequences is based on stratigraphic
relations and fossil flora, as well as a K-Ar age of 10.1 0.5
Ma for a tuffaceous horizon near the base of the Paruro Formation (Carlotto et al., 1997).
Oligocene and Miocene volcanic rocks in the region and
nearby areas are largely dominated by the calc-alkaline sequences of the Western Cordillera (Inner-Western Cordillera
of Sandeman et al., 1995) and Altiplano, and include the
Tacaza (Oligocene) and Sillapaca (Miocene) groups. In addition to these, a series of scattered, small shoshonitic volcanic
centers of Pliocene to Quaternary age occur in the region
(Figs. 2 and 4; Wasteneys, 1990; Carlier et al., 1996; Carlotto,
1998). The Tacaza Group consists dominantly of trachyandesite, andesite, and rhyolite tuff (Klinck et al., 1986;
Wasteneys, 1990; Carlotto, 1998), with shoshonitic rocks
being important in the Santa Luca area, southeast of Yauri
(Clark et al., 1990; Sandeman et al., 1995). Shoshonitic volcanism in the Santa Luca area took place between ~32 and
24 Ma (Fig. 4; Clark et al., 1990; Sandeman et al., 1995),

1580

PLIOCENE

1581

PALEOCENE

LATEST
TOQUEPALA
PLUTONISM/
VOLCANISM

ATASPACA
PLUTONS

MOQUEGUA FM

HUAYLILLAS FM

CHUNTACALA FM

CAPILLUNE FM

BARROSO GP

AMBATO GP

B
WESTERN
CORDILLERA

ANDAHUAYLASYAURI
BATHOLITH

PUNO GP

ANTA
FM

TACAZA GP

SANTA
LUCIA
FM

SILLAPACA
GP

BARROS O
GP

EARLY TER TIARY


REDBED S

KAYRA FM

Red bed Copper

SAN JERONIMO
GP

SONCCO FM

PUNACANCHA FM

PARURO FM

SANTO TOMAS
IGNIMBRITES

CUZCO-SICUANI
VOLCANOE S

C
WESTERN CORDILLERA/
ALTIPLANO

D
EASTERN
CORDILLERA

15
PERU

PICOTANI
GP

QUENAMAR I
GP

MACUSANI FM

ARCO-AJA
FM

B
BOLIVIA

CHILE

70

Dominantly mafic cumulate


calc-alkaline intrusions

Dominantly intermediate
compositioncalc-alkaline
intrusions

Dominantly syenogranitic
intrusions

Peraluminous monzogranitic
intrusions

Evaporites

Dominantly mollasic
sedimentation

Mixed dacitic/basaltic-andesitic
or lamprophyric volcanism

Dominantly rhyolitic to
rhyodacitic volcanism

Dominantly dacitic volcanism

Dominantly andesitic volcanism

Apparent non-volcanic interval

FIG. 4. Summary stratigraphic columns for representative Eocene to present-day volcanic, sedimentary, and intrusive units of the study area and nearby southeastern
Peru transects. Columns A, B, and D simplified after Sandeman et al. (1995) and references therein and A.H. Clark (pers. commun., 2002). Column C for the study
area compiled after Carlotto (1998), with additions after Carlier et al. (1989, 1996) and this study. In column C, note the spatial and temporal relationships between
batholithic plutons, volcanic rocks of the Anta Formation, and the sedimentary red bed sequences of the San Jernimo Group.

60

55

50

40

37

34

30

28.5

24

20

16

11

10

1.8

PLEISTOCENE 0

Age (Ma)

MIOCENE

LATE

MIDDLE

EARLY

LATE

EARLY

LATE

MIDDLE

OLIGOCENE

EOCENE

0361-0128/98/000/000-00 $6.00

EARLY

A
ARC FRONT

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

1581

1582

PERELL ET AL.

whereas farther south, along the arc front, Tacaza-equivalent


pyroclastic flows intercalated with the molassic Moquegua
Formation commenced at ~26 Ma (A. H. Clark, pers. commun., 2002). In the Andahuaylas-Yauri region, similar age
(~29 Ma) shoshonitic rocks have been interpreted to be part
of uppermost Anta Formation (see above and Carlotto, 1998),
thereby implying some degree of temporal overlap with
Tacaza rocks (Fig. 4). Sillapaca Group rocks include mainly
dacite flows with subordinate andesite in the southeastern
part of the study region (Carlotto, 1998) and subvolcanic
dacite plugs and ash-flow tuff in the Santa Luca area (Fig. 4;
Clark et al., 1990). Rocks from Santa Luca yield ages of between ~22 and 14 Ma (Clark et al., 1990; Sandeman et al.,
1995); elsewhere in the Puno region a second effusive event,
also assigned to the Sillapaca Group, returns ages of between
~14 to 12 Ma (Klinck et al., 1986).
If the correlations above are accepted (Fig. 4), it may be
speculated that the temporal overlap of the Oligocene to
Miocene volcanism of the Western Cordillera, Altiplano, and
Eastern Cordillera (Sandeman et al., 1995) would also apply
to the Andahuaylas-Yauri region. This interpretation is consistent with the suggestion by Sandeman et al. (1995) that a
>350-km-wide arc was episodically active throughout southern Peru during late Oligocene and Miocene times. In the
Andahuaylas-Yauri region, however, volcanism seems to have
been intermittently active since the middle Eocene (Carlier
et al., 1996, 2000; Carlotto, 1998; Carlotto et al., 1999).

and possess amphibole > biotite as the dominant ferromagnesian phases, with local pyroxene in the more mafic members. They are regularly distributed throughout the region
and constitute the main mass of the batholith. Contact aureoles within country rocks are extremely irregular in shape,
size, and composition, although garnet skarn is typically
formed in calcareous rocks (e.g., Ferrobamba Formation) and
biotite and cordierite hornfels are developed where the more
pelitic facies of the Mesozoic formations are present (Carlotto, 1998).
The age of the batholith is constrained by regional stratigraphic relations and geochronologic data (Table 2; Fig. 5b).
Batholith rocks intrude mostly Mesozoic and early Cenozoic
marine and continental strata as well as the middle Eocene to
early Oligocene Anta Formation (Fig. 4). In addition, several
K-Ar ages reported by Carlier et al. (1996), Carlotto (1998),
and Perell et al. (2002), together with a number of ages obtained during the course of the present study, confirm a middle Eocene to early Oligocene age (~48-32 Ma) for the bulk
of the batholith (Fig. 5). The geochronologic data support the
inference by Bonhomme and Carlier (1990) that cumulate
rocks are older (~48-43 Ma) and that intermediate composition rocks are younger (~40-32 Ma), thereby corroborating
the concept that batholith emplacement took place in at least
two main stages. The data also suggest, however, that considerable time overlap existed between the more mafic and the
more felsic intrusions of the younger group (Fig. 5b).

The Andahuaylas-Yauri batholith


The northeastern border of the Western Cordillera in the
study area is underlain by large bodies of intrusive rocks collectively known as the Andahuaylas-Yauri batholith (Carlier et
al., 1989; Bonhomme and Carlier, 1990). It is also known locally as the Abancay (Marocco, 1975, 1978) or Apurmac
batholith (Pecho, 1981; Mendvil and Dvila, 1994). The
name Andahuaylas-Yauri batholith is used in this paper, following Bonhomme and Carlier (1990). The batholith is composed of a multitude of intrusions that crop out discontinuously for >300 km between the towns of Andahuaylas in the
northwest and Yauri in the southeast. Its width varies between ~25 km in the Tintaya area and ~130 km along the
Chalhuanca-Abancay transect (Fig. 5a).
In general terms, the batholith includes early-stage intrusions of cumulates (gabbro, troctolite, olivine gabbro, gabbrodiorite, and diorite) followed by rocks of intermediate composition (monzodiorite, quartz diorite, quartz monzodiorite,
and granodiorite) (Fig. 5c; Carlier et al., 1989; Bonhomme
and Carlier, 1990; Carlotto, 1998). Subvolcanic rocks of dominantly granodioritic/dacitic composition, locally associated
with porphyry-style mineralization, represent the terminal
stage (see below). Early-stage cumulate rocks are exposed
mainly along the northern border of the batholith (Fig. 5) between Curahuasi and Limatambo (Carlier et al., 1989; Ligarda et al., 1993), where petrologic work by Carlier et al.
(1989, 1996) determined that they constitute typical calcalkaline cumulates crystallized at the bottoms of shallow magma
chambers, with temperatures of emplacement of ~1,000C
and pressure conditions of ~2 to 3 kbars. The intrusions of the
intermediate stage are lighter gray in color, display mediumto coarse-grained, equigranular to slightly porphyritic textures,

Other intrusions
Post-batholith intrusive activity in the region is characterized by a series of small syenitic stocks that have yielded K-Ar
ages of ~28 Ma in the Curahuasi area (Carlotto, 1998). These
intrusions are part of a larger alkalic magmatic province that
also includes the basanites, phonotephrites, and trachytes of
the Ayaviri region, with ages between 29 and 26 Ma (Carlier
et al., 1996, 2000).

0361-0128/98/000/000-00 $6.00

Structural geology
The structure of the region is, in general terms, poorly constrained and understood. Although some exceptions exist
(Marocco, 1975; Pecho, 1981; Cabrera et al., 1991; Carlotto
et al., 1996b; Carlotto, 1998), regional maps lack the detailed
structural data that would help to understand the regional
tectonics as a whole. The northeastern border of the Western
Cordillera is dominated by Mesozoic to Cenozoic sequences
that have been moderately to intensely deformed in large,
northwest-trending folds with dominantly northerly vergence
(Fig. 2). Intense folding in the region typically involves carbonate and shaly sequences (Ferrobamba Formation and
equivalent units) that wrap around cores of quartz arenite of
the Yura Group. Low- and high-angle thrusts locally accompany the most intense deformation and folding, particularly in
the southern quadrangles of the region (Pecho, 1981), with
most of the mapped thrusts displaying northerly vergence.
This style has similarities to thin-skinned fold-thrust belts
elsewhere (e.g., Benavides-Cceres, 1999), as no involvement
of pre-Mesozoic basement is apparent.
The limit between the Western Cordillera and the Altiplano is characterized by two main northwest-trending fault
systems (Limatambo-Ayaviri and Abancay-Yauri) with exposed

1582

MIOCENE
EARLY

OLIGOCENE
EARLY LATE

LATE

EOCENE
MIDDLE

0361-0128/98/000/000-00 $6.00

EARLY

1583

(1)

(1)

(1)

CURAHUASI

7300

35.80.9

ABANCAY

37.91.4

50 km

7230

SANTO TOMAS

34.2 0.9

LAS BAMBAS

39.71.9

43.21.1

39.81.5

43.31.9(1)

COTABAMBAS

35.13.1(1)

MACHU PICHU

(1) K-Ar (Carlotto, 1998)

Batholith Plutons

CHALHUANCA

ANDAHUAYLAS

DIORIT E / GABBRO (CUMULATES)

DIORITE

MONZODIORITE

QUARTZ MONZODIORITE

GRANODIORITE

(1) (1)

7330

1430

1400

(1)
(1)

(1)

ANDAHUAYLAS-YAURI
BATHOLITH

13

7200

35.70.9

KATANGA

YAURI

7130

31.60.8

LIVITACA

40.31.0

15

TINTAYA

SICUANI

43.71.1(1)

POMACANCHIS

14

Monzogranite
Granodiorite
Tonalite
Quartz Monzonite
Quartz Monzodiorite
Quartz Diorite
Monzodiorite
Diorite / Gabbro

CUZCO

5
6
7
10
11
12
16
17

1430

1400

16

11

12
17

Curahuasi
Livitaca
Pomacanchis
Katanga
Tintaya
Cotabambas
Las Bambas

FIG. 5. Distribution and age of the Andahuaylas-Yauri batholith in the study area. a. Displays the main body of the batholith (Fig. 2) and the location of the K-Ar age
data from the present study (Table 2) and Carlotto (1998). b. Available K-Ar age data relative to volcanism of the Anta Formation and sedimentation of the San Jernimo Group as in column C of Figure 4. c. Composition of the main phases of the Andahuaylas-Yauri batholith on a QAP diagram (Streckeisen, 1976), based primarily
on work by the writers with additions after Pecho (1981), Carlier et al. (1989), and Carlotto (1998). Main localities studied are identified for better comprehension (see
text for descriptions).

50

40

37

34

30

28.5

24

20

Ma

PUNACANCHA
FORMATION

ANTA Fm
SAN JERONIMO Gp

STRATIGRAPHY

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

1583

1584

PERELL ET AL.
TABLE 2. K-Ar Ages of Various Intrusive Phases of the Andahuaylas-Yauri Batholith

Sample no.

Latitude

Longitude

Mineral

K (%)

Radiogenic Ar (nl/g)

Ar (at. %)

Age 2

LIVIKAR 02
PORKAR 02
KATKAR 05
COTKAR 01
COTKAR 02
PROGKAR 01
PROGKAR 02
LAHUKAR 01
CHALCOKAR 02

1418'58"
1429'34"
1426'38"
1345'03"
1341'11"
1406'02"
1400'54"
1425'08"
1403'37"

7144'10"
7156'02"
7154'30"
7221'23"
7221'14"
7228'35"
7228'28"
7300'45"
7218'21"

Biotite
Biotite
Biotite
Biotite
Amphibole1
Biotite1
Amphibole
Biotite
Amphibole

7.041
7.212
5.066
7.556
0.812
6.833
0.331
7.493
0.504

11.155
10.100
6.278
1.335
1.271
9.171
0.516
10.544
0.751

31
25
13
17
34
26
53
27
36

40.3 1.0
35.7 0.9
31.6 0.8
43.2 1.1
39.8 1.5
34.2 0.9
39.7 1.9
35.8 0.9
37.9 1.4

Constants: = 4.962 10-10y1; = 0.581 1010y1; 40Ar/36Ar = 295.5; 40K = 0.01167 at. percent
See Figure 5 for sample location
1 Some degree of alteration to chlorite present

lengths of >300 km (Fig. 2). Both are made up of several segments or smaller faults with individual continuous runs of >50
km that display high-angle reverse and strike-slip movements.
In the vicinity of the Abancay deflection (Marocco, 1978),
these structures transpose Paleozoic plutonic rocks over
younger cover sequences. Farther east, near Curahuasi, they
place deep cumulate facies of the Andahuaylas-Yauri
batholith on top of either younger intrusions of the same
batholith or over volcanic horizons of the Anta Formation
(Carlotto, 1998). Farther southeast, in the area of Santa
Luca, high-angle reverse structures belonging to the southeastern extension of the Abancay-Yauri fault are interpreted
to have been associated with a major fold-thrust deformation
event (Jaillard and Santander, 1992). The ~300-km-long, 10
to 50-km-wide corridor defined by the Limatambo-Ayaviri
and Abancay-Yauri fault systems is occupied by the synorogenic rocks of the Anta Formation and the San Jernimo
Group. The two main fault systems are inferred to have been
active during Mesozoic time and to have largely controlled
the shape and extension of the Cuzco-Puno high in the region
(Carlotto, 1998); they would therefore constitute structures
reactivated during Andean deformation (Jaillard and Santander, 1992; Benavides-Cceres, 1999).
The Altiplano is characterized by the synorogenic sequences that filled the basins of the San Jernimo Group and
the Punacancha and Paruro formations. These sequences display intense synsedimentary deformation structures including
tight folding and fault-controlled progressive unconformities
(Carlotto, 1998).
Tectono-Magmatic Synthesis
The main part of the region under consideration seems to
have been affected by several Late Cretaceous to Pliocene
tectonic events (Marocco, 1975; Pecho, 1981; Cabrera et al.,
1991; Carlotto et al., 1996b) of which the Eocene to early
Oligocene (Incaic) and Oligocene to Miocene (Quechua)
pulses are the most important. Important sedimentary, tectonic, and magmatic activity occurred in the Eocene and
Oligocene. The red beds of the San Jernimo Group were deposited in structurally controlled, northeast-trending synorogenic basins localized at the boundary between the Eastern
and Western Cordillera. Fluvial sedimentation is thought to
have progressed from south to north (Fig. 6). The presence of
several progressive unconformities in the sedimentary se0361-0128/98/000/000-00 $6.00

quences is interpreted to indicate that successive compressive


events were modifying the original pull-apart transtensional
architecture into contractional basins (Crdova, 1986; Noblet
et al., 1987; Marocco and Noblet, 1990; Chvez et al., 1996;
Carlotto, 1998). Locally, important volcanism (Anta Formation) accompanied deposition of the San Jernimo red bed
sequences.
The bulk of the deformation, interpreted to have begun at
~42 Ma (Carlotto, 1998) and thus broadly synchronous with
the Incaic orogeny of central Peru (Noble et al., 1974, 1979;
Mgard et al., 1984; Mgard, 1987; Farrar et al., 1988;
Sbrier et al., 1988; Sbrier and Soler 1991), is also thought
to have been the most important event of compressive deformation in the region. Paleogeographic reconstructions (Fig.
6) suggest that this northeast-directed deformation was responsible for the development of the basins that accommodated middle Eocene to early Oligocene volcanism and sedimentation. Reactivation of older, basin-bounding structures
(e.g., Cuzco-Puno high) into major high-angle reverse faults
that favored the uplift of the Andahuaylas-Yauri batholith,
also took place during late Eocene to early Oligocene time
(~4032 Ma; Carlier et al., 1996; Carlotto, 1998; see below).
This summary suggests that the Incaic orogeny in the study
region constitutes a long-lived period of semicontinuous deformation of ~20 to 15 m.y., between the middle Eocene and
the early Oligocene. Several distinct deformation events and
associated shortening and uplift are, however, apparent and
may have accompanied emplacement of the various phases of
the Andahuaylas-Yauri batholith in at least two main events,
at ~48 to 43 and ~40 to 32 Ma (Bonhomme and Carlier, 1990;
Carlier et al., 1996). Evidence from outside the study region
(Portugal, 1974; Jaillard and Santander, 1992; Carlotto et al.,
1996b; Carlotto, 1998) strongly suggests that magmatism,
folding, uplift, and erosion were all integral components of
the Incaic orogeny in the Western Cordillera and Altiplano, a
contrasting view to that of Clark et al. (1990) and Sandeman
et al. (1995) for a contiguous transect (1320S) farther
southeast.
Porphyry Copper Geology
Distribution
The Andahuaylas-Yauri belt extends for ~300 km and is
defined by 31 prospects and deposits with porphyry-style

1584

1585

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU


71

73
MACHU PICHU

Ri

o
A

ur

Ab
ANDAHUAYLAS

an

ca y

im

Fau

ac

Fault

ault
Limatambo F
CUZCO

lt
ABANCAY

ot

ab

URCOS

am

ba

s F
a

ult

14

14

Po

CHALHUANCA

SICUANI

ma

LIVITACA

ca

Fa

av

ir

ult

i
Fa

ul

a: CONGLOMERATES b: VOLCANIC ROCKS

MAIN PLUTONS OF
ANDAHUAYLAS-YAURI BATHOLITH

hi

ault
uri F

b ANTA FORMATION

Ay

nc

Ya

SANTO TOMAS

SAN JERONIMO GROUP


BASINS

YAURI

MESOZOIC ROCKS

50 km
PALEOZOIC BASEMENT
SEDIMENT PROVENANCE

71

73

FIG. 6. Schematic paleogeographic reconstruction of the study area during late Eocene to early Oligocene time. Note the
intimate spatial relationship between batholithic plutons, volcanic rocks of the Anta Formation, and the sedimentary basins
of the San Jernimo Group in this reconstruction. Also shown are the main fault systems and high-angle thrusts that are interpreted to have controlled both uplift of the batholith at the deformation front south of Cuzco (the Cuzco-Puno high) and
the San Jernimo basins. Main localities are shown for reference. Modified after Carlotto (1998).

alteration and mineralization, including 19 systems grouped


in 5 main clusters plus 12 separate porphyry centers (Fig. 7a).
Of these, 3 systems (Leonor, Leticia, and Aceropata; Fig. 7a)
are excluded from this review due to a lack of data. The belt
has a known maximum width of 130 km in a northeast-southwest direction, with deposits exposed at elevations between
3,400 and 4,700 m (Fig. 7b). However, beyond the limits of
the belt in the Eastern Cordillera, porphyry Cu-Mo mineralization is exposed at elevations as low as 2,800 m, as at Aurora
(Fig. 7b).
A salient feature of the belt is the spatial distribution of
porphyry copper stocks around the edges of the main intrusions that make up the Andahuaylas-Yauri batholith, as exemplified by the Katanga, Morosayhuas, and Las Bambas
clusters and by the Pea Alta, Lahuani, Alicia, Leonor, and
Winicocha systems. Only at Panchita, Portada, Leticia,
Cotabambas, and Antapaccay was the earlier plutonic phase
effectively penetrated by the later porphyry stocks. Isolated
systems, such as those of the Tintaya cluster and Trapiche,
Los Chancas, and Chaccaro, far from the main batholithic
0361-0128/98/000/000-00 $6.00

intrusions, are nevertheless related to smaller plutons and


outliers of the same batholith.
Host rocks
Country rocks for selected porphyry systems of the belt are
indicated in Figure 8 and Table 3. Wall rocks to the Las Bambas, Katanga, and Tintaya clusters include either intrusive
batholithic rocks or sedimentary rocks preserved as roof-pendants in the batholith. Within the Mesozoic units, it is apparent that lower Ferrobamba Formation horizons, near their
contact with the Mara Formation, are the preferred host
rocks (Tintaya deposits; Fierro et al., 1997; Zweng et al., 1997;
Jones et al., 2000), whereas the Chuquibambilla and Soraya
formations constitute common host rocks where the Yura
Group predominates (Fig. 8), in the southwestern part of the
belt (Los Chancas; Corrales, 2001). The Cotabambas cluster
is restricted to diorite and granodiorite of the AndahuaylasYauri batholith (Perell et al., 2002), whereas both batholith
intrusions and Anta Formation volcanic horizons make up the
host rocks of the Morosayhuas cluster farther north (Fig. 8).

1585

1586

PERELL ET AL.

71

72

73

AURORA (4)

13
Elevation
meters a.s.l.
IN
SE

5000

17
25

ON

FO

MOROSAYHUAS
CLUSTER

29

30

CT
I

4500

SE

6
3
24

4000
32

LLOCLLACSA (21)

18

19

2
23

31 14 16
26

28 13

33

11

27

22

3500

MAKI (22)

COTABAMBAS
CLUSTER

10

20

QENCO (27)

CHA-CHA (11)

12 21

8
1

15
CHILCACCASA (12)
HUACLLE (15)
CCALLA (7)

LETICIA (20)

3000

ACEROPATA (1)

AZULCCACCA (5)

CCARAYOC (8)

50 km

2500
CHACCAR O (10)

CHALCOBAMBA (9)

FERROBAMBA (14)

14
ALICIA (2)
LOS CHANCAS (18)
PEA ALTA (24)

SULFOBAMBA (29)

LAS BAMBAS
CLUSTE R

LEONOR (19)

CRISTO DE LOS
ANDES (6)

PANCHITA (25)

KATANGA
CLUSTER

LAHUANI (17)

WINICOCHA (33)

MONTE ROJO (23)

PORTADA (26)

SAN JOSE (30)

TRAPICHE (32)
KATANGA (16)

TINTAYA (31)

50 km

TINTAYA
CLUSTER

ANTAPACCAY (3)

15

COROCCOHUAYCO (13)
QUECHUA (28)

FIG. 7. Distribution of the porphyry copper deposits and prospects referred to in this study. a. Illustrates the location of
the main clusters at Morosayhuas, Katanga, Cotabambas, Las Bambas, and Tintaya, together with other separate deposits
and prospects. Numbers in parentheses are keyed to the section of Figure 7b. The Aurora prospect is also shown for reference. b. Simplified section A-A displaying the distribution of the systems relative to present-day elevation above sea level.

Geometry
Porphyry copper-bearing stocks of the Andahuaylas-Yauri
belt are centered on multiple-pulse porphyritic intrusions
(Table 3) that commonly display both dike- and cylinder-like
geometries. In plan view, the stocks generally range from
~0.25 to 0.6 km2, but also include the much smaller examples
of the Morosayhuas cluster where the stocks can be as small
as 150 50m, as at Qenqo. In general, the form of the stocks
0361-0128/98/000/000-00 $6.00

is difficult to determine, owing to (1) post-mineralization


moraine cover (Katanga), (2) late mineral dikes (Cotabambas and Tintaya; Fig. 9), (3) syn- to post-mineralization
faults (Cotabambas and Antapaccay; Fig. 9), and (4) textural
and compositional similarities between porphyry copperbearing stocks and wall rocks (Winicocha, Morosayhuas,
Chalcobamba, and Panchita). Moreover, other systems seem
to have developed complex geometries from the outset,
such as the rootless nature of the stocks at Chabuca and

1586

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

LABRA/
CHUQUIBAMBILLA
Fms

WINICOCHA
CRISTO DE LOS ANDES

PEA ALTA

COTABAMBAS

ALICIA

LAHUANI

YURA Gp

SORAYA/
HUALHUANI
Fms

LOS CHANCAS

MURCO/MARA/
HUAMBO Fms

QUECHUA

FERROBAMBA/
ARCURQUINA/
AYAVACAS
Fms

TINTAYA

ANTAPACCAY CHACCARO

NEOCOMIAN - TURONIAN

PUQUIN/
ANTA-ANTA
Fms

LAS BAMBAS

MOROSAYHUAS

ANTA/
KAYRA/
SONCCO
Fms

CHILCA/
QUILQUE
Fms

MAASTRICHTIAN

BAJOCIAN - TITHONIAN

ANDAHUAYLAS - YAURI
BATHOLIT H

KATANGA

SAN JERONIMO
Gp

PALEOCENE

EOCENE-EARL Y
OLIGOCENE

PUNACANCHA
Fm

CACHIOS/
PISTE Fms

PELITE
LAGUNILLAS
Gp

SANDSTONE

CONGLOMERATE

QUARTZITE

GYPSUM

LIMESTONE

ANDESITE

FIG. 8. Schematic diagram illustrating the tentative location of selected


porphyry systems of the belt relative to the main Mesozoic to Cenozoic stratigraphic units of the region and the Andahuaylas-Yauri batholith.

Coroccohuaycco (Fierro et al., 1997; Zweng et al., 1997) and


the bedding-parallel attitude of the multiple sills at Quechua
(J. Perell, unpub. data, 2003), in the Tintaya cluster.
Systems dominated by tabular, dike-like geometries include
the Cotabambas (Fig. 9c), Las Bambas, and Morosayhuas
clusters, and the Pea Alta and Quechua composite stocks.
Typical cylindrical host intrusions include Los Chancas (Corrales, 2001), Cristo de los Andes, Chaccaro, and Alicia (Fig.
9a), with plan dimensions of between 300 and 600 m.
Composition
Porphyry copper-bearing stocks of the belt are, like the
major intrusions of the Andahuaylas-Yauri batholith, typically
calc-alkaline in composition (Carlier et al., 1989; Bonhomme
and Carlier, 1990). Although it is difficult to assign precise
petrographic names to altered and mineralized intrusive
rocks, most of the porphyry-related intrusions studied possess
an intermediate composition, with dacite and/or granodiorite
predominanting (Fig. 10c). Exceptions include some quartz
monzodioritic stocks at Cotabambas (Perell et al., 2002; Fig.
10c), the monzogranitic composition of Panchita, the monzonitic nature of some stocks in the Katanga cluster (MMAJ,
1983), and the dominantly quartz monzonitic to monzonitic
composition of the Tintaya cluster, including Quechua
(Fierro et al., 1997; Zweng et al., 1997; Jones et al., 2000;
Fierro et al., 2002).
0361-0128/98/000/000-00 $6.00

1587

Biotite and amphibole are by far the most abundant ferromagnesian phenocrysts in all the porphyries studied, although
pyroxene is also present locally at Katanga. Proportions of biotite to amphibole vary greatly among the different systems,
with biotite being more abundant at Alicia, Los Chancas,
Pea Alta, Panchita, Trapiche, and in the Las Bambas and
Katanga clusters. Amphibole dominates at Cristo de los
Andes, Portada, Chaccaro, and in the Cotabambas and Tintaya clusters. Other important phenocryst populations are
largely dominated by plagioclase (30 to 80 vol %) and subordinate quartz eyes and orthoclase (~10 vol % each) (Fig.
10c). Groundmass mineralogy is dominated by quartz, plagioclase and orthoclase in microfelsitic aggregates, which locally
contain interstitial biotite.
In most deposits, the bulk of the mineralization seems to be
genetically associated with one single phase of intrusion, as at
Alicia, Cristo de los Andes, Pea Alta, Portada, and the Morosayhuas and Cotabambas clusters (Table 3). Two phases are
apparent at Los Chancas and Lahuani, and up to six phases
have been described at Antapaccay (Jones et al., 2000). Similarly, inter- to late-mineral porphyry intrusions constitute integral parts of all systems in the belt. These later intrusions
vary from one central phase at Alicia (Fig. 9), through two at
Ccalla (Cotabambas) and Chabuca (Tintaya), to at least three
at Antapaccay. Earliest inter-mineral porphyries exhibit similar textures, compositions, and alteration products to the
main intrusions, making distinction between them difficult.
However, they tend to possess weaker versions of the same alteration and mineralization types. Later inter-mineral and
younger phases display different compositions and textures,
and are characterized by much weaker alteration, in addition
to lacking significant hydrofracturing. Inter-mineral dikes of
roughly the same composition as that of the main stocks are
present at Cotabambas, Alicia, Chaccaro, Antapaccay,
Quechua, Katanga, Lahuani, Las Bambas, Morosayhuas, and
Tintaya, whereas younger, compositionally and texturally distinct phases occur at Tintaya, Antapaccay, Los Chancas,
Katanga, Lahuani, and Pea Alta. These younger phases,
which may or may not include postmineral intrusions, are
commonly dominated by andesitic, dacitic, and microdioritic
dikes. Postmineral mafic dikes are common at Tintaya (Fierro
et al., 1997; Zweng et al., 1997).
The size, shape and location of inter- to late-mineral intrusions, with respect to the main stock, exert a marked influence on the geometry of both the main stocks and the mineralized zones in porphyry deposits of the belt. In most cases,
inter- to late-mineral phases are dike-like in form, as at
Lahuani and Pea Alta, and in the Tintaya, Katanga, and
Cotabambas clusters, but are cylindrical in shape at Alicia
and irregular at Morosayhuas and Chaccaro. In some deposits, as at Ccalla at Cotabambas, the earliest inter-mineral
phases are centrally located with respect to the main-phase
stock, whereas later-mineral intrusions occupy peripheral
positions in association with late-stage dome and dike
swarms of dacitic composition (Fig. 9c). At Alicia, however,
the cylinder-like, inter-mineral intrusion occupies a central
position (Fig. 9a) and at Tintaya (e.g., the Chabuca deposit)
the dikes cut the early-stage porphyry stock and related mineralization almost at a 90-degree angle and extend far beyond mineralized zones (Zweng et al., 1997). Modification of

1587

0361-0128/98/000/000-00 $6.00

Six phases:
monzonite
to quartz
monzonite
One phase?:
quartz
monzonite
One phase:
granodiorite
to quartz
monzodiorite
Several phases:
dacite
Several phases:
dacite and/or
granodiorite

Antappacay
(Tintaya)

1588

One phase:
dacite

Winicocha

Cristo de los
Andes
Pea Alta

Lahuani

Chaccaro

Alicia

Two phases:
granodiorite
to quartz
monzonite
One phase:
granodiorite
and/or dacite
One phase:
dacite
One phase:
rhyodacite to
dacite
One phase:
dacite
One phase:
dacite/
rhyodacite

Los Chancas

Qenco/Maki
One phase:
(Morosayhuas)
diorite and
quartz diorite

San Jos
(Katanga)

Katanga
(Katanga)

Ferrobamba
(Las Bambas)
Chalcobamba
(Las Bambas)

Ccalla
(Cotabambas)

Several phases:
monzonite
and dacite
Two main
phases: dacite

Two phases:
monzonite

Chabuca
(Tintaya)

Quechua
(Tintaya)

Mineralized
intrusion(s)

Deposit or
prospect
(cluster)

N.A.

Several phases: andesite


and microdiorite dikes
Several phases:
dacite and
andesite dikes
Several phases:
andesite dikes
Several phases:
andesite dikes

One central
phase: dacite

N.A.

Several phases: dacite


and andesite dikes
Several phases:
dacite and
andesite dikes;
granodiorite stock
Several phases:
dacite and
andesite dikes
Two phases:
dacite and
andesite dikes
Several phases:
andesite dikes

Several phases:
andesite and quartz
monzonite dikes
Two phases:
granodiorite and
dacite dikes

Several phases:
diorite, dacite
an mafic dikes
Three phases:
andesite dikes

Mid- and late


mineral intrusions

Microdiorite and
dacite stocks

Upper Soraya Fm

Soraya Fm

Upper
Ferrobamba Fm
Chuquibambilla
Fm

Upper
Ferrobamba Fm

Chuquibambilla
and Soraya Fms

Lower Anta Fm;


diorite pluton

Undifferentiated
Ferrobamba Fm

Undifferentiated
Ferrobamba Fm

Lower
Ferrobamba Fm
Lower
Ferrobamba Fm

Lower Ferrobamba,
Mara, and upper
Soraya Fms
Diorite and
granodiorite
plutons

Lower Ferrobamba
Fm; diorite
pluton

Lower
Ferrobamba Fm

Wall rocks

Early potassic overprinted


by quartz-sericitic
Early potassic and calcicpotassic overprinted by
sericite-clay-chlorite and
peripheral quartz-sericitic
Early potassic? overprinted
by quartz-sericitic and
local advanced argillic

Early potassic overprinted


by sericite-clay-chlorite
Early potassic overprinted
by quartz sericitic

Early potassic overprinted by


sericite-clay-chlorite

Early potassic overprinted


by intense quartz-sericitic
and local advanced argillic
Early potassic and calcicpotassic overprinted by
sericite-clay-chlorite and
local quartz-sericitic and
advanced argillic
Early potassic with
peripheral quartz-sericitic

Early potassic

Early potassic and calcicpotassic overprinted by


sericite-clay-chlorite and
local quartz-sericitic
Early potassic overprinted
by sericite-clay-chlorite
Early potassic overprinted
by sericite-clay-chlorite

Early potassic overprinted


by quartz-sericitic

Early potassic overprinted


by albite-rich sericiteclay-chlorite
Early potassic overprinted
by sericite-clay-chlorite

Ore-related
hydrothermal alteration

Trace
chalcopyrite

Chalcopyrite

Chalcopyrite

Chalcopyrite,
molybdenite

Chalcopyrite

Chalcopyrite ~
bornite

Chalcopyrite >
bornite

Trace
chalcopyrite

Chalcopyrite

Chalcopyrite

Chalcopyrite >
bornite
Chalcopyrite ~
bornite

Chalcopyrite >
bornite

Chalcopyrite

Bornite ~
chalcopyrite

Chalcopyrite >
bornite

Ore
mineralogy

Absent

Copper oxides and


chalcocite blanket
Minor copper
oxides

Minor copper
oxides
Minor copper
oxides

Minor copper
oxides

Copper oxides
and chalcocite
blanket

Copper oxides
and chalcocite
blanket
Absent

Abundant
copper oxides

Irregular
copper oxides
Irregular
copper oxides

Irregular copper
oxides and
chalcocite blanket
Irregular copper
oxides and
chalcocite

N.A.

Irregular copper
oxides

Supergene
mineralization

TABLE 3. Geological Features of Selected Porphyry Systems of the Andahuaylas-Yauri Belt

Inter- to latemineral;
sericite-rich

Inter-mineral;
sericite-rich
N.A.

Inter-mineral;
magnetite-rich
N.A.

N.A.

N.A.

Large intermineral;
sericite-rich
Local
tourmalinerich dikes

N.A.

Contact
breccias
N.A.

Contact
breccias and
pebble dikes

Pebble dikes

Large postmineral
diatreme

Pebble dikes

Hydrothermal
breccias

Distal skarn at
12 km

Minor exoskarn;
distal jasperoids
at ~2 km
Distal skarn at
23 km
Distal skarn at
1.5 km

Minor exoskarn

Dominant:
exoskarn

N.A.

Distal skarn at
12 km

Present:
exo>endoskarn

Dominant:
exo>endoskarn

Important:
exo>endoskarn
Dominant:
exo>endoskarn

Distal skarn at
23 km

Locally important
exoskarn

Minor exoskarn

Dominant:
exo>endoskarn

Skarn
mineralization

1588
PERELL ET AL.

1589

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

ALICIA

ANTAPACCAY
7120
ANTAPACCAY NORTH

7159

1405

1500

Late-mineral Diatreme
Late Porphyry

200 m

Late Porphyry

Main Porphyry

Main Porphyry

Pre-mineral Diorite

Skarn

Skarn

Reverse Fault

Zone of Intense
Stockwork Veining

Anticline

ANTAPACCAY
SOUTH

COTABAMBAS

500 m

SAN JOSE

7222

Cerro Saiwa

4400m
HUACLLE

4300m

4200m

4100m

1344

CCALLA

200 m

CCARAYOC

AZULCCACCA

Late Dome and Dikes

1 km

a
b

a
b

Porphyry-related
Intrusions
Andahuaylas-Yauri Batholith
a: Diorite
b: Granodiorite
Skarn

Leached Capping

Late Andesite Dike

Supergene Chalcocite Blanket


a: Intersected by Drilling
b: Projected

Hydrothermal Breccia
Main Porphyry
Drill Hole

FIG. 9. Main geologic attributes of selected porphyry copper systems of the Andahuaylas-Yauri belt. a. Displays the cylindrical form of the porphyry copper-bearing stock at Alicia and the central location of the late-mineral porphyry dike, as
mapped by the writers. b. Illustrates the structurally controlled nature of the porphyry copper systems at Antapaccay and the
large postmineral diatreme breccia (simplified after Jones et al., 2000 and Fierro et al., 2002). c. Displays the cluster at
Cotabambas and the structurally controlled nature of the stocks at Ccalla and Azulccacca, together with the peripheral, latemineral dome and its dike swarm (simplified after Perell et al., 2002). d. Schematic cross section through the San Jos porphyry system at Katanga displaying the distribution of the main geologic units and the location of the supergene enrichment
zone. Based on data provided by the Metal and Mining Agency of Japan (1983) and mapping by the writers.
0361-0128/98/000/000-00 $6.00

1589

1590

(1)
(1)

(1)

(1)

1430

1400

(1)

LOS CHANCAS
PANCHITA

WNW

LAHUANI

TRAPICH E

7330

200 km

CHALHUANCA

50 km

CRISTO DE
LOS ANDES

SANTO TOMAS

SULFOBAMBA

KATANGA

ALICIA

7200

PORTADA

CHACCARO

CHALCOBAMBA
FERROBAMBA

LAS BAMBAS
CLUSTER

7230

CHILCACCASA

CUZCO

5
10

20

35

7130

ANTAPACCAY

KATANGA
CLUSTER

SAN JOSE

90

90

65

11

10
2
3
8
5 1

DACITE

60

90

QUECHUA

COROCCOHUAYCO

TINTAYA
CLUSTER

SICUANI

TINTAYA

YAURI

RHYOLITE

WINICOCHA

LIVITACA

MONTE ROJO

60

ALICIA
CRISTO DE LOS ANDES
CHACCARO
KATANGA CLUSTER
PORTADA
PEA ALTA
PANCHITA
WINICOCHA
LAS BAMBAS CLUSTER
TRAPICHE
COTABAMBAS CLUSTER

1330

1
2
3
4
5
6
7
8
9
10
11

COTABAMBAS
CLUSTER

AZULCCACCA

CCALLA

CCARAYOC

HUACLLE

MAKI

QENCO

LLOCLLACSA

MOROSAYHUAS
CLUSTER
CHA-CHA

MACHU PICHU

LETICI A

ACEROPATA

SSE

Skarn Fe-(Cu, Au)

TRAPICHE

LEONOR

PEA ALTA

7300

LAHUANI

PANCHITA

ABANCAY

7300

PORPHYR Y COPPER
ALTERATION-MINERALIZ ATION

CRISTO DE ANDES
SULFOBAMBA
CHALCOBAMBA
FERROBAMBA

LOS CHANCAS

7330
ANDAHUAYLAS

ANDAHUAYLAS-YAURI
BATHOLITH

(1) (1)

(1)

QUECHUA

20

FIG. 10. a. Distribution of porphyry copper clusters and systems relative to the Andahuaylas-Yauri batholith and Fe-Cu skarn occurrences. Note the preferred location of the porphyry clusters along the edges of main batholithic bodies. b. Age distribution of selected porphyry copper deposits and prospects of the belt (Table 4) relative to the Andahuaylas-Yauri batholith and the volcanic and sedimentary stratigraphy of the region. c. Dominant composition of selected porphyry copper-bearing
stocks of the belt according to their modal mineral contents on a QAP diagram (Streckeisen, 1978).

STRATIGRAPHY

50

40

37

34

30

28.5

PUNACANCHA
FORMATION

ANTA Fm
SAN JERONIMO Gp

24
ALICIA
CCALLA
CHILCACCASA
CHACCARO

20
SAN JOSE
MONTE ROJO
WINICOCHA

OTHER INTRUSION
INTER-LATE PHASE PORPHYR Y
AND/OR ALTERATION
MAIN PHASE ALTERATION
(2) Re-Os (Mathur et al., 2001)
(3) K-Ar (Noble et al., 1984)

KATANGA

GRANODIORITE
QUARTZ MONZODIORITE
MONZODIORITE
DIORITE
DIORIT E / GABBRO (CUMULATES)
(1) K-AR (Carlotto, 1998)

PORTADA

MIOCENE
EARLY

OLIGOCENE
EARLY LATE

LATE

EOCENE
MIDDLE

EARLY

0361-0128/98/000/000-00 $6.00
PEA ALTA

TINTAYA (3)
TINTAYA (2)

Ma

1590
PERELL ET AL.

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

ore zones at Cotabambas, Tintaya, and Las Bambas by intrusion of the inter- to late-mineral bodies is appreciable.
Hydrothermal alteration and mineralization
Six distinct types of alteration-mineralization are recognizable in porphyry systems of the Andahuaylas-Yauri belt.
These make up the potassic, propylitic, sericitic (phyllic), advanced argillic, and calc-silicate types of Meyer and Hemley
(1967) and subsequent investigators (Lowell and Guilbert,
1970; Guilbert and Lowell, 1974), as well as an alteration type
characterized by sericite, chlorite, and clays (illite-smectite).
The latter was originally termed SCC-type by Sillitoe and
Gappe (1984) in the Philippines porphyry copper deposits
and is now referred to as intermediate argillic alteration by
Sillitoe (2000). An additional, less widespread alteration-mineralization type includes the mixed calcic-potassic assemblages observed at several deposits and prospects.
Potassic alteration: With a few exceptions (Morosayhuas,
Winicocha), potassic alteration is the principal alteration type
directly associated with mineralization in Andahuaylas-Yauri
porphyry systems (Table 3). In all cases, potassic alteration occurs early in the evolution of each system and consists of
quartz, biotite, and K-feldspar. Hydrothermal biotite replaces
ferromagnesian components, typically magmatic hornblende
and, less commonly, magmatic biotite. It also occurs in the
groundmass of porphyry stocks and in veinlets, either alone or
accompanied by other silicate phases. Early, typically barren
biotite seams and veinlets occcur at several systems, including
Pea Alta and Cotabambas, and can be compared with the
early biotite veins described by Gustafson and Quiroga (1995)
at El Salvador, Chile. In most deposits and prospects, including the Cotabambas, Tintaya, Las Bambas, and Katanga clusters, and at Lahuani, Alicia, and Los Chancas, biotite is accompanied by K-feldspar, which at Cotabambas and Tintaya
constitutes a volumetrically significant alteration mineral. For
example, the most intense potassic alteration at Ccalla is dominated by aggregates of quartz and K-feldspar, with local development of graphic textures and complete destruction of
original rock textures. K-feldspar also occurs in a variety of
veinlet types with quartz and biotite, within the veinlets or as
alteration halos, and as partial replacements of original plagioclase sites. Calcite, apatite, anhydrite, and magnetite are
additional minerals in potassic alteration assemblages and are
also common constituents of veinlet assemblages. Conspicuous magnetite accompanies potassic alteration in gold-rich
porphyry systems of the belt, and at Cotabambas attains >5
vol percent (Perell et al., 2002).
Major quantities of quartz were introduced as either uni- or
multidirectional veinlets during potassic alteration in all deposits and prospects, but most characteristically at Cotabambas, Antappaccay, San Jos, Ferrobamba, Chalcobamba, Maki,
Llocllasca, and Winicocha. In at least five systems, including
Maki at Morosayhuas, San Jos at Katanga, Ccalla and Azulccacca at Cotabambas, and Winicocha, the quartz veinlets coalesce to form massive bodies of nearly pure quartz. Where
overprinting by either quartz-sericitic or sericite-clay-chlorite
alteration is intense, these massive bodies are the only remnants of the early-stage potassic alteration. In common with
porphyry systems elsewhere (e.g., Gustafson and Hunt, 1975;
Gustafson and Quiroga, 1995; Sillitoe, 2000), a variety of
0361-0128/98/000/000-00 $6.00

1591

quartz veinlets, introduced in several generations, characterize potassic alteration in porphyry deposits and prospects of
the Andahuaylas-Yauri belt. Typical assemblages and textures
compare closely with the A- and B-type veinlets described by
Gustafson and Hunt (1975) from El Salvador porphyry copper deposit, Chile. A-type veinlets carry significant mineralization in the form of chalcopyrite and/or bornite at a number
of deposits, including the Ccalla and Azulccacca centers at
Cotabambas, Pea Alta, Ferrobamba, and Chalcobamba at
Las Bambas, and Antapaccay at Tintaya. At Alicia and Chalcobamba, B-type veinlets are characterized by semicontinous
centerlines filled by millimeter- to centimeter-sized grains of
bornite and chalcopyrite, whereas at Pea Alta, Lahuani, Los
Chancas, and Quechua, they are dominated by chalcopyrite
and molybdenite. A-veinlets also occur in copper-poor, goldbearing porphyry systems, such as those from the Morosayhuas cluster (see below), where they contribute minor
amounts of chalcopyrite. Gold-rich porphyry copper deposits
of the belt, such as the Ccalla and Azulccacca centers at
Cotabambas, contain appreciable amounts of magnetitebearing veinlets that are similar to the M-type veinlets of
Clark and Arancibia (1995) and to the A- and C-type veinlets
described by Cox (1985) at Tanam, Puerto Rico.
Calcic-potassic alteration: Calcic-potassic alteration is represented at Cotabambas, Morosayhuas, and Pea Alta (Table
3). The assemblage is characterized by veinlets of quartz, actinolite, and hornblende, with important K-feldspar, biotite,
apatite and calcite, and volumetrically minor amounts of
clinopyroxene and epidote. In general, plagioclase is variably
altered to K-feldspar, calcite, and/or epidote, whereas magmatic biotite and amphibole are selectively replaced by needles of actinolite, commonly intergrown with apatite. Magmatic pyroxene is altered to aggregates of actinolite-apatite
and actinolite-biotite, and magmatic hornblende is converted
to mixtures of clinopyroxene, biotite, and hornblende as in
several systems of the Morosayhuas cluster. Alteration halos
to various veinlet sets include K-feldspar, actinolite, biotite,
and chlorite. Magnetite is a common constituent, and chalcopyrite, as part of this association, gives rise to ore-grade CuAu mineralization.
Sericite-clay-chlorite alteration: Several deposits and
prospects of the belt, including Alicia, Chaccaro, Pea Alta,
Las Bambas, Morosayhuas, Cotabambas, and Tintaya, possess
significant sericite-clay-chlorite alteration as part of their ore
zones (Table 3). This assemblage imparts a pale-green overprint to potassic alteration and gives a soft aspect to the rock
(cf. Sillitoe and Gappe, 1984). It generally modifies, but with
some degree of preservation, original rock textures. Sericiteclay-chlorite alteration varies in both intensity and mineralogy, although assemblages defined for systems of the belt
always include one or more associations of sericite (finegrained muscovite), illite, smectite, chlorite, calcite, quartz,
and varied proportions of epidote, halloysite, and albite. Plagioclase (both phenocrysts and groundmass) is replaced by a
pale-green, greasy sericite assemblage which also includes illite and, locally, smectite. Amphibole and biotite, the latter of
magmatic and/or hydrothermal origin, are characteristically
replaced by chlorite. Calcite is common as a replacement of
plagioclase, and in some deposits and prospects, including
Cotabambas and Chaccaro, it is a major constituent of the

1591

1592

PERELL ET AL.

assemblage. Albite is locally important as a replacement of


plagioclase, as at Morosayhuas and at the Chabuca Este deposit at Tintaya (Zweng et al., 1997).
Quartz veinlets include various associations with chlorite
and calcite, and halos of green sericite, halloysite, and mixedlayer illite-smectite are common (Cotabambas, Chaccaro).
Where quartz veining is intense and halos coalesce, the rock
is completely replaced by fine- to very fine-grained mosaics of
quartz, sericite, and mixed-layer illite-smectite that obliterate
original host rock textures, as at Ccalla, San Jos, Maki, Llocllacsa, and Winicocha. Chalcopyrite is locally present in
some veinlet assemblages and may constitute monomineralic
veinlets with chlorite and quartz, but in general, chalcopyrite
contents are lower than in earlier potassic alteration-mineralization. Bornite, if present in earlier assemblages, rarely survives sericite-clay chlorite alteration. Pyrite is typically present in the form of veinlets and disseminations (Chaccaro,
Maki), and is volumetrically important at Cotabambas. Magnetite is variably transformed to martite, and specular
hematite is a characteristic constituent of the assemblage.
Quartz-sericitic alteration: Well-defined zones of quartzsericitic alteration accompany ore in several systems of the
belt at San Jos, Cristo de Los Andes, Quechua and, possibly,
at Winicocha and Chilcaccasa. Moderate amounts are also
present at Cotabambas (Ccalla), Los Chancas, Chaccaro,
Lahuani, Pea Alta, and Morosayhuas (Table 3). The quartzsericitic assemblages typically comprise white, texturally destructive aggregates of quartz, sericite (fine-grained muscovite), and illite, accompanied by several percent pyrite. In
all systems mentioned above, quartz-sericitic alteration typically overprints earlier-formed potassic or, as at Cotabambas,
sericite-clay-chlorite alteration. Broad quartz-sericitic alteration halos around potassic cores, common in many porphyry
Cu-Mo deposits worldwide (e.g., Lowell and Guilbert, 1970),
are not widely developed in systems of the Andahuaylas-Yauri
belt, although they are inferred at Los Chancas (Corrales,
2001). At the Ccalla deposit in Cotabambas, structurally controlled patches of quartz-sericitic alteration abut intermediate
argillic assemblages in the upper parts of the system (Perell
et al., 2002) and contributed to the formation of an irregular
chalcocite blanket. A similar situation is also observed at
Cristo de los Andes, Quechua and at the San Jos deposit in
the Katanga cluster (MMAJ, 1983).
D-type veinlets (Gustafson and Hunt, 1975) are typically
associated with overprinting quartz-sericitic alteration in
most systems of the belt where this style of alteration occurs.
D veinlets fill planar, continuous, centimeter-wide structures
with pyrite and quartz, which develop quartz-sericitic halos.
Tourmaline, a common constituent of sericitic alteration in
many parts of the world, is rarely developed in AndahuaylasYauri porphyry systems, occurring only at Trapiche and Morosayhuas.
Advanced argillic alteration: Hypogene advanced argillic
alteration is not recognized as a common assemblage in porphyry systems of the Andahuaylas-Yauri belt. However,
sericite-rich alteration that also contains pyrophyllite and
kaolinite-group minerals is present at San Jos, Winicocha,
and Maki (Table 3). At San Jos, advanced argillic alteration
is intimately associated with transgressive structures, whereas
at Maki it is controlled by permeability contrasts at the
0361-0128/98/000/000-00 $6.00

contact between quartz diorite and volcanosedimentary country rocks. At Maki and San Jos, advanced argillic alteration is
superimposed on the porphyry stocks and associated potassic
and sericite-clay-chlorite alteration, whereas at Winicocha it
is developed at higher elevations and constitutes the roots of
a porphyry copper lithocap.
Propylitic alteration: Propylitic alteration in AndahuaylasYauri belt porphyry systems (chlorite, epidote, and calcite) is
found mainly as part of the outer halo confined to noncarbonate wall rocks. In other systems, as at Cotabambas, Chaccaro, Lahuani, and Las Bambas, propylitic alteration occurs
within porphyry copper ore zones in late-mineral stocks and
dikes. In both cases, disseminated and veinlet pyrite, in
amounts of ~1 vol percent, is common.
Calc-silicate alteration: Calc-silicate alteration is represented in many deposits and prospects of the belt. Indeed, associated mineralization has constituted the main source of
Cu-Au ore at the Tintaya (Terrones, 1958; Santa Cruz et al.
1979; Noble et al., 1984; Zweng et al., 1997) and Katanga
(MMAJ, 1983) mines, and it is an important contributor to
mineralization at the Las Bambas skarn-porphyry cluster and
the Quechua deposit (E. Tejada, pers. commun., 2003). In
addition, proximal calc-silicate assemblages and associated
skarn-type mineralization are present in most systems of the
belt, excluding Cotabambas, Cristo de los Andes, Pea Alta,
and Morosayhuas. However, at distances of ~3 km, all of the
deposits have distal skarn-type assemblages in roof-pendants
of Ferrobamba Formation and equivalent units.
Garnet, diopside, epidote, and actinolite are the characteristic calc-silicate assemblages (Terrones, 1958; Santa Cruz et
al., 1979). At Tintaya (Fierro et al., 1997; Zweng et al., 1997);
calc-silicate alteration and mineralization occur in endoskarn
and exoskarn facies, and as products of prograde (anhydrous)
and retrograde (hydrous) events (Table 3). The bulk of the Cu
(-Au, -Mo) mineralization at Tintaya and Las Bambas was introduced during prograde events, typically as chalcopyrite
and, less commonly, bornite, whereas at the smaller Alicia system, bornite, with or without chalcopyrite, is the dominant
Cu and Au contributor. Distal skarn mineralization in porphyry-centered systems of the belt is similar to that elsewhere
(Einaudi et al., 1981), in that it is richer in Pb and Zn (e.g.,
Morosayhuas). Another expression of the distal environment
is the structurally and lithologically controlled, yellow-brown
jasperoid developed in limestone beyond the skarn front at
Tintaya, Las Bambas, Katanga, and Lahuani, which at Tintaya
is reported to contain up to 1 ppm Au (Zweng et al., 1997).
The presence of jasperoid in several deposits and prospects is
evidence that they constitute integral parts of porphyry-centered systems in the region. Moreover, at Lahuani (Table 3),
distal replacement of calcareous shale by structurally controlled, As-anomalous jasperoidal mantos resembles the Carlin-style gold environment described around some porphyry
centers (Sillitoe and Bonham, 1990).
Hydrothermal breccias
Hydrothermal breccias are poorly documented in Andahuaylas-Yauri porphyry systems (Table 3). During this
study, they were identified at most deposits and prospects, an
observation supported by descriptions of Antapaccay (Jones
et al., 2000; Fierro et al., 2002) and Tintaya (Fierro et al.,

1592

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

1997; Zweng et al., 1997). Most of the breccias are volumetrically small, with that at Antapaccay probably constituting
the largest single mass in any porphyry system in the belt
(Fig. 9). As at Antapaccay, all the observed breccias postdate
main-stage mineralization, although contact (igneous) breccias associated with the emplacement of early and intermineral porphyry stocks are clearly intermineral in timing. This is
particularly evident where the breccias are cut by mineralized
veinlets, as at Cotabambas and San Jos. Most mapped hydrothermal breccias are either dike-like in form or occur as
narrow zones at intrusive contacts. Dikelike breccias possess
strong structural control and conform to pebble dikes as is
common in porphyry systems worldwide. They typically consist of centimeter-sized, subrounded lithic clasts supported by
volumetrically important matrices of finely comminuted (rock
flour) material. Illite, chlorite, and fine-grained (dusty) pyrite
are typical constituents of the matrices. Larger expressions of
a similar style of brecciation, as at Winicocha, include
rounded to subrounded, exfoliated clasts, tens of centimeters
in size, in a sericitic matrix.
Ore zone geometry
Most Andahuaylas-Yauri porphyry deposits and prospects
possess mineralization that is variably hosted by porphyry
stocks and their immediate country rocks. The following examples document the variety observed in the belt.
1. Mineralized skarns are present in country rocks where
porphyry stocks intrude carbonate rocks of the Ferrobamba
Formation and equivalent units, as at Tintaya, Alicia, and
Chalcobamba. Significant mineralization, however, is also
hosted by both porphyry stocks and wall rocks at Ferrobamba and San Jos, despite the fact that country rocks
there are dominated by carbonate horizons of the Ferrobamba Formation.
2. Porphyry stocks constitute the main host to ore where
country rocks are dominated by the terrigenous facies of the
Yura Group and equivalent units, as at Lahuani, Cristo de los
Andes, Los Chancas and Quechua, or by volcaniclastic and
red bed horizons of the Anta Formation, as in the Morosayhuas cluster.
3. Ore seems to be evenly distributed between porphyry
stocks and wall rocks where intrusions of the AndahuaylasYauri batholith constitute the dominant country rock, as at
Cotabambas.
The Tintaya cluster further exemplifies the diversity of
mineralization styles and ore hosts, including (1) skarns at the
various Chabuca deposits and Coroccohuayco, and associated
with low-grade porphyry-style mineralization, (2) porphyry
stocks, with minor skarn mineralization at Quechua, and (3)
porphyry stocks and dioritic country rocks with small amounts
of skarn at Antappaccay (Jones et al., 2000; Fierro et al.,
2002).
Metal contents
Porphyry copper deposits and prospects of the Andahuaylas-Yauri belt range from gold-rich, molybdenum-poor
examples (Cotabambas), through deposits carrying both
gold and molybdenum (Tintaya, Los Chancas), to relatively
0361-0128/98/000/000-00 $6.00

1593

molybdenum-rich, gold-poor end-members (Lahuani). Goldonly porphyry systems, although poorly explored, are present
in at least two areas at Morosayhuas and Winicocha.
The Ccalla and Azulccacca centers at Cotabambas are the
best examples of the gold-rich category of porphyry copper
deposits (Perell et al., 2002) of the belt, with an average Au
grade >0.3 ppm and appreciable volumes averaging >0.4
ppm. Molybdenum contents are low, <100 ppm. Other goldbearing porphyry copper systems of the belt include Los
Chancas (Corrales, 2001), Antappaccay (Jones et al., 2000;
Fierro et al., 2002), the Chabuca deposits at Tintaya (Zweng
et al., 1997), and although less well defined, the Ferrobamba
and Chalcobamba systems at Las Bambas. All of these deposits possess gold grades in the 0.1 to 0.3 ppm range (Table
1) and, because smaller, but higher grade volumes are present
(e.g., Antappaccay; Fierro et al., 2002), they can be considered as members of the Cu-Au clan of porphyry deposits (Sillitoe, 2000). Los Chancas and Tintaya are reported to contain
appreciable molybdenum grades (Table 1) and would therefore constitute members of the Cu-Au-Mo category of Cox
and Singer (1986). Another Mo-bearing center in the belt is
Lahuani, where quartz eye-bearing rhyodacitic intrusions
hosting heavily veined zones with development of brain
rock-type texture (i.e., unidirectional solidification texture)
are distinctive.
Although still poorly known, gold-only (e.g., Vila and Sillitoe, 1991; Sillitoe, 2000) porphyry systems are interpreted to
be present in the Morosayhuas cluster and at the isolated
Winicocha system. Where best studied, at Morosayhuas, the
porphyry centers possess most of the features that characterize the porphyry gold mineralization of the Maricunga belt,
Chile (Vila and Sillitoe, 1991; Muntean and Einaudi, 2000)
including (1) subvolcanic quartz diorite to dacite porphyry
stocks, (2) unidirectional, sheeted veinlets dominated by dark
gray, banded, gold-bearing quartz-magnetite associations, (3)
moderate to intense, albite-rich, sericite-clay-chlorite alteration, (4) quartz-magnetite-biotite veinlets of A type, (5) Au
values typically averaging between 0.3 and 1 ppm, (6) Cu values of up to several hundred ppm, and (7) low Mo contents
(<10 ppm). At the Maki and Qenco centers at Morosayhuas,
the presence of both A-type veinlets rich in magnetite and biotite and later, banded quartz-magnetite veinlets typically
cutting the former, are features that compare closely with the
porphyry gold mineralization at Refugio (Muntean and Einaudi, 2000).
Age of the Andahuaylas-Yauri Belt
A reconnaissance K-Ar study has been conducted on 18
systems in the belt (Table 4; Fig. 10). Additional geochronologic data available from the literature include the K-Ar ages
reported by Noble et al. (1984) for Tintaya and Chalcobamba,
by Yoshikawa et al. (1976) for Quechua, and the new Re-Os
ages of Mathur et al. (2001) for Tintaya (Fig. 10b). K-Ar ages
were mainly determined for hydrothermal alteration silicates,
dominantly biotite, associated with main stage potassic alteration and mineralization as at Panchita, Pea Alta, Cristo de
los Andes, Chalcobamba, Ferrobamba, Alicia, Ccalla, Portada, Monte Rojo, Chaccaro, and Los Chancas. Where such
alteration proved unsuitable for K-Ar dating, mainly due to
intense overprinting by chlorite or quartz-sericitic alteration,

1593

1594

PERELL ET AL.
TABLE 4. K-Ar Ages of Alteration Minerals from Selected Porphyry Systems of the Andahuaylas-Yauri Belt

Deposit or prospect
(cluster)
Chilcaccasa
(Morosayhuas)
Ccalla
(Cotabambas)
Monte Rojo
(Katanga)
San Jos
(Katanga)
Katanga Pit
(Katanga)
Ferrobamba
(Las Bambas)
Chalcobamba
(Las Bambas)
Sulfobamba
(Las Bambas)
Chaccaro
Los Chancas
Alicia
Portada
Winicocha
Lahuani
Trapiche
Pea Alta
Panchita
Cristo de los Andes

Mineral

K (%)

Radiogenic Ar (nl/g)

Ar (at. %)

Age 2

Sericite2

7.537

10.410

35.2 0.9

Secondary biotite

7.393

10.375

10

35.7 0.9

Secondary biotite1

6.373

7.916

17

31.7 0.8

Amphibole2

0.753

0.982

68

33.2 1.9

Amphibole2

1.068

1.231

31

29.4 1.0

Secondary biotite1

6.465

9.271

18

36.5 1.0

Secondary biotite1

6.434

9.002

13

35.6 0.9

Magmatic biotite1
Amphibole2
Secondary biotite2
Secondary biotite
Secondary biotite1
Whole rock (sericite2)
Secondary biotite
Secondary biotite
Secondary biotite
Secondary biotite
Biotite

5.176
0.743
7.664
7.162
6.742
5.055
7.442
7.374
7.557
7.328
6.737

7.146
0.985
9.608
10.392
9.515
5.692
10.482
8.747
11.743
10.367
9.810

7
38
23
11
21
29
19
27
21
30
25

35.2 0.9
33.8 1.2
32.0 0.8
36.9 0.9
35.9 0.9
28.7 0.8
35.9 0.9
30.3 0.8
39.5 1.1
36.0 1.0
37.1 1.0

Constants: = 4.962 10-10y-1; = 0.581 10-10y-1; 40Ar/36Ar = 295.5; 40K = 0.01167at. %


See Figures 7 and 10c for sample location and plots
1
Some degree of alteration to chlorite present
2
Mid- to late-mineral porphyry phase or hydrothermal alteration event

dating was conducted on inter- to late-mineral dikes or hydrothermal alteration events. Amphibole was dated from
inter- to late-mineral porphyry stocks and dikes at Trapiche,
Katanga, and San Jos, whereas sericitic alteration was dated
at Chilcaccasa and Winicocha. All K-Ar dates are considered
here as minimum ages because (1) the K-Ar method records
cooling rather than crystallization of the dated silicates (e.g.,
biotite) and (2) the K-Ar system offers no way of experimentally testing whether later Ar loss occurred to produce artificially young ages (J. Dilles, pers. commun., 2003).
Overall, the data confirm the presence of a widespread, late
Eocene to earliest late Oligocene porphyry copper event in
the belt (Noble et al., 1984), with ages ranging between approximately 39.5 1.1 Ma at Pea Alta and 28.7 0.8 Ma at
Winicocha (Table 4; Fig. 10b). This age range can be expanded further into the middle Eocene when the Re-Os age
of 41.9 0.2 Ma for Tintaya (Mathur et al., 2001) is considered. If, following Table 4, only those ages from unaltered
secondary biotite are taken into account, it can be inferred
that much of the main-stage potassic alteration in the belt
formed between approximately 42 and 35 Ma, i.e., during the
middle to late Eocene. The K-Ar age of 35.2 0.9 Ma (Table
4; Fig 10b) for magmatic biotite from a skarn-related intrusion at Sulfobamba further confirms the age range of the belt
for both porphyry- and nonporphyry-related mineralization.
No sub-belts or age trends are apparent from Figures 7
and 10, and porphyry-style alteration and mineralization are
0361-0128/98/000/000-00 $6.00

inferred to have occurred almost simultaneously along a


~130-km-wide, 300-km-long belt. The Katanga cluster and
nearby Winicocha system, however, consistently return the
youngest ages in the belt and seem to have been active at the
end of the metallogenic episode responsible for the Andahuaylas-Yauri belt. Although both biotite and amphibole
ages from this cluster are modified by a chloritic overprint
that presumably caused Ar loss, the plutons of the area return the youngest ages of the entire Andahuaylas-Yauri
batholith. Thus, on the basis of available K-Ar data, it is suggested that a distinct, localized plutonic and porphyry copper
event took place in the Katanga area during the early
Oligocene, with minimum ages between approximately 35
and 30 Ma.
Supergene Effects
The depth of partial to complete oxidation of sulfides in
porphyry deposits and prospects of the belt is commonly 30
to 50 m but, locally, extends to 150 m. As expected from the
rugged topography of much of the region, oxidation tends to
be thicker under ridge crests and nearly absent beneath valley floors. Most of the porphyry systems lack economically
significant zones of supergene enrichment, because of the
relatively low pyrite contents, the poorly developed nature of
quartz-sericitic alteration, and the high neutralization capacities of both potassic alteration zones and carbonate country
rocks. Consequently, most cappings are immature, typically

1594

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

goethitic in composition, with some containing appreciable


copper in the form of malachite, chrysocolla, neotocite, pitch
limonite, and associated copper oxide minerals. In general,
these cappings have developed by in situ oxidation of lowpyrite, typically chalcopyrite-bornite mineralization with total
sulfide contents of <3 vol percent and pyrite/copper sulfide
ratios of <2:1. The porphyry-related skarn mineralization, as
at Tintaya, Las Bambas, Katanga, and Alicia, include, in addition, important gossan zones as oxidation products of magnetite and massive sulfides.
Although poorly explored to date, exceptions to the above
include those porphyry systems emplaced in rocks other than
carbonate horizons of the Ferrobamba Formation and equivalent units, most notably in quartzite and sandstone of the
Yura Group and certain phases of the Andahuaylas-Yauri
batholith (Fig. 8). For example, the small Cristo de los Andes
and the large Los Chancas (Corrales, 2001) systems, emplaced into quartzose sandstone of the upper Soraya Formation, contain irregular, stratigraphically and structurally controlled immature chalcocite blankets developed in
quartz-sericitic alteration. Similarly, chalcocite enrichment at
Cotabambas occurs in intimate association with pyrite-rich,
intermediate argillic and quartz-sericitic alteration in both
diorite wall rock of the Andahuaylas-Yauri batholith and the
porphyry stocks.
In general, the presence of appreciable pyrite as a component of intermediate argillic and quartz-sericitic alteration, rather than host-rock composition, seems to be the
most important control on supergene chalcocite formation.
This is evidenced by the San Jos prospect at Katanga,
where a chalcocite blanket formed in quartz-sericitic assemblages althougth the immediate country rocks are dominated by carbonates of the Ferrobamba Formation (MMAJ,
1983; Fig. 9d). In all cases, sooty chalcocite is the main host
for copper in the supergene enrichment zones, and is accompanied by varying proportions of covellite. Cuprite,
tenorite, and native copper are also present. Both leached
cappings and supergene blankets are accompanied by widespread kaolinization of feldspars, martitization of magnetite,
and removal of anhydrite and, within the limits of groundwater penetration, gypsum. Supergene alunite may be locally formed in leached cappings overlying the more pyritic
parts of the systems, as at Ccalla, Cotabambas (Perell et al.,
2002).
In contrast to some other regions (e.g., northern Chile; Sillitoe and McKee, 1996), the age of formation of supergene
chalcocite blankets and leached cappings in the belt is poorly
constrained. In general, the close parallelism of most of the
chalcocite blankets to the present surface, together with their
partial exposure and overall immaturity, strongly favors a
Pliocene or younger age for the supergene processes; a conclusion supported by additional lines of evidence. For example, at the Ccalla deposit at Cotabambas, supergene alunite
from the top part of the leached capping returned a K-Ar age
of 3.3 0.2 Ma or late Pliocene (Perell et al., 2002). Supergene enrichment blankets located at higher elevations, as at
San Jos, Cristo de los Andes, and Quechua (~4,200 m), may
be associated either with the same late Pliocene event or with
younger topographic rejuvenation events common throughout the region (e.g., Cabrera et al., 1991).
0361-0128/98/000/000-00 $6.00

1595

Discussion
Metallogenic implications
It is apparent from the regional geology and age relationships described above that mineralization of porphyry type in
the Andahuaylas-Yauri belt broadly overlapped with the various pulses of the Andahuaylas-Yauri batholith (Fig. 10). In
particular, a close association is observed between porphyrytype alteration-mineralization and the intermediate and late
stages of quartz monzodiorite and granodiorite of the
batholith (Noble et al., 1984; Carlier et al., 1989; Perell et
al., 2002). The age relationship between porphyry copper emplacement, coeval volcanism of the Anta Formation, and syntectonic sedimentation of the San Jernimo Group red beds
are illustrated in Figures 6 and 10. All were broadly synchronous with the regional thin-skinned shortening and uplift associated with the Incaic orogeny between ~42 and 30 Ma. In
central Peru, the age of the Incaic deformation is generally
accepted to be ~41 Ma (Noble et al., 1979), whereas in southeastern Peru and northeastern Bolivia it is bracketed between
~41 and 38 Ma (Farrar et al., 1988; Sandeman et al., 1995).
In northern Chile, Incaic compression is generally assigned
an age between ~42 and 39 Ma (Hammerschmidt et al., 1992;
Mpodozis et al., 1999), although other authors consider it to
be part of a longer event of at least 20 m.y., between 50 and
30 Ma (e.g., Maksaev and Zentilli, 1999). To the east of the
Andahuaylas-Yauri belt, major thick-skinned compression, involving basement uplift along the Zongo-San Gabn
tectonothermal zone, took place at ~40 Ma (Farrar et al.,
1988), an Incaic event of crustal shortening interpreted to
have caused ramping of the (proto-) Cordillera Oriental over
the foreland (Sandeman et al., 1995).
Geodynamic evolution
A model for the tectonomagmatic evolution of the Andahuaylas-Yauri batholith and its associated porphyry-style
mineralization is formulated here, based on the geological relationships documented in this paper together with data in
Noble et al. (1984), Carlier et al. (1996), Carlotto (1998) and
references therein. The recent models for the geodynamic
evolution of the Central Andes (southern Peru, northern
Chile, and northern Bolivia) from both geologic (Clark et al.,
1990; Kennan et al., 1995; Sandeman et al., 1995) and geophysical perspectives (James and Sacks, 1999) are fundamental to the discussion that follows. These models concur in that
a period of slab flattening began in southern Peru between 50
and 45 Ma, became stable at ~42 to 40 Ma, and broadened
southward, along the arc, into northern Chile. Thus, by ~35
Ma, the entire central Andes region at the latitude of the Bolivian orocline is thought to have been undergoing flat subduction (James and Sacks, 1999).
Slab flattening is, furthermore, inferred to have produced
the crustal shortening, tectonism, and uplift assigned to the
major Incaic orogeny of the Central Andes. In the area of
study, located approximately 300 km from the trench, earlystage, mafic, cumulate pulses of the Andahuaylas-Yauri
batholith are thought to have been generated in the asthenospheric wedge between ~48 and 43 Ma and rapidly ascended into the crust (Bonhomme and Carlier, 1990) (Fig.
11a). Establishment of complete flat subduction conditions at

1595

1596

PERELL ET AL.
EARLY-STAGE ANTA ARC ?

WC
?

Depth (km)

50

moho

oc

ean

100

ic l

itho

150

200

50 - 45 Ma)

EC

sph

ere

a
0

300

200

100

400

500

Distance (km)

(
INITIAL STAGE OF
PORPHYRY COPPER EMPLACEMENT

ZONGO-SAN GABN ZONE


(PROJECTED)

ANTA ARC

WC

41 - 38 Ma)

EC

Depth (km)

50

moho

oce

ani

100

c lit

hos

phe

150

re

b
200
0

300

200

100

400

500

600

Distance (km)
MAIN STAGE OF PORPHYRY COPPER EMPLACEMENT
END-STAGE ANTA ARC

WC

Depth (km)

50

oce

ani

100

c lit

38 - 32 Ma)

EC

moho

hos

phe

re

150

200

c
0

100

200

300

400

500

Distance (km)
WC: WESTERN CORDILLERA
A : ALTIPLANO
EC : EASTERN CORDILLERA

ANDAHUAYLAS - YAURI BATHOLITH


a

a, b, c : EARLY, INTERMEDIATE,
AND LATE PHASES

FIG. 11. Schematic sequence of cross sections illustrating the formation of the Central Andes at the latitude of the Andahuaylas-Yauri belt and contiguous southeastern Peru, between the Eocene and the earliest Oligocene. Geodynamic setting largely based on models after Clark et al. (1990), Sandeman et al. (1995), and James and Sacks (1999). Present-day positions of the Western Cordillera, Altiplano, and Eastern Cordillera shown for reference. a. Illustrates the middle Eocene
(5045 Ma) magmatic arc along the eastern edge of the Western Cordillera. b. Displays emplacement of the early phases of
the batholith, followed by the intrusion of the intermediate-composition plutons (between 41 and 38 Ma) that make up much
of the present-day body of the batholith. Synchronous magma ascent and deformation of the uppermost crust (Incaic
orogeny) in conjunction with slab flattening is implied. Volcanism of the Anta Formation (Anta Arc) is important at the Western Cordillera-Altiplano transition. Deformation along the Zongo-San Gabn zone, interpreted to have occurred at the craton-orogen interface (Farrar et al., 1988; Clark, 1993; Sandeman et al., 1995), is shown for reference. c. Illustrates the main
stage of porphyry copper formation along the belt together with the terminal stages of Anta Formation volcanism (3832 Ma)
(see Fig. 12 for details).
0361-0128/98/000/000-00 $6.00

1596

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

~40 Ma is interpreted to have increased deformation and


shortening of the upper crust (James and Sacks, 1999),
thereby impeding magma ascent and favoring evolution of
large magma chambers (e.g., Sillitoe, 1998; Kay et al., 1999).
Thermal and mechanical gradients, associated with large-volume magma emplacement, would have eventually given rise
to the tectonic and gravitational conditions in the upper crust
(e.g., Kimbrough et al., 2001) that favored rapid tectonic (surface) uplift and denudation, and subsequent intrusion of the
more differentiated, intermediate-stage phases of the
batholith, between ~40 and 35 Ma (Figs. 11b and c). In the
proposed model, synchronous surface uplift and unroofing assisted with the ascent of these smaller, confined magma
chambers in the uppermost crust and triggered porphyry copper emplacement at the appropriate depths (Fig. 12).
The intermediate-stage phases of the Andahuaylas-Yauri
batholith, therefore, must have ascended as a large composite
unit and presumably reached the depth of porphyry copper
formation broadly at the same time over its entire extent,
judging by the similar ages of hydrothermal biotite formation
at Lahuani, Cotabambas, Portada, and Panchita (Table 4; Fig.
10). It should also be noted that porphyry copper formation
was recurrent and, locally, much younger systems developed
in zones that had been the locus of earlier hydrothermal activity, as in the area of Lahuani, Trapiche, Los Chancas, and
Leonor (Fig. 10).
Volcanism seems to have been subdued along the belt during much of the transition from normal to flat subduction, although, apparently, it never became completely extinct (Carlier et al., 2000) as occurred in nearby transects (Clark et al.,
1990; Sandeman et al., 1995). An example is provided by the
arc that formed the Anta Formation south of Cuzco (Fig. 6),
which is inferred to have been fed from magma chambers associated with the Andahuaylas-Yauri batholith.
Figure 13a displays a hypothetical map of the Wadati-Benioff zone during the late Eocene to Oligocene period of flat
subduction in the central Andes as reconstructed by James

Porphyry Copper
Stock

and Sacks (1999). In such a geodynamic model, the late


Eocene to early Oligocene Andahuaylas-Yauri batholith and
its associated porphyry copper mineralization appear to have
been emplaced at the inflection point between flat subduction to the south and normal subduction to the north. Although speculative, this location is remarkably similar to the
position of the late Miocene to Pliocene porphyry copper belt
of central Chile (Fig. 13b), which hosts the world class deposits at Los Pelambres, Los Bronces-Ro Blanco, and El Teniente (Skewes and Stern, 1994, 1995). Moreover, progressive flattening of the subducting slab during the middle
Miocene in central Chile resulted in eastward migration of
the volcanic front accompanied by intense crustal thickening,
deformation, uplift, and erosion (Jordan et al., 1983; Kay et
al., 1991; Skewes and Stern, 1994, 1995; Kurtz et al., 1997),
an overall geodynamic setting that compares closely with slab
flattening and deformation inferred for the AndahuaylasYauri belt during the Incaic orogeny.
Possible extensions of the Andahuaylas-Yauri belt
The northern and southern extensions of the AndahuaylasYauri belt are not well constrained. Indeed, most reconstructions for late Eocene to early Oligocene magmatism and mineralization imply significant segmentation along the belt at
the latitude of the Bolivian Orocline in the central Andes (Sillitoe, 1988, 1989; Clark et al., 1990; Clark, 1993; James and
Sacks, 1999; Petersen, 1999). To the north, the belt apparently stops at the Abancay Deflection (Fig. 14). Farther
north, late Eocene to Oligocene magmatism is present in central Peru (Noble et al., 1984; Soler and Bonhomme, 1988;
Clark et al., 1990; Noble and McKee, 1999; Petersen, 1999),
and northeast of Lima the Quicay high-sulfidation gold deposit has been related to magmatism associated with Incaic
deformation (Noble and McKee, 1999).
The extension of the belt into southeastern Peru is complicated by the presence of the widespread, post-Oligocene volcanic cover of the region. The following lines of evidence,

Andesitic Volcanism
(Anta Formation)

Rapid
Uplift

20 km

Molasse Deposits

Mesozoic Sequences
Fold-Thrust
Belt
50 km

Early-Stage
Cumulate Gabbros
and Diorites

Intermediate-Stage
Granodiorites and
Quartz Monzodiorites

FIG. 12. Schematic illustration of the spatial and temporal relationships between batholith ascent, rapid regional uplift,
porphyry copper emplacement, compressive deformation, volcanism, and synorogenic sedimentation. The relatively shallow
emplacement of the main, intermediate-stage body of the batholith is implied. Figure borrows from Skewes and Stern (1994)
and Sillitoe (1997, 1998).
0361-0128/98/000/000-00 $6.00

1597

1597

1598

PERELL ET AL.

75

Lima

ANDAHUAYLAS-YAURI
BATHOLITH

50
5

65B

70

RA

ZI L

Cuzco

75

ZONGO - SAN GABAN ZONE


15

Puno
Arequipa

BOLIVIA

La Paz

PERU

CHILE

FLAT-SLAB
SEGMENT

20

100 km

UA
Y

Arica

200
PA
R

AG

100
Calama

65

70

75

ARGENTINA

Recent Volcanoes
Late Tertiary Batholith Antofagasta
125

CENTRAL
VOLCANI C
ZONE

Benioff Zone Contours (km)

25

Chile T

Giant Porphyry
Copper deposit

rench

Sierras Pampeanas

Tucumn
Copiap
250

La Serena

FLAT-SLAB/
NORMAL SLAB
ALONG-ARC
PROPAGATION

La Rioja
125

FLAT-SLAB
SEGMENT

30

200

LOS PELAMBRES-EL PACHN


150
San Juan
JUAN F ER NNDEZ R ID

GE

Mendoza
Santiago
EL TENIENTE

San Luis
RO BLANCO-LOS BRONCES

100 km

SOUTHERN
VOLCANI C
ZONE
35

50 100

FIG. 13. Comparison between the geodynamic settings of the Andahuaylas-Yauri and central Chile porphyry copper belts.
a. Illustrates the setting of the Andahuaylas-Yauri batholith at the inflection corridor between flat and normal subduction
during the late Eocene and the Oligocene, as modeled by James and Sacks (1999). b. Displays the setting of the central Chile
Miocene plutons and late Miocene to Pliocene porphyry copper belt at the inflection of the present-day subduction zone
(Cahill and Isacks, 1992; Ramos et al., 2002). Plutons of the partially unroofed late Tertiary batholith taken from Parada et
al. (1988) and Kurtz et al. (1997). Note the broad similarities between both settings and the presence of the Zongo-San
Gabn zone and the Sierras Pampeanas basement highs in each case.
0361-0128/98/000/000-00 $6.00

1598

1599

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU


80

70
MIDDLE EOCENE-EARLY OLIGOCENE
PORPHYRY COPPER DEPOSIT
SELECTED DEPOSIT OF THE
LATE PALEOCENE-EARLY EOCENE BELT

BRAZIL
PERU

(1) CUAJONE; (2) QUELLAVECO; (3) TOQUEPALA;


(4) CERRO COLORADO; (5) SPENCE

C
EN

POST-EARLY OLIGOCENE VOLCANIC


AND SEDIMENTARY COVER
SEQUENCES

TR

QUICAY (38-37)

RU
PE

10

10

(30-32) : AGE RANGE


MOROSAYHUAS (35)

COTABAMBAS (36-35)

LIMA

LAS BAMBAS (36-35)

ECTIO
?

FL
Y DE
NCA

ABA

ZONGO-SAN GABAN

LOS CHANCAS (32)


KATANGA (33-30)

BOLIVIA

TINTAYA (42; 35-33)

LA PAZ
(1)
(2)

SANTA LUCIA (32-30)

ATASPACA (40)
(3)

TICNAMAR (41)
QUEEN ELIZABETH (36)

BRAZIL

PER U

20

(4)

20

QUEBRADA BLANCACOLLAHUASI-UJINA (35-32)

LIMA

ESPERANZATELEGRAFO (42-41)

EL ABRA-CONCHI (37-36)

BOLIVIA

(5)

CHUQUICAMATAMM-TOMIC (34-31)
GABY (42-40)

LA PAZ

CENTINELAPOLO SUR (44)

SALTA
TACA-TACA (34-33)

CHIMBORAZO
(39-38*)

ESCONDIDA - ESCONDIDA
NORTE-ZALDIVAR (38-35*; 35-31)
EXPLORADORA (33-32)

CHILE

SALTA

EL SALVADOR (44-41)

CHILE

Area with Eocene to


Oligocene (Incaic)
deformation and/or
terrigenous sedimentation

TREN

CH

Inner Arc

POTRERILLOS (36-35)
LA FORTUNA (35-32)

CHILE

APOLINARIO (35)

30
LOICA (35)

30

ARGENTINA

ARGENTINA

SANTIAGO

SANTIAGO

500 km

500 km

a
80

70

FIG. 14. a. Schematic illustration of the Andahuaylas-Yauri belt and its possible northward and southward extensions. Segmentation along the belt is apparent at the Abancay Deflection. The extension of the belt into northern Chile, if real (see
text for discussion), is masked by younger volcanic and sedimentary cover sequences. Geologic elements borrowed from
Maksaev and Zentilli (1988), Sillitoe (1988, 1990), Clark et al. (1990, 1998), Maksaev (1990), Clark (1993), Perell et al.
(1996), Petersen et al. (1996), Cornejo et al. (1997), Cuadra et al. (1997), Clark et al. (1998) and references therein, Noble
and McKee (1999), Petersen (1999), Rojas et al. (1999), Gustafson et al. (2001), Zappettini et al. (2001), and J. Perell,
unpub. data for Esperanza (2000), Telgrafo, Polo Sur, and Conchi. Asterisks indicate ages of intrusions at La Escondida,
Zaldvar, and Chimborazo, based on U-Pb (zircon) dating by Richards et al. (1999). All other ages are for hydrothermal alteration assemblages. b. (inset) Relationship between Eocene to early Oligocene (Incaic) porphyry copper mineralization,
synorogenic sedimentation in Chile (Mpodozis et al., 1999; Tomlinson et al., 1999; J. Perell, unpub. data, 2000) and Argentina (Jordan and Alonso, 1987; Adelman and Grler, 1999; Hernndez et al., 1999; Kraemer et al., 1999; Coutand et al.,
2001). The location of the inner Arc in Bolivia and southeastern Peru (sensu James and Sacks, 1999) is shown for reference.
According to James and Sacks (1999), the inner arc is characterized by high heat flow, high degree of crustal deformation,
and high electrical conductivity, features that they consider to have been inherited from the period of flat subduction beneath the central Andes during late Eocene to early Oligocene time. Their definition of inner arc includes the inner arc domain and parts of the inner Cordillera Occidental of Clark et al. (1990) and Sandeman et al. (1995).
0361-0128/98/000/000-00 $6.00

1599

1600

PERELL ET AL.

however, imply that late Eocene to early Oligocene deformation and magmatism occurred in the area and that the extension of the belt into Chile is represented at Ataspaca and,
probably, Santa Luca.
1. Important thin-skinned deformation of fold-thrust belt
type took place during the late Eocene (~40 Ma) in the Santa
Luca area, and was accompanied by the synorogenic, molasse-type sedimentation of the Puno Group (Portugal, 1974;
Jaillard and Santander, 1992). This association is remarkably
similar to that described for the San Jernimo Group in the
Andahuaylas-Yauri belt near Cuzco.
2. Magmatism at Santa Luca was active in the early
Oligocene (~3032 Ma; Clark et al., 1990) and was broadly
coeval with the ~30-Ma flows interbedded in San Jernimo
Group redbeds, and with the ~30- to 35-Ma magmatism, alteration, and mineralization of the Katanga area, in the Andahuaylas-Yauri belt (Fig. 10).
3. Magmatism and Incaic deformation in the Santa Luca
area are interpreted to have occurred along the same crustal
discontinuity that chaneled the deformational front in the Andahuaylas-Yauri belt near Cuzco (Jaillard and Santander,
1992), i.e., the Cuzco-Puno high (Figs. 3 and 6).
4. Southeasterly extensions of the regional late Eocene to
early Oligocene structures present in the Andahuaylas-Yauri
region appear to extend into southeastern Peru (Sempere et
al., 2002) and across the border into northernmost Chile
(Garca et al., 2002).
5. In the Tarata district, near the border with Chile, the
~45- to 39-Ma magmatic and hydrothermal activity recorded
at Ataspaca (Clark et al., 1990) represents the connection
with the late Eocene to early Oligocene porphyry copper belt
of northern Chile.
These observations suggest that there may have been continuity between the Andahuaylas-Yauri and the northern
Chile porphyry copper belts during middle (late) Eocene to
early Oligocene time (~4530 Ma; Fig. 14). Interruption of
the belt at the latitude of the Arica deflection may be apparent, and enhanced by the widespread younger volcanism that
mantles much of the region. Alternatively, the interruption
may be attributable to the changing tectonomagmatic framework in response to along-arc flat subduction propagation in
southern Peru during the late Eocene to Oligocene (James
and Sacks, 1999). Although the rapidly changing metallogenesis of the Miocene to Pliocene magmatic belt of central
Chile (Kay et al., 1999; Kay and Mpodozis, 2002) may be considered as a more modern analog (Fig. 13), no such along-arc
migration is apparent in the late Eocene to early Oligocene
porphyry copper belt of northern Chile. On the contrary, as
shown in Figure 14a, much of the porphyry copper mineralization there, irrespective of size, seems to have taken place
during a well-defined interval of ~10 to 12 Ma duration (cf.
Maksaev and Zentilli, 1988; Maksaev, 1990).
By analogy with the Andahuaylas-Yauri belt, and given that
a genetic association between the late Eocene to early
Oligocene porphyry copper belt of northern Chile and Incaic
compression, uplift, and denudation is widely accepted
(Maksaev and Zentilli, 1999), it is here speculated that late
Eocene to early Oligocene porphyry copper mineralization
0361-0128/98/000/000-00 $6.00

of northern Chile might have also taken place under similar


conditions of subduction flattening (Mpodozis and Perell,
2003). Furthermore, subduction flattening, perhaps triggered
by fast convergence (Pardo Casas and Molnar, 1987; Soler
and Bonhomme, 1990) between ~45 and 35 Ma, might have
occurred essentially simultaneously along the late Eocene
magmatic arc of southern Peru and northern Chile. Indeed,
current knowledge (Fig. 14b) suggests that Eocene to
Oligocene Incaic compression, shortening, and sedimentation extended eastward as far as the Altiplano and Eastern
Cordillera of Bolivia (Farrar et al., 1988; Sempere et al.,
1997) and as far south as 26S in the Puna of northwestern
Argentina (Jordan and Alonso, 1987; Alonso, 1992; Kraemer
et al., 1999; Coutand et al., 2001).
Summary and Conclusions
The Andahuaylas-Yauri belt is defined by 31 deposits and
prospects with porphyry-style alteration and mineralization,
accompanied by hundreds of occurrences of magnetite-bearing, skarn-type Fe-Cu mineralization. A salient feature of the
belt is the spatial association of porphyry stocks and related
mineralization with the Andahuaylas-Yauri batholith, a large
composite body of calc-alkaline intrusions of middle Eocene
to early Oligocene age (~4832 Ma). Porphyry stocks are
dominated by multiphase, calc-alkaline, biotite- and amphibole-bearing intrusions of dacitic and granodioritic composition. Hydrothermal alteration is typical of porphyry ore deposits elsewhere and includes potassic, propylitic,
quartz-sericitic, sericite-clay-chlorite, calcic-potassic, and advanced argillic alteration types. Calc-silicate assemblages with
skarn-type mineralization occur where carbonate country
rocks of the Ferrobamba Formation predominate.
Porphyry copper deposits and prospects of the belt range
from gold-rich, molybdenum-poor examples, through systems containing both gold and molybdenum, to relatively
molybdenum-enriched, gold-depleted endmembers. Goldonly porphyry systems with similarities to the gold porphyries
of the Maricunga belt of northern Chile also occur. Gold-rich
members of the belt do not possess any unique features that
distinguish them from gold-poor, molybdenum-rich counterparts, except perhaps for their appreciably higher contents of
hydrothermal magnetite and the presence of amphibole- and
pyroxene-bearing alteration assemblages in addition to potassic alteration.
Available geochronologic data confirm that much of the
porphyry alteration and mineralization in the belt formed between ~40 and 35 Ma, i.e., during the late Eocene, although
the complete spectrum of age ranges between ~42 and 28 Ma
(middle Eocene to earliest late Oligocene). No sub-belts or
age trends are apparent in the data, which suggest that porphyry mineralization took place during a well-defined interval
of time along the 130-km-wide, 300-km-long belt, broadly simultaneously with emplacement of intermediate-stage
phases of the Andahuaylas-Yauri batholith (~4032 Ma). The
data also support the broad inference that gold-rich and goldpoor members of the belt formed synchronously, irrespective
of their locations, country rocks, and porphyry stock compositions. However, evidence from the belt does not support any
apparent connection between metallogenesis and crustal
properties (Titley, 1990).

1600

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU

There is ample evidence to support a genetic relationship


between emplacement of the intermediate-stage phases of
the Andahuaylas-Yauri batholith, porphyry copper formation,
and regional shortening, uplift, and synorogenic clastic sedimentation associated with the compressive Incaic orogeny
between ~40 and 32 Ma. This overall geologic setting is strikingly similar to that of the late Eocene to early Oligocene porphyry copper belt of northern Chile, where compelling evidence defines a close link between porphyry copper
formation and Incaic compression, crustal thickening, and
synchronous exhumation along the Domeyko Cordillera
(Maksaev and Zentilli, 1999).
Major cumulative supergene chalcocite enrichment blankets are absent from porphyry systems of the AndahuaylasYauri belt because of the low pyrite contents, the poorly developed nature of quartz-sericitic alteration, the high
neutralization capacities of both potassic alteration and carbonate country rocks, and the climatic regime of the region.
Under the appropriate conditions, however, supergene chalcocite in the belt does form and, although producing immature blankets, can give rise to potentially economic enriched
mineralization. Geomorphologic and geochronologic evidence from the region indicates that the formation of chalcocite blankets has occurred at least since the late Pliocene,
thereby postdating, as in other regions of the central Andes
(e.g., the Faralln Negro district and nearby Agua Rica deposit, Argentina; McBride, 1972, in Sasso and Clark, 1998;
Perell et al., 1998), by >10 m.y. the last global scale secondary enrichment event that terminated with central Andean climatic dessication at ~15 Ma (Sillitoe and McKee,
1996; Brimhall and Mote, 1997). In the Andahuaylas-Yauri
belt, conditions appropriate for supergene copper enrichment include the presence of country rocks other than carbonate rocks, such as quartz arenite of the Yura Group, appreciable pyrite contents in sericite-clay-chlorite and/or
quartz-sercititic alteration, and the pluvial regime characteristic of the elevated cordilleran topography of the region.
Nevertheless, given the predominance of carbonate sequences in the belt, giant (cf. Clark, 1993) copper deposits
contained in porphyries and porphyry-related skarns, and not
chalcocite blankets, should continue to be the principal exploration targets in the belt.
Acknowledgments
We would like to thank Gabriel Carlier, Ricardo Martini,
Rodrigo Arcos, Toms Vila, Victor Torres, Carlos Seminario,
Orlando Pariona, Edmundo Martnez, Jos Crdenas, Sandra
Romanville, Jorge Fierro, Gerardo Soto, Eduardo Tejada, and
Christophe Noblet for their contributions on various aspects
of the geology of the belt. Percy Garca assisted with the geologic database for the region and Hctor Poblete drafted the
figures that accompany this paper. We are also grateful for the
support of the Anaconda Per and Antofagasta Minerals staffs
in Lima and Santiago, respectively and the field assistance of
Luis Castillo, Luis Gago, Elmer Prado, Mximo Condori, and
many others. Luca Cuitio is thanked for her petrographic
work and Carlos Prez de Arce for K-Ar geochronology. Dick
Sillitoe and Constantino (Cocho) Mpodozis are thanked for
their comments on an early version of the manuscript. We are
indebted to Alan Clark for his extraordinarily detailed reviews
0361-0128/98/000/000-00 $6.00

1601

that led to numerous improvements. Constructive comments


by Economic Geology reviewers Dick Tosdal and John Dilles,
as well as the indefatigable editor are highly appreciated.
Nevertheless, responsibility for the opinions expressed and
any errors is ours alone. Special thanks are due to Antofagasta
Minerals for supporting basic research beyond day-to-day exploration work and for permission to publish.
January 6, June 4, 2003
REFERENCES
Adelman, D., and Grler, K., 1999, Evolucin de las cuencas negenas de la
Puna Austral, ejemplificado por el rea del Salar de Antofalla, noroeste de
Argentina, in Gonzlez Bonorino, G., Omarini, R., and Viramonte, J., eds.,
Geologa del Noroeste Argentino: Relatorio XIV Congreso Geolgico
Argentino, Salta, Argentina, September 1924, 1999, v. I, p. 361363.
Aizawa, N., and Tomizawa, N., 1986, Prospecting of Coroccohuayco copper
deposit, Per-with special reference to the southern skarn zone: Mining
Geology, v. 36, p. 83100 (in Japanese with English abstract).
Alonso, R.N., 1992, Estratigrafa del Cenozoico de la cuenca de Pastos
Grandes (Puna Saltea) con nfasis en la Formacin Sijes y sus boratos:
Revista de la Asociacin Geolgica Argentina, v. 47, no. 2, p. 189199.
BHP Company Limited, 1999, Quarterly report on exploration and development, MarchMay, 1999: Melbourne, Australia, press release, June 17,
1999.
Baksi, A.K., 1982, A note on the calculation of errors in conventional K-Ar
dating: Chemical Geology, v. 35, p. 167172.
Bellido, E., Girard, D., and Paredes, J., 1972, Mapa metalognico del Per:
Per, Servicio de Geologa y Minera, scale 1: 2,500,000.
Benavides-Cceres, V., 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.
Bonhomme, M.G., and Carlier, G., 1990, Relations entre magmatisme et
minralisations dans le Batholite dAndahuaylas-Yauri (Sud Prou):
Donnes gochronologiques: 2nd International Symposium on Andean
Geodynamics, Grenoble, France, p. 329331.
Brimhall, G.H., and Mote, T.I., 1997, Optimal secondary mineralization in
the Andes: Vadose response to global Cenozoic cooling events, glaciation,
eustacy, and dessication [abs]: Geological Society of America, Annual
Meeting, Salt Lake City, Abstracts with Programs, p. A17.
Cabrera, J., Sbrier, M., and Mercier, J.L., 1991, Plio-Quaternary geodynamic evolution of a segment of the Peruvian Andean Cordillera located
above the change in the subduction geometry: The Cuzco region: Tectonophysics, v. 190, p. 331362.
Cahill, T., and Isacks, B.L., 1992, Seismicity and shape of the subducted
Nazca plate: Journal of Geophysical Research, v. 97, B12, p. 17,50317,529.
Crdenas, J., Carlotto, V., Flores, T., Mamani, M., Rodriguez, R., Latorre, O.,
and Jaimes, F., 1997, Estratigrafa de la regin de Sicuani y sus
correlaciones con las regiones de Cuzco y Puno: IX Congreso Peruano de
Geologa, Sociedad Geolgica del Per, Extended Abstracts, Volumen
Especial 1, p. 255259.
Crdenas, J., Carlotto, V., Vallenas, V., Chvez, R., and Gil, W., 1999, Las
areniscas cuprferas de las capas rojas del Grupo San Jernimo (Eoceno
Medio-Oligoceno Inferior) de la regin de Cuzco y Sicuani: XVIII Curso
Internacional de Postgrado en Metalogenia, Quito, Ecuador, Universidad
Central del Ecuador, June 1999, Actas, p. 155159.
Carlier, G., Carlotto, V., Ligarda, R., and Manrique, E., 1989, Estudio
metalogentico de la subprovincia cuproaurfera Tintaya-Bambas:
Convenio de Cooperacin Cientfica UNI-ORSTOM, Informe Final 19841988, Lima, Per, p. 143248.
Carlier, G., Lorand, J.P., Bonhomme, M., and Carlotto, V., 1996, A reappraisal of the Cenozoic Inner Arc magmatism in southern Peru: Consequences for the evolution of the Central Andes for the past 50 Ma: Third
International Symposium on Andean Geodynamics, St. Malo, France, Extended Abstracts volume, p. 551554.
Carlier, G., Carlotto, V., Fornari, M., Sempere, T., and Crdenas, J., 2000, El
magmatismo Cenozoico del sur del Per y sus implicaciones geodinmicas
[abs.]: X Congreso Peruano de Geologa, Sociedad Geolgica del Per,
Abstracts, Volumen Especial 2, p. 11.
Carlotto, V., 1998, volution Andine et Raccourcissement au niveau de
Cusco (13-16S) Prou: Enregistrement sdimentaire, chronologie,

1601

1602

PERELL ET AL.

controles palogographiques, volution cinmatique: Unpublished Ph.D.


thesis, Grenoble, France, Universit Joseph Fourier, 159 p.
Carlotto, V., Jaillard, E., and Mascles, G., 1993, Sedimentation, paleogeography and tectonics of the Cuzco area between Kimmeridgian?Paleocene
times: Relations with the south Peruvian margin: Second International
Symposium on Andean Geodynamics, Oxford, United Kingdom, Extended
Abstracts volume, p. 287290.
Carlotto, V., Carlier, G., Crdenas, J., Gil, W., and Chvez, R., 1996a, The red
beds of the San Jernimo Group (Cuzco, Peru), marker of the Inca 1 tectonic event: Third International Symposium on Andean Geodynamics, St.
Malo, France, October 46, 1999, Extended Abstracts volume, p. 303306.
Carlotto, V., Gil, W., Crdenas, J., and Chvez, R., 1996b, Geologa de los
cuadrngulos de Urubamba y Calca, Hojas 27r y 27s: Lima, Instituto
Geolgico Minero y Metalrgico, Carta Geolgica Nacional, Boletn no. 65,
serie A, 245 p.
Carlotto, V., Jaillard, E., Carlier, G., and Mascle, G., 1997, Las cuencas sinorognicas (Eoceno-Mioceno) de la terminacin NO del Altiplano (Cusco):
IX Congreso Peruano de Geologa, Sociedad Geolgica del Per, Extended
Abstracts, Volumen Especial 1, p. 267271.
Carlotto, V., Carlier, G., Jaillard, E., Sempere, T., and Mascle, G. 1999, Sedimentary and structural evolution of the Eocene-Oligocene Capas Rojas
basin: Evidence for a Late Eocene lithospheric delamination event in the
southern Peruvian Altiplano: Fourth International Symposium on Andean
Geodynamics, Goettingen, Germany, Extended Abstract Volume, p.
141146.
Carman, G.D., Nicholson, P., Ianos, S., Bernuy, E., Salinas, C., Ormsby, W.,
and Perkins, J., 2000, Geology and exploration progress of the Accha zinc
deposit, Cusco, Peru: ProExplo99, Primer Volumen de Monografa de
Yacimientos Minerales Peruanos: Historia, Exploracin y Geologa. Lima,
Instituto de Ingenieros de Minas del Per, Volumen Luis Hoschschild
Plaut, p. 245250
Chvez, R., Gil, W., Carlotto, V., Crdenas, J., Jaillard, E., 1996, The Altiplano-Eastern Cordillera limit in the Urubamba region (Cusco, Peru):
Third International Symposiun on Andean Geodynamics, St. Malo, France,
Extended Abstracts volume, p. 319322.
Clark, A.H., 1993, Are outsized porphyry copper deposits either anatomically
or environmentally distinctive?: Society of Economic Geologists Special
Publication 2, p. 213283.
Clark, A.H., and Arancibia, O.N., 1995, The occurrence, paragenesis, and
implications of magnetite-rich alteration-mineralization in calc-alkaline
porphyry copper deposits, in Clark, A.H., ed., Giant Ore Deposits II: Controls on the Scale of Orogenic Magmatic-Hydrothermal Mineralization:
Proceedings of the Second Giant Ore Deposits Workshop, April 2527,
1995, Kingston, Ontario, Canada, Queens University, p. 511581.
Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., Arenas, M., France,
L.J., McBride, S.L., Woodman, P.l., Wasteney, H.A., Sandeman, H.A., and
Archibald, D.A., 1990, Geologic and geochronologic constraints on the
metallogenic evolution of the Andes of southeastern Peru: ECONOMIC GEOLOGY, v. 85, p. 15201583.
Clark, A.H., Archibald, D.A., Lee, A.W., Farrar, E., and Hodgson, C.J., 1998,
Laser probe 40Ar/ 39Ar ages of early- and late-stage alteration assemblages,
Rosario porphyry copper-molybdenum deposit, Collahuasi district, I Region, Chile: ECONOMIC GEOLOGY, v. 93, p. 326337.
Crdova, E., 1986, Un bassin intramontagneaux andin pruvienne. Les
Couches Rouges du bassin de Cuzco (Maastrichtien-Palocne):
Unpublished Ph.D. thesis, Pau, France, Univerit de Pau, 272 p.
Cornejo, P., Tosdal, R.M., Mpodozis, C., Tomlinson, A.J., Rivera, O., and
Fanning, C.M., 1997, El Salvador, Chile porphyry copper deposit revisited:
Geologic and geochronologic framework: International Geology Review, v.
39, p. 2254.
Corrales, E., 2001, Proyecto Los Chancas: II International Congress of
Prospectors and Explorers, ProExplo 2001, Lima, Per, April 2427, 2001,
Instituto de Ingenieros de Minas del Per, 9 p. (CD-ROM)
Coutand, I., Cobbold, P.R., de Urreiztieta, M., Gautier, P., Chauvin, A.,
Gapais, D., Rossello, E., and Lpez-Gamund, O., 2001, Style and history
of Andean deformation, Puna plateau, northwestern Argentina: Tectonics,
v. 20, no. 2, p. 210234.
Cox, D.P., and Singer, D.A., eds., 1986, Mineral deposit models: U.S. Geological Survey Bulletin 1693, 379 p.
Cuadra, P., Zentilli, M., Puig, A., and Tidy, E., 1997, Dataciones radiomtricas recientes en Radomiro Tomic: VIII Congreso Geolgico Chileno,
Universidad Catlica del Norte, Departamento de Ciencias Geolgicas,
Antofagasta, Chile, October 1317, 1997, Actas, v. II, p. 916919.

0361-0128/98/000/000-00 $6.00

Dalrymple, G.B., and Lamphere, M.A., 1966, Potasium argon dating: Principles, techniques and applications to geochronology: San Francisco, Freeman and Company, 258 p.
Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits:
ECONOMIC GEOLOGY 75th ANNIVERSARY VOLUME, p. 317391.
Farrar, E., Clark, A.H., Kontak, D.J., and Archibald, D.A., 1988, Zongo-San
Gabn Zone: Eocene foreland boundary of the Central Andean orogen,
northwestern Bolivia and southeast Peru: Geology, v. 16, p. 5558.
Fierro, J., Zweng, P.L., Gamarra, H., and Garate, G., 1997, Chabuca Este
Cu- (Au, Ag) skarn deposit at Tintaya: IX Congreso Geolgico Peruano de
Geologa, Sociedad Geolgica del Per, Extended Abstracts, Volumen
Especial 1, p. 3739.
Fierro, J., Jones, B., and Lenzi, G., 2002, Los prfidos de Cu-Au de
Antapaccay en el distrito mineralizado de Tintaya, Per [abs.]: XI Congreso
Peruano de Geologa, Sociedad Geolgica del Per, Lima, Per,
September 2528, 2002, Abstracts, p. 235.
Fornari, M., Mamani, M., Ibarra, I., and Carlier, G., 2002, Datacin del
perodo volcnico Tacaza en el Altiplano de Per y Bolivia [abs.]: XI
Congreso Peruano de Geologa, Sociedad Geolgica del Per, Lima, Per,
September 2528, 2002, Abstracts, p. 21.
Garca M., Hrail G., Charrier, R., Mascle, G., Fornari, M., and Prez de
Arce, C., 2002, Oligocene-Neogene tectonic evolution of the Altiplano of
northern Chile: 5th International Symposium on Andean Geodynamics,
Toulouse, France, September 1618, 2002, Extended Abstracts volume, p.
235238.
Guilbert, J.M., and Lowell, J.D., 1974, Variations in zoning patterns in porphyry ore deposits: Canadian Institute of Mining and Metallurgy Transactions, v. 77, p. 105115.
Gustafson, L.B., and Hunt, J.P., 1975, The porphyry copper deposit at El Salvador, Chile: ECONOMIC GEOLOGY, v. 70, p. 857912.
Gustafson, L.B., and Quiroga, J., 1995, Patterns of mineralization and alteration below the porphyry copper orebody at El Salvador, Chile: ECONOMIC
GEOLOGY, v. 90, p. 216.
Gustafson, L.B., Orquera, W., McWilliams, M., Castro, M., Olivares, O.,
Rojas, G., Maluenda, J., and Mendez, M., 2001, Multiple centers of mineralization in the El Indio Muerto district, El Salvador, Chile: ECONOMIC
GEOLOGY, v. 96, p. 325350.
Hammerschmidt, K., Dobel, R., and Friederichsen, H., 1992, Implication of
40
Ar/39Ar dating of early Tertiary volcanic rocks from the north Chilean Precordillera: Tectonophysics, v. 202, p. 5581.
Haq, B.U., and van Eysinga, F.W.B., 1987, Geological time table: Amsterdam, The Netherlands, Elsevier Science Publishers. (Table)
Hrail, G., Rochat. P., Baby, P., Aranibar, O., Lavenu, A., and Mascle, G.,
1993, El Altiplano norte de Bolivia: evolucin geolgica terciaria, in
Charrier, R., ed. Actas del II Simposio Internacional de Estudios
Altiplnicos, Arica, Chile, October 19-21, 1993: Santiago, Chile,
Universidad de Chile, Vicerrectora Acadmica y Estudiantil,
Departamento de Postgrado y Posttulo, p. 3344
Hernndez, R.M., Galli, C.I., and Reynolds, J., 1999, Estratigrafa del
Terciario en el noroeste argentino, in Gonzlez Bonorino, G., Omarini, R.,
and Viramonte, J., eds., Geologa del Noroeste Argentino: Relatorio XIV
Congreso Geolgico Argentino, Salta, Argentina, September 1924, 1999,
v. I, p. 316328.
Jaillard, E., 1994, Kimmeridgian to Paleocene tectonic and geodynamic evolution of the Peruvian (and Ecuadorian) margin, in Salfity, J.A., ed., Cretaceous Tectonics in the Andes: Braunschweig/Wiesbaden, Fried, Vieweg
and Sohn, Earth Evolution Sciences, p. 101107.
Jaillard, E., and Santander, G., 1992, La tectnica polifsica en escamas de la
zona de Maazo-Lagunillas (Puno, sur del Per): Bulletin de lInstitute
Franais dEtudes Andines, Lima, v. 21, p. 3758.
Jaillard, E., and Soler, P., 1996, The Cretaceous to Early Paleogene tectonic
evolution of the northern central Andes and its relations to geodynamics:
Tectonophysics, v. 259, p. 4153.
Jaillard, E., Carlotto, V., Crdenas, J., Chvez, R., and Gil, W., 1993, La
nappe des couches rouges de Cuzco (sud du Prou): Mise en vidence
stratigraphique, interprtations tectoniques et palogographiques:
Comptes Rendus de lAcademie de Sciences de Paris, v. 316, no. 2, p.
379386.
Jaillard, E., Grambast-Fessard, N., Feist, M, and Carlotto, V., 1994, Senonian-Paleocene charophyte succession of the Peruvian Andes: Cretaceous
Research, v. 15, p. 445456.
Jaillard, E., Hrail, G., Monfret, T., Daz-Martnez, E., Baby, P., Lavenu, A.,
and Dumont, J.F., 2000, Tectonic evolution of the Andes of Ecuador, Peru,

1602

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU


Bolivia and northernmost Chile. in Cordani, U.G., Milani, E.J., Thomaz
Filho, A., and Campos, D.A., eds., Tectonic Evolution of South America:
31st International Geological Congress, Rio de Janeiro, Brazil, August 617,
2000, Proceedings, p. 481559.
Jaimes, F., Romero, D., Carlotto, V., Marocco, R., and Crdenas, J., 1997, La
tectnica y la evolucin geodinmica de la cuenca Paruro (Mioceno
Superior) en la regin de Cusco: IX Congreso Peruano de Geologa,
Sociedad Geolgica del Per, Extended Abstracts, Volumen Especial 1, p.
337340.
James, D.E., 1971, Andean crustal and upper mantle structure: Journal Geophysical Research, v. 84, p. 32463271.
James, D.E., and Sacks, I.S., 1999, Cenozoic formation of the Central Andes:
A geophysical perspective, in Skinner, B.J., ed., Geology and Ore Deposits
of the Central Andes: Society of Economic Geologists Special Publication
7, p. 125.
Jones, B., Fierro, J., and Lenzi, G., 2000, Antapaccay project geology [abs.]:
Seminario Internacional Yacimientos tipo Prfido de Cu-Au, Lima, 2000,
Facultad de Ingeniera Geolgica, Minera y Metalrgica, Promocin de
Gelogos 2000 Abstracts, v. II, 1 p.
Jordan, T.E., and Alonso, R.N., 1987, Cenozoic stratigraphy and basin tectonics of the Andes mountains, 20-28 South latitute: American Association of Petroleum Geologists Bulletin, v. 71, no. 1, p. 4964.
Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ramos, V.A.,
and Ando, C.J., 1983, Andean tectonics related to geometry of subducted
Nazca plate: Geological Society of America Bulletin, v. 94, p. 341361.
Kay, S.M., and Mpodozis, C., 2002, Magmatism as a probe to the Neogene
shallowing of the Nazca plate beneath the modern Chilean flat-slab: Journal of South American Earth Sciences, v. 15, p. 3957.
Kay, S.M., Mpodozis, C., Ramos, V.A., and Munizaga, F., 1991, Magma
source variations for Tertiary magmatic rocks associated with a shallowing
subduction zone and a thickening crust in the Central Andes (28 to 33S),
in Harmon, R.S., and Rapela, C.W., eds., Andean Magmatism and its Tectonic Setting: Geological Soiety of America Special Paper 265, p. 113137.
Kay, S.M., Mpodozis, C., and Coira, B., 1999, Neogene magmatism, tectonism, and mineral deposits of the Central Andes (22 to 33 S Latitude), in
Skinner, B.J., ed., Geology and Ore Deposits of the Central Andes: Society
of Economic Geologists Special Publication 7, p. 2759.
Kennan, L., Lamb, S., and Rundle, C., 1995, K-Ar dates from the Altiplano
and Cordillera Oriental of Bolivia: Implications for Cenozoic stratigraphy
and tectonics: Journal of South American Earth Sciences, v. 18, p. 163186.
Kimbrough, D.l., Smith, D.P., Mahoney, J.B., Grove, M., Gastil, R.G., Ortega-Rivera, A., and Fanning, C.M., 2001, Forearc-basin sedimentary response to rapid Late Cretaceous batholith emplacement in the Peninsular
Ranges of southern Baja California: Geology, v. 29, p. 491494.
Klinck, B.A., Ellison, R.A., and Hawkins, M.P., compilers, 1986, The geology
of the Cordillera Occidental and Altiplano west of Lake Titicaca, southern
Peru: Lima, British Geological Survey, Instituto Geolgico Minero y Metalrgico, 353 p.
Kraemer, B., Adelmann, D., Alten, M., Schnurr, W., Erpstein, K., Kiefer, E.,
van den Bogaard, P., and Grler, K., 1999, Incorporation of the Paleogene
foreland into the Neogene Puna plateau: The Salar de Antofalla area, NW
Argentina: Journal of South American Earth Sciences, v. 12, p. 157182.
Kurtz, A.C., Kay, S.M., Charrier, R., and Farrar, E., 1997, Geochronology of
Miocene plutons and exhumation history of the El Teniente region, central
Chile (3435): Revista Geolgica de Chile, v. 24, no. 1, p. 7590.
Ligarda, R., Carlier, G., and Carlotto, V., 1993, Petrogenesis and occurrences
of gabbroic rocks in the limit Eastern Cordillera-high plateau in the Abancay deflection area (Curahuasi-south Peru): Second International Symposium on Andean Geodynamics, Oxford, United Kingdom, Extended Abstracts volume, p. 393397.
Lowell, J.D., and Guilbert, J.M., 1970, Lateral and vertical alteration-mineralization zoning in porphyry ore deposits: ECONOMIC GEOLOGY, v. 65, p.
373408.
Maksaev, V., 1990, Metallogeny, geological evolution and thermochronology
of the Chilean Andes between latitutes 21 and 26 South, and the origin
of major porphyry copper deposits: Unpublished Ph.D. thesis, Halifax,
Nova Scotia, Canada, Dalhousie University, 544p.
Maksaev, V., and Zentilli, M., 1988, Marco metalognico regional de los
megadepsitos de tipo prfido cuprfero del norte grande de Chile: V
Congreso Geolgico Chileno, Actas, v. 1, p. B181B212.
1999, Fission track thermochronology of the Domeyko Cordillera, northern Chile: Implications for Andean tectonics and porphyry copper metallogenesis: Exploration and Mining Geology, v. 8, nos. 1 and 2, p. 6589.

0361-0128/98/000/000-00 $6.00

1603

Marocco, R., 1975, Geologa de los cuadrngulos de Andahuaylas, Abancay y


Cotabambas: Lima, Instituto de Geologa, Minera y Metalurgia, no. 27, 51
p.
1978, Un segment EW de la cordillre des Andes Pruviennes: La
dflexion dAbancay. Etude gologique de la Cordillre Orientale et des
Hauts-plateaux entre Cuzco et San Miguel (Sud du Prou): Paris, France,
Travaux et Documents de lORSTOM, no. 94, 195 p.
Marocco, R., and Noblet, C., 1990, Sedimentation, tectonism and volcanism
relationships in two Andean basins of southern Peru: Geologische Rundschau, v. 79, p. 111120.
Mathur, R., Ruiz, J., and Munizaga, F., 2001, Insights into Andean metallogenesis from the perspective of Re-Os analyses of sulfides: III South American Symposium on Isotope Geology, Sociedad Geolgica de Chile, Santiago, Chile, Extended Abstracts Volume, p. 500503. (CD-ROM)
Mgard, F., 1987, Cordilleran and marginal Andes: A review of Andean geology north of the Arica elbow (18S), in Monger, J.W.H., and Franheteau,
J., eds., Circum-Pacific Orogenic Belts and Evolution of the Pacific Ocean
Basin: American Geophysical Union, Geodynamic Series, v.18, p. 7195.
Mgard, F., Noble, D.C., McKee, E.H., and Bellon, H., 1984, Multiple
pulses of Neogene compressive deformation in the Ayacucho intermontane
basin, Andes of central Peru: Geological Society of America Bulletin, v. 95,
p. 11081117.
Mendvil, S., and Dvila, D., 1994, Geologa de los cuadrngulos de Cuzco y
Livitaca (Hojas 28-S y 29-S): Lima, Instituto Geolgico, Minero y
Metalrgico, Carta Geolgica Nacional, Boletn no. 52, 115 p.
Meyer, C., and Hemley, J.J., 1967, Wall rock alteration. in Barnes, H.L., ed.,
Geochemistry of Hydrothermal Ore Deposits: New York, Holt, Rinehart,
and Winston, p. 166235.
MMAJ (Metal Mining Agency of Japan), 1983, Peru-Cuzco region: Katanga
area: Overseas Geologic Investigation Bulletin, 21 p. (in Japanese)
Mpodozis, C., and Perell, J., 2003, Porphyry copper metallogeny of the middle Eocene-early Oligocene arc of western South America: relationships
with volcanism and arc segmentation [abs.]: X Congreso Geolgico
Chileno, Concepcin, Chile, October 610, 2003, CD-ROM.
Mpodozis, C., Arriagada, C., and Roperch, P., 1999, Cretaceous to Paleogene
geology of the Salar de Atacama basin, northern Chile: A reapprisal of the
Purilactis Group stratigraphy: Fourth International Symposium on Andean
Geodynamics, Goettingen, Germany, October 46, 1999, Extended Abstracts volume, p. 523526.
Muntean, J.L., and Einaudi, M.T., 2000, Porphyry gold deposits of the Refugio district, Maricunga belt, northern Chile: ECONOMIC GEOLOGY, v. 95, p.
14451472.
Noble, D.C., and McKee, E.H., 1999, The Miocene metallogenic belt of central and northern Per, in Skinner, B.J., ed., Geology and Ore Deposits of
the Central Andes: Society of Economic Geologists Special Publication 7,
p.155193.
Noble, D.C., McKee, E.H., Farrar, E., and Petersen, U., 1974, Epidosic volcanism and tectonism in the Andes of Peru: Earth Planetary Science Letters, v. 21, p. 213221.
Noble, D.C., McKee, E.H., and Mgard, F., 1979, Early Tertiary Incaic
tectonism, uplift, and volcanic activity, Andes of central Peru: Geological
Society 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: ECONOMIC GEOLOGY, v. 79, p. 172176.
Noblet, C., Marocco, R., and Delfaud, J., 1987, Analyse sdimentologique
des Couches Rouges du bassin intramontagneux de Sicuani (Sud du
Prou): Bulletin de lInstitut Franais dEtudes Andines, Lima, v. 16, p.
5578.
Parada, M.A, Rivano, S., Seplveda, P., Herv, M., Herv, F., Puig, A., Munizaga, F., Brook, M., Pankhurst, R., and Snelling, N., 1988, Mesozoic and
Cenozoic plutonic development in the Andes of central Chile (3030
3230S): Journal of South American Earth Sciences, v. 1, p. 249260.
Pardo Casas, F., and Molnar, P., 1987, Relative motion of the Nazca (Farallon) and South America plates since Late Cretaceous time: Tectonics, v. 6,
p. 233248.
Pecho, V., 1981, Geologa de los cuadrngulos de Chalhuanca, Antabamba y
Santo Toms: Lima, Instituto de Geologa, Minera y Metalurgia, Boletn
no. 35, 67p.
Perell, J., Urza, F., Cabello, J., and Ortiz, F., 1996, Clustered, gold-bearing
Oligocene porphyry copper and associated epithermal mineralization at La
Fortuna, Vallenar region, northern Chile, in Camus, F., Sillitoe, R.H., and

1603

1604

PERELL ET AL.

Petersen, R., eds., Andean Copper deposits: New Discoveries, Mineralization, Styles and Metallogeny: Society of Economic Geologists Special Publication 5, p. 8190.
Perell, J., Rojas, N., Devaux, C., Fava, L., Etchart, E., and Harman, P.,
1998, Discovery of the Agua Rica porphyry Cu-Mo-Au deposit, Catamarca
province, northwestern Argentina, Part II: Geology: Australian Mineral
Foundation Conference, Perth, Western Australia, November 30December 1, 1998, Proceedings, p. 117132.
Perell, J., Neyra, C., Zrate, A., Posso, H., Ramos, P., Martini, R., Fuster, N.,
Muhr, R., and Caballero, A., 2002, Porphyry Cu-Au mineralization at
Cotabambas, Cuzco region, Peru [abs.]: XI Congreso Peruano de Geologa,
Sociedad Geolgica del Per, Lima, Per, September 2528, 2002,
Abstracts, p. 242.
Petersen, C.R., Rivera, S.L., and Peri, M.A., 1996, Chimborazo copper deposit, Region II, Chile: Exploration and geology, in Camus, F., Sillitoe,
R.H., and Petersen, R., eds., Andean Copper Deposits: New Discoveries,
Mineralization, Styles and Metallogeny: Society of Economic Geologists
Special Publication 5, p. 7180.
Petersen, U., 1999, Magmatic and metallogenic evolution of the Central
Andes: Society of Economic Geologists Special Publications, v. 7, p. 109153.
Portugal, J.A., 1974, Mesozoic and Cenozoic stratigraphy of Puno, Peru:
American Association of Petroleum Geologists Bulletin, v. 58, p. 982999.
Ramos, V.A., and Aleman, A. 2000, Tectonic evolution of the Andes, in Cordani, U.G., Milani, E.J., Thomaz Filho, A., and Campos, D.A., eds., Tectonic Evolution of South America, 31st International Geological Congress,
Rio de Janeiro, Brazil, August 617, 2000, p. 635685.
Ramos, V.A., Cristallini, E.O., and Prez, D.J., 2002, The Pampean flat-slab
of the Central Andes: Journal of South American Earth Sciences, v. 15, p.
5978.
Richards, J.P., Noble, S.R., and Pringle, M.S., 1999, A revised late Eocene
age for porphyry copper magmatism in the Escondida area, northern Chile:
ECONOMIC GEOLOGY, v. 94, p. 12311247.
Rojas, N., Drobe, J., Lane, R., and Bonafede, D., 1999, Descubrimiento y
geologa del prfido cuprfero Taca Taca de Abajo, Provincia de Salta,
Argentina [abs.]: XIV Congreso Geolgico Argentino, Salta, Argentina,
September 1924, 1999, Actas, v. I, p. 117.
Romero, D., Jaimes, F., Carlotto, V., Marocco, R., and Crdenas, J., 1997,
Evolucin sedimentolgica y paleogeogrfica de la cuenca Paruro:
Mioceno superior (Regin de Cusco): IX Congreso Peruano de Geologa,
Sociedad Geolgica del Per, Extended Abstracts, Volumen Especial 1, p.
387391.
Sandeman, H.A., Clark, A.H., and Farrar, E., 1995, An integrated tectonomagmatic model for the evolution of the southern Peruvian Andes (1320S) since 55 Ma: International Geology Review, v. 37, p. 10391073.
Santa Cruz, S., Guerrero, T., Castilla, F., Caro, E., and Candiotti, H., 1979,
Geologa de yacimientos de cobre en skarn en la regin sur-oriental del
Per: Boletn Sociedad Geolgica del Per, v. 60, p. 153174.
Sasso, A.M., and Clark, A.H., 1998, Magmatic, hydrothermal and tectonic
evolution and implications for Cu-Au metallogeny in the Andean backarc: Society of Economic Geologists Newsletter, no. 34, July, 1998, p. 1,
818.
Sbrier, M., and Soler, P., 1991, Tectonics and magmatism in the Peruvian
Andes from Late Oligocene time to Present, in Harmon, R.S., and Rapela,
C.W., eds., Andean Magmatism and its Tectonic Setting: Geological Society of America, Special Paper 265, p. 259278.
Sbrier, M., Lavenu, A., Fornari, M., and Soulas, J., 1988, Tectonics and uplift in the Central Andes (Peru, Bolivia and northern Chile) from Eocene
to present: Godynamique, v. 3, p. 139161.
Sempere, T., Butler, R.F., Richards, D.F., Marshall, L.G., Sharp, W., and
Swisher III, C.C., 1997, Stratigraphy and chronology of Upper CretaceousLower Paleogene strata in Bolivia and northwest Argentina: Geological Society of America Bulletin, v. 109, no. 6, p. 709727.
Sempere, T., Jacay, J., Fornari, M., Roperch, P., Acosta, H., Bedoya, C.,
Cerpa, L., Flores, A., Husson, L., Ibarra, I., Latorre, O., Mamani, M.,
Meza, P., Odonne, F., Ors, Y., Pino, A., and Rodrguez, R., 2002, Lithospheric-scale transcurrent fault systems in Andean southern Peru: Fifth International Symposium on Andean Geodynamics, Toulouse, France, September 1618, 2002, Extended Abstracts volume, p. 610608.
Sillitoe, R.H., 1976, Andean mineralization: A model for the metallogeny of
convergent plate margins, in Strong, D.F., ed., Metallogeny and Plate Tectonics: Geological Association of Canada Special Paper 14, p. 59100.
1988, Epochs of intrusion-related copper mineralization in the Andes:
Journal of South American Earth Sciences, v. 1, p. 89108.

0361-0128/98/000/000-00 $6.00

1990, Copper deposits and Andean evolution, in Ericksen, G.E.,


Caas-Pinochet, M.T., and Reinemund, J.A., eds., Geology of the Andes
and its Relation to Hydrocarbon and Mineral Resources: Circum-Pacific
Council for Energy and Mineral Resources, Earth Science Series, v. 11,
p. 285311.
1997, Characteristics and controls of the largest porphyry copper-gold
and epithermal gold deposits in the circum-Pacific region: Australian Journal of Earth Sciences, v. 44, p. 373388.
1998, Major regional factors favouring large size, high hypogene grade,
elevated gold content and supergene oxidation and enrichment of porphyry
copper deposits: Australian Mineral Foundation Conference, Perth, Western Australia, November 30-December 1, 1998, Proceedings, p. 2134.
2000, Gold-rich porphyry deposits: Descriptive and genetic models and
their role in exploration and discovery: Reviews in ECONOMIC GEOLOGY, v.
13, p. 315345.
Sillitoe, R.H., and Bonham, H.F., 1990, Sediment-hosted gold deposits: Distal products of magmatic-hydrothermal systems: Geology, v. 18, p. 157161.
Sillitoe, R.H., and Gappe, I.M., Jr., 1984, Philippine porphyry copper deposits: Geologic setting and characteristics: United Nations Economic and
Social Commission Asia-Pacific, Committee Coordination Joint Prospecting Mineral Resources Asian Offshore Areas Technical Publication 14, 89p.
Sillitoe, R.H., and McKee, E.H., 1996, Age of supergene oxidation and enrichment in the Chilean porphyry copper province: ECONOMIC GEOLOGY,
v. 91, p. 164179.
Skewes, M.A., and Stern, C.R., 1994, Tectonic trigger for the formation of
Late Miocene Cu-rich breccia pipes in the Andes of central Chile: Geology,
v. 22, p. 551554.
1995, Genesis of the giant Late Miocene to Pliocene copper deposits of
central Chile in the context of Andean magmatic and tectonic evolution: International Geology Review, v. 37, p. 893909.
Soler, P., and Bonhomme, M.G., 1988, New K-Ar age determinations of intrusive rocks from the Cordillera Occidental and Altiplano of central Peru:
Identification of magmatic pulses and episodes of mineralization: Journal of
South American Earth Sciences, v. 1, no.2, p. 169177.
1990, Relation of magmatic activity to plate dynamics in central Peru
from Late Cretaceous to present, in Kay, S.M., and Rapela, C.W., eds., Plutonism from Antartica to Alaska: Geological Society of America Special
Paper 241, p. 173192.
Steiger, R.H., and Jaeger, E., 1977, Subcommission on geochronology convention on the use of decay constants in geo- and cosmochronology: Earth
and Planetary Science Letters, v. 36, no. 3, p. 359362.
Streckeisen, A., 1976, To each plutonic rock its proper name: Earth Science
Reviews, v. 12, p. 133.
1978, IUGS subcommision on the systematics of igneous reocks. Clasification and nomenclature of volcanic rocks, lamprophyres, carbonatites,
and melilite rocks. Recommendations and suggestions: Neues Jarbuch fr
Mineralogie, v. 143, p. 114.
Terrones, A.J., 1958, Structural control of contact metasomatic deposits in
the Peruvian cordillera: American Institute Mining Metallurgy Transactions, v. 90, section B, p. B15B36.
Titley, S.R., 1990, Contrasting metallogenesis and regional settings of Circum-Pacific Cu-Au porphyry systems: Pacific Rim Congress 90, Gold
Coast, Queensland, 1990, Melbourne, Autralasian Institute of Mining and
Metallurgy, Proceedings, v. 11, p. 127133.
Tomlinson, A.J., Cornejo, P., and Mpodozis, C., 1999, Hoja Potrerillos,
Regin de Atacama: Servicio Nacional de Geologa y Minera, Subdireccin
Nacional de Geologa, Mapas Geolgicos, no. 14, scale 1: 100,000.
Travisany, V., 1979, Consideraciones genticas sobre el yacimiento
estratiforme San Bartolo: II Congreso Geolgico Chileno, Actas, v. 1, p.
C149C159.
Vicente, J.C., Beaudouin, B., Chvez, A., and Len, I., 1982, La cuenca de
Arequipa (Sur Per) durante el Jursico-Cretcico inferior: Congreso
Latinoamericano de Geologa, 5th, Buenos Aires, Argentina, Actas, v. 1, p.
121153.
Vila, T., and Sillitoe, R.H., 1991, Gold-rich porphyry systems in the Maricunga belt, northern Chile: ECONOMIC GEOLOGY, v. 86, p. 12711286.
Wasteneys, H.A.H.P., 1990, Epithermal silver mineralization associated with
a mid-Tertiary diatreme: Santa Brbara, Santa Luca district, Puno, Peru:
Unpublished Ph.D. thesis, Kingston, Ontario, Canada, Queens University,
367 p.
Yoshikawa, S., Sakai, S., and Sato, H., 1976, Discovery of Quechua deposit
and its characteristics: Mining Geology, v. 26, p. 143152. (in Japanese with
English abstract)

1604

PORPHYRY-STYLE ALTERATION AND MINERALIZATION, ANDAHUAYLAS-YAURI BELT, PERU


Zappettini, E., Miranda-Angles, V., Rodrguez, C., Palacios, O., Cocking, R.,
Godeas, M., Uribe-Zeballos, H., Vivallo, W., Paz, M.M., Seggiaro, R.,
Heushmidt, B., Gardeweg, M., Boulangger, E., Korzeniewski, L.,
Mpodozis, C., Carpio, M., and Rubiolo, D., 2001, Mapa metalognico de la
regin fronteriza entre Argentina, Bolivia, Chile y Per (14S-28S):
Santiago, Chile, Servicio Nacional de Geologa y Minera, Publicacin
Geolgica Multinacional, no. 2, 222 p.

0361-0128/98/000/000-00 $6.00

1605

Zweng, P.L., Yagua, J., Fierro, J., Gamarra, H., Jordn, L., Brooks, J., Yurko,
E., and Mulhollen, R., 1997, The Cu- (Au-Ag) skarn deposits at Tintaya,
Peru: IX Congreso Peruano de Geologa, Sociedad Geolgica del Per,
Lima, Peru, August 1997, Extended Abstracts, Volumen Especial 1, p.
237242.

1605

0361-0128/98/000/000-00 $6.00

1606

You might also like