You are on page 1of 38

Ore Geology Reviews 65 (2015) 327364

Contents lists available at ScienceDirect

Ore Geology Reviews


journal homepage: www.elsevier.com/locate/oregeorev

Review

Physiographic and tectonic settings of high-suldation epithermal


goldsilver deposits of the Andes and their controls on
mineralizing processes
Thomas Bissig a,, Alan H. Clark b, Amelia Rainbow b, Allan Montgomery c
a
b
c

Mineral Deposit Research Unit, University of British Columbia, 2020-2207 Main Mall, Vancouver, BC V6T 1Z4, Canada
Queen's University, Bruce Wing/Miller Hall, Kingston, ON K7L 3N6, Canada
Riverside Resources Suite 1110, 1111 West Georgia Street, Vancouver, BC V6E 4M3, Canada

a r t i c l e

i n f o

Article history:
Received 6 July 2014
Received in revised form 10 September 2014
Accepted 17 September 2014
Available online 28 September 2014
Keywords:
High-suldation
Epithermal
Andes
Landscape evolution
Erosion
Uplift
Flat subduction
Neogene

Corresponding author.
E-mail address: tbissig@eos.ubc.ca (T. Bissig).

http://dx.doi.org/10.1016/j.oregeorev.2014.09.027
0169-1368/ 2014 Elsevier B.V. All rights reserved.

a b s t r a c t
Gold and silver ores in the vast majority of Andean high-suldation epithermal AuAg deposits occur at high
present day elevations and typically 200500 m below low relief landforms situated at 3500 to 5200 m a.s.l.
Most deposits are middle Miocene and younger and include, El Indio, Tambo, PascuaLama, Veladero (El Indio
belt, Chile/Argentina), Cerro de Pasco (Central Peru), Pierina, Lagunas Norte, Yanacocha (northern Peru),
Quimsacocha (Ecuador), and the CaliforniaVetas mining district (Santander, Colombia), jointly accounting for
N 130 Moz Au resources. Slightly older examples are only preserved in the Atacama Desert and include the middle
Eocene El Guanaco and El Hueso and the late Oligocene/early Miocene La Coipa deposits. The absence of Paleocene and older high-suldation epithermal deposits can be explained by limited preservation potential imposed
by transpressional tectonics within overall contractile episodes and surface uplift. These conditions prevailed predominantly in segments of shallow-angle subduction of the Nazca or Caribbean plate below the South American
continent, a tectonic setting also common for porphyry-style Cu (Au, Mo) deposits. Stratovolcanoes are uncommon ore hosts and volcanic rocks coincident with mineralization are in most cases volumetrically restricted or
absent, recording the terminal stages of local arc magmatism. However, dacitic domes are important at,
e.g., Yanacocha and La Coipa. At Lagunas Norte, a small stratovolcano largely pre-dating but temporally overlapping with mineralization occurs immediately east of the deposit and volcanic sector collapse may have occurred
during hydrothermal activity.
Mineralization is typically located near the backscarp of pediments or the heads of valleys incising now highelevation, low-relief surfaces. In the CaliforniaVetas Mining District and El Indio belt, hydrothermal alunite
ages become generally younger upstream along the incising valleys, indicating that hydrothermal activity and,
by inference, ore deposition were facilitated by erosion. The lowering of the water table and reduction of hydrostatic and lithostatic pressure at these sites of high local relief are believed to have enhanced both boiling and
mixing of magmatic with meteoric uids, ultimately enhancing ore deposition.
The host rock composition, permeability and location of the water table control the distribution of alteration
zones and ore. Intermediate volcanic rocks are the most common ore-hosts but they typically pre-date mineralization by several Ma. However, high-suldation epithermal mineralization can be hosted in any conceivable
rock type including high grade metamorphic rocks (CaliforniaVetas mining district), signicantly older plutonic
rocks (PascuaLama) or quartzites (Lagunas Norte). Large vuggy quartz alteration zones and commonly oxidized
low-grade large-tonnage mineralization are best developed in relatively permeable volcaniclastic rocks or hydrothermal breccia bodies, whereas coherent volcanic, plutonic, or metamorphic rocks may host fault- and brecciacontrolled ores. The near-surface steam-heated zone can attain a thickness of several hundred meters in dry
climates (e.g. Veladero, PascuaLama, Tambo) but is typically poorly developed and less than 20 m thick in
humid climatic zones.
The physiographic and tectonic settings of high-suldation epithermal deposits are distinct from low-suldation
epithermal districts such as those of Patagonia, El Pen (Chile) or Fruta del Norte (Ecuador). The latter range to
signicantly older ages (Jurassic to early Eocene) occur at mainly lower elevations and were emplaced in extensional settings. A temporal coincidence between uplift, erosion and mineralizing processes as well as a spatial and
temporal association with porphyry style mineralization is not evident for these low-suldation districts.
2014 Elsevier B.V. All rights reserved.

328

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Contents
1.
2.
3.
4.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Epithermal deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Distribution of high-suldation deposits in the Andes . . . . . . . . . . . . . . . . . . . . . . . . . . . .
El Indio belt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.1.
The Tambo deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.
El Indio deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3.
PascuaLama . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.4.
Veladero . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.
Maricunga belt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1.
La Coipa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2.
La Pepa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.
High-suldation epithermal deposits of the Domeyko fault system . . . . . . . . . . . . . . . . . . . . . .
6.1.
El Hueso . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2.
El Guanaco . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
7.
Late Miocene high-suldation epithermal Au deposits of the western Cordillera of northern Chile and southern Peru
8.
Central to northern Peruvian at slab segment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.
High-suldation epithermal deposits of the central Peruvian polymetallic belt . . . . . . . . . . . . . . . . .
9.1.
Julcani . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.2.
The MarcapuntaColquijirca district . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.3.
Cerro de Pasco . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
9.4.
Quicay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10.
The high-suldation epithermal Au (Cu, Ag) deposits of northwestern Peru . . . . . . . . . . . . . . . . . .
10.1.
Pierina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10.2.
Lagunas Norte . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10.3.
Yanacocha . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
10.4.
Tantahuatay, Sipan and La Zanja . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
11.
The northern Andes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
11.1.
Quimsacocha . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
11.2.
California Vetas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
12.
Summary and comparison to low-suldation deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.
Controls of geomorphic processes and climate on mineralization . . . . . . . . . . . . . . . . . . . . . . .
14.
Igneous rocks, volcanology and magmatic uids related to high-suldation epithermal deposits . . . . . . . . .
15.
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
The Andes are the world's most endowed region with respect to
giant magmatic-hydrothermal ore deposits (Cooke et al., 2005). They
host the largest-known porphyry copper deposits (e.g., Rio BlancoLos
BroncesLos Sulfatos, El Teniente, Chuquicamata) as well as many of
the world's largest epithermal AuAg deposits (e.g., Yanacocha, Lagunas
Norte, PascuaLama, Veladero: Sillitoe, 2008). The vast majority of
Andean epithermal deposits containing N 10 Moz Au are of highsuldation type. These deposits have a close link to a magmatic source
for uids, volatiles and metals (e.g., Deyell et al., 2004; Rye, 1993) but
form at depths of typically less than 1 km (e.g. Sillitoe, 2010) and consequently mineralizing processes are inuenced by the near-surface physicochemical environment. The main focus of this review is on deposits
and districts where the bulk of the precious metal is contained in
the epithermal environment, i.e., the shallow part of magmatichydrothermal systems, and concentrates on the physiographic environment of epithermal mineralization. This paper does not discuss major
porphyry Cu deposits in detail, although the shallow portions of many
of these have been overprinted by epithermal mineralization or alteration (e.g., Masterman et al., 2004; Ossandn et al., 2001). Similarly,
the deposits hosting Sn, W, Ag and Au ores in the eastern Cordillera of
Bolivia and Peru are not discussed. Following a general summary of
epithermal deposit types and their terminology, this article presents a
comprehensive overview of the major high-suldation epithermal districts and mineral belts of the Andes. It focuses on the links between
landscape evolution, climatic setting, volcanology and tectonics, and

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

328
328
329
330
331
331
331
337
338
340
343
343
343
344
344
345
345
345
345
345
346
346
346
346
348
351
351
353
353
355
357
360
360
361
361

discusses the inuence these factors can have on both mineralizing


processes and the preservation of the deposits.

2. Epithermal deposits
Epithermal deposits are usually classied into sub-types based on either ore sulde assemblage or characteristic associated alteration; both
schemes have inherent limitations. Some of the most widely referenced
review papers on the topic (Hedenquist et al., 2000; Sillitoe and
Hedenquist, 2003; Simmons et al., 2005) prefer a classication into
high, low and intermediate-suldation types. This classication scheme
can, however, be problematic, because sulde assemblages may be difcult to classify in a eld exploration setting, particularly if the deposit
has been oxidized. Moreover, sulde assemblages within a single deposit may represent precipitation over the entire breadth of suldation
state, from high to intermediate and low-suldation, depending on
uidwall rock interaction (e.g., El Indio: Heather et al., 2003a, 2003b;
Cerro de Pasco: Baumgartner et al., 2008; Lagunas Norte: Cerpa et al.,
2013). Alternative classication is based on dominant alteration
and gangue assemblages and includes quartzadulariasericite and
quartzalunite or acid sulfate type epithermal deposits (Heald et al.
1987, Tosdal et al., 2009), the former typically including low to
intermediate-suldation sulde assemblages and the latter associated
with high-suldation deposits. The limitation of this classication
scheme is that, particularly in high-suldation deposits, alteration may
pre-date and may not be directly related to mineralization (see below).

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

For this paper we use the widely applied classication scheme of


high-suldation (normally associated with acidic quartzalunite
alteration) and low-suldation type (associated with near neutral
quartzadularia gangue and illiteadularia alteration) deposits. The
term intermediate suldation is used to describe suldation state but
is not used in the sense of a distinct subtype of epithermal deposit due
to the inherent variation of sulde assemblages within a single deposit
or vein.
In most high-suldation epithermal deposits, two hydrothermal
stages can be distinguished (e.g., Simmons et al., 2005). An early stage
of intense acid leaching of the wall-rock by a magmatic vapor that condensed into groundwater, cooled to temperatures below 300 C and
was acidied by dissociation of acids such as HCl, or H2SO4
(Hedenquist and Taran, 2013); and, a subsequent stage of weaklyacidic uid incursion, from which the bulk of the suldes and precious
metals, together with euhedral quartz, precipitated (e.g., Heinrich
et al., 2004, 2007; Holley, 2012; Stoffregen, 1987). The intense acid
leaching generates the zoning from vuggy quartz to quartzalunite
and more distal kaolinitic alteration. This assemblage is commonly referred to as a lithocap because it caps potential porphyry Cu mineralization (Sillitoe, 1995). Although mineralization is widely related to the
inltration of the less acidic uids into the precursor lithocap
(e.g., Holley, 2012; Stoffregen, 1987), it has been shown that the acidic
magmatic vapors can also transport gold into the shallow epithermal
environment (Scher et al., 2013; Taran et al., 2000). At PascuaLama
such uids were responsible for at least some of the gold mineralization
(Chouinard et al., 2005). Conversely, there are examples of highsuldation epithermal deposits where vuggy quartz alteration is largely
absent (e.g., La Bodega, Colombia: Rodriguez, 2014).
The precipitation mechanisms for precious metals in highsuldation epithermal deposits are not as well understood as for lowsuldation systems. For the latter, ample textural and uid inclusion evidence leaves little doubt that uid boiling leads to precipitation of
metals due to the loss of the dominant ligands (H2S) to the gas phase
(Moncada et al., 2012; Simmons et al., 2005). In contrast, uid inclusion
studies of high-suldation deposits are difcult, as the appropriate
transparent minerals are rare. Several possible precipitation mechanisms such as boiling, and uid mixing of a magma derived uid with
oxidized groundwater have been proposed (e.g., Rye et al., 1992;
Jannas et al., 1999). Recent work on the La Bodega deposit in
Colombia, where the high-suldation sulde assemblage is hosted in
veins with similar textures as described for low-suldation deposits
(see below), supports that boiling is a viable mechanism to precipitate
suldes (Rodriguez, 2014). Moreover, steam-heated alteration, present
above many Andean high-suldation systems, forms by H2S-rich steam
that condenses and oxidizes to form sulfuric acid in the vadose zone,
and is considered evidence for uid boiling at depth (Simmons et al.,
2005).
A close association of high-suldation epithermal deposits with a
parent intrusion and porphyry-style mineralization has been proposed
by Hedenquist and Loewenstern (1994) and documented in detail
for the LepantoFar Southeast porphyry system in the Philippines
(Hedenquist et al., 1998). In earlier literature, stratovolcanoes, calderas
and extrusive or subvolcanic dacitic to andesitic domes have been
considered the dominant hosts, and large volcanic centers have been
presumed to be temporally and genetically associated with highsuldation epithermal deposits (e.g. Hayba et al. 1985, Arribas, 1995).
Indeed, stable S, H and O isotopic evidence (e.g., Deyell et al., 2005a,
2005b; Rye, 1993, 2005) as well as thermodynamic considerations
(Heinrich, 2005, 2007) clearly indicate that magma-derived uids or
condensed magmatic vapors constitute the dominant component of
the mineralizing uids. However, as will be discussed below, better geological understanding and new geochronological information, together
with exploration discoveries made since 1995, suggest that stratovolcanoes and calderas are uncommon hosts to epithermal ore in the
Andes.

329

3. Distribution of high-suldation deposits in the Andes


High-suldation epithermal deposits occur mostly in the Central
Andes (~ 34 to 3 Lat. S) with few examples in the Northern Andes
(~3 Lat. S to 11.3 Lat. N) and no known example in the southern volcanic zone of the Central Andes (south of 34 Lat. S) and the southern,
a.k.a. Patagonian Andes, south of 46 Lat. S (Fig. 1). The overwhelming
majority of high-suldation epithermal deposits are Miocene or younger, and are located in segments of the Andean arc where magmatism
is now absent or subdued (e.g., Bissig et al., 2008; Kay and Mpodozis,
2001; Rosenbaum et al., 2005). The lack of magmatism is attributed to
two zones of shallow angle subduction of the Nazca plate below the
South American continent in the Central Andes (the Pampean at slab
in northern Chile and the Peruvian at slab in central and northern
Peru: Ramos, 2009) as well as the Bucaramanga at slab where Caribbean plate is subducting below northern Colombia (Vargas and Mann,
2013). Flat subduction coincides with shortening and crustal thickening
of the overriding continental plate (Martinod et al., 2010) and has been
spatially linked to the subduction of aseismic ridges or oceanic plateaus
on the oceanic plate (e.g. Gutscher et al., 1999b; Yaez et al., 2001).
However the causative relationship between ridge impingement and
onset of at subduction remains a matter of debate (Skinner and
Clayton, 2013). Thus, crustal shortening, uplift and porphyry Cu
mineralization pre-dates the arrival of the subducting aseismic Juan
Fernndez ridge in Central Chile (Deckart et al., 2013), a situation
comparable to northern Peru where high-suldation epithermal gold
mineralization at Lagunas Norte occurred at 17 Ma, coincident with uplift, but several Ma prior to the inferred onset of Nazca ridge subduction
(Montgomery, 2012, see below).
The Northern Andes are characterized by a series of oceanic terranes
accreted to the South American continent dominantly between the Mesozoic and Miocene (Cediel et al., 2003). The Central Andes, in contrast,
consist of a collage of terranes accreted during the Proterozoic-toPaleozoic (Ramos, 2008). Here, Andean type subduction has dominated
the active continental margin since Permo-Triassic times (Ramos,
2009), and in both the Northern Andes south of ~ 5.5 Lat. N and the
Central Andes, the Nazca plate is currently subducting eastward
beneath the South American continent.
High-suldation epithermal deposits occur in the same segments of
the Andean arc as porphyry-style deposits (e.g. Sillitoe, 2008, 2010).
However, a direct temporal, spatial and genetic relationship to porphyry
deposits containing economic Cu mineralization, as demonstrated at Far
Southeast and Lepanto (Hedenquist et al., 1998), has not been documented for any Andean high-suldation epithermal deposits discussed
herein, although the two mineralization styles may overlap either
spatially or temporally (e.g. Yanacocha district: Longo et al., 2010, La
Pepa, Maricunga belt: Muntean and Einaudi, 2001).
All high-suldation epithermal deposits of the Andes now
crop out at high elevation, and with few exceptions, above
~ 3500 m a.s.l. Most are also located near the water-shed along the
crest of the orogen, or what appears to have been the water-shed
at the time of hydrothermal activity, in areas dominated by relatively
at landforms which are interpreted as uplifted sub-planar
paleosurfaces. In the semi-arid and arid portions of the Andes,
these paleosurfaces probably represent pediplains (regionally interconnected pediment surfaces) formed under semi-arid conditions
and several pulses of pediment erosion separated by uplift pulses
can commonly be distinguished (Bissig and Riquelme, 2009; Bissig
et al., 2002a; Riquelme et al., 2007; Rodriguez et al., 2014; Sillitoe
et al., 1968), resulting in a series of pediment surfaces, vertically separated up to several 100 m by steeper back-scarps. Similar relationships between high-level sub-planar landforms incised by broad,
at bottomed valleys are present in southern Peru (Quang et al.,
2005; Tosdal et al., 1984) and northern Peru (Montgomery, 2012).
The major high-suldation deposits and districts are described within their geomorphologic and volcanologic context from south to north

330

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

70 W

80 W

90 W

60 W

50 W

epithermal deposits

20 N

10 N

23/24
CoR

des

n An

CaR

22 Norther al Andes
ntr
Ce

IP
21 20 19
18
17
16 15
14
13
12

10 S

1: Tambo
2: El Indio
3: Pascua-Lama
4: Veladero
5: Famatina
6: La Coipa
7: El Hueso
8: El Guanaco
9: Choquelimpie
10: Santa Rosa
11: Tucari
12: Julcani
13: Marcapunta/Colquijirca
14: Cerro de Pasco
15: Quicay
16: Pierina
17: Lagunas Norte
18: Yanacocha
19: Tantahuatay
20: Sipn
21: La Zanja
22: Quimsacocha
23: La Bodega
24: Angostura
Active volcano
Flat slab segments
Aseismic ridges
Subducted aseismic
ridges/plateaus

11
10
9
10 S

NR

IR
30 S

7
6
4 35
2
1

JFR

Fig. 1. Map of South America, showing locations of major high-suldation epithermal deposits discussed in this paper. Also indicated are segments of at subduction (dashed black lines),
volcanoes, aseismic ridges and inferred subducted portions of aseismic ridges and oceanic plateaus. The boundary between the northern and central Andes is indicated. Abbreviations:
CoR: Cocos Ridge; CaR: Carnegie Ridge; IP: Inca Plateau; NR: Nazca Ridge; IR: Iquique Ridge; JFR: Juan Fernandez Ridge.

in the following section. Key characteristics, locations, ages and approximate contained resources are summarized in Table 1.
4. El Indio belt
The El Indio belt straddles the ChileArgentina Border between 29
and 30 Lat. S. which corresponds to the middle of the Pampean at
slab segment. The belt is named after the El Indio Au (Ag, Cu) deposit
which was mined between 1978 and 2001. The belt also includes the
small Tambo and the giant Veladero and PascuaLama deposits as
well as other prospects (Fig. 2), jointly accounting for ~ 40 Moz of Au
resources. Additionally, the belt contains numerous alteration zones
ranging in age from Eocene to late Miocene but, with the exception of
Veladero (12.710.3 Ma; Holley, 2012), economic mineralization is

only known to have occurred between 9.5 and 5 Ma (Bissig et al.,


2001; A.H. Clark, unpubl. data). Alteration and mineralization ages quoted herein for the El Indio belt are all 40Ar/39Ar plateau ages on alunite or
sericite mostly with analytical errors of b1 to 5% on individual dates. It is
worth noting that the small epithermal CuAu district at Famantina
(Table 1), is located in the same Andean segment, albeit some 220 km
E of Veladero in Argentina, and is at 5 to 3.8 Ma just slightly younger
than El Indio (Lozada-Caldern et al., 1994; Pudack et al., 2009).
Volcanism in the Oligocene and early Miocene was voluminous in
the El Indio belt (Bissig et al., 2001; Martin et al., 1995; Winocur et al.,
2014) initially generating the 2723 Ma Tilito Formation dacitic-toandesitic pyroclastic and volcaniclastic rocks and small volumes of
basalts in an extensional setting (Winocur et al., 2014). The Tilito
Formation is unconformably overlain by the dominantly-andesitic

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Escabroso (2117 Ma) and Cerro de las Trtolas (16.614 Ma) Formations (Fig. 2), both associated with large volcanic edices and
subvolcanic intrusions (Bissig et al., 2001; Martin et al., 1995). Erupted
magma volumes decreased markedly after ca. 14 Ma and volcanism
was conned to isolated eruptive centers at ~ 12.711 Ma (Vacas
Heladas formation andesite to dacite) and the rhyodacitic Pascua
(87.6 Ma) and Vallecito (65.5 Ma) Formations (Bissig et al., 2001).
An upper Pliocene rhyolite dome (Cerro de Vidrio) has been documented ~7 km E of the Veladero deposit (Bissig et al., 2001, 2002a, 2002b).
The voluminous upper Oligocene to middle Miocene andesites show
an increasing modal abundance of phenocrystic hornblende, relative to
clinopyroxene, with time (Bissig et al., 2003). Further, igneous rocks
younger than ca. 14 Ma are characterized by elevated Sr/Y (N40) and
depleted HREE contents, indicative of garnet fractionation at the base
of the crust and elevated pressure equivalent to N 45 km crustal thickness (Bissig et al., 2003). This change in both the erupted volumes and
chemistry of the igneous rocks is attributed to the onset of at subduction and associated increased coupling between the Nazca slab and
overlying continental plate, which led to contractile deformation and
crustal thickening (Bissig et al., 2003; Kay et al., 1999). Epithermal mineralization is generally hosted in Cerro de las Trtolas Formation and
older rocks but occurred several Ma after the emplacement of the host
rocks. Mineralization of the Filo Federico ore zone of the Veladero
deposit is partly hosted in volcaniclastic deposits of the Vacas Heladas
Formation (Holley, 2012).
The geomorphology in the El Indio belt preserves evidence of a history of episodic uplift and pediplain erosion throughout the middle to
late Miocene (Bissig et al., 2002a; Figs. 3, 4, 5, 6). Relics of three lowrelief erosional surfaces, vertically separated from each other by
200400 m, have been documented. These are the 1715 Ma
FronteraDeidad surface, the 1412.5 Ma AzufrerasTorta surface and
the 106 Ma Los Ros surface (Bissig et al., 2002a). The FronteraDeidad
surface dominating the higher interuves is a relic of a regionally extensive pediplain on the western Andean slope at the latitude of the El Indio
belt (Aguilar et al., 2011; Rodriguez et al., 2014), subsequently incised
by the AzufrerasTorta and Los Ros pediment surfaces. These nested
paleosurfaces in the main Cordillera are now separated from the landforms of the Coastal Cordillera by the north striking, west verging VicuaSan Flix fault system (Aguilar et al., 2011, 2013). Vertical
displacement along this fault system starting in the earlymiddle Miocene is thought to be responsible for the three distinct paleosurfaces observed in the El Indio belt (Aguilar et al., 2011). Steam-heated alteration
associated with epithermal mineralization in all deposits of the El Indio
belt is almost exclusively exposed on the AzufrerasTorta surface
(Figs. 3,4,5, 6), whereas the top of the mineralized zone typically occurs
~200 m below this surface (Bissig et al., 2002a).

4.1. The Tambo deposit


At Tambo, mineralization is hosted mainly in three diatreme
breccia bodies (Wendy, Kimberly and Canto Sur; Figs. 3, 4), as well
as a number of satellite breccia bodies and replacement veins, all of
which were emplaced into dacitic pyroclastic and volcanoclastic
rocks assigned to the upper Oligocene Tilito Formation. The ore
zone has an overall vertical extent of about 300 m (Deyell et al.,
2005a; Jannas et al., 1999). The paleosurface level is located at
4500 m a.s.l. for Kimberly and Wendy and has been downfaulted
~ 100 m relative to Canto Sur, located at 4600 m a.s.l. elevation
(Deyell et al., 2005a). Alunite, some intergrown with native Au, has
been dated at 8.8 to 8 Ma although at Canto Sur it yields ages as
young as 7.1 Ma (Bissig et al., 2001; Deyell et al., 2005a). Alunite
ages coincide with the inferred age of the Los Ros surface incision
(106 Ma) and the Tambo area is situated near the back scarp of a
Los Ros valley where it incises the older AzufrerasTorta surface
(Figs. 3, 4).

331

4.2. El Indio deposit


Mineralization at El Indio, in contrast to Tambo, is hosted in veins
(Jannas et al., 1999) and includes sulde-rich massive enargite pyrite
veins as well as quartz-rich veins with tetrahedrite/tennantite and
chalcopyrite. For some veins, an along-strike zonation from high- to
intermediate-suldation sulde assemblages, and from pyrophyllite
sericitediaspore alunite to illite-dominated alteration has been documented (Heather et al., 2003a; K.B. Heather, pers. commun. 2012). El
Indio is world-renowned due to exceptional bonanza grades of up to
N1% Au in the quartz-rich, intermediate-suldation El Indio Sur 3500
vein which permitted shipping of the ore directly to the smelter (direct
shipping ore) in the early days of exploitation. Gold occurred mainly as
native Au, auricupride and as tellurides (Heather et al., 2003a; Jannas
et al., 1990, 1999). As is the case for Tambo, the top of the mineralized
zone at El Indio is located about 200 m below the AzufrerasTorta Surface, here located at ~4400 m a.s.l. Steam-heated alteration is exposed
locally on Cerro Campana above the western part of the deposit
(Fig. 3). The mineralized zone has a vertical extent of 500 m extending
down to 3700 m a.s.l, below which the gold grade decreases markedly
(Bissig et al., 2002a; Jannas et al., 1999). The veins are best developed
in dacitic pyroclastic rocks of the Tilito Formation, which in the area of
the deposit are uncomformably overlain by andesitic lavas and
volcaniclastic rocks of the lower Miocene Escabroso Group. Most veins
also cross-cut the Escabroso Group but typically horsetail out above
the contact (Bissig et al., 2002a; Heather et al., 2003a). Low volumes
of ca. 8 Ma dacitic hypabyssal intrusions spatially associated with mineralization have also been reported (Heather et al., 2003b). The mineralization ages range from 8 to 5 Ma and are inferred from sericite
immediately adjacent to and alunite from within veins (Bissig et al.,
2001; A.H. Clark and K.B. Heather, unpubl. data) and a late stage 3.5
2.5 Ma barren hydrothermal overprint has also been identied (A.H.
Clark and K. B. Heather, unpubl. sericite 40Ar/39Ar data). The age range
indicates that multiple hydrothermal pulses related to successive pressure release, likely facilitated by erosion, from a sizeable batholith scale
magma system were responsible for mineralization. Contrary to the
conclusion of Jannas et al. (1999) enargite-rich veins are not systematically older than gold-rich veins, since the youngest data come
from 6.2 Ma alunite in the enargite-rich Campana B vein (Bissig et al.,
2001) and sericite from the 5 Ma Viento Oeste/Cuarzo Uno high to
intermediate-suldation veins whereas the El Indio Sur 3500 vein
yielded the oldest sericite ages of ca. 8 Ma (A.H. Clark and K.B. Heather
unpubl. data). The geochronological data demonstrate a signicant
age gap of at least 7 Ma between volcanic rock deposition (~ 26
15 Ma) and mineralization (85 Ma). Only small amounts of igneous
rocks contemporaneous with mineralization are known and these include ca. 8 Ma dacitic plugs and 5.5 Ma Vallecito Formation rhyolite
tuffs in the vicinity of El Indio (Bissig et al., 2001; Heather et al.,
2003a; A.H. Clark and K.B. Heather unpubl. data). Like Tambo, it is
plausible that El Indio formed near the back-scarp of the incising Ro
Malo valley (part of the Los Ros surface) and is hosted below the
AzufrerasTorta surface (Bissig et al., 2002a; Fig. 3).
4.3. PascuaLama
PascuaLama is the largest deposit in the El Indio belt and is located
some 50 km N of El Indio on the ChileArgentina border. It is hosted
within an uplifted Paleozoic basement block (Figs. 5, 6), with mineralization centered on a series of diatreme breccia complexes. The most important of these is Brecha Central, which cuts Paleozoic and Jurassic
granitoids (Chouinard et al., 2005). Hydrothermal activity is constrained
to 9.18 Ma based on 40Ar/39Ar dating of alunite and jarosite; alunite
from pyriteenargite veins constraining the gold mineralization to
8.8 0.6 Ma (Bissig et al., 2001; Deyell et al., 2005b). Slightly older alunite of 9.5 Ma is documented from the Penelope ore zone some 2 km
SE of Brecha Central (Bissig et al., 2001; Deyell et al., 2005b). No igneous

332

Table 1
Summary of main HS deposits.
Mineralization
Total
resources
(approx.)

Host rocks

Volcanology

Selected references

8.5 to 8 Ma

0.8 Moz
Au

Hydrothermal breccia hosted AuAg


mineralization associated with
intense acid sulfate alteration.
Subordinate replacement veins.

Late Oligocene Tilito


Formation dacite tuffs

Bissig et al. (2002a,


2002b); Deyell et al.
(2005a); Jannas et al.
(1999)

40

8 to 5 Ma
Ar/39Ar,
alunite, sericite

4.2 Moz
Au

Largely vein hosted Au (Cu, Ag)


mineralization. Most gold produced
from quartz rich veins with sulde
assemblage of intermediate
suldation state, most Cu produced
from massive enargite veins.
Dominant alteration is advanced
argillic to phyllic (sericite, diaspore,
locally pyrophyllite, alunite).
Mineralization in two principal ore
zones and associated with intense
acid sulfate alteration and
silicication: Amable (12.1 to
12.7 Ma) and Filo Federico (10.7
11 Ma). Apart from minor argentite
no primary suldes recognized. Most
Ag in jarosite at Amable.
Breccia and stockwork vein-hosted
mineralization associated with high
suldation state sulde assemblage
and intense acid sulfate alteration
Vein hosted mineralization above
porphyry center. Ore mineralogy
indicates high-suldation state and
includes famatinite, enargite, pyrite
and lesser, tennantite, tetrahedrite,
sphalerite, gold, tellurides,covellite,
and chalcopyrite.
Primary suldes include pyrite
enargite, at depth also bornite,
covellite, fahlore, galena and
spalerite. Much of the ore is oxidized

Late Oligocene Tilito


Formation dacite tuffs and
overlying early Miocene
Escabroso Group andesite
lavas and breccias

No igneous rocks age equivalent to


mineralization recognized, dacite
tuffs of the Tilito Formation pre-date
mineralization by N 14 m.y. Minor,
undated pre-mineral diorite
porphyry has locally been
intersected below mineralization.
Phreatomagmatic activity pre to syn
mineralization evidenced by breccia
bodies.
No igneous rocks age equivalent to
mineralization recognized, except for
possible 8 Ma dacite dyke. Host rocks
pre-date mineralization by ~7 to
17 m.y.

Elevation of Mineraliza-tion
Elevation of
mineralization paleosurface age (method)
(if known)
(m.a.s.l.)

Deposit/
district

Lat.

Long.

Tambo

29.8

69.95 43004100

4500

El Indio

29.74 69.97 42003700

4400

29.36 69.95 43003900

4600

12.710.7 Ma
(40Ar/39Ar,
alunite)

12.2 Moz
Au; 180
Moz Ag

Pascua-Lama

29.32 70.01 48504550

50005150

9.18 Ma
(40Ar/39Ar,
alunite)

17.6 Moz
Au; 585
Moz Ag

Famatina

29.00 67.78 47504350

3.8 Ma
(40Ar/39Ar,
sericite)

0.4 Moz
Au; 3
Moz Ag

La Coipa

26.81 69.27 41503800

4350

2017 Ma K/Ar,
alunite

3.5 Moz
Au; 230
Moz Ag

El Hueso

26.49 69.53 39503700

n/a

4036.2 Ma
(40Ar/39Ar,
alunite)

0.84 Moz
Au; 3.9
Moz Ag

Mineralization is associated with


illitesericite and quartz alteration.
Sulde minerals include pyrite,
chalcopyrite, galena, native bismuth
and BiPb sulde minerals. This
assemblage is overprinted by
advanced argillic alteration (alunite,
pyrophyllite) and pyrite. Most gold
was introduced during the rst stage

Cerro de las Trtolas


Formation (1614 Ma)
volcano sedimentary breccias,
subvolcanic domes

Permo-Triassic to Jurassic
granitic rocks

Volcano sedimentary host rocks


pre-date hydrothermal breccias and
mineralization by ~ 24 Ma. Locally
12.1 Ma andesite dykes cutting
mineralization at Amable zone

Phreatomagmatic breccias, pre and


syn mineralization. Post mineral
dacitic dyke (7.8 Ma) cutting
mineralization.
Dacitic porphyries dated at 5 Ma are
Cambrian weakly
metamorphosed marine shale associated with porphyry
mineralization but no extrusive
and siltstones of the Negro
volcanic rocks known from the
Peinado formation
district but ash-ow tuffs from the
eastern ank of the Famatina range
may be age equivalent to porphyries.
Black shales and sandstones of Domes and lava ows of
the Triassic Ternera formation intermediate composition were
dated at 2422 as well as 1614.7 Ma
and overlying Oligocene
dacitic domes and pyroclastic and thus both pre- and post date
rocks. Diatreme breccias are a hydrothermal alunite ages. No
common ore host throughout volcanic rocks coincident with
mineralization known
the district as well
Calcareous silt and
Host rocks pre-date mineralization
sandstones, limestones of the by ~15 m.y. No age equivalent
Asientos Formation (u.
volcanic rocks known but
Jurassic) and overlying
subvolcanic porphyries are
Paleocene andesite lavas and
widespread and associated with
volcaniclastic rocks
porphyry Cu mineralization at
Potrerillos

Charchaie et al. (2007);


Holley (2012)

Bissig et al. (2002a,


2002b); Chouinard et al.
(2005); Deyell et al.
(2005b)
Lozada Caldern et al.
(1994); Pudack et al.
(2009).

Cecioni and Dick (1992);


Oviedo et al. (1991)

Marsh et al. (1997)

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Veladero

Bissig et al. (2002a,


2002b); Heather et al.
(2003a); Jannas et al.
(1999)

25.1

n/a

4842 Ma
(K/Ar, alunite)

1.6 Moz
Au

Choquelimpie 18.49 69.39 48604700

n/a

6.6 Ma?? (K/Ar,


host rock)

0.5 Moz
Au

Santa Rosa
(Aruntani)

16.63 70.02 49504750

n/a

7.2 Ma
(40Ar/39Ar,
alunite)

0.5 Moz
Au

Tucari
(Aruntani)

16.57 70.19 52004900

n/a

4.6 Ma
(40Ar/39Ar,
alunite)

2.1 Moz
Au

Julcani

12.95 74.95 45004000

~9.7 Ma (K/Ar
on biotite)

8 Moz Ag

Marcapunta/
Colquijirca

10.78 76.27 44504850

n/a

11.910.6 Ma
(40Ar/39Ar,
alunite)

Quicay

10.7

n/a

37.5 Ma (K/Ar)
alunite (?)

2345 kt
Zn, 910
kt Pb,
1510 kt
Cu,
6620 t
Ag, 64 t
Au
n/a

76.26 4350?

Vein hosted mineralization of


enargite, pyrite and minor
chalcopyrite with quartz gangue in
andesite, smaller irregular veins in
overlying pyroclastic rocks. Oxidaton
and supergene enrichment affected
the ore to ~ 300 m depth and is
characterized by the presence of a
variety of rare arsenate minerals.
Mineralization associated with
hydrothermal breccia bodies in the
core of a late Miocene stratovolcano
Mineralization is hosted in vuggy
quartz zones centered on magmatic
hydrothermal breccia bodies and is
surrounded by quartzalunite
alteration zones. Primary suldes
include pyrite and enargite but the
deposit has been oxidized to a depth
of 300400 m below surface. Highgrade mineralization is controlled by
NNE oriented fractures.
Mineralization is hosted in vuggy
quartz zones centered on
hydrothermal breccias and is
surrounded by quartzalunite
alteration zones. Primary suldes
include pyrite and enargite but the
deposit has been oxidized to a depth
of 300400 m below surface.
Pre-ore acid sulfate alteration with
vuggy quartz core and surrounding
quartzalunite zone in the center of
the district, cut by discontinuous
quartztourmalinepyrite breccias.
Mineralization is hosted by laterally
zoned veins with quartzpyrite
wolframiteenargitetetrahedrite
galena and barite rich zones. Zoning
is from high suldaton near the core
to distal intermediate suldation
assemblages).
Zoned polymetallic district with a
high-suldation AuAg prospect and
distal carbonate replacement
PbZnAg mineralization.

Mineralization is hosted in an
oxidized diatremedome complex.
Highest grades up to 3 g/t are hosted
in a central vuggy quartz zone.

Andesite lava ows and


overlying dacitic pyroclastic
rocks. Both units were dated
by KAr at ~60.260.7 Ma

Volcanic host rocks pre-date


mineralization by 12 or more m.y. No
volcanic rocks age equivalent to
mineralization known.

Puig et al. (1988);


http://
www.australgold.com.au
accessed Nov 2013.

Late Miocene andesite lavas


and dacite subvolcanic dome

Andesitic stratovolcano.
Mineralization largely hosted in late
dacitic subvolcanic dome
Mineralization occurs adjacent to
rhyolitic ow/dome complex and is
closely associated with magmatichydrothermal breccias.

Groepper et al. (1991)

Miocene (97 Ma) felsic


pyroclastic deposits are
dominant ore host.

Barreda et al. (2004);


Morche et al. (2008)

Intrusions and domes of


dacitic composition.

Mineralization is spatially associated


with magmatic-hydrothermal
breccias emplaced at the margin of
dacitic domes. The latter were
emplaced in the core of an andesitic
stratovolcano.

Barreda et al. (2004);


Morche et al. (2008)

10.1 Ma andesitic to dacitic


domes overlying paleozoic
phyllites of the Excelsior
group

Voluminous extrusive domes of


andesitic to dacitic composition,
covering 16 km2. Syn mineralization
9.7 Ma dykes and post
mineralization 7 Ma rhyolite dome
are also documented.

Deen et al. (1994);


Noble and Silberman,
1984

Eocene limestone and marls


and 12.912.4 Ma diatreme
dome complex.

Mineralization is spatially associated


with a 12.912.4 Ma diatreme dome
complex (Marcapunta) pre-dating
alunite by N0.8 m.y. A dacite dyke
dated at 14.1 Ma has been
documented W of Marcapunta. No
other igneous rocks recognized.

Bendezu et al. (2008);


Vidal and Ligarda
(2004);

Intermediate volcanic rock


and hydrothermal breccias

Subvolcanic to extrusive dome


complex of uncertain age hosting
mineralization. Apparently no postmineral volcanic rocks.

Rossell et al. (2006);


Ingemmet.gob.pe,
accessed Jan 2014.

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

69.53 24002900

El Guanaco

(continued on next page)

333

334

Long.

Elevation of Mineraliza-tion
Elevation of
mineralization paleosurface age (method)
(if known)
(m.a.s.l.)

Mineralization
Total
resources
(approx.)

Host rocks

Volcanology

Selected references

Late Triassic to early Jurassic


Pucara group limestone and
mid Miocene diatreme
breccias.

Mineralization occurs at eastern


border of a large 15.4 to 15.15 Ma
diatremedome complex pre-dating
stage II mineralization by N 0.8 Ma
but with unclear temporal
relationship to stage I mineralization.
No volcanic rocks pre-dating the
dome complex known.
Late dacite ow dome complexes
immediately pre-date
mineralization. No post mineral
volcanic rocks reported.

Baumgartner et al.
(2008, 2009)

Deposit/
district

Lat.

Cerro de
Pasco

10.68 76.27 43503900

n/a

15.414.4
(40Ar/39Ar,
alunite)

12,250 kt
Zn, 3500
kt Pb,
15,750 t
Ag, 1000
kt Cu

Pierina

9.46

77.59

n/a

14.214.7
(40Ar/39Ar,
alunite)

Oligocene to Middle Miocene


(29.314.8 Ma) andesitic to
dacitic tuffs, lavas and
subvolcanic domes of the
Huaraz group

Lagunas
Norte

7.95

78.25

n/a

17.416.5 Ma
(40Ar/39Ar,
alunite)

9 Moz Au Mineralization hosted in permeable


pumiceous rock units and associated
with intense acid sulfate alteration,
including vuggy quartz and quartz
alunite. Suldes include enargite,
pyrite as well as galena, sphalerite
and stibnite, and lesser
paragenetically late digenite and
covellite.
13.1 Moz Disseminated and fracture hosted
Au
mineralization, centered on diatreme
breccias. Mineralization in volcanic
rocks hosted in vuggy quartz and
quartzalunite zones. Dominant
sulifdes include enargite, pyrite and
digenite.

Yanacocha

78.58 41003500

~4200

118.2 Ma
(40Ar/39Ar
Alunite)

N70 Moz
Au

Calipuy group (19.515.1 Ma)


andesitic to dacitic lavas and
pyroclastic deposits and
14.58.4 Ma Yanacocha unit
dacite lavas and pyroclastic
rocks

Early pyrrhotite-pyritegalenaFe
rich sphalerite (stage I) and later
CuAg (AuPbZn) enargitepyrite
veins as well as carbonate
replacement PbZn ore (stage II).

Massive and vuggy quartz zones


hosting pyrite enargite
tennantitecovellite

Lower Cretaceous Chim


Formation quartzites (two
thirds) and early Miocene
dacite tuffs and breccias (one
third).

Volcanic rocks hosting


mineralization were emplaced
between early and main
mineralization stage. Volcanic rocks
were likely derived from diatreme
breccias which also host
mineralization. Andesite lava and
dome adjacent to main diatreme has
same age as mineralization.
Mineralization overlaps in age with
voluminous andesitic to dacitic
volcanic rocks. Most mineralization
(N47 Moz Au) emplaced in the
waning stages of volcanism
represented by the felsic Coriwachay
dacite domes (10.98.4 Ma).

Fifarek and Rye (2005);


Rainbow et al. (2005,
2006)

Cerpa et al. (2013);


Montgomery (2012).

Longo et al. (2010)

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Table 1 (continued)

6.9
6.83

78.79 ~3500
78.89 37003400

n/a
n/a

n/a
n/a

n/a
0.73 Moz
Au, 6
Moz Ag

Tantahuatay

6.72

78.67 40503800

n/a

12.4 Ma
(K/Ar alunite)

Quimsacocha

3.04

79.22 37003450

3900

7.57.3 Ma
(40Ar/39Ar
Alunite)

Gold is mostly associated to pyrite


5.1 Moz
Au; 108.3 enargite hosted in intensely quartz
pyrophyllitealunite and vuggy to
Moz Ag
massive quartz alteration. Much of
the ore is oxidized
Stratabound Au-Ag mineralization
3.6 Moz
associated with pyrite and enargite,
Au, 27
hosted in vuggy quartz zones
Moz Ag

La Bodega/La
Mascota

7.38

72.9

29002300

~3600

3.51.7 Ma
(40Ar/39Ar,
alunite, sericite)

3.5 Moz
Au

Angostura

7.39

72.88 35002600

~3600

41.9 Ma
(40Ar/39Ar,
alunite, sericite)

2.7 Moz
Au

n/a
n/a

Dominantly structurally controlled


breccias cemented by quartz, alunite
and high-suldation sulde
assemblage, overprinting earlier
quartzpyrite chalcopyrite veins
in phyllic alteration. Most AuAg in
late stages. Banded, colloform and
lattice textures normally typical for
low to intermediate suldation
deposits are common.
Widespread sheeted quartz
pyrite chalcopyrite veins
associated with pervasive quartz
sericite alteration. Locally cut by
enargiteCu sulde bearing breccias
and veins associated with quartz
alunite alteration. The latter hosts
highest grade Au zones.

n/a
n/a

Miocene andesite domes and


to a lesser degree underlying
pyroclastic rocks

Mineralization located in
pyroclastic units between less
permeable coherent andesite
lava ows of the Quimsacocha
Formation

Proterozoic gneisses and late


Triassic leucogranite.

Proterozoic gneisses and late


Triassic leucogranite.

n/a
n/a

Gustafson et al. (2004)


Compaa minera
Buenaventura 2012
Annual report, Gustafson
et al. (2004)
Gustafson et al. (2004);
Volcanic domes hosting the
Noble and Mckee
mineralization are not age
(1999); Compaa
constrained. Post mineral dyke of
minera Buenaventura
restricted volume was dated at
2012 Annual report
8.6 Ma (K/Ar, biotite).
MacDonald et al. (2011,
Andesite lavas and intercalated
pyroclasitc units associated to a large 2012)
stratovolcano complex. Andesite
is ~ 9 Ma, pre-dating mineralization
by 1.5 Ma. A central caldera and
dacite domes were emplaced at
6.7 Ma, 0.5 Ma after mineralization
Rodriguez (2014)
No igneous rocks age equivalent to
mineralization recognized. No
juvenile clasts in breccias observed.

No igneous rocks age equivalent to


mineralization recognized.

Felder et al. 2005,


Rodriguez (2014).

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Sipan
La Zanja

335

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

6,800,000

336

El Indio belt geology

Cerro del Toro (6168 m)

6-5.5 Vallecito Fm. rhyolite

Au mineralization

6,780,000

2 Ma Cerro de Vidrio rhyolite dome

8-7.6 Ma Pascua Fm. dacite


12.7-11 Ma Vacas Heladas Fm. dacite
21-14 Ma Escabroso and Cerro
de las Trtolas Fm. andesite
and subvolcanic diorite

Cordille
ra Ortg
a

6,760,000

27-23 Ma Tilito Fm. and equivalents


dacite, andesite and minor basalt

Pascua-Lama

Paleozoic and Mesozoic basement


Thrust/reverse fault (symbol on
hanging wall side)
Fault (unspecified)
Lineament from satellite image
High-sulfidation epithermal deposit
Major Mountains

6,720,000

Cordille
ra Zanc

Sancarrn

Cordille
ra de la

arrn

Brea

6,740,000

Veladero

Tambo

6,700,000

Cerro Doa Ana


(5625 m)

Cordille
ra Cola
ngil

El Indio

Cerro de las Trtolas (6160 m)

6,680,000

N
20 km

380,000

400,000

420,000

440,000

460,000

Fig. 2. Geological map of the El Indio belt and location of principal epithermal deposits. Compiled from Martin et al. (1995), Bissig et al. (2001), Zappettini (2008), and Winocur et al. (2014).
Coordinates in UTM Zone 19S, Provisional South American Datum 56.

rocks contemporaneous with mineralization have been conrmed but


low volumes of juvenile igneous material probably played a role during
emplacement of the diatreme complexes and rare, locally weakly
argillically altered, 7.8 Ma rhyodacitic dykes post-dating mineralization
have been recognized at Pascua (Bissig et al., 2001).
The alteration and sulde paragenesis reveals that unusually acidic
uids were responsible for alteration and some of the mineralization.
Gold mineralization is associated with pyrite and enargite, as well as
alunite, jarosite and szomolnokite (Chouinard et al., 2005). Jarosite
is locally overprinted by hypogene aluniteenargite bearing veins

(Chouinard et al., 2005). Based on light stable isotope evidence, jarosite


precipitated from mixed meteoric and magmatic uids whereas alunite
has a magmatic uid signature (Deyell et al., 2005b). This indicates that
uid sources varied through time, plausibly as a function of a uctuating
but overall decreasing water table and short lived magma-derived uid
pulses, as also documented for El Indio (see above).
Near PascuaLama, the FronteraDeidad surface reaches elevations
of 52005300 m a.s.l. (Fig. 5) The AzufrerasTorta surface is incised
into this high-elevation pediplain at 49505100 m and extensive
steam-heated alteration and clastic deposits, both probably related to

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

337

El Indio/Tambo Area, Chile & Argentina


6958'0"W

6956'0"W

2942'0"S
2944'0"S

Cerro Campana

6952'0"W

6950'0"W

Argentina

Chile

Ro del Medio

6954'0"W

LR
AT

6710000

700'0"W

2940'0"S

702'0"W

El Indio

LR

2946'0"S

AT

AT
Canto Sur

2948'0"S

Kimberly

FD
Veta Veronica

Tambo

Wendy
6700000

2950'0"S

AT

2952'0"S

LR

5 km

400000

410000

Elevation (m)

Slope (Degrees)

420000

Paleosurfaces

High : 6107

Los Ros (LR); 10-6 Ma


0 - 10

Azufreras-Torta (AT); 14-12.5 Ma

Low : 1775

Frontera-Deidad (FD); 17-15 Ma

Fig. 3. El Indio Tambo area topography and landscape elements. Paleosurfaces after Bissig et al. (2002a). UTM Zone 19S, WGS84.

phreatomagmatic eruptions derived from the mineralized breccia pipes,


are exposed on this landform (Bissig et al., 2002a). Mineralization occurs
below the steam-heated zone at 4850 to 4600 m (Chouinard et al., 2005).
As at El Indio and Tambo, a valley (Ro Turbio valley: Fig. 5) incised the
AzufrerasTorta surface from the east. Ferricrete-cemented conglomerates at the valley-bottom display glacial striations, showing that this valley pre-dated Plio-Pleistocene glacial erosion. Further, steam-heated
alteration also overprints the eastern side of the deposit to lower

elevations (to 4800 m) than to the west (to 5000 m: Chouinard et al.,
2005) indicating that incision of the Ro Turbio valley from the east
may have occurred during hydrothermal activity.
4.4. Veladero
Veladero is situated about 8 km to the SE of Pascua in a NNE striking
graben (Charchai et al., 2007). Gold and silver mineralization is

338

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Veta Vernica

Co
Ele
fan
te
Co
T de la
rto
las s

Deidad

Steam-heated alteration
Frontera Deidad surface (17-15 Ma)
Azufreras-Torta surface (14-12.5 Ma)
Los Ros surface (10-6 Ma)

Tambo
(Kimberly pit)

Azufreras

C Doa Ana

Canto Sur pit

Looking S from Canto Sur, Tambo

Fig. 4. Landscape in the Tambo area. A) Panoramic view to the south, B) line drawing of the photograph in A) showing the principal landscape elements, locations of steam-heated
alteration zones as well as the Kimberly and Canto Sur orebodies. See Fig. 3 for locations.

hosted in two principal breccia bodies: Amable and Filo Federico


(Fig. 6). Ore mineralogy comprises electrum and silver-bearing jarosite
which were deposited together with euhedral quartz, and are
superimposed on earlier acidic residual quartz alteration (Holley,
2012). Jarosite is abundant, especially at Amable, whereas suldes are
exceedingly rare. At Amable, jarosite forms euhedral crystals and
grows in open spaces and veins, and is locally overgrown by barite.
Stable-isotope evidence indicates that most jarosite precipitated from
meteoric uids above the water table, with sulfur derived from precursor suldes. However, one location in the Fabiana zone to the east of
Veladero (Figs. 5, 6) shows unambiguous hypogene jarosite, replacing
plagioclase and biotite phenocrysts in the host rock and displaying
stable isotope signatures of magmatic uids (Holley, 2012).
Mineralization at Amable is constrained between a 12.7 Ma alunite
40
Ar/39Ar age and a 12.14 Ma U/Pb zircon date of a narrow andesite
dyke that cross-cuts mineralization (Holley, 2012). Mineralization at
Filo Federico is at 11.1 to 10.3 Ma (40Ar/39Ar, alunite, n = 4) somewhat
younger than at Amable (Bissig et al., 2001; Holley, 2012). Jarosite ages
at Amable range from 11.8 to 8.6 Ma and overlap with hypogene alunite
and gold deposition at Filo Federico (Holley, 2012). These data are
interpreted as evidence for an overall lowering of the water table with
time, episodically interrupted by pulses of magma derived uids.
Jarosite at Veladero almost entirely pre-dated hypogene alunite alteration and sulde mineralization at nearby PascuaLama.
The host rocks of the Amable ore zone consist of crudely stratied
volcaniclastic breccias assigned to the Cerro de las Trtolas formation,
unconformably overlying Tilito Formation andesites and dacites as
well as Permian felsic tuffs (Charchai et al., 2007; Holley, 2012). The
Filo Federico ore zone is partly hosted in volcaniclastic deposits assigned
to the Vacas Heladas Formation (Holley, 2012). Mineralization is closely
associated with hydrothermal breccias intruding the volcaniclastic deposits and occurs between 4400 and 3900 m a.s.l. in a subhorizontal,
N S elongated ore body, extending extends to lower elevations at
Filo Federico than at Amable. A steam-heated alteration zone of up to
200 m thickness overlies the orebodies and denes the paleosurface at
~44004600 m a.s.l. At Filo Federico, steam-heated alteration is localized along fracture zones, and overprints the mineralized zone down
to more than 500 m below the inferred paleosurface (Holley, 2012).
Here, the FronteraDeidad surface is marked by the lower contact of
the Cerro de las Trtolas formation which here has been downfaulted

from 5300 m to about 4100 m in the NS oriented Ro de las Taguas


graben (Charchai et al., 2007, Fig. 6), whereas the AzufrerasTorta
surface has been cut into Cerro de las Trtolas volcaniclastic deposits
at an elevation of ~ 4500 m (Charchai et al., 2007, Fig. 6), i.e. at the
time of, or immediately prior to, hydrothermal activity in that area.
Conversely, the Filo Federico mineralization probably formed at a time
when incision of the Los Ros surface was well advanced, as evidenced
by the steam-heated overprint of the ore zone and the mineralization
extending to lower elevations than at Amable. The lowering of the
water table in the Veladero area during Los Ros surface incision is also
reected in the formation of jarosite in the vadose zone at Amable
between 11.8 and 8.5 Ma.
5. Maricunga belt
The Maricunga belt is dened as the region of the Andean Cordillera
between 26 and 28 Lat S (Vila and Sillitoe, 1991), located at the northern margin of the Pampean at slab segment. Gold-rich porphyry type
deposits (e.g., Cerro Casale (a.k.a. Adelbarn), Caspiche, Lobo, Marte)
are the dening mineralization style and the gold-rich nature of
porphyry-style mineralization is thought to be related to the shallow
emplacement of the porphyry intrusions (Muntean and Einaudi,
2001). Although barren quartzalunite alteration is commonly found
in the shallow portions of these porphyry systems, La Pepa is the only
porphyry deposit where signicant gold mineralization hosted in
quartzalunite veins that overprint porphyry-style mineralization has
been documented (Muntean and Einaudi, 2001). In contrast, La Coipa,
located at the northern end of the Maricunga belt, hosts AgAu mineralization exclusively as epithermal mineralization of largely highsuldation type (Cecioni and Dick, 1992; Oviedo et al., 1991).
Porphyry and epithermal deposits of the Maricunga belt occur in
two partly overlapping parallel NS striking belts, with mineralization
traditionally considered to have been emplaced during two distinct episodes at 2520 Ma and 1513 Ma (Sillitoe et al., 1991). However, more
recently published, but still limited geochronology requires modication of this model. The Caspiche porphyry Au deposit was dated at
~25.4 0.01 Ma by Re/Os on molybdenite (Sillitoe et al., 2013) whereas
the Caserones porphyry Cu deposit in the southern Maricunga belt
formed at 2018 Ma (Mpodozis and Kay, 2003). Further, alunite K/Ar
ages from the La Coipa district range from ~ 20 to ~ 17 Ma (Oviedo

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

339

Pascua-Lama/Veladero Area, Chile & Argentina


700'0"W

6955'0"W

Chile

6760000

Argentina

AT

AT
2920'0"S

Pascua-Lama

Ro

Tu

rbi

o
Lama Central

Penelope

LR
Veladero

Fabiana

AT

Cerro de Vidrio

6750000

AT

FD

2925'0"S

LR
AT
AT
0
400000

Elevation (m)

1.5

6 km

410000

Slope (Degrees)

High : 5678

Paleosurfaces
Los Ros (LR); 10-6 Ma

0 - 13

Azufreras-Torta (AT); 14-12.5 Ma


Frontera-Deidad (FD); 17-15 Ma

Low : 2877
Fig. 5. Veladero and Pascua area topographic map and landscape elements. Paleosurfaces after Bissig et al. (2002a) and Charchai et al. (2007). UTM Zone 19S, WGS84.

et al., 1991) which disagree with the range between 25 and 20 Ma


established on the basis of alunite K/Ar ages (Sillitoe et al., 1991) and
biotite and hornblende K/Ar ages for dacite domes that were thought
to be closely related to mineralization (Oviedo et al., 1991; Sillitoe
et al., 1991).
The large-scale morphotectonic features of the Andes are different at
the latitude of the Maricunga belt to those of the El Indio belt, probably
inuenced by the differing Paleozoic and older basement architecture

(Ramos, 2008). Thus, the Precordillera (a.k.a. Cordillera Domeyko),


resulting from the late Eocene Incaic deformation is present north of
29 Lat. S but absent south of it (Rodriguez et al., 2014), a location
which roughly coincides with the inferred southern limit of the
Antofalla massif and the northern limit of the Chilenia terrane (Ramos,
2008). Oligocene extensional tectonics are documented from south of
the Vallenar Orocline (Abanico basin, Tilito Fromation: Winocur et al.,
2014; Charrier et al., 2002) whereas a contractile tectonic regime at

340

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Cerro Pelado
Amable

Filo Federico

Penelope
Lama Central

Ro Turbio

Cerro Nevado

B
Pascua/Lama

nb
grabe

oundin

g fault

Fabiana
Steam-heated alteration
Frontera Deidad surface (17-15 Ma)
Azufreras-Torta surface (14-12.5 Ma)
Los Ros surface (10-6 Ma)

Looking W from Fabiana

Fig. 6. Landscape in the Veladero and PascuaLama area. A) panoramic view from the Fabiana prospect toward the west. B) Line drawing of the landscape shown in A. Landscape elements
are shown as well as steam-heated alteration zones. See also Fig. 5 for locations.

roughly the same time is documented from the Cordillera Claudio Gay at
the latitude of the Maricunga belt during the late Oligocene (Mpodozis
and Clavero, 2002). The dominant landscape elements at the latitude of
the northern Maricunga belt include the probably Oligocene relict Sierra
Checos de Cobre surface which is incised by the upper Oligocene
to ~ 18 Ma Asientos pediplain in the eastern Precordillera (Bissig and
Riquelme, 2009; Mortimer, 1973). The Oligocene to early Miocene landscape was incised by the middle- to early-upper Miocene Atacama
pediplain and late Miocene canyons (Bissig and Riquelme, 2009)
which record a progressive tilting of the western Andean slope and uplift of the Andean Cordillera since the early Miocene (Riquelme et al.,
2007).

5.1. La Coipa
The La Coipa district contains multiple epithermal deposits of mostly
high-suldation type including Ladera Farelln, Can Can, Coipa Norte,
Purn, Purn Sur and Pompeya (Fig. 7). With the exception of Purn
and Pompeya, mineralization is hosted in the Triassic La Ternera formation, consisting of black shales and sandstones, as well as in overlying
upper Oligocene (2422 Ma) dacitic tuffs, volcaniclastic breccias and
domes (Cecioni and Dick, 1992; Oviedo et al., 1991). Hydrothermal
breccias hosting ore are probably penecontemporaneous with mineralization. Alunite K/Ar ages range between ca. 20 and 17 Ma (Oviedo et al.,
1991). A relict, large, upper Oligocene volcanic center, Cerros Bravos

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

341

La Coipa Area, Chile


6916'0"W

6914'0"W

6912'0"W

2646'0"S

7040000

6918'0"W

Purn
L. Oligocene/E. Miocene paleosurface
Purn Sur
Coipa Norte

2648'0"S

Pompeya
Can Can

2652'0"S

7030000

2650'0"S

Ladera Farelln

0.5

470000

Elevation (m)

5 km

480000

Slope (Degrees)

High : 4907
0 - 10

Outline of post-mineral
dacite domes and lavas

Low : 3013
Fig. 7. Topographic map and landscape elements of the La Coipa district. Some geological information from Cornejo et al. (1998). UTM Zone 19S, WGS84.

(Cornejo et al., 1993), is located 16 km N of Ladera Farelln. Unaltered,


1614.7 Ma dacitic domes and lavas post-date mineralization (Oviedo
et al., 1991) and locally overlie the mineralized zones at Pompeya
(Figs. 7, 8). The Pompeya and the Purn ore bodies were discovered
after 2010, and only Purn has been mined. Pompeya is located ~3 km
NE of Ladera Farelln and is of high-suldation type. Purn, 8 km
NE of Ladera Farelln features high grade PbZn mineralization
representing an intermediate suldation sulde assemblage, an absence

of advanced argillic alteration and the occurrence of various carbonates


including rhodochrosite as alteration minerals (S. Gamonal, pers.
commun. 2014). At Purn Sur, some 2.5 km south of Purn, native
sulfur associated with steam-heated alteration was historically mined.
Steam-heated alteration is also present at Pompeya, Purn and Coipa
Norte and is generally preserved at 4250 to 4400 m a.s.l. (Figs. 7, 8).
Mineralization at Ladera Farelln, Can Can and Coipa Norte is largely
controlled by subvertical fractures when hosted in Triassic shales, but

342

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

16 Ma dacite lava

16 Ma dacite dome

Pompeya

B
Steam-heated
alteration zone

Late Oligocene/early Miocene erosion surface

Ladera Farelln

Can Can

Coipa Norte

16 Ma
dacite dome

Fig. 8. Landscape and physical setting of mineralization in the La Coipa district. A) Panoramic view to the southwest from Pompeya. Dark unaltered rocks at left margin of photograph
belong to a post mineral dacite dome. B) Line drawing of photograph shown in A) highlighting the relationship of landscape and mineralization. Steam-heated alteration is exposed on
a late Oligocene to early Miocene erosion surface (light blue). C) Panoramic view looking northwest at the Ladera Farelln, Can Can and Coipa Norte ore bodies. D) Line drawing of the
photograph shown in C). Light blue indicates the late Oligocene to early Miocene paleosurface.

expands into mushroom-shaped zones in the overlying volcanic rocks.


The latter hosts higher Ag grades, whereas the former is relatively
Au-rich (Oviedo et al., 1991). The top of the exposed orebodies lies
around 4300 m a.s.l., with sulde oxidation extending downwards

to ~ 3800 m a.s.l. (Cecioni and Dick, 1992). The dominant alteration


is vuggyquartz and quartzalunite in the volcanic rocks and quartz
alunite dickite and pyrophyllite in the shale and sandstones.
There is a close association of alunite vein stockworks with gold

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

mineralization (Oviedo et al., 1991). The suldes observed at depth include pyrite, chalcocite, covellite, bornite, chalcopyrite, tennantite,
tetrahedrite, galena and sphalerite (Oviedo et al., 1991), representative
of precipitation at intermediate to high-suldation states. No porphyrystyle mineralization related to epithermal mineralization is known at La
Coipa.
The geomorphology of the La Coipa district is dominated by a planar,
slightly west-inclined landscape constructed on the upper Oligocene
pyroclastic rocks and domes, but covered by the post mineral
1614.7 Ma dacite lavas and extrusive domes which form the topographic highs in the district (Cornejo et al., 1998; Figs. 7, 8). This planar
surface is located at present day elevations of 44004200 m a.s.l., and
appears to have been the dominant paleosurface at the time of
epithermal mineralization as steam-heated mineralization is commonly
exposed on it (Fig. 8). The paleosurface that plausibly controlled
epithermal activity can be correlated with the upper Oligocene
to 18 Ma Asientos pediplain mapped on the northern side of the
Cerros Bravos volcanic complex (Bissig and Riquelme, 2009). The uplift
and subsequent erosion of the Asientos pediplain is probably related to
the late Oligocene contractile tectonic phase described from the Cordillera Claudio Gay some 50 km E of La Coipa (Mpodozis and Clavero,
2002). Gold and silver mineralization at Ladera Farelln and Pompeya
is located near the upper reaches of a NE oriented valley, incised into
the probable upper Oligocene to lower Miocene paleosurface, but the
timing of valley erosion is currently unknown.
5.2. La Pepa
Within the Maricunga belt, besides la Coipa, only La Pepa contains
signicant high-suldation epithermal mineralization. Here it is hosted
in NNW-trending quartz alunite veins that cross-cut porphyry-Au style
mineralization centered on the Cavancha porphyry (Muntean and
Einaudi, 2001). Similar quartzalunite veins are widespread in other
porphyry Au deposits of the Maricunga belt, but only those at La Pepa
contain gold grades high enough to sustain small-scale historic production (Muntean and Einaudi, 2001). La Pepa is also the only one of these
deposits where alunite precipitation (between 23.5 and 23.25 Ma,
based on 40Ar/39Ar plateau ages), did not overlap temporally with the
potassic alteration, here dated at 23.81 0.08 Ma (hydrothermal biotite
40
Ar/39Ar plateau age, Cavancha porphyry; Muntean and Einaudi, 2001).
The late Oligocene landscape at La Pepa is obscured by extensive upper
Miocene ignimbrite deposits through which La Pepa is exposed in an
erosional window (Muntean and Einaudi, 2001).
6. High-suldation epithermal deposits of the Domeyko fault system
Two noteworthy middle to late Eocene high-suldation epithermal
deposits, El Hueso and El Guanaco, are located in the southern Atacama
Desert. These deposits occur near the Domeyko fault system of the
Chilean Precordillera, which also contains the behemothian Escondida,
Chuquicamata, and Collahuasi porphyry Cu districts. All of these
deposits were emplaced during, or immediately after the late Eocene
Incaic orogenic event which has been ascribed to an inferred episode
of slab attening, crustal thickening, volcanic lull and subsequent eastward shift of the volcanic arc (Mpodozis and Cornejo, 2012; O'Driscoll
et al., 2012).
Regional uplift and formation of large scale erosional surfaces of
Incaic age are implied from constraints on supergene activity and
large scale morphotectonic features such as the Precordillera which
formed during the Incaic orogeny in the southern Atacama Desert of
northern Chile. Uplift in the Eocene to elevations similar to the present
day is also evidenced from stable-isotope paleoaltimetry for the Puna
plateau (Canavan et al., 2014) as well as for the pre-Cordillera at ~26
Lat. S (Bissig and Riquelme, 2010). Supergene oxidation and Cu enrichment in Paleocene to middle Eocene porphyry deposits in response to
uplift and erosion, initiated as early as 45 to 36 Ma at Cerro Verde

343

(Peru), Cerro Colorado, Spence and El Salvador (Bissig and Riquelme,


2010; Bouzari and Clark, 2002; Gustafson and Hunt, 1975; Quang
et al., 2003, 2005). The late Eocene climate was probably semi-arid
and considerably wetter than the present day since sufcient water is
necessary for economically signicant supergene enrichment and
erosion.
6.1. El Hueso
Arguably, nowhere in the Andes is the interrelationship between
major orogenic episodes and porphyry and epithermal mineralization
better demonstrated than in the PotrerillosEl Hueso district (Cornejo
et al., 1993; Marsh et al., 1997; Niemeyer and Munizaga, 2008;
Thompson et al., 2004). At El Hueso, epithermal mineralization is located about 3 km E of the Potrerillos porphyry Cu deposit, in the
hangingwall of the SE-verging Potrerillos Mine thrust fault which was
active during and immediately after porphyry and epithermal mineralization (Marsh et al., 1997; Niemeyer and Munizaga, 2008). The district
also hosts the Jernimo carbonate-hosted epithermal gold deposit
(Thompson et al., 2004), located some 1.5 km to the east of El Hueso,
in the footwall of the Potrerillos Mine Thrust fault.
Epithermal alteration and mineralization at El Hueso are hosted in
calcareous silt and sandstones of the Upper Jurassic Asientos formation,
as well as the uncomformably-overlying andesitic-to-rhyolitic lavas and
volcaniclastic rocks of the Paleocene Hornitos Formation (Marsh, 1997).
Hydothermal activity occurred in two stages. An early mineralization
stage spatially associated with a 40.8 Ma porphyry stock was dated at
ca. 40.2 Ma by hydrothermal sericite and alteration at this time was
dominated by sericite and illite, but incorporated only minor advanced
argillic alteration. A second stage of barren quartzalunite alteration
and silicication occurred at ca. 36.2 Ma (King, 1992; Marsh et al.,
1997; Olson, 1984). The early stage deposited gold, accompanied by
12% pyrite, stibnite, chalcopyrite, galena and arsenopyrite, all hosted
in quartz veinlets (Marsh et al., 1997). The principal porphyry CuAu
mineralization centered on the Cobre porphyry was emplaced about
3 km west of El Hueso and immediately after the epithermal mineralization and alteration at 35.535.9 Ma (Marsh et al., 1997). Structural analysis of the district suggests that the emplacement depth of porphyry
CuAu mineralization at the Cobre porphyry was shallow, as little as
1 km below the paleosurface (Niemeyer and Munizaga, 2008). Although
a number of Eocene porphyry stocks intrude the Jurassic and Paleocene
rocks in the district, no age-equivalent volcanic rocks are recognized
(Marsh et al., 1997). In the El Salvador porphyry Cu district, located
some 20 km NW of Potrerillos, hypogene mineralization occurred at
4341 Ma (Gustafson et al., 2001), but the deposit was exposed to oxidation and supergene modication at ca. 36 Ma (Bissig and Riquelme,
2010; Gustafson and Hunt, 1975; Mote et al., 2001). Thus, uplift and erosion exhumed porphyry style mineraliztion at El Salvador about 6 Ma
after hypogene mineralization, i.e. at the time of hydrothermal activity
at el Hueso and Potrerillos. The Precordillera (a.k.a. Cordillera Domeyko)
at the latitude of the present day southern Atacama Desert is located at
the southewestern boundary of the Puna Plateau and probably formed
part of the Andean crest from late Eocene until the late Oligocene, before
late Oligocene thrusting in the Cordillera Claudio Gay, some 70 km east
of the Precordillera, established the main Andean range as known today
(Bissig and Riquelme, 2010).
The geomorphology of the area around El Hueso is dominated by a
subplanar erosion surface, now situated at 39004000 m elevation
which has been cut into the Paleocene Hornitos formation (Fig. 9).
This erosion surface is interpreted as a pediplain and its age is uncertain
but probably represents a somewhat erosion-degraded relic of a
pediplain that formed in response to uplift related to the late Eocene
to early Oligocene Incaic orogeny (Fig. 9). In a broad context, El Hueso
is located below a topographic high, near a scarp that separates it
from topographically lower areas to the west and may have been
emplaced in a physiographic setting comparable, albeit less well

344

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Cerro El Hueso

El Hueso epithermal Au Deposit

Late Eocene to Oligocene paleosurface

>18 Ma Asientos pediplain with gravel cover

Fig. 9. Panoramic view of the physiographic setting of the El Hueso epithermal Au deposit. Photograph taken from the East looking West. The Potrerillos porphyry Cu (Au) deposit is
located behind the Horizon West of El Hueso and is not visible in this view. See Bissig and Riquelme (2009).

preserved, to e.g., PascuaLama or El Indio. Given the widespread preservation of late Eocene to early Oligocene epithermal alteration and
mineralization styles at most about 200300 m of erosion occurred
since hydrothermal activity in the El Hueso area. Erosion modication
of the original Incaic erosion surface probably occurred after 33 Ma as
it affects the granodiorite stock of Cerro El Hueso dated at ca. 33 Ma
(Cornejo et al., 1993), but prior to 18 Ma, which is the age of the
~100200 m lower Asientos pediplain cutting into the high level post
Incaic surface at El Hueso (Bissig and Riquelme, 2009). The base level
of the Asientos pediplain is the Salar de Pedernales which was
established in the late Oligocene (Bissig and Riquelme, 2009). The
Oligocene to early Miocene landscape elements and underlying shallow
level epithermal mineralization in the El Hueso area are preserved due
to the arid climate. Erosion since the early Miocene was mostly concentrated in canyons and average denudation rates were low (~14 m/Ma:
Riquelme et al., 2008).

6.2. El Guanaco
El Guanaco is a second example of a high-suldation epithermal deposit spatially associated with the Domeyko fault system. It is located
30 km W of the crest of the Precordillera at 25.1 Lat. S. Mineralization
is hosted in roughly E to ENE-trending quartz replacement veins with
a narrow advanced argillic and argillic alteration halo. Suldes include
enargite, pyrite and subordinate chalcopyrite. Veins were emplaced
in andesite lavas and overlying dacite tuffs of Paleocene age (ca.
5460 Ma; Puig et al., 1988). The veins are well-dened, subvertical,
tabular structures when hosted in the andesite, but assume an irregular
behavior in the overlying, more permeable felsic pyroclastic rocks.
Supergene oxidation affected the deposit up to 300 m depth and gave
rise to a large variety of arsenate minerals, El Guanaco being the typelocality for some of these (Witzke et al., 2006).
At El Guanaco, hydrothermal alunite yields K/Ar ages between ~48
and 42 Ma (Puig et al., 1988). No age-equivalent intrusive or volcanic
rock or porphyry-style mineralization has been documented in the district. El Guanaco is the oldest-known high-suldation epithermal deposit of the Andes, but overlaps in age with the El Salvador and
Esperanza porphyry Cu (Au) districts (Gustafson et al., 2001; Perell
et al., 2004), similarly located west of the main Domeyko fault system.
Mineralization at El Guanaco is located below Cerro Estrella, a local
topographic high at ~ 2900 m a.s.l., and extends to depths of at least
2500 m a.s.l. Cerro Estrella rises about 300 m above surrounding relict
pediment surfaces of unknown age. However, El Guanaco's physiographic setting is consistent with a position below a high-level Incaic
or pre-Incaic erosion surface being incised by pediments either during
or after hydrothermal activity.

7. Late Miocene high-suldation epithermal Au deposits of the


western Cordillera of northern Chile and southern Peru
High-suldation epithermal deposits are rare in the Central Volcanic
zone, between ~25 Lat S in northern Chile and 15 Lat S in Central Peru,
an area that coincides with the major ChilePeru orogenic bend. Here,
the subduction angle is at ~ 30 steeper than in the present at-slab segments of northern Chile and central and northern Peru (Cahill and
Isacks, 1992). The current slab geometry resulted from steepening of
an Eocene shallow angle slab (Sandeman et al., 1995). The Andes at
the ChilePeru oroclinal bend are metallogenetically diverse and include W, Sn, Ag, Pb, Zn and U deposits in the Inner Arc (partly coinciding
with the Cordillera Oriental) ranging from Jurassic to Pleistocene age, as
well as porphyry and skarn type mineralization of Paleocene and
Eocene ages (Clark et al., 1990; Perell et al., 2003). However, a few
Miocene epithermal gold deposits occur in the Cordillera Occidental of
northern Chile and southern Peru, all of which are located at local topographic highs with mineralization exposed at elevations of 4800 to
5200 m a.s.l. Signicant deposits include the 6.6 Ma Choquelimpie
high-suldation deposit in Chile (Groepper et al., 1991) and the Santa
Rosa and Tucar high-suldation deposits of the Aruntani district in
southern Peru (Barreda et al., 2004; Morche et al., 2008). Hypogene
alunites from Santa Rosa and Tucari have been dated at 6.4 and
4.6 Ma, respectively (Morche et al., 2008).
Mineralization at Choquelimpie is hosted in the core of a stratovolcano interpreted as broadly cogenetic with epithermal mineralization
(Groepper et al., 1991). At Aruntani, multiple episodes of volcanism,
represented by dacitic plugs and subvolcanic intrusions as well as andesitic lavas both pre-dated and overlapped with epithermal mineralization (Barreda et al., 2004). Thus, the volcanologic setting of these
deposits is distinct from the deposits farther south in that mineralization is temporally more closely associated with voluminous volcanism
and does not post-date the age of their host rocks by 5 or more Ma.
The geodynamic setting of late Miocene high-suldation deposits of
the Cordillera Occidental also differs from the current at slab segments
in that magmatism is associated with contraction of the arc due to
slab steepening rather than arc-broadening and cessation related to
shallowing subuduction angles as observed in the Pampean and
Peruvian at slab segments (cf. Bissig et al., 2003, 2008; Kay and
Mpodozis, 2001; Sandeman et al., 1995).
The landscape evolution documented on the western Andean slope
indicates that multiple stages of pediment bevelling due to uplift of
the Incaic paleosurface occurred between ca. 24 Ma and 8 Ma (Quang
et al., 2005; Tosdal et al., 1984) followed by a major pulse of Altiplano
uplift and incision of deep Canyons (Thouret et al., 2007). The majority
of workers agree that most of the Altiplano uplift to average elevations
of 3800 m, occurred during the Miocene. Evidence from both

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

geomorphology (Hoke et al., 2007; Jordan et al., 2010; Schildgen et al.,


2007) and paleoaltimetry (based on stable isotope and paleobotany;
Garzione et al., 2008) suggest that at least 1000 m or more of this elevation gain occurred after 10 Ma, but prior to about 5 Ma. Thus, epithermal
gold deposits of the Central Volcanic zone formed during a major uplift
pulse.
8. Central to northern Peruvian at slab segment
The Andean segment extending from ~15 Lat S in Peru, to 2 Lat S in
southern Ecuador, lacks recent arc volcanism, which, as in Chile, has
been attributed to at-slab subduction spatially coinciding with subduction of aseismic ridges and oceanic plateaus (e.g., Gutscher et al., 1999a;
Martinod et al., 2010; Skinner and Clayton, 2013). The high-suldation
epithermal deposits of this region occur in two distinct belts (Noble
and McKee, 1999). The southeastern belt, located east of the Cordillera
Occidental, which forms the Continental divide, and extends from
Julcani (~13 Lat S) to the latitude of Antamina (~9.5 Lat S). It coincides
with the central Peruvian polymetallic province and contains a variety
of porphyry-related and carbonate hosted deposits emplaced at different paleodepths (Bissig and Tosdal, 2009; Escalante, 2008; Escalante
et al., 2010; Love et al., 2004). This segment includes the cordilleran
base metal lode deposits (i.e. sulde rich polymetallic deposits:
Bendez et al., 2003) of MarcapuntaColquijirca and Cerro de Pasco
which are centered on high-suldation epithermal deposits emplaced
in reactive host rocks (Bendez and Fontbot, 2009; Bendez et al.,
2008). The northwestern metallogenetic belt has its southern terminus
at Pierina at 9.5 Lat S, west of the Cordillera Blanca and is dominated by
high-suldation epithermal and porphyry CuAu deposits but also
includes a number of smaller vein hosted AgAu deposits such as
Quiruvilca (Gustafson et al., 2004; Noble and McKee, 1999). It includes,
from south to north, Pierina, Lagunas Norte and Yanacocha, the latter of
which contains upwards of 70 Moz contained Au (Longo et al., 2010;
Teal and Benavides, 2010) and which is by far the largest epithermal
deposit cluster of the Andes.
Subduction of the Nazca aseismic ridge in Peru commenced in the
middle Miocene at 1410 Ma (Hampel, 2002) and is the inferred
cause of at subduction (e.g., Martinod et al., 2010). However, estimates
on the timing of crustal thickening and uplift of the Cordillera Occidental in central and northern Peru vary from the early to the middle Miocene and do not coincide everywhere with the onset of aseismic ridge
subduction and at subduction. Thus, Noble et al. (1990) on the basis
of the age of tuff deposits in deeply incised paleovalleys in northern
Peru, suggested that uplift and erosion occurred in the earlylate
Miocene. Crustal thickening inferred from whole rock geochemical
data of volcanic rocks in the central Peruvian polymetallic province is
inferred to have occurred around 1215 Ma (Bissig and Tosdal, 2009).
In contrast, Montgomery (2012) proposed that a major episode of uplift
commenced at about 17 Ma, i.e., before the inferred onset of ridge subduction, but coincident with high-suldation epithermal mineralization
at Lagunas Norte (Cerpa et al., 2013; Montgomery, 2012).
9. High-suldation epithermal deposits of the central Peruvian
polymetallic belt
The central Peruvian polymetallic belt contains a number of late
Miocene silver and base metal-rich vein deposits that were emplaced
in the epithermal environment. These include: San Cristbal (Beuchat
et al., 2004), Morococha (Catchpole et al., 2011) and Uchucchacua
(Bussell et al., 1990; Escalante, 2008). At Morococha an evolution from
the porphyry to the epithermal environment occurred concurrently
with erosion, and late Miocene epithermal Ag mineralization probably
occurred only a few 100 m below the paleosurface (Catchpole et al.,
2011). All of these epithermal deposits have sulde assemblages of intermediate suldation state and veins are largely hosted in Mesozoic
carbonaceous rocks as well as Triassic volcanosedimentary rocks.

345

However, the polymetallic belt also contains a number of epithermal


deposits with important high-suldation characteristics; these are
described in the following.
9.1. Julcani
The Julcani district, mined since colonial times, consists of severalvein hosted deposits from which Ag and base metals, as well as subordinate Au, have been produced. The mineralization is hosted in coalescing
domes of andesitic-to-dacitic composition and postdates a large zone of
vuggy quartz and quartz alunite alteration in the center of the district, as
well as quartztourmalinepyrite breccia bodies (Deen et al., 1994;
Petersen et al., 1977). Veins are zoned along strike from higher
suldation state enargitetennantite bearing assemblages near the center of the district to lower-suldation assemblages containing galena
and tennantitetetrahedrite more distally (Petersen et al., 1977).
Magmatism largely predated the epithermal mineralization but a
9.7 Ma syn-mineral dyke and small, 7 Ma, post-mineral rhyolite
domes have been recognized (Deen et al., 1994; Petersen et al., 1977).
Julcani is hosted below a sub-planar, slightly east-inclined, land surface
situated at elevations of 40004500 m a.s.l.
9.2. The MarcapuntaColquijirca district
Marcapunta is a high-suldation epithermal Au (Ag, Cu) deposit.
The district also contains cordilleran base metal mineralization at the
Smelter deposit, northward contiguous to Marcapunta, and more distally the Colquijirca AgPbZn deposit, some 5 km north of Marcapunta.
The San Gregorio deposit is another cordilleran base metal deposit
southerly adjacent to Marcapunta (Bendez et al., 2008). Vuggy quartz
alteration, phreatomagmatic breccias and disseminated gold mineralization occur in the central Marcapunta dome, whereas stratabound
base metal and silver rich mineralization at Smelter and Colquijirca,
extending as far as 5 km north from Marcapunta, are hosted in Eocene
limestones and marls (Bendez et al., 2008). There is a district-scale
lateral zoning from proximal high-suldation mineralization containing
enargite and pyrite to intermediate pyrite, chalcopyrite and tennantitebearing ore, to a distal pyrite, sphalerite and galena dominated sulde
assemblage, the later reecting an intermediate suldation state
(Bendez et al., 2008; Vidal and Ligarda, 2004). Alunite associated
with the central high-suldation gold mineralization yielded ages of
11.9 to 11.1 Ma, whereas that associated with the base metal mineralization yielded slightly younger ages of 10.8 to 10.5 Ma (Bendez et al.,
2008). The dacitic dome complex hosting some of the epithermal
mineralization was emplaced at 12.9 to 12.1 Ma (40Ar/39Ar biotite
ages). No syn- or post mineral volcanic or intrusive rocks are exposed
in the district.
Colquijirca is located near the eastern margin of a regionally extensive low-relief plain situated at an elevation of 4200 to 4300 m a.s.l.
whereas the Cerro Marcapunta summit at 4450 m a.s.l. constitutes the
highest feature in the deposit area. Vuggy quartz hosted gold mineralization extends from 4450 m to about 4000 m a.s.l. or 450 m below surface, whereas distal base metal mineralization occurs at depths less than
~400 m below surface. Proximal enargite-rich mineralization adjacent
to the dome complex, however, has been drilled to a depth of N600 m
below surface and attains thickness of up to 100 m (Bendez et al.,
2008; Vidal and Ligarda, 2004).
9.3. Cerro de Pasco
Cerro de Pasco is located about 11 km N of Marcapunta and shares
many similarities to the geologic setting and mineralization style of
Colquijirca. The deposit is spatially related to a 15.4 to 15.2 Ma dacitic
diatreme dome complex which was emplaced at the boundary of Devonian phyllites to the west and Triassic-to-Jurassic limestones to the east
(Baumgartner et al., 2008, 2009). Mineralization occurred in two stages

346

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

(Baumgartner et al., 2008; Einaudi, 1977). Stage 1 was focused at the


eastern margin of the diatreme dome complex and is characterized
by a low-suldation sulde assemblage consisting of a massive
quartzpyrite pyrrhotite body with associated distal carbonate replacement ore containing galena and Fe-rich sphalerite. Alteration related to stage 1 is quartzsericite. The second mineralizing stage reects
high-suldation states and has proximal EW trending enargitepyrite
veins, hosted in part by the diatreme breccia body, and distal carbonatehosted PbZnAg veins and replacement bodies containing Fe-poor
sphalerite (Baumgartner et al., 2008). Alunite is part of the alteration assemblage of stage 2 and gives 40Ar/39Ar ages of 14.5 to 14.2 Ma, whereas
stage 1 is bracketed between 15.4 and 14.5 Ma (Baumgartner et al., 2009).
Like the nearby Marcapunta and Colquijirca deposits, Cerro de Pasco
is associated with a dome complex which forms a local topographic high
above the regionally-extensive, low-relief surface to the west of it. The
pre-mining land surface was situated at about 4350 m a.s.l. and mineralization extends about 450 m below this original land surface
(Baumgartner et al., 2008).
9.4. Quicay
Quicay is a high-suldation epithermal center located 14 km W of
Cerro de Pasco. It is being mined by the Peruvian private company
Chancadora Centauro SAC and little information on its geology and
resources is publically available. According to information obtained
through the INGEMMET website Peru (Rossell et al., 2006) gold mineralization is hosted in a diatremedome complex and the highest gold
grades are found in the central vuggy quartz zones (average 3 ppm).
The ore is oxidized. A mineralization age of 37.5 Ma based on a K/Ar alunite age is suggested by Noble and McKee, (1999) which is signicantly older than Cerro de Pasco, despite the similar inferred shallow level of
formation. Quicay mineralization is centered on a small hill that contains outcrops of vuggy residual quartz. It has a summit elevation of
4350 m, which is roughly 100 m above the same low-relief land surface
described for Cerro de Pasco. If the age constraint is reliable, it can be
inferred that no signicant erosion affected the area since the Eocene.
10. The high-suldation epithermal Au (Cu, Ag) deposits
of northwestern Peru
10.1. Pierina
Pierina is located in the Cordillera Negra, Ancash, west of the Cordillera Blanca. The coeval, dominantly intermediate suldation, Santo
Toribio Ag-base metal vein systems occur ~ 5 km south of Pierina
(Rainbow, 2009), and both deposits are located beneath the shoulder
of an erosional surface overlooking the Callejn de Huaylas valley
(Figs. 10, 11). At Pierina, disseminated mineralization forms a subhorizontal body, and is largely hosted in a lithologically-controlled
vuggy quartz alteration zone, focused in a ca. 16.9 0.6 Ma
(40Ar/39Ar, weakly chlorite-altered biotite total gas age) pumice tuff
and an underlying dacitic ow dome complex (Rainbow, 2009); both
members of the Oligocene to mid-Miocene Huaraz Group (Rainbow
et al., 2005; Strusievicz et al., 2000), the upper succession of the Calipuy
Supergroup subaerial volcanic arc. Hydrothermal breccias and small
dacitic domes cut these rocks but appear to pre-date mineralization
(Rainbow et al., 2005). Goldsilver mineralization was introduced
after initial acidic alteration, and is associated with enargite, pyrite,
bismuthinitestibnite, galena and low-Fe sphalerite (Rainbow et al.,
2005). However, suldes have largely been oxidized. Sub-microscopic
Au and Ag are now hosted in a goethitehematite dominated oxide assemblage (Rainbow et al., 2005), the formation of which was facilitated
by microbial activity during supergene oxidation. During this process,
the local reduction of supergene uids led to the formation of Aubearing acanthite (Rainbow et al., 2006). Supergene minerals at Pierina
do not include jarosite.

Hydrothermal alunite 40Ar/39Ar ages range from 15.08 0.09 to


13.89 0.13 Ma (n = 19), clustering in two pulses around 15 Ma and
14.4 Ma (Rainbow, 2009), whereas an age of rare, vug lling porcellaneous alunite from the oxide zone, yielded a large-error 14.12 1.59 Ma
plateau age and may be of supergene origin (Rainbow, 2009). This
suggests that oxidation closely followed hypogene mineralization. The
host rocks of the Huaraz group range in age from 29.3 to 14.8 Ma
(Rainbow et al., 2005; Strusievicz et al., 2000).
The mineralized part of Pierina is located between about 4000 and
3800 m a.s.l., starting about 100 below the nearest topographic highs.
A horizontally extensive steam-heated alteration blanket has not been
documented at Pierina, and the presence of any steam-heated alunite
remains controversial (Fifarek and Rye, 2005; Rainbow et al., 2005,
2006). However, paragenetic relationships and stable-isotope geochemistry demonstrate that meteoric waters played an increasingly important role in deposit formation, from early alteration to subsequent
oxidation, and that these uids became progressively less isotopically
exchanged over time (Rainbow et al., 2006). This suggests that the
water table was progressively lowered during the lifetime of the hydrothermal system (Fifarek and Rye, 2005; Rainbow et al., 2005, 2006).
In the PierinaSanto Toribio area, components of middle Miocene
planar erosional landforms both preceding and broadly contemporaneous with mineralization at Pierina can be recognized (Fig. 11). The deposit formed at the crest of the Cordillera Occidental and Calipuy arc
facing the Amazonian low-lands to the east, the latter not yet separated
hydrographically from the Cordillera Negra as the intervening Cordillera
Blanca was only uplifted in the late Miocene (Gonzlez and Pffner,
2012; Petford and Atherton, 1992). An erosion surface, marked by an
angular unconformity, now tilted and dipping at ~ 20 to the ENE, underlies the andesites below the Pierina deposit. The andesites have
been dated at ca. 21 Ma, suggesting that this angular unconformity records uplift and erosion in the late Oligocene to early Miocene, probably
representing the Aymar orogenic event (Sbrier et al., 1988).
The upper slope of the Cordillera Negra, immediately south of the
Santo Toribio deposit is faceted by four erosional surfaces, one of
which is constrained by the overlying 14.10 1.3314.99 0.50 Ma
(40Ar/39Ar hornblende plateau ages) lava ows of the unaltered Santo
Toribio Formation andesite package. This constrained surface also
intersects the extensive area of phyllic alteration surrounding the
Santo Toribio vein system. This shows that hydrothermal activity (at
both Santo Toribio and Pierina, dated at ca. 15 to 14.4 Ma) was
penecontemporaneous with both andesite eruption and the incision of
the pediment. Volcanism terminated after 14 Ma in the area. The mineralization at Pierina (and Lagunas Nortesee below), the cessation of
volcanism and onset of uplift around 14 Ma pre-dated the arrival of
the Nazca ridge at the subduction zone by at least 2 Ma (Hampel,
2002). Uplift in response to crustal thickening in central and northern
Peru has, instead, been linked to increased mid Miocene plate convergence and may not be directly related to initiation of at subuduction
(Montgomery, 2012; PardoCasas and Molnar, 1987).
10.2. Lagunas Norte
Lagunas Norte, La Libertad, is located about 200 km NNW of Pierina
and 100 km SSE of Yanacocha in the northeastern mineral belt of northern Peru. Mineralization is largely hosted in quartzites with scarce interspersed coal beds of the Lower Cretaceous Chim Formation and, to a
lesser degree, in overlying dacitic pyroclastic and volcaniclastic rocks
of the Lagunas Norte Formation (Fig. 12) Montgomery, 2012; Cerpa
et al., 2013). Lagunas Norte Formation volcanic units are volumetrically
minor and are restricted in distribution to the immediate Lagunas Norte
deposit area (Fig. 12) Mineralization is centered on at least 2 diatremes
which are also considered the source of the pyroclasitc rocks overlying
the quartzites (Cerpa et al., 2013; Montgomery, 2012). Hydrothermal
alunite 40Ar/39Ar data constrain the age of mineralization at Lagunas
Norte to between 17.4 and 16.5 Ma (Cerpa et al., 2013; Montgomery,

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

347

Pierina Area, Peru


7740'0"W

7735'0"W

7730'0"W

7725'0"W

Pierina

Co

rd

930'0"S

ra

Bl

8940000

ille

Santo Toribio

an

ca

no

rm

al

fa

ul

Huaraz

8930000
8920000

940'0"S

egra

935'0"S

Cordillera N

2.5

7.5

10 km

-450000

Elevation (m)

-440000

-430000

-420000

Slope (Degrees)

High : 6280
0 - 10

Low : 1565
Fig. 10. Topographic map at landscape elements and main physiographic features of the Pierina area. The approximate trace of the Cordillera Blanca normal fault is shown. Note that this
fault was not active prior to the late Miocene and areas to the east of it have been uplifting relative to the Cordillera Negra since the Late Miocene to the present (Gonzlez and Pffner,
2012; Petford and Atherton, 1992). UTM Zone 18S, WGS84.

2012), whereas a minimum age of 16.5 Ma and maximum age of


17.2 Ma for volcanic units of the Lagunas Norte Formation have been determined (Montgomery, 2012). Extensive lower Miocene (21.1 Ma to
16.4 Ma; Montgomery, 2012) andesitic-to-dacitic volcanic rocks of the
Sauco Volcanic Complex (age- and compositional-equivalent to Huaraz
Group strata of the Pierina district; Strusievicz et al., 2000; Montgomery,

2012) were deposited east of Lagunas Norte. These rocks largely predate hydrothermal activity at Lagunas Norte (Fig. 12) but mapping
together with geomorphology and geochronology suggests that the
Sauco volcanic complex is the result of sector collapse contemporaneous with mineralization at Lagunas Norte. At Lagunas Norte a reduction
of erupted magma volumes and increased SiO2 content over time is

348

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Quebrada Pacchac
Quebrada Huellap (with heap leach facility)
Santo Toribio
Callejn de Huaylas
15.59 0.23

14.10 1.33

15.21 0.23
14.81 1.27

14.60 0.08

Looking S from Pierina


Surface IV (~3970 m)
Surface III (~4040 m)
Surface II (~4260 m)
Surface I (~4500 m)

andesite
hydrothermal alteration

hornblende Ar-Ar plateau age


sericite Ar-Ar plateau age
sericite Ar-Ar total fusion age
alunite Ar-Ar plateau age

Fig. 11. Field photograph of the landscape and physical setting of the Pierina area. A) Panorama photographs taken at Pierina looking south across Santo Toribio. B) Line drawing of
principal landscape elements and locations. See text for further detail.

documented leading up to the main period of high-suldation


epithermal activity, with restricted, dacitic eruptions overlapping temporally and spatially with mineralization (Montgomery, 2012). The volcanic rocks at Lagunas Norte exhibit alteration zonation typical of highsuldation epithermal deposits with a vuggy quartz core surrounded by
quartzalunite and kaolinitedickite alteration zones. The quartzites, in
contrast, are non-reactive, and display only subtle alteration that includes kaolinite, minor pyrophyllite and traces of alunite (Cerpa et al.,
2013; Montgomery, 2012). Mineralization in the quartzites is fracturecontrolled, and hypogene ore minerals include digenite, enargite and
pyrite, the latter hosting some of the gold. Locally, within coal beds,
low-suldation sulde assemblages are present (Cerpa et al., 2013).
The ore is oxidized to a depth of over 80 m.
The orebody forms a tabular body between ~4250 to 4000 m a.s.l.,
although sulde mineralization extends to greater depths. It occurs
below relics of the extensive ca. 2625 Ma subplanar erosional Pampa
La Julia surface (Montgomery, 2012) into which the ca. 1816 Ma Ro
Chicama surface, consisting of steep-walled, but at-bottomed valleypediment was cut. The Pampa La Julia surface may be correlative with
the upper Oligocene to lower Miocene angular unconformity recognized 200 km to the SSE at Pierina in the Huaraz District (see above), attributed to the regional Aymar orogenic event (Sbrier et al., 1988).
The eastern Sauco Volcanic Complex in the Lagunas Norte district
erupted onto the older Pampa La Julia surface, whereas the Lagunas
Norte deposit is laterally contiguous with pronounced scarp between
this older surface and the younger Rio Chicama valley pediment
(Figs. 13, 14). Active erosion and scarp retreat during hydrothermal activity are constrained in age by truncated a 17.05 Ma alunite veins of at
the western margin of the Lagunas Norte deposit and 16.75 Ma volcanic
rocks emplaced on the Rio Chicama surface immediately west of the deposit (Figs. 13, 14; Montgomery, 2012). Incision of the Rio Chicama

surface at Lagunas Norte occurred during, and was likely a response


to, the initial stages of an episode of major, mid- to late Miocene
(ca. 175 Ma) regional uplift (e.g., Montario et al., 2005a, 2005b) and
crustal thickening, during which, as much as 23 km of surface uplift
may have occurred (Montgomery, 2012).
Hydrothermal activity and ore deposition at Lagunas Norte occurred
during the latter stages of development of the Sauco Volcanic Complex,
centered 4 km east of the deposit, during waning volcanism, and immediately post-dated collapse of the volcanic edice. Regional contractile
tectonism and uplift may have ultimately triggered the collapse of this
volcanic edice, and may also have led to marked decrease in magmatic
activity which set the stage for development of the Lagunas Norte
ore body (Montgomery, 2012). Inferred emplacement of a shallow
intrusion subjacent to Lagunas Norte within this evolving tectonomagmatic setting, and its subsequent cooling, downward contraction
and associated uid release, is envisaged to have resulted in early acid
leaching and subsequent ore deposition at Lagunas Norte. Incision of
the Rio Chicama pediment at the margin of the Lagunas Norte deposit
at this time, would have favored decompression and uid release, and
subsequent uid mixing and/or boiling at the site of ore deposition.
10.3. Yanacocha
Gold mineralization occurs in about 10 individual centers within
broad NE-trending zones of residual quartz, massive quartz, quartz
alunitepyrophyllite and distal kaolinite alteration (Longo et al., 2010;
Teal and Benavides, 2010). Early pervasive alteration also includes
quartzpyrophyllite featuring the wormy or gusano texture characteristic of the lower portion of the lithocap environment (Sillitoe, 2010;
Teal and Benavides, 2010). Gold was introduced in several pulses and
is largely associated with permeable pyroclastic rocks, subvertical

5 km

810000 E

349

805000 E

795000 E

78 15' W

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

9130000 N

Unmapped

Lagunas Norte

CSVC

9120000 N
Shulcahuanga dome

9115000 N

Quiruvilca mine

8 S

Quaternary
Undifferenated glacial deposits
Middle Miocene
Las Princesas volcanic domes. 16.3-15.2 Ma

Quiruvilca-Tres Cruces Volcanic Complex (> 25 -18.4 Ma)


Andesite domes and lavas
Cretaceous
Undifferenated sedimentary rocks

Lagunas Norte Formaon, dacite to andesite,


domes lavas and tuffs. 17.3-16.4 Ma
Sauco Volcanic Complex (21.1-17.1 Ma)
Dacite to andesite lavas and domes
(CSVC - Central Sauco Vent complex)
Andesite volcaniclasc rocks

Structures
Reverse fault (observed/inferred)
Normal fault (observed/inferred)
Fault (observed/inferred)
Syncline
Ancline

Early Andesite Dome/Lava


Fig. 12. Simplied geological map of the Lagunas Norte area. Geology and age constraints from Montgomery (2012). Coordinates in UTM Zone 18S, Provisional South American Datum 56.

fractures or breccia zones as well as margins of domes (Teal and


Benavides, 2010). The highest gold grades are associated with replacive
cream-colored chalcedony veins containing barite and ne grained rutile
and Fe oxides. These are interpreted as representing precipitation under
intermediate-suldation conditions (Teal and Benavides, 2010). Much of
the ore mined to date came from the oxidized portion of the deposit
where gold is associated with supergene Fe oxides. Sulde assemblages
associated with precursor hypogene epithermal mineralization include
enargite and covellite and, at depth, chalcopyrite (Teal and Benavides,
2010). Late-stage veins of base metal suldes associated with rhodochrosite have locally been reported from the Cerro Yanacocha deposit and
reect higher pH and intermediate-suldation state uids (Teal and
Benavides, 2010). Besides high-suldation and intermediate-suldation
epithermal mineralization, porphyry style CuAu ore has been described
from Kupfertal, an area exposed at ~3800 m a.s.l., 300 m below the summit of Cerro Yanacocha (Gustafson et al., 2004; Teal and Benavides,
2010).

The rocks hosting the epithermal mineralization largely belong to


the 14.511.2 Yanacocha Volcanics (Longo et al., 2010) which consist
of andesitic-to-dacitic lavas and pyoclastics and associated subvolcanic
rocks. As at Lagunas Norte, the volcanic pile records a reduction of
erupted magma volumes and increased SiO2 content over time, with
over 90% of the volcanic rocks erupted prior to 11 Ma (Longo et al.,
2010). Hydrothermal alunite ages (13.6 to 8.2 Ma) temporally overlap
with the Yanacocha Volcanics, but less than 20% of the gold is associated
with hydrothermal activity older than 11 Ma. The bulk of the gold
was introduced after 10.9 Ma and temporally overlaps with the
volumetrically-minor Coriwachay dacite domes dated at 10.9 to
8.4 Ma (Longo et al., 2010). Thus, high-suldation epithermal gold mineralization clearly post-dated much of the volcanism and occurred in
multiple pulses over a period of 23 Ma. Additionally, hydrothermal biotite associated with porphyry CuAu mineralization at Kupfertal
yielded an 40Ar39Ar age of 10.7 Ma. Although both porphyry and
high-sulifdation epithermal style mineralization are present and

350

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Lagunas Norte Area, Peru


78150W

7550S

RCh surface
18-16 Ma

Sauco volcanic complex


9120000

PLJ surface
26-25 Ma

Lagunas Norte

Q Constructional
surface

800S

PLJ surface
26-25 Ma

5 km
800000

Elevation (m)

810000

Slope (Degrees)

High : 4254
0-8

Low : 2275
Fig. 13. Topographic map, at landscape elements and main physiographic features of the Lagunas Norte area. PLJ: Pampa La Julia 2625 Ma paleosurface, RCh: Ro Chicama 1816 Ma
paleosurface, Q: Quesquenda constructional surface composed of lower Miocene volcanic rocks, including those of the Sauco volcanic complex which is indicated. Surfaces and location
of Sauco Volcanic complex from Montgomery (2012). UTM Zone 18S, WGS84.

spatially overlapping, they are probably formed in distinct magmatichydrothermal settings (cf. La Pepa, Muntean and Einaudi, 2001).
The Yanacocha deposits underlie a 3900 to 4000 m a.s.l. subplanar
land surface (Fig. 15) evident in digital elevation models, with local topographic highs, such as Cerro Yanacocha, reaching 4150 m. The high
planation surface has a similar elevation to the Pampa la Julia surface

in the Lagunas Norte district (Montgomery, 2012) and could conceivably be age-equivalent. The area hosting mineralization is incised by
at bottomed valleys of ~ 3500 m elevation to the west and south
which themselves have been incised up to 250 m by deep canyons. No
detailed documentation and direct age constraints on erosional surfaces
are available. However, steam-heated alteration is locally preserved

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

351

view looking southwest

Lower Miocene
Sauco volcanic
complex

Lagunas Norte
(early stages of open pit)

~ 16-18 Ma Ro Chicama valley-pediment network


~ 25-26 Ma Pampa La Julia pediplain

Alexa Zone
17.05 Ma alunite Shulcahuanga dome 16.95 Ma
Lagunas Norte orebody

Lagunas Norte Fm.

Looking south

Tres Amigos dome 16.75 Ma

4200 m a.s.l.

Pampa La Julia surface


Chim Fm.
4000 m a.s.l.

Tres Amigos
lava flows

Chim Fm.

Ro Chicama valley-pediment

Fig. 14. Field photographs of the landscape and physical setting of the Lagunas Norte area.

above mineralization (Teal and Benavides, 2010), suggesting that only


limited erosion has occurred since the late Miocene. Although glaciation
affected parts of the district above 3900 m a.s.l., glacial erosion was only
locally signicant and led to the formation of the detrital La Quinua gold
deposit hosted in glacial till (Mallette et al., 2004).
10.4. Tantahuatay, Sipan and La Zanja
Besides Yanacocha, northern Peru hosts a number of porphyry
CuAu and high-suldation epithermal deposits of latemiddle to late
Miocene age (Gustafson et al., 2004; Noble and McKee, 1999; Noble
et al., 2004). Most of these occur within 50 km of Yanacocha (Fig. 15).
The high-suldation epithermal deposits include Tantahuatay, Sipan
and La Zanja, whereas Cu Au Mo porphyry style mineralization
includes El Galeno and Cerro Corona (Gustafson et al., 2004). At
Tantahuatay, mineralization is hosted in an intensely vuggy to, locally,
massive quartz and quartzpyrophyllitealunite alteration zone, hosted
in andesitic domes and underlying pyroclastic rocks (Gustafson et al.,
2004). The andesites constitute the local topographic highs at
4050 m a.s.l. Gold is associated with pyrite and enargite and is concentrated in areas of secondary permeability (vuggy quartz or fracture

zones). The deposit has been affected to a more signicant degree


by glaciation than Yanacocha, and it is thought that a signicant
part of the mineralization has been eroded (Gustafson et al., 2004).
Little detailed geological information has been published for La Zanja,
operated by Compaia Minera Buenaventura and for Sipn. However,
Tantahuatay as well as Sipn and La Zanja lie beneath a high-elevation
planar landform similar to- and apparently contiguous with the one at
Yanacocha (Fig. 15). This surface is slightly inclined to the west and
Sipan and La Zanja are exposed at about 400 m lower elevation than
Tantahuatay and Yanacocha, the latter being located about 3040 km
to the E of Sipan and La Zanja.

11. The northern Andes


A major break in the basement architecture in southern Ecuador
separates the Northern Andes (3.5 Lat S. to 11 Lat N.) from the Central
Andes (Cediel et al., 2003; Gansser, 1973) and is reected in the differing metallogeny of the two regions. The western Cordillera of the northern Andes of Ecuador and Colombia is comprised of terranes accreted
during the Cretaceous and Miocene, whereas the basement of the

352

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Yanacocha/Sipan/Tantahuatay/La Zanja Area, Peru


7850'0"W

7845'0"W

7840'0"W

7835'0"W

9260000

7855'0"W

650'0"S

Cerro Corona P. Cu-Au

9250000

645'0"S

Tantahuatay

9240000

La Zanja

9230000

655'0"S

Sipan

0 1 2

730000

9220000

70'0"S

Yanacocha District

10 km

740000

Elevation (m)

750000

760000

770000

Slope (Degrees)

High : 4203
0-8

Porphyry Cu-Au deposit

Low : 1041
Fig. 15. Topographic map, at landscape elements and geomorphologic setting of Yanacocha and other epithermal deposits of northern Peru. UTM Zone 18S, WGS84.

Central Andes was assembled during the Paleozoic and earlier (Cediel
et al., 2003; Ramos, 2009). In the northern Andes, in contrast to the
Central Andes, high-suldation epithermal deposits are scarce, although
numerous other gold-rich deposits are known. The latter occur along
the eastern margin of the western Cordillera in northern Ecuador and
particularly throughout Colombia (Leal Meja, 2011; Leal Meja et al.,
2011; Schuette, 2010) and include the late Miocene gold-only La Colosa
porphyry deposit (Lodder et al., 2010) as well as the Marmato (Tassinari

et al., 2008) and Buritic (Lesage, 2011) low-suldation epithermal deposits. Large high-suldation epithermal deposits include Quimsacocha
in Ecuador and the CaliforniaVetas district in the eastern Cordillera of
Colombia. The northern Andes are now dominated by humid tropical
climate and thus are wetter than the western cordillera of the Central
Andes. Areas at high elevations record 12 m annual rainfall, but some
of the adjacent low-lying areas record even higher precipitation rates
of up to 10 m/yr (Alvarez Villa et al., 2011).

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

The uplift history and landscape evolution of the Northern Andes is


less well established than in the Central Andes and the geomorphology
is inuenced by the terrane architecture. However, extensive planation
surfaces have been documented in the Ecuadorian Andes (Coltorti and
Ollier, 2000) and for the central Cordillera in Colombia (RestrepoMoreno et al., 2009). In Ecuador, relict planation surfaces are commonly
exposed at 35003800 m a.s.l. The nature of these and the timing of uplift are controversial. Coltorti and Ollier (2000) interpret them have
formed as planation surfaces near sea-level as recently as early Pliocene
and having been uplifted rapidly since the Pliocene. Conversely,
Steinmann et al. (1999) documented the onset of major uplift in the
earlylate Miocene on the basis of stratigraphic relationships in
intramontane basins. A major tectonic change and increased uplift and
contractile deformation in the Andes has been associated with the arrival of the Carnegie Ridge at the subduction trench but age estimates for
this event vary widely between 15 Ma (Spikings et al., 2001) and
2 Ma (e.g., Gutscher et al., 1999b). The Colombian Andes, comprise
three separate ranges, the Cordilleras Occidental, Central and Oriental,
each of which experienced differing geomorphologic histories. Planar
surfaces have been described from the Central Cordillera and the Eastern Cordillera. A high plateau, the Antioqueo Plateau, is located at
the northern limit of the central Cordillera (6 1 Lat. N). It consists
of an extensive upper Oligocene to lower Miocene planar erosion surface which was cut from the NE into an older, uplifted early Eocene
landscape (Restrepo-Moreno et al., 2009; Villagomez and Spikings,
2013). The Antioqueo plateau is separated from higher landscape elements by a relict back-scarp and may have formed originally as a
pediplain. A renewed late Miocene to early Pliocene uplift pulse and associated incision of deep canyons is identied on the basis of apatite
UTh/He data (Villagmez and Spikings, 2013). The magnitude of late
Miocene uplift and exhumation increases south of the Antioqueo Plateau and contiguous central Cordillera (Villagmez and Spikings,
2013) but there the planar nature of the uplifted surface is obscured
by recent voluminous volcanic deposits and presumably by more
intense landscape dissection in response to uplift. In contrast to the
central Cordillera, the Eastern Cordillera was established in the late
Oligocene when the tectonic regime changed from extensional to contractile and the frontal thrust separating it from the Llanos Foreland
was established (Horton et al., 2010). However, based on paleobotanical
evidence, much of the uplift to the present day elevations of around
26003600 m a.s.l. probably took place in the late Miocene-toPliocene (Gregory-Wodzicky, 2000). The northern extension of the
eastern Cordillera which hosts the California Vetas Mining District,
may have experienced uplift pulses in the Paleocene and early Miocene
but, most importantly, in the late Miocene-to-Pleistocene (Villagmez
et al., 2011).
11.1. Quimsacocha
Quimsacocha is located in the northern Andean block (sensu Gansser,
1973), about 20 km NW of the NE-trending Girn fault which constitutes
an important terrane boundary separating the Upper Cretaceous Pin
terrane to the northwest, from the Lower Creteaceous Anaime terrane
to the southeast (Ramos, 2009; Schuette et al., 2012). The Quimsacocha
deposit adheres closely to the model for high-suldation epithermal deposits (MacDonald et al., 2011). Mineralization is hosted within a large
vuggy quartz, massive quartz and quartzalunite pyrophyllite alteration zone which is centered on the N-striking Ro Falso fault zone. The
ore is conned in pyroclastic strata between subhorizontal andesite
lava ows of the 9 Ma Quimsacocha Formation (all age dates are
40
Ar/39Ar plateau ages: MacDonald et al., 2011, 2012). The host rocks
underwent intense acid leaching resulting in vuggy quartz alteration,
prior to the deposition of pyrite, enargite and associated gold and
minor quartz (MacDonald et al., 2011). Alunite associated with the
vuggy quartz and advanced argillic alteration is dated at 7.6 to 7.3 Ma
and biotite from a post-mineral dacite dome yielded an age of 6.7 Ma

353

(MacDonald et al., 2012). Steam-heated alteration zones of limited thickness have locally been preserved in the highest parts of the deposit. The
ore-controlling Rio Falso fault is located along the eastern margin of a
caldera in which the post mineral dacitic domes lie but is a districtscale feature and does not constitute the caldera margin (Fig. 16).
Mineralization at Quimsacocha is located between 3500 and
3700 m a.s.l. The paleosurface delimited by the steam-heated alteration
zone is located at 38003900 m a.s.l. The deposit is located near the
eastern margin of a large plateau at 3800 100 m elevation with
steep margins relative to the adjacent valleys (Fig. 16; Coltorti and
Ollier, 2000). Based on sedimentological evidence and zircon ssiontrack age data for tuff layers (Hungerbuehler et al., 2002; Steinmann
et al., 1999), the area was at sea-level as recently as 1511 Ma, with uplift to almost 4000 m a.s.l. occurring in the Late Miocene, broadly coeval
with mineralization. Quimsacocha also overlies the subducting Carnegie
ridge and no recent volcanism has been documented from the area
(Chiaradia et al., 2004; Schuette et al., 2010).
11.2. California Vetas
The California Vetas Mining District is located in the northeastern
Cordillera of Colombia, some 30 km N of the city of Bucaramanga. The
modern climate in the area is tropical and has two pronounced rainy
seasons per year, with annual rainfall of around 2000 mm. The location
corresponds to the southern tip of the triangular Maracaibo block which
is bounded by the NNW-striking sinistral Santa MartaBucaramanga
fault and the NE-striking dextral Bocon fault. Mineralization is hosted
by Grenvillia-aged Bucaramanga gneisses as well as upper Triassic
to lower Jurassic peraluminous granites (Mantilla Figueroa et al.,
2013). Locally, at the El Cuatro prospect (Fig. 17), small volumes of
metaluminous, coarsely-porphyritic granodiorite dykes of late Miocene
age (10.98.4 Ma: Mantilla F. et al., 2009, 2011; Mantilla Figueroa et al.,
2013; Bissig et al., 2014) are associated with porphyry Mo mineralization (Bissig et al., 2012).
Gold mineralization is mainly vein-hosted and is largely of highsuldation epithermal type, overprinting earlier porphyry-style Mo
mineralization (Rodriguez, 2014). Numerous artisanal mines have been
operating since colonial times, but most current resources are contained
in the La Bodega/La Mascota and contiguous Angostura deposits. Mineralization was largely controlled by the NE-trending La Baja fault trend,
Angostura focused in a dextral strike-slip sigmoidal loop (Fig. 17;
Rodriguez, 2014). Several hydrothermal stages can be distinguished.
Early quartzpyrite chalcopyrite veins associated with pervasive
sericitic alteration represent precipitation in the porphyry environment
and are associated with low-grade Au mineralization (b1 g/t). These
are overprinted by several stages of veins and fault-controlled breccias
which introduced the bulk of the gold mineralization (Rodriguez,
2014). Ore minerals of the epithermal stages include covellite, bornite,
chalcopyrite, hubnerite, enargite and pyrite, all of which are associated
with native gold and goldsilver tellurides, as well as late Fe-poor sphalerite. Epithermal veins and breccias have quartz and alunite gangue that
form banded, colloform and cockade textures (Rodriguez, 2014), such
as are more typically associated with low-suldation type deposits
(Simmons et al., 2005). Conversely, vuggy residual quartz and aeriallyextensive, pervasive, quartzalunitekaolinite alteration typical of
high-suldation deposits is largely absent in the CaliforniaVetas district.
Both uid inclusion and textural evidence indicates that boiling was an
important ore depositional mechanism, occurring at 200250 C
(Rodriguez, 2014). Early muscovite alteration associated with quartz
pyrite veins at Angostura and La Bodega gives 40Ar/39Ar ages of ca. 4 to
3.5 Ma and one alunite sample from Los Laches, at ~ 3500 m a.s.l.,
some 800 m above the muscovite sample locations, was dated at
4.02 0.06 Ma and likely was deposited close to the paleosurface
(Fig. 17, Rodriguez, 2014). The age of mineralization inferred from
alunite in mineralized veins and breccias ranges from 2.6 to 2 Ma at La
Bodega and Angostura, but alunite the La Plata, San Celestino and El

354

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Quimsacocha Area, Ecuador


79100W

Quimsacocha
caldera

Ro Falso fault zone

300S

9670000

79150W

350S

9660000

Quimsacocha Deposit

0
690000

Elevation (m)

5 km

700000

Slope (Degrees)

High : 4079
0-8

Low : 2925
Fig. 16. Topographic map of the Quimsaochca area. The approximate trace and orientation of the Ro Falso fault is indicated.

Cuatro prospects downstream from La Bodega (Fig. 17) yielded ages of


3.5 to 3.2 Ma. Post-mineralization alunite associated with sphalerite
yielded ages from 1.9 to 1.6 with Ma (Rodriguez, 2014). Paragenetically
late U mineralization has been reported from San Celestino (Polania,
1980) 2 km SW of La Bodega, along the same mineralized trend. No
igneous rocks similar in age to the epithermal mineralization have
been documented, but stable-isotope data indicate a magmatic source
for the mineralizing uid (Rodriguez, 2014).

The geomorphology in the La Bodega and Angostura area is characterized by a deeply-incised, steep-walled valley paralleling the NEtrending dextral La Baja fault zone (Figs. 18, 19). Angostura is located
at the upper termination of this valley, La Bodega/La Mascota at 2800
to 2350 m, and Angostura between 3500 and 2700 m a.s.l. More
subdued topography characterizes the topographically-high areas
from ~ 3400 to 4000 m a.s.l. (Figs. 18, 19) representing the highest
parts of the northeastern cordillera of Colombia. This suggests that,

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

722'0"N

4.02 +/- 0.06 Ma


2.77 +/- 0.15 Ma

Fault
Rio Cucutil
la

fau
ill
a

Veta de Barro
ura
ost
Los Laches
Ang
La Picota
o

3.91 +/- 0.15 Ma

Perezosa
PaLa
ez f
ault

La Mascota

2.1- 1.8 Ma (n=2)

2.5 - 1.6 Ma (n=9)

El Cuatro

Pie de Gallo

ult
fa

La Bodega

San Celestino

ra
tu

e ralCu
cu
t

An
go
s

La

3.4 +/- 0.06 Ma

7252'0"W
lt

7254'0"W

Ba
ja f
aul
t

7256'0"W

355

3.26+/- 0.30 Ma
3.23+/- 0.06 Ma

1305000

La Plata
3.43 +/- 0.07 Ma

Violetal

Geologic Units
Pliocene

California

Hydrothermal breccia/veins

Late Miocene
Hydrothermal breccia

Faults & Contacts


Normal fault
Normal fault, inferred
Thrust fault

720'0"N

Strike-slip, dextral
Porphyry

Non conformity

Late Cretaceous

Santander

Tambor Formation

Late Triassic to Early Jurassic

Bucaramanga
Antioquia

Rosablanca Formation

California Vetas
Mining District

Bolivar

Arauca

Town
Prospect/Artisanal Mine
40Ar/39Ar ages

Diorite to Granodiorite

Alunite

Leucogranite

Sericite

Boyaca

Casanare

1125000

1300000

Proterozoic
Bucaramanga Gneiss

1130000

1135000

Fig. 17. Geological map of the La Baja Trend, California Vetas Mining District, Santander Colombia. Prospect locations as well as alunite and sericite ages are shown. Modied from
Rodriguez (2014). Coordinates given are geographic and Colombian Gauss with Bogota Observatory datum.

despite the wet climate, the topographically high areas are not in erosional equilibrium with the present day base-level, as in northern
Chile and Peru. This high elevation surface is interpreted as a relict
lower to middle Miocene paleosurface.
Based on apatite ssion track data, uplift and valley incision commenced around 17 Ma (Van Der Lelij, 2013) but the most important uplift pulse probably occurred in the Pliocene (Gregory-Wodzicky, 2000;
Shagam et al., 1984; Villagomez et al., 2011). Thus, uplift, erosion and
epithermal mineralization overlapping in age and the fact that the alunite ages along the La Baja trend become younger upstream strongly
suggests that erosion is directly stimulating epithermal mineralization.
12. Summary and comparison to low-suldation deposits
High-suldation epithermal deposits of the Andes occur over a geographically wide area, yet were emplaced in remarkably predictable
geologic and geomorphologic settings. Virtually all high-suldation
epithermal deposits formed during episodes of major uplift in the respective Andean segment in which they are hosted. The uplifted landforms hosting epithermal mineralization are mostly extensive low

relief surfaces, in many cases demonstrably pediplains, located at the


crest of the Andes at the time of mineralization. The low-relief surfaces
were commonly subject to erosion and valley pediments or valleys were
being incised during mineralization. A direct temporal and spatial
relationship of erosion and mineralization has been documented for a
number of epithermal districts including the El Indio belt deposits,
Pierina, Lagunas Norte and CaliforniaVetas.
Most high-suldation epithermal deposits were emplaced above
segments of at subduction where no volcanism is presently observed.
The few exceptions to the latter occur in the western Cordillera near the
Andean orocline in southern Peru and northern Chile. The vast majority
of deposits are 17 Ma or younger, but exceptions include the Eocene El
Hueso, and Guanaco deposits in the southern Atacama Desert near the
Domeyko fault system, and possibly the poorly documented Quicay
deposit as well as the late Oligocene to early Miocene La Coipa deposit
for which age constraints are ambiguous. Large volcanic edices are uncommon hosts to ore. Host rocks to high-suldation systems commonly
pre-date mineralization by several Ma to as much as 1 Ga, as at
La Bodega and Angostura in Colombia. However, examples exist
where the ages of ore-hosting volcanic rocks overlap with those of

356

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

California Vetas Mining District, Colombia


72500W

7250N

72550W

Angostura

La Bodega/La Mascota
El Cuatro
San Celestino

La Plata

7200N

California

Vetas

Elevation (m)

5 km

Slope (Degrees)

High : 4272
0 - 16

Low : 1358

High elevation low relief paleosurface

Town
Prospect

Fig. 18. Topographic map of the CaliforniaVetas Mining District, showing steep valleys incised into a high-elevation relatively low relief landscape. WGS84.

alteration and mineralization (e.g., Yanacocha, Lagunas Norte, Aruntani)


but where this temporal overlap exists, the volcanic rocks contemporaneous with mineralization are, with exception of Aruntani, of low
volume and volcanism ceased shortly after mineralization. Volcanic
rocks of signicant volume covering and/or post-dating mineralization
are reported from La Pepa and La Coipa.

Andean low-suldation epithermal deposits are commonly smaller


and many occur in geologic settings distinct from those of their highsuldation counterparts. Some of the larger examples include the Jurassic Fruta del Norte deposit in Ecuador (Henderson, 2009), the Paleocene
El Pen deposit of northern Chile (Warren et al., 2004, 2007). The Late
Jurassic to Early Cretaceous deposits of the Deseado and Patagonian

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

California
San Celestino
El Cuatro

La Mascota
La Bodega

Fig. 19. Photograph looking SW from Angostura, showing the approximate locations of
prospects. Note the steeply incised La Baja valley and the less rugged higher elevation
terrain.

Massifs (Dietrich et al., 2011; Echavarria et al., 2005; Fernndez et al.,


2008) are, strictly speaking, located outside the Andes but the Cerro
Bayo low-suldation deposit considered the westernmost deposit of
this low suldation epithermal province is located in the Chilean
Patagonian Andes (Poblete et al., 2014). Mineralization here took
place in several pulses between 144 and 113 Ma and partly overlaps
in age with the slightly older low-suldation epithermal deposits of
the Deseado and north Patagonian massifs (Poblete et al., 2014).
The low-suldation deposits mentioned above occur at elevations
below 2400 m a.s.l., are hosted by dominantly rhyolitic volcanic and
volcaniclastic rocks, lack a spatial or temporal association with porphyry
systems (as dened by Sillitoe, 2010) and occur in extensional tectonic
settings without documented contractile deformation and surface uplift
during mineralization. However, there are also examples of epithermal
deposits containing low-suldation mineralization that occur within
larger magmatic-hydrothermal systems in arc segments also containing
porphyry style or high-suldation epithermal mineralization. Examples
of these include Cerro de Pasco (Peru: Baumgartner et al., 2008),
Marmato and Buritic (Colombia: Tassinari et al., 2008; Lesage, 2011)
which are all mid to late Miocene in age. The low-suldation nature of
epithermal deposits of some contractile arc settings may be attributed
to more reactive, locally-reducing host or basement rock characteristics
(e.g., Marmato: Tassinari et al., 2008). Thus, as proposed earlier (Sillitoe
and Hedenquist, 2003) low-suldation deposits may be associated with
both, overall contractile arc as well as rift settings. Those associated with
rift settings and rhyolitic magmatism evidently are more likely to be
preserved over longer time intervals than epithermal deposits in
contractile settings.
13. Controls of geomorphic processes and climate on mineralization
Epithermal deposits are emplaced at shallow crustal levels, in the
case of high-suldation epithermal deposits at depths commonly less
than 500 m (see above). Such shallow hydrothermal environments
are directly inuenced by surface processes and climatic conditions.
Precious metals in the epithermal environment are typically precipitated by processes of boiling or uid mixing (Simmons et al., 2005). The
depth of boiling in the epithermal environment is controlled by the
hydrostatic pressure which depends on the elevation of the water
table. A lowering of the water table during hydrothermal activity such
as would occur during catastrophic erosion or volcanic sector collapse
would enhance boiling and can be important for the efciency of precious metal deposition (Bissig et al., 2002a; Simmons, 1991). This can

357

lead to the superposition of epithermal mineralization on the porphyry


environment during the life-span of a single magmatic-hydrothermal
system, a process commonly referred to as telescoping (Sillitoe,
1994). At the giant Ladolam deposit in Papua New Guinea, volcanic sector collapse eliminated more than 500 m of rock cover very rapidly, and
is thought to be integral to epithermal ore formation (Blackwell et al.,
2014; Carman, 2003; Sillitoe, 1994). While Ladolam has a lowsuldation sulde assemblage and is associated with alkalic magmatism
and has, thus, uid chemistry differing from that of Andean highsuldation epithermal systems, the gold transport and precipitation
processes are still comparable, Au bisulde complexes being probably
important in both (e.g., Heinrich et al., 2004). Sector collapse and
telescoping have also been proposed for Marte and other porphyry
Au deposits in the Maricunga belt (Muntean and Einaudi, 2001;
Sillitoe, 1994), indicating that catastrophic erosion during hydrothermal
activity is not unique to Ladolam. However, in the absence of large
volcanic landforms which are prone to rapid degradation, erosion in response to uplift events can enhance mineralizing processes (Bissig et al.,
2002a) and may exert a rst-order control on exsolution of uids from
magmas. This hypothesis is supported by the fact that Eocene and Miocene porphyry and high-suldation epithermal deposits in the Central
Andes were emplaced during contractile deformation, uplift and largescale erosion on the western Andean slope, which occurred during the
Eocene Incaic, late Oligoceneearly Miocene Aymar and Miocene Quechua orogenic phases. On a more local scale, valley or pediment incision
during hydrothermal activity can lead to lowering of the water-table
near the head of the incising valleys and, by generating local topography, may lead to increased lateral groundwater ow and enhanced
mixing of magmatic with meteoric uids. This process can enhance boiling and uid mixing especially near the shoulders or back-scarps of incising pediments (e.g. El Indio belt: Bissig et al., 2002a, Lagunas Norte:
Montgomery, 2012) or fault induced steep topographic gradients (potentially the case in the Famatina district: Losada-Caldern et al.,
1994; Pudack et al., 2009). The time scale of valley incision is well within
the proposed duration of hydrothermal activity of individual hydrothermal systems, which is in the order of 10 to 100 k.y. (e.g., Simmons, 2002;
Simmons and Brown, 2006). Erosion can in part inuence the locus of
mineralization over time. The example of Veladero and Pascua as well
as the deposits along the La Baja trend in the California Vetas district illustrate this relationship (Figs. 20, 21). In both cases, older mineralization occurs downstream from younger mineralization. Perhaps the
best temporal control is available for the La Baja trend where at
3.5 Ma, alunite and associated Au mineralization formed to the west
and downstream of contemporaneous higher temperature phyllic alteration, which formed under 500 m or more of cover rock (Rodriguez,
2014; Fig. 20). At a later stage, the La Baja valley incised northeastward
and subsequent epithermal hydrothermal alteration and mineralization
overprinted the phyllic alteration at La Bodega and Angostura (Fig. 20).
In the case of Veladero and Pascua, no constraints for phyllic alteration
overprinted by high-suldation epithermal mineralization are available.
However, mineralization at the Filo Federico zone at Veladero is at
1110.3 Ma distinctly older than 9.5 Ma mineralization at the Penelope
ore zone and 9.18.1 Ma mineralization at Pascua which occur upstream
from Filo Federico (Fig. 21). At the time of mineralization at Pascua,
Veladero was already subject to oxidation, as suggested by the jarosite
ages of Veladero, roughly contemporaneous with hypogene alunite formation at Pascua (Deyell et al., 2005b; Holley, 2012).
Andean high-suldation epithermal deposits occur over a wide
range of climatic zones from humid tropical to hyper-arid climate. The
Miocene climate of the Central Andes was probably slightly more
humid than at present but probably not fundamentally different as
South America was located at similar latitudes as at present (Clarke,
2006). Thus, the availability of meteoric uids during hydrothermal activity does not seem to be the principal controlling factor for mineralization. This is in agreement with the growing number of stable-isotope
studies, all of which propose a dominantly magmatic source for the

358

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

La Bodega

10 Ma

NE
Los Laches

Angostura

Paleosurface

es
en
td
ay
su
rfa
ce

Perezosa

La Bodega

La Mascota

El Cuatro

San Celestino

3,000

Pr

Elevation (m)

3,500

La Plata

California

SW

2,500

Cu?

2,000
0

Elevation (m)

3,500

Mo Mo
Cu?

500

1,000 1,500 2,000 2,500 3,000 3,500 4,000 4,500 5,000 5,500 6,000 6,500 7,000 7,500

Distance (m)

4 - 3.25 Ma

4 Ma

3,000

3.

e
al

os

ur

e
ac

~
2,500

2
3.

Au
Ag

2,000
0

500

rfa

ce

Au
Au

3.25 Ma

3.9 Ma Cu

3.4 Ma

Au Au
Ag Ag

Au
Cu

1,000 1,500 2,000 2,500 3,000 3,500 4,000 4,500 5,000 5,500 6,000 6,500 7,000 7,500

Elevation (m)

le

3.25 Ma
3.5 Ma

3,500

pa

u
os

Distance (m)

~ 2.5 - 2 Ma

2.7 Ma

Advanced argillic
Sericite/illite quartz
Potassic, K-feldspar, biotite
Porphyries
Ar/Ar on alunite age Alunite < 2 Ma
Ar/Ar on sericite
Re/Os molybdenite
U/Pb zircon

3,000

2,500

2.1 Ma
2.2 Ma
2.6-2.3 Ma

U
2,000
0

500

Au
Ag 1.9 Ma
Zn

1.6 Ma

Au
Ag
(Cu)

Au

1.8 Ma Ag

(Cu)

1,000 1,500 2,000 2,500 3,000 3,500 4,000 4,500 5,000 5,500 6,000 6,500 7,000 7,500

Distance (m)
Fig. 20. Schematic relationships between erosion, tectonics and mineralization along the La Baja trend, CaliforniaVetas Mining District, Colombia. Based on data from Rodriguez (2014);
Refer to Figs. 17 and 18 for context. A) snap-shot of the landscape conguration in relation to hydrothermal systems at ~10 Ma, at the time of porphyry intrusion and Mo (Cu) mineralization. B) Distribution of hydrothermal alteration in relation to the eroding La Baja valley at ~3.5 to 3.25 Ma. Note that at La Bodega and Angostura sericite forms at a depth of
7001000 m below surface, whereas further downstream at alunite, spatially associated with Au mineralization forms at only a few 100 m depth at the same time. Alunite also forms
near the paleosurface at Los Laches which is the topographically highest prospect of the area. C) At 2.52 Ma, during the main stage of epithermal Au (Ag, Cu) mineralization at La Bodega,
La Mascota and Angostura, erosion is further advanced and overprints 3.53.9 Ma sericite. Note, Los Laches has not been affected by signicant erosion since the early Pliocene and 4 Ma
and 2.7 Ma alunite formed at the same place.

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

359

NW
Pascua
9.1-8 Ma

SE
Penelope
9.5 Ma

Ro de las Taguas
N-S valley

5000
4750

Lama Central
barren alteration
13.3 Ma

4500
4250

Filo Federico
11.1-10.3 Ma

4000
Frontera-Deidad surface
Azufreras-Torta surface
Los Ros surface

Steam-heated alteration

Present day Ro Turbio profile

16-~12.5 Ma volcaniclastic rocks


Mineralized zones

Future mineralized zones

Jarosite formation

B
5000
4750
4500
4250
4000

Valley profile at 9.5 Ma


Present day Ro Turbio profile

C
5000
4750
4500
4250
4000

Valley profile at 10.5 Ma


Present day Ro Turbio profile

Fig. 21. Schematic relationship between erosion, tectonics and mineralization at Veladero and PascuaLama, Argentina, Chile. The cartoon is based on data from Bissig et al. (2002a),
Charchai et al. (2007) and Holley (2012). A) Present-day conguration of landscape elements, faults, mineralized zones. B) Snap-shot of the landscape conguration at 9.5 Ma, i.e., at
the initiation of hydrothermal activity at Penelope. At that time, Veladero was already subject to oxidation and jarosite formation. C) Snap-shot of the landscape conguration at
10.5 Ma, during hypogene mineralization at Filo Federico.

precious metal-bearing uids (e.g., Cerpa et al., 2013; Deyell et al., 2004;
Rainbow et al., 2005; Rye, 2005). However, the hydrothermal systems
overall are inuenced by the climate. Thus, the position of the water
table is inuenced by the climate and the extent of near surface
steam-heated alteration formed in the vadose zone varies. Up to several
100 m thick steam-heated blankets overlying and overprinting mineralization are observed in the El Indio belt and La Coipa. These are
interpreted as evidence for dry climate at the time of mineralization
and general water table lowering during hydrothermal activity
(e.g., Bissig et al., 2002a; Holley, 2012), but short term water table uctuations due to episodic incursion of magmatic uid or periods of more
humid climate has been inferred from hypogene alunite or barite
overprinting jarosite (Chouinard et al., 2005; Deyell et al., 2005b;
Holley, 2012). In contrast to deposits of northern Chile, in the more
humid Miocene climates of Peru, Ecuador or Colombia where the
water-table was probably much closer to the land-surface, steamheated alteration blankets tend to be less well developed or even lacking (McDonald et al., 2011; Rainbow et al., 2005; Rodriguez, 2014;
Teal and Benavides, 2010). Although this difference may be due to
better preservation under dry climatic conditions, the persistence of
high-elevation paleosurfaces in northern Peru and Ecuador that predate hydrothermal activity suggests that steam-heated blankets should
have been forming but were probably never as extensive in those areas
as in dryer climatic zones, and could therefore have been removed by
modest erosion.

Most high-suldation epithermal deposits were emplaced between


17 and 6 Ma, irrespective of the climatic zone in which they are located.
There is no clearcut overall correlation of younger age with better preservation within this age range. The location and age of currently exposed Miocene high-suldation epithermal deposits is, thus, not only
a function of the preservation potential but also due to tectonomagmatic factors. However, the Eocene deposits of El Guanaco and El
Hueso are located in one of the most arid climatic zone of the planet,
whereas the youngest deposits La Bodega and Angostura in Colombia
are located in the wettest climatic zone in which high-suldation
epithermal deposits are known in the Andes. This indicates that a
longer-term control on preservation potential does exist. The fact that
no high-suldation epithermal deposits older than Eocene are known,
together with the fact that they generally were emplaced near surface
during uplift indicates that the preservation of high-suldation
epithermal deposits older than the Eocene is unlikely except in the driest climatic zones. Oxidation of primary sulde assemblages is an important factor for the economic viability of high-suldation epithermal
deposits, as it may liberate encapsulated gold. Oxidation occurs where
suldes are exposed to atmospheric oxygen and water and is favored
by decreasing water table as well as sufcient rock permeability. Some
examples of high-suldation epithermal deposits of Northern Peru as
well as the Pampean at slab are oxidized to considerable depth
(e.g., Yanacocha, Lagunas Norte, Pierina, Veladero), whereas others
(e.g., PascuaLama, Cerro de Pasco) have undergone considerably less

360

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Although low-density magmatic vapor may be responsible for early


vuggy quartz alteration, later high-suldation epithermal mineralization hosted in vuggy quartz likely precipitated from contracted magmatic vapor exsolved from a deeper porphyry intrusion (Heinrich
et al., 2004; Pudack et al., 2009). This evolution from shallow toward
deeper locus of uid exsolution is consistent with waning magmatism
and increasing SiO2 content of magmas; as is evident for the majority
of Andean high-suldation epithermal districts. Large-scale tectonic
changes such as the onset of at subduction, increased plate coupling,
contractile deformation and uplift result in decreasing volumes of
eruptive magmatism and increasing depth of porphyry emplacement
and may favor high-suldation epithermal Au and associated deeper
porphyry Cu over porphyry CuAu mineralization.

extensive supergene oxidation. In the case of PascuaLama and


Veladero as well as the La Coipa district, oxidized and non-oxidized
deposits occur within b 15 km from each other, which indicates that
the degree of oxidation cannot simply ascribed to climatic factors.
Rock-permeability or physiographic setting as well as the posthydrothermal history inuence the oxidation and hence economic
viability of epithermal deposits as well.
14. Igneous rocks, volcanology and magmatic uids related
to high-suldation epithermal deposits
Although magmas are the primary contributors of metals and volatiles in porphyry systems (e.g., Audtat and Simon, 2012), extrusive
magmatism is commonly waning, of low volume or absent during the
formation of high-suldation epithermal AuAg deposits in the Andes.
Conversely, Au-rich porphyry deposits in the Maricunga belt and
in northern Peru are typically emplaced in volcanic centers or associated
with shallowly emplaced magmatic bodies (e.g. Cerro Casale: Muntean
and Einaudi, 2001). Murakami et al. (2010) showed that Cu/Au ratios increase with greater depth of emplacement of porphyry systems and
suggested that high-suldation epithermal Au mineralization occurs
where signicant amounts of Au has been physically separated from
Cu, a process which is apparently favored where the depth of intrusion
of the magma providing the magmatic volatiles is N 3 km (Murakami
et al., 2010). A compelling explanation for this is that precious metals
are better transported in high-density magmatic vapors released from
magmas emplaced at more than 34 km depth than in low-density vapors exsolved at depths of less than 2 km (Heinrich, 2005, 2007; Fig. 22).
Fluid evolution at depth can lead to porphyry Cu mineralization but
allows for efcient Au transport in contracted magmatic vapor. If
condensed magmatic vapors are efciently transported and focused, a
high-suldation epithermal Au deposit may form several km above.

High-suldation epithermal deposits of the Andes formed in predictable geological and geomorphological settings. They are located at
high elevation and largely near the crest of the Andean cordillera in segments where volcanism is currently absent or subdued and where subduction angles are shallow. All were emplaced during major periods of
contractile deformation and uplift and most coincided with the terminal
stages of local arc magmatism. The vast majority of deposits are between ca. 17 and 5 Ma old. Although the tectonic setting in which
these deposits form makes them prone to erosion, their restricted age
range is not only a function of preservation potential but also attributed
to favorable tectonomagmatic settings at the time of hydrothermal activity. Eocene high-suldation epithermal deposits are only known
from the driest regions of the Atacama Desert but they plausibly formed
under similar tectonomagmatic conditions as the Miocene ones. Conversely, the youngest deposits, La Bodega and Angostura, Colombia,

B
v

1 km

15. Conclusions

vvvv
vvvv

Au

v
v

v
Au

Cu

Au

Au
v

vvvv

vvvv
v

2 km

3 km
Cu
Mo

4 km

Cu
Mo

Cu
Mo

Steam-heated alteration

Diatreme breccia

High-sulfidation epithermal Au

Water table

Porphyry Cu-Au

vvvv

Volcanic rock

Porphyry Cu-Mo
Fig. 22. Schematic relationships between depth of intrusion volcanic setting and mineralizatiion style, inspired by Murakami et al. (2010) and relationships observed in the Andes. Three
scenarios are shown. A) Porphyry AuCu hosted in a volcanic edice such as observed in the Maricunga belt (e.g. Cerro Casale). Depth of intrusion and uid exsolution is b2 km, magmatic
vapor cannot dissolve signicant Au and temperature gradient to surrounding rocks is steep, leading to bulk co-precipitation of Cu and Au. B) High-suldation epithermal deposit forms
N34 km above porphyry intrusion. At that depth, magmatic vapor has a higher density and is, after contraction, capable of transporting signicant Au as bisulde complexes at low uid
temperatures (Heinrich et al., 2004), whereas Cu and Mo precipitate in the higher-T porphyry environment. Erosion at surface stimulates precipitation of precious metals but through
reduction of lithostatic load may enhance uid release from the magma and the structural pathways permitting efcient separation of magmatic vapor derived uids from the
porphyry Cu environment (this scenario reects the situation at La Bodega and Angostura, Colombia as well as El Indio and PascuaLama). C) A scenario where multiple pulses of
both, porphyry CuAu and high-suldation epithermal mineralization occurred. This scenario reects the situation at Yanacocha. Volcanism in the form of ow-domes and pyroclastic
ow deposits occurred intermittedly as well.

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

were emplaced in the late Pliocene in an area with pluvial tropical


climate with low preservation potential.
High-suldation epithermal AuAg deposits can be emplaced in a
range of host rocks, and volcanic rocks both pre- or syn-hydrothermal
activities, are not a pre-requisite for epithermal mineralization. However, a magma from which the mineralizing magmatic uids are derived
must have present, albeit uid exsolution typically occurred N34 km
below the surface at the time of hydrothermal activity.
Signicant erosion during hydrothermal activity is documented for a
number of districts (e.g., Lagunas Norte, Pierina, El Indio belt, California
Vetas Mining District) and mineralization ages generally become younger upstream in both, the El Indio belt and the California Vetas Mining
District. Fluid boiling and mixing with meteoric water are enhanced
near topographic breaks such as the heads of incising valleys or backscarps of pediments, ultimately stimulating mineralization in those
locations.

Acknowledgments
This paper reviews and summarizes many years of work on Andean
epithermal systems, much of it carried out at Queen's University. The
authors would like to express their special thanks to Barrick Gold
Corp. and former chief geologist Jay Hodgson for the continued support
and funding of TB's, AM's and AR's PhD theses at Queen's University.
Barrick also funded projects at MDRU and Universidad Catlica del
Norte, Chile in which the senior author continues to be involved. Without Barrick's support, our understanding of Andean high-suldation
epithermal deposits would be far less advanced.
We also acknowledge the contributions and student support from
NSERC in the form of grants to AHC and to Kurt Kyser, as well as from
many other companies who supported research at Queen's University
and MDRU, including (but not limited to) IamGold, Ventana Gold, Eco
Oro minerals and Kinross.
Sara Jenkins is acknowledged for help with the gures and we thank
the OGR editors Franco Piranjo and Tim Horscroft for the invitation of
this review article. Comments provided by reviewers Stuart Simmons
and Dick Tosdal helped improve clarity of the paper.

References
Aguilar, G., Riquelme, R., Martinod, J., Darrozes, J., Maire, E., 2011. Variability in erosion
rates related to the state of landscape transience in the semi-arid Chilean Andes.
Earth Surf. Process. Landf. 36, 17361748.
Aguilar, G., Riquelme, R., Martinod, J., Darrozes, J., 2013. Role of climate and tectonics in
the geomorphologic evolution of the Semiarid Chilean Andes between 2732 S.
Andean Geol. 40, 79101.
lvarezVilla, O.D., Vlez, J.I., Poveda, G., 2011. Improved longterm mean annual rainfall
elds for Colombia. Int. J. Climatol. 31, 21942212.
Arribas, A., 1995. Characteristics of high-suldation epithermal deposits, and their
relation to magmatic uid. Mineral. Assoc. Can. Short Course Ser. 23, 419454.
Audtat, A., Simon, A.C., 2012. Magmatic controls on porphyry copper deposits. In:
Hedenquist, J.W., Harris, M., Camus, F. (Eds.), Geology and Genesis of Major Copper
Deposits and Districts of the World: A Tribute to Richard H. Sillitoe. Society of
Economic Geologists Special Publication 16, pp. 573618.
Barreda, J., Loayza, D., Juarez, P., Torres, R., 2004. Depsitos epitermales de alta sulfuracin
en el distrito minero Aruntani, Moquegua, XII Congreso Peruano de Geologa, Lima,
Peru. pp. 605608.
Baumgartner, R., Fontbote, L., Vennemann, T., 2008. Mineral zoning and geochemistry of
epithermal polymetallic ZnPbAgCuBi mineralization at Cerro de Pasco, Peru.
Econ. Geol. 103, 493537.
Baumgartner, R., Fontbote, L., Spikings, R., Ovtcharova, M., Schaltegger, U., Schneider, J.,
Page, L., Gutjahr, M., 2009. Bracketing the age of magmatic-hydrothermal activity at
the Cerro de Pasco epithermal polymetallic deposit, central Peru; a UPb and
40
Ar/39Ar study. Econ. Geol. 104, 479504.
Bendez, R., Fontbot, L., 2009. Cordilleran Epithermal CuZnPb(AuAg) Mineralization in the Colquijirca District, Central Peru: Deposit-Scale Mineralogical Patterns.
Econ. Geol. 104, 905944.
Bendez, R., Fontbot, L., Cosca, M., 2003. Relative age of Cordilleran base metal lode and
replacement deposits, and high suldation Au(Ag) epithermal mineralization in the
Colquijirca mining district, central Peru. Mineral. Deposita 38, 683694.
Bendez, R., Page, L., Spikings, R., Pecskay, Z., Fontbot, L., 2008. New 40Ar/39Ar alunite
ages from the Colquijirca district, Peru: evidence of a long period of magmatic SO2

361

degassing during formation of epithermal AuAg and Cordilleran polymetallic ores.


Mineral. Deposita 43, 777789.
Beuchat, S., Moritz, R., Pettke, T., 2004. Fluid evolution in the WCuZnPb San Cristobal
vein, Peru: uid inclusion and stable isotope evidence. Chem. Geol. 210, 201224.
Bissig, T., Riquelme, R., 2009. Contrasting landscape evolution and development of supergene enrichment in the El Salvador porphyry Cu and PotrerillosEl Hueso CuAu districts, Northern Chile. In: Titley, S. (Ed.), Society of Economic Geologists Special
Publication No. 14: Supergene Environments. Processes and Products, pp. 5968.
Bissig, T., Riquelme, R., 2010. Andean uplift and climate evolution in the southern
Atacama Desert deduced from geomorphology and supergene alunite-group
minerals. Earth Planet. Sci. Lett. 299, 447457.
Bissig, T., Tosdal, R.M., 2009. Petrogenetic and Metallogenetic Relationships in the Eastern
Cordillera Occidental of Central Peru. J. Geol. 117, 499518.
Bissig, T., Lee, J.K.W., Clark, A.H., Heather, K.B., 2001. The cenozoic history of volcanism
and hydrothermal alteration in the central Andean at-slab region: New 40Ar39Ar
constraints from the El IndioPascua Au (Ag, Cu) belt, 29 20 30 30 S. Int.
Geol. Rev. 43, 312340.
Bissig, T., Clark, A.H., Lee, J.K.W., Hodgson, C.J., 2002a. Miocene landscape evolution and
geomorphologic controls on epithermal processes in the El IndioPascua AuAgCu
belt, Chile and Argentina. Econ. Geol. Bull. Soc. 97, 971996.
Bissig, T., Clark, A.H., Lee, J.K., 2002b. Cerro de Vidrio rhyolitic dome: evidence for Late
Pliocene volcanism in the central Andean at-slab region, LamaVeladero district,
29 20 S, San Juan Province, Argentina. J. S. Am. Earth Sci. 15, 571576.
Bissig, T., Clark, A.H., Lee, J.K.W., von Quadt, A., 2003. Petrogenetic and metallogenetic
responses to Miocene slab attening: new constraints from the El IndioPascua
AuAgCu belt, Chile/Argentina. Mineral. Deposita 38, 844862.
Bissig, T., Ullrich, T.D., Tosdal, R.M., Friedman, R., Ebert, S., 2008. The timespace distribution of Eocene to Miocene magmatism in the central Peruvian polymetallic province
and its metallogenetic implications. J. S. Am. Earth Sci. 26, 1635.
Bissig, T., Mantilla, F.L.C., Rodriguez, M.A., Raley, C.A., Hart, C.J.R., 2012. The California
Vetas District, Eastern Cordillera, Santander, Colombia: Late Miocene Porphyry and
Epithermal Mineralization Hosted in Proterozoic Gneisses and Late TriassicEarly
Jurassic Intrusions, XVI Congreso Peruano de Geologa & SEG2010 Conference, Lima,
Peru, (Poster 11).
Bissig, T., Mantilla Figueroa, L.C., Hart, C.J., 2014. Petrochemistry of igneous rocks of
the CaliforniaVetas mining district, Santander, Colombia: Implications for
northern Andean tectonics and porphyry Cu (Mo, Au) metallogeny. Lithos
200, 355367.
Blackwell, J.L., Cooke, D.R., McPhie, J., Simpson, K.A., 2014. Lithofacies Associations and
Evolution of the Volcanic Host Succession to the Minie Ore Zone: Ladolam Gold
Deposit, Lihir Island, Papua New Guinea. Econ. Geol. 109, 11371160.
Bouzari, F., Clark, A.H., 2002. Anatomy, evolution, and metallogenic signicance of the supergene orebody of the Cerro Colorado porphyry copper deposit, I region, northern
Chile. Econ. Geol. 97, 17011740.
Bussell, M.A., Alpers, C.N., Petersen, U., Shepherd, T.J., Bermudez, C., Baxter, A.N., 1990. The
AgMnPbZn vein, replacement, and skarn deposits of Uchucchacua, Peru; studies
of structure, mineralogy, metal zoning, Sr isotopes, and uid inclusions. Econ. Geol.
85, 13481383.
Cahill, T., Isacks, B.L., 1992. Seismicity and shape of the subducted Nazca plate. J. Geophys.
Res. 97, 1750317517 (17529).
Canavan, R.R., Carrapa, B., Clementz, M.T., Quade, J., DeCelles, P.G., Schoenbohm, L.M.,
2014. Early Cenozoic uplift of the Puna Plateau, Central Andes, based on stable
isotope paleoaltimetry of hydrated volcanic glass. Geology 42, 447450.
Carman, G.D., 2003. Geology, mineralization, and hydrothermal evolution of the Ladolam
gold deposit, Lihir Island, Papua New Guinea. In: Simmons, S.F., Graham, I. (Eds.),
Volcanic. Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of Processes
Within the Earth, Special Publication-Society of Economic Geologists 10, pp. 247284.
Catchpole, H., Kouzmanov, K., Fontbote, L., Guillong, M., Heinrich, C.A., 2011. Fluid evolution in zoned cordilleran polymetallic veins; insights from microthermometry and
LAICP-MS of uid inclusions. Chem. Geol. 281, 293304.
Cecioni, A.J., Dick, L.A., 1992. Geologa del yacimiento epitermal de oro y plata Can Can,
Franja de Maricunga, precordillera de Copiapo, Chile. Andean Geol. 19, 317.
Cediel, F., Shaw, R.P., Cceres, C., Bartolini, C., 2003. Tectonic assembly of the Northern
Andean block. The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats.
Basin Form. Plateau Tectonics 79, 815848.
Cerpa, L.M., Bissig, T., Kyser, K., McEwan, C., Macassi, A., Rios, H.W., 2013. Lithologic controls on mineralization at the Lagunas Norte high suldation epithermal gold deposit,
northern Peru. Mineral. Deposita 48, 653673.
Charchaie, D., Tosdal, R.M., Mortensen, J.K., 2007. Geologic framework of the Veladero
high-suldation epithermal deposit area, Cordillera Frontal, Argentina. Econ. Geol.
Bull. Soc. 102, 171192.
Charrier, R., Baeza, O., Elgueta, S., Flynn, J.J., Gans, P., Kay, S.M., Muoz, N., Wyss, A.R.,
Zurita, E., 2002. Evidence for Cenozoic extensional basin development and tectonic
inversion south of the at-slab segment, southern Central Andes, Chile (3336S.L.).
J. S. Am. Earth Sci. 15, 117139.
Chiaradia, M., Fontbote, L., Beate, B., 2004. Cenozoic continental arc magmatism and
associated mineralization in Ecuador. Mineral. Deposita 39, 204222.
Chouinard, A., Williams-Jones, A.E., Leonardson, R.W., Hodgson, C.J., Silva, P., Tellez, C.,
Vega, J., Rojas, F., 2005. Geology and genesis of the multistage high-suldation
epithermal Pascua AuAgCu deposit, Chile and Argentina. Econ. Geol. 100, 463490.
Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., France, L.J., McBride, S.L., Woodman, P.L.,
Wasteneys, H.A., Sandeman, H.A., Douglas, A., Archibald, D.A., 1990. Geologic and
geochronologic constraints on the metallogenic evolution of the Andes of southeastern Peru. Econ. Geol. 85, 15201583.
Clarke, J.D.A., 2006. Antiquity of aridity in the Chilean Atacama Desert. Geomorphology
73, 101114.

362

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Coltorti, M., Ollier, C., 2000. Geomorphic and tectonic evolution of the Ecuadorian Andes.
Geomorphology 32, 119.
Cooke, D.R., Hollings, P., Walshe, J.L., 2005. Giant porphyry deposits: characteristics,
distribution, and tectonic controls. Econ. Geol. 100, 801818.
Cornejo, P., Mpodozis, C., Ramrez, C., Tomlinson, A., 1993. Estudio geolgico de la Regin de
Potrerillos y El Salvador (2627 Lat. S). Servicio Nacional de Geologa y MineraCorporacin del Cobre, informe registrado IR-9301, 12 cuadrngulos escala 1:50.000,
258 p.
Cornejo, P., Mpodozis, C., Tomlinson, A.J., 1998. Hoja Salar de Maricunga, Regin de
Atacama. Servicio Nacional de Geologa y Minera, Mapas Geolgicos N 7, 1 mapa
escala 1:100.000. Santiago.
Deckart, K., Clark, A., Cuadra, P., Fanning, M., 2013. Renement of the timespace
evolution of the giant Mio-Pliocene Ro BlancoLos Bronces porphyry CuMo cluster,
Central Chile: new UPb (SHRIMP II) and ReOs geochronology and 40Ar/39Ar
thermochronology data. Mineral. Deposita 48, 5779.
Deen, J.A., Rye, R.O., Munoz, J.L., Drexler, J.W., 1994. The magmatic hydrothermal system
at Julcani, Peru: evidence from uid inclusions and hydrogen and oxygen isotopes.
Econ. Geol. 89, 19241938.
Deyell, C., Bissig, T., Rye, R., 2004. Isotopic evidence for magmatic-dominated epithermal
processes in the El IndioPascua AuCuAg Belt and relationship to geomorphologic
setting. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and Updates. Society of Economic Geologists Special Publication 11,
pp. 5574.
Deyell, C.L., Rye, R.O., Landis, G.P., Bissig, T., 2005a. Alunite and the role of magmatic uids
in the Tambo high-suldation deposit, El IndioPascua belt, Chile. Chem. Geol. 215,
185218.
Deyell, C.L., Leonardson, R., Rye, R.O., Thompson, J.F.H., Bissig, T., Cooke, D.R., 2005b.
Alunite in the PascuaLama high-suldation deposit: Constraints on alteration and
ore deposition using stable isotope geochemistry. Econ. Geol. 100, 131148.
Dietrich, A., Gutierrez, R., Nelson, E., Layer, P., 2011. Geology of the epithermal AgAu
Huevos Verdes vein system and San Jos district, Deseado massif, Patagonia,
Argentina. Mineral. Deposita 117.
Echavarra, L.E., Schalamuk, I.B., Etcheverry, R.O., 2005. Geologic and tectonic setting of
Deseado Massif epithermal deposits, Argentina, based on El DoradoMonserrat. J. S.
Am. Earth Sci. 19, 415432.
Einaudi, M.T., 1977. Environment of ore deposition at Cerro de Pasco, Peru. Econ. Geol. 72,
893924.
Escalante, A.D., 2008. Patterns of Distal Alteration Zonation Around Antamina CuZn
Skarn and Uchucchacua Ag-base Metal Vein Deposits, Peru: Mineralogical, Chemical
and Isotopic Evidence for Fluid Composition, and Inltration, and Implications for
Mineral Exploration. The University of British Columbia, Vancouver, Canada, p. 788
(Unpublished PhD Thesis).
Escalante, A., Dipple, G.M., Barker, S.L.L., Tosdal, R., 2010. Dening trace element alteration
halos to skarn deposits hosted in heterogeneous carbonate rocks; case study from the
CuZn Antamina skarn deposit, Peru. J. Geochem. Explor. 105, 117136.
Felder, F., Ortz, G., Campos, C., Monsalve, I., Silva, A., 2005. Angostura Project: a high
suldation gold-silver deposit located in the Santander Complex of north eastern Colombia. Proceedings Congreso Internacional de Prospectores y Exploradores (ProExplo). Instituto de Ingenieros de Minas del Per, Lima p. 15.
Fernndez, R.R., Blesa, A., Moreira, P., Echeveste, H., Mykietiuk, K., Andrada de Palomera, P.,
Tessone, M., 2008. Los depsitos de oro y plata vinculados al magmatismo jursico de
la Patagonia: revisin y perspectivas para la exploracin. Rev. Asoc. Geol. Argent. 63,
665681.
Fifarek, R.H., Rye, R.O., 2005. Stable isotope geochemistry of the Pierina high-suldation
AuAg deposit, Peru; inuence of hydrodynamics on SO2H2S sulfur isotopic
exchange in magmatic-steam and steam-heated environments. Chem. Geol. 215,
253279.
Gansser, A., 1973. Facts and theories on the Andes (Twenty-sixth William Smith lecture).
J. Geol. Soc. Lond. 129, 93131.
Garzione, C.N., Hoke, G.D., Libarkin, J.C., Withers, S., MacFaddon, B., Eiler, J., Ghosh, P.,
Mulch, A., 2008. Rise of the Andes. Science 320, 13041307.
Gonzalez, L., Pffner, O.A., 2012. Morphologic evolution of the Central Andes of Peru. Int.
J. Earth Sci. 101, 307321.
Gregory-Wodzicki, K.M., 2000. Uplift history of the Central and Northern Andes: A
review. Bull. Geol. Soc. Am. 112, 10911105.
Groepper, H., Calvo, M., Crespo, H., Bisso, C.R., Cuadra, W.A., Dunkerley, P.M., Aguirre, E.,
1991. The epithermal goldsilver deposit of Choquelimpie, northern Chile. Econ.
Geol. Bull. Soc. 86, 12061221.
Gustafson, L.B., Hunt, J.P., 1975. The porphyry copper deposit at El Salvador, Chile. Econ.
Geol. 70, 857912.
Gustafson, L.B., Orquera, W., McWilliams, M., Castro, M., Olivares, O., Rojas, G., Maluenda,
J., Mendez, M., 2001. Multiple Centers of Mineralization in the Indio Muerto District,
El Salvador, Chile. Econ. Geol. 96, 325350.
Gustafson, L.B., Vidal, C., Pinto, R., Noble, D., 2004. Porphyry-epithermal transition,
Cajamarca region, northern Peru. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.),
Andean Metallogeny: New Discoveries, Concepts, and Updates. Society of Economic
Geologists Special Publication 11, pp. 279299.
Gutscher, M.A., Olivet, J.L., Aslanian, D., Eissen, J.P., Maury, R., 1999a. The lost Inca
Plateau: Cause of at subduction beneath Peru? Earth Planet. Sci. Lett. 171, 335341.
Gutscher, M.A., Malavieille, J., Lallemand, S., Collot, J.Y., 1999b. Tectonic segmentation of
the North Andean margin: impact of the Carnegie Ridge collision. Earth Planet. Sci.
Lett. 168, 255270.
Hampel, A., 2002. The migration history of the Nazca Ridge along the Peruvian active
margin: a re-evaluation. Earth Planet. Sci. Lett. 203, 665679.
Hayba, D.O., Bethke, P.M., Heald, P., Foley, N.K., 1985. The geological, mineralogical and
geochemical characteristics of volcanic-hosted epithermal deposits. In: Berger, B.R.,

Bethke, P.M. (Eds.), Geology and Geochemistry of Epithermal Systems. Society of Economic Geologists, Reviews in Economic Geology No.2, 129168.
Heald, P., Foley, N.K., Hayba, D.O., 1987. Comparative anatomy of volcanic-hosted
epithermal deposits; acid-sulfate and adularia-sericite types. Econ. Geol. 82, 126.
Heather, K.B., Leach, T., Clark, A.H., 2003a. Geology of the El Indio AuCuAg Deposit: The
Final Chapter? 10 Congreso Geolgico Chileno. Universidad de Concepcin,
Concepcin, Chile (CD-ROM).
Heather, K.B., Bissig, T., Staff of Exploraciones Barrick Chile Ltda, 2003b. Regional Geology
of the El Indio Mineral District, North-central Chile. 10 Congreso Geolgico Chileno.
Universidad de Concepcin, Concepcin, Chile (CD-ROM).
Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of hydrothermal ore deposits. Nature 370, 519527.
Hedenquist, J.W., Taran, Y.A., 2013. Modeling the Formation of Advanced Argillic
Lithocaps: Volcanic Vapor Condensation Above Porphyry Intrusions. Econ. Geol.
108, 15231540.
Hedenquist, J.W., Arribas Jr., A., Reynolds, T.J., 1998. Evolution of an intrusion-centered
hydrothermal system; Far Southeast-Lepanto porphyry and epithermal CuAu
deposits, Philippines. Econ. Geol. Bull. Soc. 93, 373404.
Hedenquist, J.W., Arribas, R.A., Gonzalez-Urien, E., 2000. Exploration for epithermal gold
deposits. Rev. Econ. Geol. 13, 245277.
Heinrich, C.A., 2005. The physical and chemical evolution of low-salinity magmatic uids
at the porphyry to epithermal transition: a thermodynamic study. Mineral. Deposita
39, 864889.
Heinrich, C.A., 2007. Fluiduid interactions in magmatic-hydrothermal ore formation.
Rev. Mineral. Geochem. 65, 363387.
Heinrich, C.A., Driesner, T., Stefnsson, A., Seward, T.M., 2004. Magmatic vapor contraction
and the transport of gold from the porphyry environment to epithermal ore deposits.
Geology 32, 761764.
Henderson, R.D., 2009. Fruta del Norte Project Ecuador NI 43-101 Technical Report. p. 135.
Hoke, G.D., Isacks, B.L., Jordan, T.E., Blanco, N., Tomlinson, A.J., Ramezani, J., 2007. Geomorphic evidence for post-10 Ma uplift of the western ank of the Central Andes 18 30
22S. Tectonics 26, TC5021. http://dx.doi.org/10.1029/2006TC002082.
Holley, E.A., 2012. The Veladero High-suldation Epithermal AuAg deposit, Argentina:
Volcanic Stratigraphy, Alteration, Mineralization, and Quartz Paragenesis. Colorado
School of Mines, Golden, Colorado, p. 226 (Unpublished PhD thesis).
Horton, B.K., Parra, M., Saylor, J.E., Nie, J., Mora, A., Torres, V., Strecker, M.R., 2010. Resolving uplift of the northern Andes using detrital zircon age signatures. GSA Today 20,
410.
Hungerbuehler, D., Steinmann, M., Winkler, W., Seward, D., Egueez, A., Peterson, D.E.,
Helg, U., Hammer, C., 2002. Neogene stratigraphy and Andean geodynamics of southern Ecuador. Earth Sci. Rev. 57, 75124.
Jannas, R.R., Beane, R.E., Ahler, B.A., Brosnahan, D.R., 1990. Gold and copper mineralization
at the El Indio deposit, Chile. J. Geochem. Explor. 36, 233266.
Jannas, R.R., Bowers, T.S., Petersen, U., Beane, R.E., 1999. High-suldation deposit types in
the El Indio District, Chile. Spec. Publ. Soc. Econ. Geol. 7, 219266.
Jordan, T.E., Nester, P.L., Blanco, N., Hoke, G.D., Davila, F., Tomlinson, A.J., 2010. Uplift of the
AltiplanoPuna Plateau; a view from the west. Tectonics 29. http://dx.doi.org/10.
1029/2010TC002661.
Kay, S.M., Mpodozis, C., 2001. Central Andean Ore deposits linked to evolving shallow
subduction systems and thickening crust. GSA Today 11, 49.
Kay, S.M., Mpodozis, C., Coira, B., 1999. Neogene magmatism, tectonism, and mineral deposits of the Central Andes (22 to 33 S latitude). In: Skinner, B.K. (Ed.), Geology and
Ore Deposits of the Central Andes. Soceity of Economic Geologists Special Publication
7, pp. 2759.
King, A.R., 1992. Magmatism, Structure and Mineralization in the Maricunga Belt. Imperial
College London (University of London), N. Chile.
Leal-Meja, H., 2011. Phanerozoic Gold Metallogeny in the Colombian Andes A Tectonomagmatic Approach. Universitat de Barcelona, p. 1000.
Leal-Meja, H., Melgarejo, I., Draper, J., Shaw, R., 2011. Phanerozoic gold metallogeny in
the Colombian Andes, Proceedings Let's talk Ore Deposits. SGA Biennial Meeting,
Antofagasta, Chile.
Lesage, G., 2011. Geochronology, Petrography, Geochemical Constraints, and Fluid
Characterization of the Buritic Gold Deposit, Antioquia Department, Colombia,
Masters Abstracts International.
Lodder, C., Padilla, R., Shaw, R., Garzon, T., Palacio, E., Jahoda, R., 2010. Discovery history of
the La Colosa gold porphyry deposit, Cajamarca, Colombia. In: Goldfarb, R.J., Marsh, E.
E., Monecke, T. (Eds.), The Challenge of Finding New Mineral Resources: Global
Metallogeny, Innovative Exploration, and New Discoveries. Society of Economic
Geologists special publication vol. I, pp. 1928 (15).
Longo, A.A., Dilles, J.H., Grunder, A.L., Duncan, R., 2010. Evolution of calc-alkaline volcanism and associated hydrothermal gold deposits at Yanacocha, Peru. Econ. Geol.
105, 11911241.
Love, D.A., Clark, A.H., Glover, J.K., 2004. The lithologic, stratigraphic, and structural setting
of the giant Antamina copperzinc skarn deposit, Ancash, Peru. Econ. Geol. 99,
887916.
Lozada Caldern, A.J., McBride, S.L., Bloom, M.S., 1994. The geology and 40Ar/39Ar geochronology of magmatic activity and related mineralization in the Nevados del Famatina
mining district, La Rioja province, Argentina. J. S. Am. Earth Sci. 7, 924.
MacDonald, P.J., Bissig, T., Hart, C.J., Barreno, J., Viera, F., 2011. The hydrothermal evolution
of the Quimsacocha high suldation AuAgCu deposit, Azuay Province, Ecuador,
11th Biennial Conference Society for Geology Applied to Ore Deposits, Antofagasta,
Chile.
MacDonald, P.J., Bissig, T., Hart, C.J., Barreno, J., Viera, F., Mantilla, G., Rogers, J., 2012.
Hydrothermal evolution of the Rio Falso Fault, Quimsacocha high suldation
AuAgCu District, Azuay Province, Ecuador. XVI Congreso Peruano de Geologa
& SEG2010 Conference, Lima, Peru, abstract and oral presentation.

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364


Mallette, P., Rojas, R., Guttierrez, A.R., 2004. Geology, mineralization, and genesis of the La
Quinua gold deposit, Yanacocha district, northern Peru. In: Sillitoe, R.H., Perello, J.,
Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and Updates.
Society of Economic Geologists Special Publication 11, pp. 301312.
Mantilla Figueroa, L.C., Bissig, T., Valencia, V., Hart, C.J.R., 2013. The magmatic history of
the VetasCalifornia mining district, Santander Massif, Eastern Cordillera, Colombia.
J. S. Am. Earth Sci. 45, 235249.
Mantilla, L.C., Valencia, V.A., Barra, F., Pinto, J., Colegial, J., 2009. GEocronologa UPb de los
cuerpos porfriticos del distrito aurofero de VetasCalifornia (Dpto de Santander,
Colombia). Bol. Geol. 31, 3143.
Mantilla, F.L.C., Mendoza, F.H., Bissig, T., Craig, H., 2011. Nuevas evidencias sobre el
magmatismo Miocenico en el distrito minero de VetasCalifornia (Macizo de
Santander, Cordillera Oriental, Colombia). Bol. Geol. 33, 4358.
Marsh, T.M., 1997. Geochronology, Thermochronology, and Isotope Systematics of the
CuAu and AuAg Deposits of the Potrerillos District, Atacama Region, Chile. Stanford
University, p. 341, (Unpubl. PhD thesis).
Marsh, T.M., Einaudi, M.T., McWilliams, M., 1997. 40Ar/39Ar geochronology of CuAu and
AuAg mineralization in the Potrerillos District, Chile. Econ. Geol. 92, 784806.
Martin, M., Clavero, J., Mpodozis, C., Cutio, L., 1995. Estudio geolgico regional de la
franja El Indio. Cordillera de Coquimbo: Informe registrado IR-95-6 Servicio Nacional
de Geologa y Minera, Chile, and Compaa Minera San Jos.
Martinod, J., Husson, L., Roperch, P., Guillaume, B., Espurt, N., 2010. Horizontal subduction
zones, convergence velocity and the building of the Andes. Earth Planet. Sci. Lett. 299,
299309.
Masterman, G.J., Cooke, D.R., Berry, R.F., Clark, A.H., Archibald, D.A., Mathur, R., Walshe, J.L.,
Duran, M., 2004. 40Ar/39Ar and ReOs geochronology of porphyry coppermolybdenum
deposits and related coppersilver veins in the Collahuasi District, northern Chile. Econ.
Geol. Bull. Soc. 99, 673690.
Moncada, D., Mutchler, S., Nieto, A., Reynolds, T.J., Rimstidt, J.D., Bodnar, R.J., 2012.
Mineral textures and uid inclusion petrography of the epithermal AgAu
deposits at Guanajuato, Mexico: Application to exploration. J. Geochem. Explor.
114, 2035.
Montario, M., Garver, J., Reiners, P., Ramage, J., Brandon, M., 2005a. Thermochronological
evidence for kilometre-scale incision of the northern Peruvian Andes. Geol. Soc. Am.
Abstr. Programs 37, 553.
Montario, M., Garver, J.I., Reiners, P., Ramage, J.M., 2005b. Timing of canyon incision of the
Ro Pativilca in response to uplift of the Andes in northern Per. Geol. Soc. Am. Abstr.
Programs 37, 76.
Montgomery, A.T., 2012. Mmetallogenetic Controls on Miocene High-sulphidation
Epithermal Gold Mineralization, Alto Chicama District, La Libertad, Northern Per.
Queen's University, Kingston, Ontario, Canada, p. 381 (Unpublished PhD thesis).
Morche, W., Velasco, C., Loayza, D., Clark, A., 2008. Late Miocene High-suldation
Epithermal Gold Deposits of the Aruntani District, Southern Peru, Congreso Geologico
del Peru. Sociedad Geologica del Peru, Lima, Peru.
Mortimer, C., 1973. The Cenozoic history of the southern Atacama Desert, Chile. J. Geol.
Soc. Lond. 129, 505526.
Mote, T.I., Becker, T.A., Renne, P., Brimhall, G.H., 2001. Chronology of exotic mineralization
at El Salvador, Chile, by 40Ar/39Ar dating of copper wad and supergene alunite. Econ.
Geol. Bull. Soc. 96, 351366.
Mpodozis, C., Clavero, J., 2002. Tertiary evolution of the southwestern edge of the Puna
Plateau: Cordillera Claudio Gay (2627 S), northern Chile. International Symposium
on Andean Geodynamics (ISAG), 5th,Toulouse, France, 2002, Extended Abstracts,
pp. 445448.
Mpodozis, C., Cornejo, P., 2012. Cenozoic Tectonics and Porphyry Copper Systems of the
Chilean Andes. In: Hedenquist, J.W., Harris, M., Camus, F. (Eds.), Geology and Genesis
of Major Copper Deposits and Districts of the World: A Tribute to Richard H. Sillitoe.
Society of Economic Geologists Special Publication 16, pp. 329360.
Mpodozis, C., Kay, S.M., 2003. Neogene tectonics, ages and mineralization along the transition zone between the El Indio and Maricunga mineral belts (Argentina and Chile
2829). 10 Congreso Geolgico Chileno. Universidad de Concepcin, Concepcin,
Chile (CD-ROM).
Muntean, J.L., Einaudi, M.T., 2001. Porphyry-Epithermal Transition: Maricunga Belt,
Northern Chile. Econ. Geol. 96, 743772.
Murakami, H., Seo, J.H., Heinrich, C.A., 2010. The relation between Cu/Au ratio and
formation depth of porphyry-style CuAu Mo deposits. Mineral. Deposita 45,
1121.
Niemeyer, H., Munizaga, R., 2008. Structural control of the emplacement of the Potrerillos
porphyry copper, central Andes of Chile. J. S. Am. Earth Sci. 26, 261270.
Noble, D.C., McKee, E.H., 1999. The Miocene metallogenic belt of central and northern
Peru. In: Skinner, B.K. (Ed.), Geology and Ore Deposits of the Central Andes. Society
of Economic Geologists Special Publication 7, pp. 155193.
Noble, D.C., Silberman, M.L., 1984. Evolucin volcnica e hidrotermal y cronologa de K-Ar
del distrito minero de Julcani. Sociedad Geolgica del Per, Volumen Jubilar, Per p.
135.
Noble, D.C., McKee, E.H., Mourier, T., Mgard, F., 1990. Cenozoic stratigraphy, magmatic
activity, compressive deformation, and uplift in northern Peru. Geol. Soc. Am. Bull.
102, 11051113.
Noble, D.C., Vidal, C.E., Perell, J., Rodriguez, P.O., 2004. SpaceTime Relationships of Some
Porphyry CuAu, Epithermal Au, and other magmatic-related mineral deposits in
Northern Peru. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny:
New Discoveries, Concepts, and Updates. Society of Economic Geologists Special
Publication 11, pp. 313319.
O'Driscoll, L.J., Richards, M.A., Humphreys, E.D., 2012. NazcaSouth America interactions
and the late Eocenelate Oligocene atslab episode in the central Andes. Tectonics 31.
Olson, S.F., 1984. Geology of the Portrerillos District, Atacama, Chile. Stanford University,
(Unpubl. PhD thesis).

363

Ossandon, C.G., Freraut, C.R., Gustafson, L.B., Lindsay, D.D., Zentilli, M., 2001. Geology of
the Chuquicamata Mine: A Progress Report. Econ. Geol. 96, 249270.
Oviedo, L., Fuster, N., Tschischow, N., Ribba, L., Zuccone, A., Grez, E., Aguilar, A., 1991.
General geology of La Coipa precious metal deposit, Atacama, Chile. Econ. Geol.
Bull. Soc. 86, 12871300.
PardoCasas, F., Molnar, P., 1987. Relative motion of the Nazca (Farallon) and South
American plates since Late Cretaceous time. Tectonics 6, 233248.
Perello, J., Carlotto, V., Zarate, A., Ramos, P., Posso, H., Neyra, C., Caballero, A., Fuster, N.,
Muhr, R., 2003. Porphyry-style alteration and mineralization of the Middle Eocene
to Early Oligocene AndahuaylasYauri Belt, Cuzco Region, Peru. Econ. Geol. 98,
15751605.
Perell, J., Brockway, H., Martini, R., 2004. Discovery and geology of the Esperanza porphyry coppergold deposit. Antofagasta Region, northern Chile: S. In: Sillitoe, R.H.,
Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and
Updates. Society of Economic Geologists Special Publication 11, pp. 167186.
Petersen, U., Noble, D., Arenas, M., Goodell, P., 1977. Geology of the Julcani mining district,
Peru. Econ. Geol. 72, 931949.
Petford, N., Atherton, M.P., 1992. Granitoid emplacement and deformation along a major
crustal lineament: the Cordillera Blanca, Peru. Tectonophysics 205, 171185.
Poblete, J.A., Bissig, T., Mortensen, J.K., Gabites, J., Friedman, R., Rodriguez, M., 2014.
The Cerro Bayo District, Chilean Patagonia: Late Jurassic to Cretaceous Magmatism
and Protracted History of Epithermal AgAu Mineralization. Econ. Geol. 109,
487502.
Polania, J.H., 1980. Die Uranvorkommen von California bei Bucaramanga (Kolumbien).
University of Stuttgart, Germany, p. 152 (Unpubl. PhD thesis).
Pudack, C., Halter, W.E., Heinrich, C.A., Pettke, T., 2009. Evolution of magmatic vapor to
gold-rich epithermal liquid: the porphyry to epithermal transition at Nevados de
Famatina, Northwest Argentina. Econ. Geol. 104, 449477.
Puig, A., Diaz, S., Cuitino, L., 1988. Sistemas hidrotermales asociados a calderas en el arco
volcanico paleogeno de la region de Antofagasta, Chile; distritos El Guanaco, Cachinal
de La Sierra y El Soldado. Rev. Geol. Chile 15, 5782.
Quang, C.X., Clark, A.H., Lee, J.K., Guilln, J., 2003. 40Ar39Ar Ages of Hypogene and Supergene Mineralization in the Cerro VerdeSanta Rosa Porphyry CuMo Cluster,
Arequipa, Peru. Econ. Geol. 98, 16831696.
Quang, C.X., Clark, A.H., Lee, J.K., Hawkes, N., 2005. Response of supergene processes to
episodic Cenozoic uplift, pediment erosion, and ignimbrite eruption in the porphyry
copper province of southern Peru. Econ. Geol. 100, 87114.
Rainbow, A., 2009. Genesis and Evolution of the Pierina High-sulphidation Epithermal
AuAg Deposit, Ancash, Per. Queen's University, Kingston, On, Canada, p. 277
(Unpublished PhD thesis).
Rainbow, A., Clark, A.H., Kyser, T.K., Gaboury, F., Hodgson, C.J., 2005. The Pierina
epithermal AuAg deposit, Ancash, Peru; paragenetic relationships, alunite textures,
and stable isotope geochemistry. Chem. Geol. 215, 235252.
Rainbow, A., Kyser, T.K., Clark, A.H., 2006. Isotopic evidence for microbial activity during
supergene oxidation of a high-suldation epithermal AuAg deposit. Geology 34,
269272.
Ramos, V.A., 2008. The basement of the Central Andes; the Arequipa and related terranes.
Annu. Rev. Earth Planet. Sci. 36, 289324.
Ramos, V.A., 2009. Anatomy and global context of the Andes: Main geologic features and
the Andean orogenic cycle. Geol. Soc. Am. Mem. 204, 3165.
Restrepo-Moreno, S.A., Foster, D.A., Stockli, D.F., Parra-Snchez, L.N., 2009. Long-term erosion and exhumation of the Altiplano Antioqueo, Northern Andes (Colombia)
from apatite (UTh)/He thermochronology. Earth Planet. Sci. Lett. 278, 112.
Riquelme, R., Herail, G., Martinod, J., Charrier, R., Darrozes, J., 2007. Late Cenozoic geomorphologic signal of Andean forearc deformation and tilting associated with the uplift
and climate changes of the southern Atacama Desert (26 S28 S). Geomorphology
86, 283306.
Riquelme, R., Darrozes, J., Maire, E., Herail, G., Soula, J.-C., 2008. Long-term denudation
rates from the Central Andes (Chile) estimated from a Digital Elevation Model
using the Black Top Hat function and Inverse Distance Weighting: implications for
the Neogene climate of the Atacama Desert. Andean Geol. 35, 105.
Rodriguez, M.A.L., 2014. Geology, Alteration, Mineralization and Hydrothermal Evolution
of the La BodegaLa Mascota Deposits, CaliforniaVetas Mining District, Eastern
Cordillera of Colombia, Northern Andes. The University of British Columbia, p. 325,
(Unpubl. MSc thesis).
Rodriguez, M.P., Aguilar, G., Urresty, C., Charrier, R., 2014. Neogene Landscape Evolution in
the Andes of North-central Chile Between 28 and 32S: Interplay Between Tectonic
and Erosional Processes: Geological Society. Special Publications, London, p. 399
http://dx.doi.org/10.1144/SP399.15.
Rosenbaum, G., Giles, D., Saxon, M., Betts, P.G., Weinberg, R.F., Duboz, C., 2005. Subduction
of the Nazca Ridge and the Inca Plateau: Insights into the formation of ore deposits in
Peru. Earth Planet. Sci. Lett. 239, 1832.
Rossell, W., De La Cruz, O., Romero, D., 2006. Visita tcnica a la Mina Quicay
Pacoyn. Instituto Geologico Minero y Metalurgico (Ingemmet), Peru (www.
slideshare.net. Technical powerpoint Presentation available for download, July
3rd, 2014).
Rye, R.O., 1993. The evolution of magmatic uids in the epithermal environment; the
stable isotope perspective. Econ. Geol. 88, 733752.
Rye, R.O., 2005. A review of the stable-isotope geochemistry of sulfate minerals in
selected igneous environments and related hydrothermal systems. Chem. Geol.
215, 536.
Rye, R.O., Bethke, P.M., Wasserman, M.D., 1992. The stable isotope geochemistry of acid
sulfate alteration. Econ. Geol. 87, 225262.
Sandeman, H.A., Clark, A.H., Farrar, E., 1995. An integrated tectono-magmatic model for
the evolution of the southern Peruvian Andes (1320 S) since 55 Ma. Int. Geol. Rev.
37, 10391073.

364

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327364

Scher, S., Williams-Jones, A.E., Williams-Jones, G., 2013. Fumarolic activity, acid-sulfate
alteration, and high suldation epithermal precious metal mineralization in the
crater of Kawah Ijen volcano, Java, Indonesia. Econ. Geol. 108, 10991118.
Schildgen, T.F., Hodges, K.V., Whipple, K.X., Reiners, P.W., Pringle, M.S., 2007. Uplift of the
western margin of the Andean Plateau revealed from canyon incision history,
southern Peru. Geology 35, 523526.
Schuette, P., Chiaradia, M., Beate, B., 2010. Geodynamic controls on Tertiary arc magmatism
in Ecuador; constraints from U/Pb zircon geochronology of OligoceneMiocene
intrusions and regional age distribution trends. Tectonophysics 489, 159176.
Schuette, P., Chiaradia, M., Barra, F., Villagomez, D., Beate, B., 2012. Metallogenic features
of Miocene porphyry Cu and porphyry-related mineral deposits in Ecuador revealed
by Re/Os, 40Ar/39Ar, and U/Pb geochronology. Mineral. Deposita 47, 383410.
Sbrier, M., Lavenu, A., Fornari, M., Soulas, J.-P., 1988. Tectonics and uplift in central Andes
(Per, Bolivia and northern Chile) from Eocene to present. Godynamique 3, 85106.
Shagam, R., Banks, P.O., Kohn, B.P., Dasch, L.E., Vargas, R., Pimentel, N., Rodrguez, G.I.,
1984. Tectonic Implications of CretaceousePliocene ssion-track ages from rocks of
the circum-Maracaibo Basin region of western Venezuela and eastern Colombia.
Geol. Soc. Am. Mem. 162, 385412.
Sillitoe, R.H., 1994. Erosion and collapse of volcanoes: Causes of telescoping in intrusioncentered ore deposits. Geology 22, 945948.
Sillitoe, R.H., 1995. Exploration of porphyry copper lithocaps. Proc. Pac. Rim Congr. 95,
527532.
Sillitoe, R.H., 2008. Special Paper: Major Gold Deposits and Belts of the North and South
American Cordillera: Distribution, Tectonomagmatic Settings, and Metallogenic
Considerations. Econ. Geol. 103, 663687.
Sillitoe, R.H., 2010. Porphyry copper systems. Econ. Geol. 105, 341.
Sillitoe, R.H., Hedenquist, J.W., 2003. Linkages between volcanotectonic settings, ore-uid
compositions, and epithermal precious metal deposits. In: Simmons, S.F., Graham, I.
(Eds.), Volcanic, Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of
Processes Within the Earth. Special PublicationSociety of Economic Geologists 10,
pp. 315343.
Sillitoe, R.H., Mortimer, C., Clark, A.H., 1968. A Chronology of Landform Evolution and
Supergene Mineral Alteration, Southern Atacama Desert. Institution of Mining and
Metallurgy Transactions, Chile, pp. 166169 (sec. B 77).
Sillitoe, R.H., McKee, E.H., Vila, T., 1991. Reconnaissance KAr geochronology of the
Maricunga goldsilver belt, northern Chile. Econ. Geol. 86, 12611270.
Sillitoe, R.H., Tolman, J., Van Kerkvoort, G., 2013. Geology of the caspiche porphyry gold
copper deposit, maricunga belt, Northern Chile. Econ. Geol. 108, 585604.
Simmons, S.F., 1991. Hydrologic implications of alteration and uid inclusion studies in
the Fresnillo District, Mexico; evidence for a brine reservoir and a descending
water table during the formation of hydrothermal AgPbZn orebodies. Econ. Geol.
86, 15791601.
Simmons, S.F., 2002. The paleosurface and paleowater table and their relationship to
epithermal mineralization. Soc. Econ. Geol. Newsl. 51, 2829.
Simmons, S.F., Brown, K.L., 2006. Gold in magmatic hydrothermal solutions and the rapid
formation of a giant ore deposit. Science 314, 288291.
Simmons, S.F., White, N.C., John, D.A., 2005. Geological characteristics of epithermal
precious and base metal deposits. In: Hedenquist, J.W., Thompson, J.F.H., Goldfarb,
R.J., Richards, J.P. (Eds.), Economic Geology; one hundredth anniversary volume,
19052005. Society of Economic Geologists, pp. 455522.
Skinner, S.M., Clayton, R.W., 2013. The lack of correlation between at slabs and bathymetric impactors in South America. Earth Planet. Sci. Lett. 371, 15.
Spikings, R.A., Winkler, W., Seward, D., Handler, R., 2001. Along-strike variations in the
thermal and tectonic response of the continental Ecuadorian Andes to the collision
with heterogeneous oceanic crust. Earth Planet. Sci. Lett. 186, 5773.
Steinmann, M., Hungerbuehler, D., Seward, D., Winkler, W., 1999. Neogene tectonic evolution and exhumation of the southern Ecuadorian Andes; a combined stratigraphy
and ssion-track approach. Tectonophysics 307, 255276.
Stoffregen, R.E., 1987. Genesis of acidsulfate alteration and AuCuAg mineralization at
Summitville, Colorado. Econ. Geol. 82, 15751591.
Strusievicz, O.R., Clark, A.H., Lee, J.K.L., Farrar, E., Slauenwhite, M., Hodgson, C.J., 2000.
Metallogenetic relationships of the Huaraz, Ancash, segment of the precious-base
metal sub-province of northern Peru. Abstr. Programs Geol. Soc. Am. 32, 504.

Taran, Y.A., Bernard, A., Gavilanes, J.-C., Africano, F., 2000. Native gold in mineral precipitates from high-temperature volcanic gases of Colima Volcano, Mexico. Appl.
Geochem. 15, 337346.
Tassinari, C.C.G., Pinzon, F.D., Buena Ventura, J., 2008. Age and sources of gold mineralization in the Marmato mining district, NW Colombia: a MiocenePliocene epizonal
gold deposit. Ore Geol. Rev. 33, 505518.
Teal, L., Benavides, A., 2010. History and geologic overview of the Yanacocha mining
district, Cajamarca, Peru. Econ. Geol. 105, 11731190.
Thompson, J.F.H., Gale, V.G., Tosdal, R.M., Wright, W.A., 2004. Characteristics and
formation of the Jeronimo carbonate-replacement gold deposit, Potrerillos district,
Chile. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and Updates. Society of Economic Geologists Special Publication 11,
pp. 7595.
Thouret, J.C., Wrner, G., Gunnell, Y., Singer, B., Zhang, X., Souriot, T., 2007. Geochronologic
and stratigraphic constraints on canyon incision and Miocene uplift of the Central
Andes in Peru. Earth Planet. Sci. Lett. 263, 151166.
Tosdal, R.M., Clark, A.H., Farrar, E., 1984. Cenozoic polyphase landscape and tectonic
evolution of the Cordillera Occidental, southernmost Peru. Geol. Soc. Am. Bull. 95,
13181332.
Tosdal, R.M., Dilles, J.H., Cooke, D.R., 2009. From source to sinks in auriferous magmatichydrothermal porphyry and epithermal deposits. Elements 5, 289295.
Van Der Lelij, R., 2013. Reconstructing North-western Gondwana With Implications for the
Evolution of the Iapetus and Rheic Oceans: A Geochronological, Thermochronological
and Geochemical Study. University of Geneva, Switzerland, p. 221 (Unpubl. PhD thesis).
Vargas, C.A., Mann, P., 2013. Tearing and Breaking Off of Subducted Slabs as the Result of
Collision of the Panama ArcIndenter with Northwestern South America. Bull.
Seismol. Soc. Am. 103, 20252046.
Vidal, C., Ligarda, R., 2004. Enargitegold deposits at Marcapunta. Colquijirca mining district, central Peru: Mineralogic and geochemical zoning in subvolcanic, limestonereplacement deposits of high-suldation epithermal type. In: Sillitoe, R.H., Perello, J.,
Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and Updates.
Society of Economic Geologists Special Publication 11, pp. 231241.
Vila, T., Sillitoe, R.H., 1991. Gold-rich porphyry systems in the Maricunga Belt, northern
Chile. Econ. Geol. Bull. Soc. 86, 12381260.
Villagmez, D., Spikings, R., 2013. Thermochronology and tectonics of the Central and
Western Cordilleras of Colombia: Early CretaceousTertiary evolution of the
Northern Andes. Lithos 160, 228249.
Villagomez, D., Spikings, R., Mora, A., Guzman, G., Ojeda, G., Cortes, E., van der Lelij, R.,
2011. Vertical tectonics at a continental crust-oceanic plateau plate boundary zone;
ssion-track thermochronology of the Sierra Nevada de Santa Marta, Colombia.
Tectonics 30.
Warren, I., Zuluaga, J.I., Robbins, C.H., Wulftange, W.H., Simmons, S.F., 2004. Geology and
geochemistry of epithermal AuAg mineralization in the El Pen district, northern
Chile. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and Updates. Society of Economic Geologists Special Publication 11,
pp. 113139.
Warren, I., Simmons, S.F., Mauk, J.L., 2007. Whole-rock geochemical techniques for evaluating hydrothermal alteration, mass changes, and compositional gradients associated
with epithermal AuAg mineralization. Econ. Geol. 102 (5), 923948.
Winocur, D.A., Litvak, V.D., Ramos, V.A., 2014. Magmatic and tectonic evolution of the
Oligocene Valle del Cura basin, Main Andes of Argentina and Chile: evidence for
generalized extension. Geol. Soc. Lond. Spec. Publ. 399. http://dx.doi.org/10.1144/
SP399.2.
Witzke, T., Kolitsch, U., Krause, W., Wiechowski, A., Medenbach, O., Kampf, A.R., Steele, I.M.,
Favreau, G., 2006. Guanacoite, Cu2Mg2 (Mg0.5Cu0.5)(OH)4 (H2O)4(AsO4), a new arsenate mineral species from the El Guanaco Mine, near Taltal, Chile; description and
crystal structure. Eur. J. Mineral. 18, 813821.
Yez, G.A., Ranero, C.R., Huene, R., Daz, J., 2001. Magnetic anomaly interpretation across
the southern central Andes (3234 S): The role of the Juan Fernndez Ridge in the
late Tertiary evolution of the margin. J. Geophys. Res. Solid Earth 106, 63256345
(19782012).
Zappettini, E., 2008. Geologa de frontera ArgentinaChile. http://sig.segemar.gov.ar/
(accessed March 2014).

You might also like