Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

Ore Geology Reviews 65 (2015) 327–364

Contents lists available at ScienceDirect

Ore Geology Reviews


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

Review

Physiographic and tectonic settings of high-sulfidation epithermal


gold–silver deposits of the Andes and their controls on
mineralizing processes
Thomas Bissig a,⁎, Alan H. Clark b, Amelia Rainbow b, Allan Montgomery c
a
Mineral Deposit Research Unit, University of British Columbia, 2020-2207 Main Mall, Vancouver, BC V6T 1Z4, Canada
b
Queen's University, Bruce Wing/Miller Hall, Kingston, ON K7L 3N6, Canada
c
Riverside Resources Suite 1110, 1111 West Georgia Street, Vancouver, BC V6E 4M3, Canada

a r t i c l e i n f o a b s t r a c t

Article history: Gold and silver ores in the vast majority of Andean high-sulfidation epithermal Au–Ag deposits occur at high
Received 6 July 2014 present day elevations and typically 200–500 m below low relief landforms situated at 3500 to 5200 m a.s.l.
Received in revised form 10 September 2014 Most deposits are middle Miocene and younger and include, El Indio, Tambo, Pascua–Lama, Veladero (El Indio
Accepted 17 September 2014
belt, Chile/Argentina), Cerro de Pasco (Central Peru), Pierina, Lagunas Norte, Yanacocha (northern Peru),
Available online 28 September 2014
Quimsacocha (Ecuador), and the California–Vetas mining district (Santander, Colombia), jointly accounting for
Keywords:
N 130 Moz Au resources. Slightly older examples are only preserved in the Atacama Desert and include the middle
High-sulfidation Eocene El Guanaco and El Hueso and the late Oligocene/early Miocene La Coipa deposits. The absence of Paleo-
Epithermal cene and older high-sulfidation epithermal deposits can be explained by limited preservation potential imposed
Andes by transpressional tectonics within overall contractile episodes and surface uplift. These conditions prevailed pre-
Landscape evolution dominantly in segments of shallow-angle subduction of the Nazca or Caribbean plate below the South American
Erosion continent, a tectonic setting also common for porphyry-style Cu (–Au, Mo) deposits. Stratovolcanoes are uncom-
Uplift mon ore hosts and volcanic rocks coincident with mineralization are in most cases volumetrically restricted or
Flat subduction
absent, recording the terminal stages of local arc magmatism. However, dacitic domes are important at,
Neogene
e.g., Yanacocha and La Coipa. At Lagunas Norte, a small stratovolcano largely pre-dating but temporally overlap-
ping 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 high-
elevation, low-relief surfaces. In the California–Vetas 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 hydro-
static and lithostatic pressure at these sites of high local relief are believed to have enhanced both boiling and
mixing of magmatic with meteoric fluids, 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 mineral-
ization by several Ma. However, high-sulfidation epithermal mineralization can be hosted in any conceivable
rock type including high grade metamorphic rocks (California–Vetas mining district), significantly older plutonic
rocks (Pascua–Lama) 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 hydro-
thermal breccia bodies, whereas coherent volcanic, plutonic, or metamorphic rocks may host fault- and breccia-
controlled ores. The near-surface steam-heated zone can attain a thickness of several hundred meters in dry
climates (e.g. Veladero, Pascua–Lama, Tambo) but is typically poorly developed and less than 20 m thick in
humid climatic zones.
The physiographic and tectonic settings of high-sulfidation epithermal deposits are distinct from low-sulfidation
epithermal districts such as those of Patagonia, El Peñón (Chile) or Fruta del Norte (Ecuador). The latter range to
significantly older ages (Jurassic to early Eocene) occur at mainly lower elevations and were emplaced in exten-
sional 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-sulfidation districts.
© 2014 Elsevier B.V. All rights reserved.

⁎ 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.
328 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
2. Epithermal deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 328
3. Distribution of high-sulfidation deposits in the Andes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
4. El Indio belt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
4.1. The Tambo deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
4.2. El Indio deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
4.3. Pascua–Lama . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
4.4. Veladero . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 337
5. Maricunga belt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
5.1. La Coipa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
5.2. La Pepa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
6. High-sulfidation epithermal deposits of the Domeyko fault system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
6.1. El Hueso . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
6.2. El Guanaco . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
7. Late Miocene high-sulfidation epithermal Au deposits of the western Cordillera of northern Chile and southern Peru . . . . . . . . . . . . . . . 344
8. Central to northern Peruvian flat slab segment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
9. High-sulfidation epithermal deposits of the central Peruvian polymetallic belt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
9.1. Julcani . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
9.2. The Marcapunta–Colquijirca district . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
9.3. Cerro de Pasco . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
9.4. Quicay . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
10. The high-sulfidation epithermal Au (–Cu, Ag) deposits of northwestern Peru . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
10.1. Pierina . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
10.2. Lagunas Norte . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
10.3. Yanacocha . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
10.4. Tantahuatay, Sipan and La Zanja . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
11. The northern Andes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 351
11.1. Quimsacocha . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
11.2. California Vetas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
12. Summary and comparison to low-sulfidation deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
13. Controls of geomorphic processes and climate on mineralization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
14. Igneous rocks, volcanology and magmatic fluids related to high-sulfidation epithermal deposits . . . . . . . . . . . . . . . . . . . . . . . . 360
15. Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361

1. Introduction discusses the influence these factors can have on both mineralizing
processes and the preservation of the deposits.
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 Blanco–Los 2. Epithermal deposits
Bronces–Los Sulfatos, El Teniente, Chuquicamata) as well as many of
the world's largest epithermal Au–Ag deposits (e.g., Yanacocha, Lagunas Epithermal deposits are usually classified into sub-types based on ei-
Norte, Pascua–Lama, Veladero: Sillitoe, 2008). The vast majority of ther ore sulfide assemblage or characteristic associated alteration; both
Andean epithermal deposits containing N 10 Moz Au are of high- schemes have inherent limitations. Some of the most widely referenced
sulfidation type. These deposits have a close link to a magmatic source review papers on the topic (Hedenquist et al., 2000; Sillitoe and
for fluids, volatiles and metals (e.g., Deyell et al., 2004; Rye, 1993) but Hedenquist, 2003; Simmons et al., 2005) prefer a classification into
form at depths of typically less than 1 km (e.g. Sillitoe, 2010) and conse- high, low and intermediate-sulfidation types. This classification scheme
quently mineralizing processes are influenced by the near-surface phys- can, however, be problematic, because sulfide assemblages may be dif-
icochemical environment. The main focus of this review is on deposits ficult to classify in a field exploration setting, particularly if the deposit
and districts where the bulk of the precious metal is contained in has been oxidized. Moreover, sulfide assemblages within a single de-
the epithermal environment, i.e., the shallow part of magmatic- posit may represent precipitation over the entire breadth of sulfidation
hydrothermal systems, and concentrates on the physiographic environ- state, from high to intermediate and low-sulfidation, depending on
ment of epithermal mineralization. This paper does not discuss major fluid–wall rock interaction (e.g., El Indio: Heather et al., 2003a, 2003b;
porphyry Cu deposits in detail, although the shallow portions of many Cerro de Pasco: Baumgartner et al., 2008; Lagunas Norte: Cerpa et al.,
of these have been overprinted by epithermal mineralization or alter- 2013). Alternative classification is based on dominant alteration
ation (e.g., Masterman et al., 2004; Ossandón et al., 2001). Similarly, and gangue assemblages and includes quartz–adularia–sericite and
the deposits hosting Sn, W, Ag and Au ores in the eastern Cordillera of quartz–alunite or acid sulfate type epithermal deposits (Heald et al.
Bolivia and Peru are not discussed. Following a general summary of 1987, Tosdal et al., 2009), the former typically including low to
epithermal deposit types and their terminology, this article presents a intermediate-sulfidation sulfide assemblages and the latter associated
comprehensive overview of the major high-sulfidation epithermal dis- with high-sulfidation deposits. The limitation of this classification
tricts and mineral belts of the Andes. It focuses on the links between scheme is that, particularly in high-sulfidation deposits, alteration may
landscape evolution, climatic setting, volcanology and tectonics, and pre-date and may not be directly related to mineralization (see below).
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 329

For this paper we use the widely applied classification scheme of 3. Distribution of high-sulfidation deposits in the Andes
high-sulfidation (normally associated with acidic quartz–alunite
alteration) and low-sulfidation type (associated with near neutral High-sulfidation epithermal deposits occur mostly in the Central
quartz–adularia gangue and illite–adularia alteration) deposits. The Andes (~ 34° to 3° Lat. S) with few examples in the Northern Andes
term intermediate sulfidation is used to describe sulfidation state but (~3° Lat. S to 11.3° Lat. N) and no known example in the southern vol-
is not used in the sense of a distinct subtype of epithermal deposit due canic zone of the Central Andes (south of 34° Lat. S) and the southern,
to the inherent variation of sulfide assemblages within a single deposit a.k.a. Patagonian Andes, south of 46° Lat. S (Fig. 1). The overwhelming
or vein. majority of high-sulfidation epithermal deposits are Miocene or youn-
In most high-sulfidation epithermal deposits, two hydrothermal ger, and are located in segments of the Andean arc where magmatism
stages can be distinguished (e.g., Simmons et al., 2005). An early stage is now absent or subdued (e.g., Bissig et al., 2008; Kay and Mpodozis,
of intense acid leaching of the wall-rock by a magmatic vapor that con- 2001; Rosenbaum et al., 2005). The lack of magmatism is attributed to
densed into groundwater, cooled to temperatures below 300 °C and two zones of shallow angle subduction of the Nazca plate below the
was acidified by dissociation of acids such as HCl, or H2SO4 South American continent in the Central Andes (the Pampean flat slab
(Hedenquist and Taran, 2013); and, a subsequent stage of weakly- in northern Chile and the Peruvian flat slab in central and northern
acidic fluid incursion, from which the bulk of the sulfides and precious Peru: Ramos, 2009) as well as the Bucaramanga flat slab where Caribbe-
metals, together with euhedral quartz, precipitated (e.g., Heinrich an plate is subducting below northern Colombia (Vargas and Mann,
et al., 2004, 2007; Holley, 2012; Stoffregen, 1987). The intense acid 2013). Flat subduction coincides with shortening and crustal thickening
leaching generates the zoning from vuggy quartz to quartz–alunite of the overriding continental plate (Martinod et al., 2010) and has been
and more distal kaolinitic alteration. This assemblage is commonly re- spatially linked to the subduction of aseismic ridges or oceanic plateaus
ferred to as a lithocap because it “caps” potential porphyry Cu mineral- on the oceanic plate (e.g. Gutscher et al., 1999b; Yañez et al., 2001).
ization (Sillitoe, 1995). Although mineralization is widely related to the However the causative relationship between ridge impingement and
infiltration of the less acidic fluids into the precursor lithocap onset of flat subduction remains a matter of debate (Skinner and
(e.g., Holley, 2012; Stoffregen, 1987), it has been shown that the acidic Clayton, 2013). Thus, crustal shortening, uplift and porphyry Cu
magmatic vapors can also transport gold into the shallow epithermal mineralization pre-dates the arrival of the subducting aseismic Juan
environment (Scher et al., 2013; Taran et al., 2000). At Pascua–Lama Fernández ridge in Central Chile (Deckart et al., 2013), a situation
such fluids were responsible for at least some of the gold mineralization comparable to northern Peru where high-sulfidation epithermal gold
(Chouinard et al., 2005). Conversely, there are examples of high- mineralization at Lagunas Norte occurred at 17 Ma, coincident with up-
sulfidation epithermal deposits where vuggy quartz alteration is largely lift, but several Ma prior to the inferred onset of Nazca ridge subduction
absent (e.g., La Bodega, Colombia: Rodriguez, 2014). (Montgomery, 2012, see below).
The precipitation mechanisms for precious metals in high- The Northern Andes are characterized by a series of oceanic terranes
sulfidation epithermal deposits are not as well understood as for low- accreted to the South American continent dominantly between the Me-
sulfidation systems. For the latter, ample textural and fluid inclusion ev- sozoic and Miocene (Cediel et al., 2003). The Central Andes, in contrast,
idence leaves little doubt that fluid boiling leads to precipitation of consist of a collage of terranes accreted during the Proterozoic-to-
metals due to the loss of the dominant ligands (H2S) to the gas phase Paleozoic (Ramos, 2008). Here, Andean type subduction has dominated
(Moncada et al., 2012; Simmons et al., 2005). In contrast, fluid inclusion the active continental margin since Permo-Triassic times (Ramos,
studies of high-sulfidation deposits are difficult, as the appropriate 2009), and in both the Northern Andes south of ~ 5.5° Lat. N and the
transparent minerals are rare. Several possible precipitation mecha- Central Andes, the Nazca plate is currently subducting eastward
nisms such as boiling, and fluid mixing of a magma derived fluid with beneath the South American continent.
oxidized groundwater have been proposed (e.g., Rye et al., 1992; High-sulfidation epithermal deposits occur in the same segments of
Jannas et al., 1999). Recent work on the La Bodega deposit in the Andean arc as porphyry-style deposits (e.g. Sillitoe, 2008, 2010).
Colombia, where the high-sulfidation sulfide assemblage is hosted in However, a direct temporal, spatial and genetic relationship to porphyry
veins with similar textures as described for low-sulfidation deposits deposits containing economic Cu mineralization, as demonstrated at Far
(see below), supports that boiling is a viable mechanism to precipitate Southeast and Lepanto (Hedenquist et al., 1998), has not been docu-
sulfides (Rodriguez, 2014). Moreover, steam-heated alteration, present mented for any Andean high-sulfidation epithermal deposits discussed
above many Andean high-sulfidation systems, forms by H2S-rich steam herein, although the two mineralization styles may overlap either
that condenses and oxidizes to form sulfuric acid in the vadose zone, spatially or temporally (e.g. Yanacocha district: Longo et al., 2010, La
and is considered evidence for fluid boiling at depth (Simmons et al., Pepa, Maricunga belt: Muntean and Einaudi, 2001).
2005). All high-sulfidation epithermal deposits of the Andes now
A close association of high-sulfidation epithermal deposits with a crop out at high elevation, and with few exceptions, above
parent intrusion and porphyry-style mineralization has been proposed ~ 3500 m a.s.l. Most are also located near the water-shed along the
by Hedenquist and Loewenstern (1994) and documented in detail crest of the orogen, or what appears to have been the water-shed
for the Lepanto–Far Southeast porphyry system in the Philippines at the time of hydrothermal activity, in areas dominated by relatively
(Hedenquist et al., 1998). In earlier literature, stratovolcanoes, calderas flat landforms which are interpreted as uplifted sub-planar
and extrusive or subvolcanic dacitic to andesitic domes have been paleosurfaces. In the semi-arid and arid portions of the Andes,
considered the dominant hosts, and large volcanic centers have been these paleosurfaces probably represent pediplains (regionally inter-
presumed to be temporally and genetically associated with high- connected pediment surfaces) formed under semi-arid conditions
sulfidation epithermal deposits (e.g. Hayba et al. 1985, Arribas, 1995). and several pulses of pediment erosion separated by uplift pulses
Indeed, stable S, H and O isotopic evidence (e.g., Deyell et al., 2005a, can commonly be distinguished (Bissig and Riquelme, 2009; Bissig
2005b; Rye, 1993, 2005) as well as thermodynamic considerations et al., 2002a; Riquelme et al., 2007; Rodriguez et al., 2014; Sillitoe
(Heinrich, 2005, 2007) clearly indicate that magma-derived fluids or et al., 1968), resulting in a series of pediment surfaces, vertically sep-
condensed magmatic vapors constitute the dominant component of arated up to several 100 m by steeper back-scarps. Similar relation-
the mineralizing fluids. However, as will be discussed below, better geo- ships between high-level sub-planar landforms incised by broad,
logical understanding and new geochronological information, together flat bottomed valleys are present in southern Peru (Quang et al.,
with exploration discoveries made since 1995, suggest that stratovol- 2005; Tosdal et al., 1984) and northern Peru (Montgomery, 2012).
canoes and calderas are uncommon hosts to epithermal ore in the The major high-sulfidation deposits and districts are described with-
Andes. in their geomorphologic and volcanologic context from south to north
330 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

90° W 80° W 70° W 60° W 50° W

epithermal deposits
1: Tambo
2: El Indio
20° N 3: Pascua-Lama
4: Veladero
5: Famatina
6: La Coipa
7: El Hueso
8: El Guanaco
9: Choquelimpie
10: Santa Rosa
10° N 11: Tucari
12: Julcani
13: Marcapunta/Colquijirca
23/24 14: Cerro de Pasco
15: Quicay
CoR 16: Pierina
17: Lagunas Norte
18: Yanacocha
0° 19: Tantahuatay
des 20: Sipán
CaR n An
22 Norther al Andes 21: La Zanja
ntr
Ce 22: Quimsacocha
23: La Bodega
IP
21 20 19 24: Angostura
18 Active volcano
17 Flat slab segments
16 15
10° S 14 Aseismic ridges
13 Subducted aseismic
12
ridges/plateaus

11
10
9
10° S

8
NR
7
IR 6
4 35
30° S 2
1

JFR

Fig. 1. Map of South America, showing locations of major high-sulfidation epithermal deposits discussed in this paper. Also indicated are segments of flat 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 approx- only known to have occurred between 9.5 and 5 Ma (Bissig et al.,
imate contained resources are summarized in Table 1. 2001; A.H. Clark, unpubl. data). Alteration and mineralization ages quot-
ed herein for the El Indio belt are all 40Ar/39Ar plateau ages on alunite or
4. El Indio belt sericite mostly with analytical errors of b1 to 5% on individual dates. It is
worth noting that the small epithermal Cu–Au district at Famantina
The El Indio belt straddles the Chile–Argentina Border between 29° (Table 1), is located in the same Andean segment, albeit some 220 km
and 30° Lat. S. which corresponds to the middle of the Pampean flat E of Veladero in Argentina, and is at 5 to 3.8 Ma just slightly younger
slab segment. The belt is named after the El Indio Au (–Ag, Cu) deposit than El Indio (Lozada-Calderón et al., 1994; Pudack et al., 2009).
which was mined between 1978 and 2001. The belt also includes the Volcanism in the Oligocene and early Miocene was voluminous in
small Tambo and the giant Veladero and Pascua–Lama deposits as the El Indio belt (Bissig et al., 2001; Martin et al., 1995; Winocur et al.,
well as other prospects (Fig. 2), jointly accounting for ~ 40 Moz of Au 2014) initially generating the 27–23 Ma Tilito Formation dacitic-to-
resources. Additionally, the belt contains numerous alteration zones andesitic pyroclastic and volcaniclastic rocks and small volumes of
ranging in age from Eocene to late Miocene but, with the exception of basalts in an extensional setting (Winocur et al., 2014). The Tilito
Veladero (12.7–10.3 Ma; Holley, 2012), economic mineralization is Formation is unconformably overlain by the dominantly-andesitic
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 331

Escabroso (21–17 Ma) and Cerro de las Tórtolas (16.6–14 Ma) Forma- 4.2. El Indio deposit
tions (Fig. 2), both associated with large volcanic edifices and
subvolcanic intrusions (Bissig et al., 2001; Martin et al., 1995). Erupted Mineralization at El Indio, in contrast to Tambo, is hosted in veins
magma volumes decreased markedly after ca. 14 Ma and volcanism (Jannas et al., 1999) and includes sulfide-rich massive enargite ± pyrite
was confined to isolated eruptive centers at ~ 12.7–11 Ma (Vacas veins as well as quartz-rich veins with tetrahedrite/tennantite and
Heladas formation andesite to dacite) and the rhyodacitic Pascua chalcopyrite. For some veins, an along-strike zonation from high- to
(8–7.6 Ma) and Vallecito (6–5.5 Ma) Formations (Bissig et al., 2001). intermediate-sulfidation sulfide assemblages, and from pyrophyllite–
An upper Pliocene rhyolite dome (Cerro de Vidrio) has been document- sericite–diaspore ± alunite to illite-dominated alteration has been doc-
ed ~7 km E of the Veladero deposit (Bissig et al., 2001, 2002a, 2002b). umented (Heather et al., 2003a; K.B. Heather, pers. commun. 2012). El
The voluminous upper Oligocene to middle Miocene andesites show Indio is world-renowned due to exceptional bonanza grades of up to
an increasing modal abundance of phenocrystic hornblende, relative to N1% Au in the quartz-rich, intermediate-sulfidation El Indio Sur 3500
clinopyroxene, with time (Bissig et al., 2003). Further, igneous rocks vein which permitted shipping of the ore directly to the smelter (direct
younger than ca. 14 Ma are characterized by elevated Sr/Y (N40) and shipping ore) in the early days of exploitation. Gold occurred mainly as
depleted HREE contents, indicative of garnet fractionation at the base native Au, auricupride and as tellurides (Heather et al., 2003a; Jannas
of the crust and elevated pressure equivalent to N 45 km crustal thick- et al., 1990, 1999). As is the case for Tambo, the top of the mineralized
ness (Bissig et al., 2003). This change in both the erupted volumes and zone at El Indio is located about 200 m below the Azufreras–Torta Sur-
chemistry of the igneous rocks is attributed to the onset of flat subduc- face, here located at ~4400 m a.s.l. Steam-heated alteration is exposed
tion and associated increased coupling between the Nazca slab and locally on Cerro Campana above the western part of the deposit
overlying continental plate, which led to contractile deformation and (Fig. 3). The mineralized zone has a vertical extent of 500 m extending
crustal thickening (Bissig et al., 2003; Kay et al., 1999). Epithermal min- down to 3700 m a.s.l, below which the gold grade decreases markedly
eralization is generally hosted in Cerro de las Tórtolas Formation and (Bissig et al., 2002a; Jannas et al., 1999). The veins are best developed
older rocks but occurred several Ma after the emplacement of the host in dacitic pyroclastic rocks of the Tilito Formation, which in the area of
rocks. Mineralization of the Filo Federico ore zone of the Veladero the deposit are uncomformably overlain by andesitic lavas and
deposit is partly hosted in volcaniclastic deposits of the Vacas Heladas volcaniclastic rocks of the lower Miocene Escabroso Group. Most veins
Formation (Holley, 2012). also cross-cut the Escabroso Group but typically horsetail out above
The geomorphology in the El Indio belt preserves evidence of a his- the contact (Bissig et al., 2002a; Heather et al., 2003a). Low volumes
tory of episodic uplift and pediplain erosion throughout the middle to of ca. 8 Ma dacitic hypabyssal intrusions spatially associated with min-
late Miocene (Bissig et al., 2002a; Figs. 3, 4, 5, 6). Relics of three low- eralization have also been reported (Heather et al., 2003b). The miner-
relief erosional surfaces, vertically separated from each other by alization ages range from 8 to 5 Ma and are inferred from sericite
200–400 m, have been documented. These are the 17–15 Ma immediately adjacent to and alunite from within veins (Bissig et al.,
Frontera–Deidad surface, the 14–12.5 Ma Azufreras–Torta surface and 2001; A.H. Clark and K.B. Heather, unpubl. data) and a late stage 3.5–
the 10–6 Ma Los Ríos surface (Bissig et al., 2002a). The Frontera–Deidad 2.5 Ma barren hydrothermal overprint has also been identified (A.H.
surface dominating the higher interfluves is a relic of a regionally exten- Clark and K. B. Heather, unpubl. sericite 40Ar/39Ar data). The age range
sive pediplain on the western Andean slope at the latitude of the El Indio indicates that multiple hydrothermal pulses related to successive pres-
belt (Aguilar et al., 2011; Rodriguez et al., 2014), subsequently incised sure release, likely facilitated by erosion, from a sizeable batholith scale
by the Azufreras–Torta and Los Ríos pediment surfaces. These nested magma system were responsible for mineralization. Contrary to the
paleosurfaces in the main Cordillera are now separated from the land- conclusion of Jannas et al. (1999) enargite-rich veins are not systemat-
forms of the Coastal Cordillera by the north striking, west verging Vicu- ically older than gold-rich veins, since the youngest data come
ña–San Félix fault system (Aguilar et al., 2011, 2013). Vertical from 6.2 Ma alunite in the enargite-rich Campana B vein (Bissig et al.,
displacement along this fault system starting in the early–middle Mio- 2001) and sericite from the 5 Ma Viento Oeste/Cuarzo Uno high to
cene is thought to be responsible for the three distinct paleosurfaces ob- intermediate-sulfidation veins whereas the El Indio Sur 3500 vein
served in the El Indio belt (Aguilar et al., 2011). Steam-heated alteration yielded the oldest sericite ages of ca. 8 Ma (A.H. Clark and K.B. Heather
associated with epithermal mineralization in all deposits of the El Indio unpubl. data). The geochronological data demonstrate a significant
belt is almost exclusively exposed on the Azufreras–Torta surface age gap of at least 7 Ma between volcanic rock deposition (~ 26–
(Figs. 3,4,5, 6), whereas the top of the mineralized zone typically occurs 15 Ma) and mineralization (8–5 Ma). Only small amounts of igneous
~200 m below this surface (Bissig et al., 2002a). rocks contemporaneous with mineralization are known and these in-
clude 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.,
4.1. The Tambo deposit 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 Río
At Tambo, mineralization is hosted mainly in three diatreme Malo valley (part of the Los Ríos surface) and is hosted below the
breccia bodies (Wendy, Kimberly and Canto Sur; Figs. 3, 4), as well Azufreras–Torta surface (Bissig et al., 2002a; Fig. 3).
as a number of satellite breccia bodies and replacement veins, all of
which were emplaced into dacitic pyroclastic and volcanoclastic 4.3. Pascua–Lama
rocks assigned to the upper Oligocene Tilito Formation. The ore
zone has an overall vertical extent of about 300 m (Deyell et al., Pascua–Lama is the largest deposit in the El Indio belt and is located
2005a; Jannas et al., 1999). The paleosurface level is located at some 50 km N of El Indio on the Chile–Argentina border. It is hosted
4500 m a.s.l. for Kimberly and Wendy and has been downfaulted within an uplifted Paleozoic basement block (Figs. 5, 6), with minerali-
~ 100 m relative to Canto Sur, located at 4600 m a.s.l. elevation zation centered on a series of diatreme breccia complexes. The most im-
(Deyell et al., 2005a). Alunite, some intergrown with native Au, has portant of these is Brecha Central, which cuts Paleozoic and Jurassic
been dated at 8.8 to 8 Ma although at Canto Sur it yields ages as granitoids (Chouinard et al., 2005). Hydrothermal activity is constrained
young as 7.1 Ma (Bissig et al., 2001; Deyell et al., 2005a). Alunite to 9.1–8 Ma based on 40Ar/39Ar dating of alunite and jarosite; alunite
ages coincide with the inferred age of the Los Ríos surface incision from pyrite–enargite veins constraining the gold mineralization to
(10–6 Ma) and the Tambo area is situated near the back scarp of a 8.8 ± 0.6 Ma (Bissig et al., 2001; Deyell et al., 2005b). Slightly older al-
Los Ríos valley where it incises the older Azufreras–Torta surface unite of 9.5 Ma is documented from the Penelope ore zone some 2 km
(Figs. 3, 4). SE of Brecha Central (Bissig et al., 2001; Deyell et al., 2005b). No igneous
332
Table 1
Summary of main HS deposits.

Deposit/ Lat. Long. Elevation of Elevation of Mineraliza-tion Total Mineralization Host rocks Volcanology Selected references
district mineralization paleosurface age (method) resources
(m.a.s.l.) (if known) (approx.)

Tambo −29.8 −69.95 4300–4100 4500 8.5 to 8 Ma 0.8 Moz Hydrothermal breccia hosted Au–Ag Late Oligocene Tilito No igneous rocks age equivalent to Bissig et al. (2002a,
Au mineralization associated with Formation dacite tuffs mineralization recognized, dacite 2002b); Deyell et al.
intense acid sulfate alteration. tuffs of the Tilito Formation pre-date (2005a); Jannas et al.
Subordinate replacement veins. mineralization by N 14 m.y. Minor, (1999)
undated pre-mineral diorite
porphyry has locally been
intersected below mineralization.
Phreatomagmatic activity pre to syn
mineralization evidenced by breccia
bodies.
El Indio −29.74 −69.97 4200–3700 4400 8 to 5 Ma 4.2 Moz Largely vein hosted Au (–Cu, Ag) Late Oligocene Tilito No igneous rocks age equivalent to Bissig et al. (2002a,
40
Ar/39Ar, Au mineralization. Most gold produced Formation dacite tuffs and mineralization recognized, except for 2002b); Heather et al.
alunite, sericite from quartz rich veins with sulfide overlying early Miocene possible 8 Ma dacite dyke. Host rocks (2003a); Jannas et al.
assemblage of intermediate Escabroso Group andesite pre-date mineralization by ~7 to (1999)
sulfidation state, most Cu produced lavas and breccias 17 m.y.

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364


from massive enargite veins.
Dominant alteration is advanced
argillic to phyllic (sericite, diaspore,
locally pyrophyllite, alunite).
Veladero −29.36 −69.95 4300–3900 4600 12.7–10.7 Ma 12.2 Moz Mineralization in two principal ore Cerro de las Tórtolas Volcano sedimentary host rocks Charchaflie et al. (2007);
(40Ar/39Ar, Au; 180 zones and associated with intense Formation (16–14 Ma) pre-date hydrothermal breccias and Holley (2012)
alunite) Moz Ag acid sulfate alteration and volcano sedimentary breccias, mineralization by ~ 2–4 Ma. Locally
silicification: Amable (12.1 to subvolcanic domes 12.1 Ma andesite dykes cutting
12.7 Ma) and Filo Federico (10.7– mineralization at Amable zone
11 Ma). Apart from minor argentite
no primary sulfides recognized. Most
Ag in jarosite at Amable.
Pascua-Lama −29.32 −70.01 4850–4550 5000–5150 9.1–8 Ma 17.6 Moz Breccia and stockwork vein-hosted Permo-Triassic to Jurassic Phreatomagmatic breccias, pre and Bissig et al. (2002a,
(40Ar/39Ar, Au; 585 mineralization associated with high granitic rocks syn mineralization. Post mineral 2002b); Chouinard et al.
alunite) Moz Ag sulfidation state sulfide assemblage dacitic dyke (7.8 Ma) cutting (2005); Deyell et al.
and intense acid sulfate alteration mineralization. (2005b)
Famatina −29.00 −67.78 4750–4350 3.8 Ma 0.4 Moz Vein hosted mineralization above Cambrian weakly Dacitic porphyries dated at 5 Ma are Lozada Calderón et al.
(40Ar/39Ar, Au; 3 porphyry center. Ore mineralogy metamorphosed marine shale associated with porphyry (1994); Pudack et al.
sericite) Moz Ag indicates high-sulfidation state and and siltstones of the Negro mineralization but no extrusive (2009).
includes famatinite, enargite, pyrite Peinado formation volcanic rocks known from the
and lesser, tennantite, tetrahedrite, district but ash-flow tuffs from the
sphalerite, gold, tellurides,covellite, eastern flank of the Famatina range
and chalcopyrite. may be age equivalent to porphyries.
La Coipa −26.81 −69.27 4150–3800 4350 20–17 Ma K/Ar, 3.5 Moz Primary sulfides include pyrite– Black shales and sandstones of Domes and lava flows of Cecioni and Dick (1992);
alunite Au; 230 enargite, at depth also bornite, the Triassic Ternera formation intermediate composition were Oviedo et al. (1991)
Moz Ag covellite, fahlore, galena and and overlying Oligocene dated at 24–22 as well as 16–14.7 Ma
spalerite. Much of the ore is oxidized 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
the district as well mineralization known
El Hueso −26.49 −69.53 3950–3700 n/a 40–36.2 Ma 0.84 Moz Mineralization is associated with Calcareous silt and Host rocks pre-date mineralization Marsh et al. (1997)
(40Ar/39Ar, Au; 3.9 illite–sericite and quartz alteration. sandstones, limestones of the by ~15 m.y. No age equivalent
alunite) Moz Ag Sulfide minerals include pyrite, Asientos Formation (u. volcanic rocks known but
chalcopyrite, galena, native bismuth Jurassic) and overlying subvolcanic porphyries are
and Bi–Pb sulfide minerals. This Paleocene andesite lavas and widespread and associated with
assemblage is overprinted by volcaniclastic rocks porphyry Cu mineralization at
advanced argillic alteration (alunite, Potrerillos
pyrophyllite) and pyrite. Most gold
was introduced during the first stage
El Guanaco −25.1 −69.53 2400–2900 n/a 48–42 Ma 1.6 Moz Vein hosted mineralization of Andesite lava flows and Volcanic host rocks pre-date Puig et al. (1988);
(K/Ar, alunite) Au enargite, pyrite and minor overlying dacitic pyroclastic mineralization by 12 or more m.y. No http://
chalcopyrite with quartz gangue in rocks. Both units were dated volcanic rocks age equivalent to www.australgold.com.au
andesite, smaller irregular veins in by K–Ar at ~60.2–60.7 Ma mineralization known. accessed Nov 2013.
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.
Choquelimpie −18.49 −69.39 4860–4700 n/a 6.6 Ma?? (K/Ar, 0.5 Moz Mineralization associated with Late Miocene andesite lavas Andesitic stratovolcano. Groepper et al. (1991)
host rock) Au hydrothermal breccia bodies in the and dacite subvolcanic dome Mineralization largely hosted in late
core of a late Miocene stratovolcano dacitic subvolcanic dome
Santa Rosa −16.63 −70.02 4950–4750 n/a 7.2 Ma 0.5 Moz Mineralization is hosted in vuggy Miocene (9–7 Ma) felsic Mineralization occurs adjacent to Barreda et al. (2004);
(Aruntani) (40Ar/39Ar, Au quartz zones centered on magmatic pyroclastic deposits are rhyolitic flow/dome complex and is Morche et al. (2008)
alunite) hydrothermal breccia bodies and is dominant ore host. closely associated with magmatic-
surrounded by quartz–alunite hydrothermal breccias.
alteration zones. Primary sulfides
include pyrite and enargite but the
deposit has been oxidized to a depth
of 300–400 m below surface. High-
grade mineralization is controlled by

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364


NNE oriented fractures.
Tucari −16.57 −70.19 5200–4900 n/a 4.6 Ma 2.1 Moz Mineralization is hosted in vuggy Intrusions and domes of Mineralization is spatially associated Barreda et al. (2004);
(Aruntani) (40Ar/39Ar, Au quartz zones centered on dacitic composition. with magmatic-hydrothermal Morche et al. (2008)
alunite) hydrothermal breccias and is breccias emplaced at the margin of
surrounded by quartz–alunite dacitic domes. The latter were
alteration zones. Primary sulfides emplaced in the core of an andesitic
include pyrite and enargite but the stratovolcano.
deposit has been oxidized to a depth
of 300–400 m below surface.
Julcani −12.95 −74.95 4500–4000 ~9.7 Ma (K/Ar 8 Moz Ag Pre-ore acid sulfate alteration with 10.1 Ma andesitic to dacitic Voluminous extrusive domes of Deen et al. (1994);
on biotite) vuggy quartz core and surrounding domes overlying paleozoic andesitic to dacitic composition, Noble and Silberman,
quartz–alunite zone in the center of phyllites of the Excelsior covering 16 km2. Syn mineralization 1984
the district, cut by discontinuous group 9.7 Ma dykes and post
quartz–tourmaline–pyrite breccias. mineralization 7 Ma rhyolite dome
Mineralization is hosted by laterally are also documented.
zoned veins with quartz–pyrite–
wolframite–enargite–tetrahedrite–
galena and barite rich zones. Zoning
is from high sulfidaton near the core
to distal intermediate sulfidation
assemblages).
Marcapunta/ −10.78 −76.27 4450–4850 n/a 11.9–10.6 Ma 2345 kt Zoned polymetallic district with a Eocene limestone and marls Mineralization is spatially associated Bendezu et al. (2008);
Colquijirca (40Ar/39Ar, Zn, 910 high-sulfidation Au–Ag prospect and and 12.9–12.4 Ma diatreme with a 12.9–12.4 Ma diatreme dome Vidal and Ligarda
alunite) kt Pb, distal carbonate replacement dome complex. complex (Marcapunta) pre-dating (2004);
1510 kt Pb–Zn–Ag mineralization. alunite by N0.8 m.y. A dacite dyke
Cu, dated at 14.1 Ma has been
6620 t documented W of Marcapunta. No
Ag, 64 t other igneous rocks recognized.
Au
Quicay −10.7 −76.26 4350–? n/a 37.5 Ma (K/Ar) n/a Mineralization is hosted in an Intermediate volcanic rock Subvolcanic to extrusive dome Rossell et al. (2006);
alunite (?) oxidized diatreme–dome complex. and hydrothermal breccias complex of uncertain age hosting Ingemmet.gob.pe,
Highest grades up to 3 g/t are hosted mineralization. Apparently no post- accessed Jan 2014.
in a central vuggy quartz zone. mineral volcanic rocks.

(continued on next page)

333
334
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364
Table 1 (continued)

Deposit/ Lat. Long. Elevation of Elevation of Mineraliza-tion Total Mineralization Host rocks Volcanology Selected references
district mineralization paleosurface age (method) resources
(m.a.s.l.) (if known) (approx.)

Cerro de −10.68 −76.27 4350–3900 n/a 15.4–14.4 12,250 kt Early pyrrhotite-pyrite–galena–Fe Late Triassic to early Jurassic Mineralization occurs at eastern Baumgartner et al.
Pasco (40Ar/39Ar, Zn, 3500 rich sphalerite (stage I) and later Pucara group limestone and border of a large 15.4 to 15.15 Ma (2008, 2009)
alunite) kt Pb, Cu–Ag (Au–Pb–Zn) enargite–pyrite mid Miocene diatreme diatreme–dome complex pre-dating
15,750 t veins as well as carbonate breccias. stage II mineralization by N 0.8 Ma
Ag, 1000 replacement Pb–Zn ore (stage II). but with unclear temporal
kt Cu relationship to stage I mineralization.
No volcanic rocks pre-dating the
dome complex known.
Pierina −9.46 −77.59 n/a 14.2–14.7 9 Moz Au Mineralization hosted in permeable Oligocene to Middle Miocene Late dacite flow dome complexes Fifarek and Rye (2005);
(40Ar/39Ar, pumiceous rock units and associated (29.3–14.8 Ma) andesitic to immediately pre-date Rainbow et al. (2005,
alunite) with intense acid sulfate alteration, dacitic tuffs, lavas and mineralization. No post mineral 2006)
including vuggy quartz and quartz subvolcanic domes of the volcanic rocks reported.
alunite. Sulfides include enargite, Huaraz group
pyrite as well as galena, sphalerite
and stibnite, and lesser
paragenetically late digenite and
covellite.
Lagunas −7.95 78.25 n/a 17.4–16.5 Ma 13.1 Moz Disseminated and fracture hosted Lower Cretaceous Chimú Volcanic rocks hosting Cerpa et al. (2013);
Norte (40Ar/39Ar, Au mineralization, centered on diatreme Formation quartzites (two mineralization were emplaced Montgomery (2012).
alunite) breccias. Mineralization in volcanic thirds) and early Miocene between early and main
rocks hosted in vuggy quartz and dacite tuffs and breccias (one mineralization stage. Volcanic rocks
quartz–alunite zones. Dominant third). were likely derived from diatreme
sulifdes include enargite, pyrite and breccias which also host
digenite. mineralization. Andesite lava and
dome adjacent to main diatreme has
same age as mineralization.
Yanacocha −7 −78.58 4100–3500 ~4200 11–8.2 Ma N70 Moz Massive and vuggy quartz zones Calipuy group (19.5–15.1 Ma) Mineralization overlaps in age with Longo et al. (2010)
(40Ar/39Ar Au hosting pyrite ± enargite– andesitic to dacitic lavas and voluminous andesitic to dacitic
Alunite) tennantite–covellite pyroclastic deposits and volcanic rocks. Most mineralization
14.5–8.4 Ma Yanacocha unit (N47 Moz Au) emplaced in the
dacite lavas and pyroclastic waning stages of volcanism
rocks represented by the felsic Coriwachay
dacite domes (10.9–8.4 Ma).
Sipan −6.9 −78.79 ~3500 n/a n/a n/a n/a n/a n/a Gustafson et al. (2004)
La Zanja −6.83 −78.89 3700–3400 n/a n/a 0.73 Moz n/a n/a n/a Compañía minera
Au, 6 Buenaventura 2012
Moz Ag Annual report, Gustafson
et al. (2004)
Tantahuatay −6.72 −78.67 4050–3800 n/a 12.4 Ma 5.1 Moz Gold is mostly associated to pyrite– Miocene andesite domes and Volcanic domes hosting the Gustafson et al. (2004);
(K/Ar alunite) Au; 108.3 enargite hosted in intensely quartz– to a lesser degree underlying mineralization are not age Noble and Mckee
Moz Ag pyrophyllite–alunite and vuggy to pyroclastic rocks constrained. Post mineral dyke of (1999); Compañía
massive quartz alteration. Much of restricted volume was dated at minera Buenaventura
the ore is oxidized 8.6 Ma (K/Ar, biotite). 2012 Annual report
Quimsacocha −3.04 −79.22 3700–3450 3900 7.5–7.3 Ma 3.6 Moz Stratabound Au-Ag mineralization Mineralization located in Andesite lavas and intercalated MacDonald et al. (2011,
(40Ar/39Ar Au, 27 associated with pyrite and enargite, pyroclastic units between less pyroclasitc units associated to a large 2012)
Alunite) Moz Ag hosted in vuggy quartz zones permeable coherent andesite stratovolcano complex. Andesite
lava flows of the Quimsacocha is ~ 9 Ma, pre-dating mineralization
Formation by 1.5 Ma. A central caldera and
dacite domes were emplaced at
6.7 Ma, 0.5 Ma after mineralization
La Bodega/La 7.38 −72.9 2900–2300 ~3600 3.5–1.7 Ma 3.5 Moz Dominantly structurally controlled Proterozoic gneisses and late No igneous rocks age equivalent to Rodriguez (2014)
Mascota (40Ar/39Ar, Au breccias cemented by quartz, alunite Triassic leucogranite. mineralization recognized. No
alunite, sericite) and high-sulfidation sulfide juvenile clasts in breccias observed.
assemblage, overprinting earlier

T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364


quartz–pyrite ± chalcopyrite veins
in phyllic alteration. Most Au–Ag in
late stages. Banded, colloform and
lattice textures normally typical for
low to intermediate sulfidation
deposits are common.
Angostura 7.39 −72.88 3500–2600 ~3600 4–1.9 Ma 2.7 Moz Widespread sheeted quartz– Proterozoic gneisses and late No igneous rocks age equivalent to Felder et al. 2005,
(40Ar/39Ar, Au pyrite ± chalcopyrite veins Triassic leucogranite. mineralization recognized. Rodriguez (2014).
alunite, sericite) associated with pervasive quartz–
sericite alteration. Locally cut by
enargite–Cu sulfide bearing breccias
and veins associated with quartz
alunite alteration. The latter hosts
highest grade Au zones.

335
336 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

6,800,000 El Indio belt geology


2 Ma Cerro de Vidrio rhyolite dome
6-5.5 Vallecito Fm. rhyolite

Au mineralization
8-7.6 Ma Pascua Fm. dacite
6,780,000

12.7-11 Ma Vacas Heladas Fm. dacite


Cerro del Toro (6168 m)
21-14 Ma Escabroso and Cerro
de las Tórtolas Fm. andesite
and subvolcanic diorite
27-23 Ma Tilito Fm. and equivalents
dacite, andesite and minor basalt

a
ra Ortíg
Paleozoic and Mesozoic basement
6,760,000

Thrust/reverse fault (symbol on

Cordille
hanging wall side)
Pascua-Lama Fault (unspecified)
Lineament from satellite image

High-sulfidation epithermal deposit


Veladero Major Mountains
6,740,000

Brea
ra de la
arrón

Sancarrón
Cordille
ra Zanc
Cordille
6,720,000

El Indio
ngüil

Cerro Doña Ana


(5625 m) Tambo
ra Cola
6,700,000

Cordille

Cerro de las Tórtolas (6160 m)


N
6,680,000

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 confirmed but (Chouinard et al., 2005). Based on light stable isotope evidence, jarosite
low volumes of juvenile igneous material probably played a role during precipitated from mixed meteoric and magmatic fluids whereas alunite
emplacement of the diatreme complexes and rare, locally weakly has a magmatic fluid signature (Deyell et al., 2005b). This indicates that
argillically altered, 7.8 Ma rhyodacitic dykes post-dating mineralization fluid sources varied through time, plausibly as a function of a fluctuating
have been recognized at Pascua (Bissig et al., 2001). but overall decreasing water table and short lived magma-derived fluid
The alteration and sulfide paragenesis reveals that unusually acidic pulses, as also documented for El Indio (see above).
fluids were responsible for alteration and some of the mineralization. Near Pascua–Lama, the Frontera–Deidad surface reaches elevations
Gold mineralization is associated with pyrite and enargite, as well as of 5200–5300 m a.s.l. (Fig. 5) The Azufreras–Torta surface is incised
alunite, jarosite and szomolnokite (Chouinard et al., 2005). Jarosite into this high-elevation pediplain at 4950–5100 m and extensive
is locally overprinted by hypogene alunite–enargite bearing veins steam-heated alteration and clastic deposits, both probably related to
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 337

El Indio/Tambo Area, Chile & Argentina


29°40'0"S 70°2'0"W 70°0'0"W 69°58'0"W 69°56'0"W 69°54'0"W 69°52'0"W 69°50'0"W

Chile Argentina
29°42'0"S

Río del Medio


LR
29°44'0"S

AT

6710000
Cerro Campana
El Indio

AT LR
29°46'0"S

AT

FD
Canto Sur
Veta Veronica
Kimberly
29°48'0"S

Tambo

Wendy
AT

6700000
29°50'0"S

LR
29°52'0"S

0 1 2 3 4 5 km

400000 410000 420000

Elevation (m) Slope (Degrees) Paleosurfaces


High : 6107
Los Ríos (LR); 10-6 Ma
0 - 10
Azufreras-Torta (AT); 14-12.5 Ma

Frontera-Deidad (FD); 17-15 Ma


Low : 1775

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, elevations (to 4800 m) than to the west (to 5000 m: Chouinard et al.,
are exposed on this landform (Bissig et al., 2002a). Mineralization occurs 2005) indicating that incision of the Río Turbio valley from the east
below the steam-heated zone at 4850 to 4600 m (Chouinard et al., 2005). may have occurred during hydrothermal activity.
As at El Indio and Tambo, a valley (Río Turbio valley: Fig. 5) incised the
Azufreras–Torta surface from the east. Ferricrete-cemented conglomer- 4.4. Veladero
ates at the valley-bottom display glacial striations, showing that this val-
ley pre-dated Plio-Pleistocene glacial erosion. Further, steam-heated Veladero is situated about 8 km to the SE of Pascua in a NNE striking
alteration also overprints the eastern side of the deposit to lower graben (Charchaflié et al., 2007). Gold and silver mineralization is
338 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

te

las s
fan

Tó de la
B

Ele

rto
o
Veta Verónica C Doña Ana

Co
Tambo

Co
Azufreras
(Kimberly pit)
Deidad

Steam-heated alteration Canto Sur pit


Frontera Deidad surface (17-15 Ma)
Azufreras-Torta surface (14-12.5 Ma)
Los Ríos surface (10-6 Ma)
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 from 5300 m to about 4100 m in the N–S oriented Río de las Taguas
(Fig. 6). Ore mineralogy comprises electrum and silver-bearing jarosite graben (Charchaflié et al., 2007, Fig. 6), whereas the Azufreras–Torta
which were deposited together with euhedral quartz, and are surface has been cut into Cerro de las Tórtolas volcaniclastic deposits
superimposed on earlier acidic residual quartz alteration (Holley, at an elevation of ~ 4500 m (Charchaflié et al., 2007, Fig. 6), i.e. at the
2012). Jarosite is abundant, especially at Amable, whereas sulfides are time of, or immediately prior to, hydrothermal activity in that area.
exceedingly rare. At Amable, jarosite forms euhedral crystals and Conversely, the Filo Federico mineralization probably formed at a time
grows in open spaces and veins, and is locally overgrown by barite. when incision of the Los Ríos surface was well advanced, as evidenced
Stable-isotope evidence indicates that most jarosite precipitated from by the steam-heated overprint of the ore zone and the mineralization
meteoric fluids above the water table, with sulfur derived from precur- extending to lower elevations than at Amable. The lowering of the
sor sulfides. However, one location in the Fabiana zone to the east of water table in the Veladero area during Los Ríos surface incision is also
Veladero (Figs. 5, 6) shows unambiguous hypogene jarosite, replacing reflected in the formation of jarosite in the vadose zone at Amable
plagioclase and biotite phenocrysts in the host rock and displaying between 11.8 and 8.5 Ma.
stable isotope signatures of magmatic fluids (Holley, 2012).
Mineralization at Amable is constrained between a 12.7 Ma alunite 5. Maricunga belt
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 The Maricunga belt is defined as the region of the Andean Cordillera
Filo Federico is at 11.1 to 10.3 Ma (40Ar/39Ar, alunite, n = 4) somewhat between 26° and 28° Lat S (Vila and Sillitoe, 1991), located at the north-
younger than at Amable (Bissig et al., 2001; Holley, 2012). Jarosite ages ern margin of the Pampean flat slab segment. Gold-rich porphyry type
at Amable range from 11.8 to 8.6 Ma and overlap with hypogene alunite deposits (e.g., Cerro Casale (a.k.a. Adelbarán), Caspiche, Lobo, Marte)
and gold deposition at Filo Federico (Holley, 2012). These data are are the defining mineralization style and the gold-rich nature of
interpreted as evidence for an overall lowering of the water table with porphyry-style mineralization is thought to be related to the shallow
time, episodically interrupted by pulses of magma derived fluids. emplacement of the porphyry intrusions (Muntean and Einaudi,
Jarosite at Veladero almost entirely pre-dated hypogene alunite alter- 2001). Although barren quartz–alunite alteration is commonly found
ation and sulfide mineralization at nearby Pascua−Lama. in the shallow portions of these porphyry systems, La Pepa is the only
The host rocks of the Amable ore zone consist of crudely stratified porphyry deposit where significant gold mineralization hosted in
volcaniclastic breccias assigned to the Cerro de las Tórtolas formation, quartz–alunite veins that overprint porphyry-style mineralization has
unconformably overlying Tilito Formation andesites and dacites as been documented (Muntean and Einaudi, 2001). In contrast, La Coipa,
well as Permian felsic tuffs (Charchaflié et al., 2007; Holley, 2012). The located at the northern end of the Maricunga belt, hosts Ag–Au miner-
Filo Federico ore zone is partly hosted in volcaniclastic deposits assigned alization exclusively as epithermal mineralization of largely high-
to the Vacas Heladas Formation (Holley, 2012). Mineralization is closely sulfidation type (Cecioni and Dick, 1992; Oviedo et al., 1991).
associated with hydrothermal breccias intruding the volcaniclastic de- Porphyry and epithermal deposits of the Maricunga belt occur in
posits and occurs between 4400 and 3900 m a.s.l. in a subhorizontal, two partly overlapping parallel N–S striking belts, with mineralization
N − S elongated ore body, extending extends to lower elevations at traditionally considered to have been emplaced during two distinct ep-
Filo Federico than at Amable. A steam-heated alteration zone of up to isodes at 25–20 Ma and 15–13 Ma (Sillitoe et al., 1991). However, more
200 m thickness overlies the orebodies and defines the paleosurface at recently published, but still limited geochronology requires modifica-
~4400–4600 m a.s.l. At Filo Federico, steam-heated alteration is local- tion of this model. The Caspiche porphyry Au deposit was dated at
ized along fracture zones, and overprints the mineralized zone down ~25.4 ± 0.01 Ma by Re/Os on molybdenite (Sillitoe et al., 2013) whereas
to more than 500 m below the inferred paleosurface (Holley, 2012). the Caserones porphyry Cu deposit in the southern Maricunga belt
Here, the Frontera–Deidad surface is marked by the lower contact of formed at 20–18 Ma (Mpodozis and Kay, 2003). Further, alunite K/Ar
the Cerro de las Tórtolas formation which here has been downfaulted ages from the La Coipa district range from ~ 20 to ~ 17 Ma (Oviedo
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 339

Pascua-Lama/Veladero Area, Chile & Argentina


70°0'0"W 69°55'0"W

Chile
Argentina

6760000
AT
AT
Pascua-Lama
29°20'0"S

Río
Tu
rbi
Penelope o
Lama Central
LR

Veladero
AT

6750000
AT Fabiana Cerro de Vidrio

FD
LR
29°25'0"S

AT
AT

0 1.5 3 6 km

400000 410000

Elevation (m) Slope (Degrees) Paleosurfaces


High : 5678 Los Ríos (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 Charchaflié et al. (2007). UTM Zone 19S, WGS84.

et al., 1991) which disagree with the range between 25 and 20 Ma (Ramos, 2008). Thus, the Precordillera (a.k.a. Cordillera Domeyko),
established on the basis of alunite K/Ar ages (Sillitoe et al., 1991) and resulting from the late Eocene Incaic deformation is present north of
biotite and hornblende K/Ar ages for dacite domes that were thought 29° Lat. S but absent south of it (Rodriguez et al., 2014), a location
to be closely related to mineralization (Oviedo et al., 1991; Sillitoe which roughly coincides with the inferred southern limit of the
et al., 1991). Antofalla massif and the northern limit of the Chilenia terrane (Ramos,
The large-scale morphotectonic features of the Andes are different at 2008). Oligocene extensional tectonics are documented from south of
the latitude of the Maricunga belt to those of the El Indio belt, probably the Vallenar Orocline (Abanico basin, Tilito Fromation: Winocur et al.,
influenced by the differing Paleozoic and older basement architecture 2014; Charrier et al., 2002) whereas a contractile tectonic regime at
340 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Cerro Pelado
Amable Filo Federico Cerro Nevado
Penelope Río Turbio
Lama Central

Pascua/Lama

s
g fault
nb oundin
grabe

Fabiana

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

Fig. 6. Landscape in the Veladero and Pascua–Lama 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 5.1. La Coipa
the latitude of the Maricunga belt during the late Oligocene (Mpodozis
and Clavero, 2002). The dominant landscape elements at the latitude of The La Coipa district contains multiple epithermal deposits of mostly
the northern Maricunga belt include the probably Oligocene relict Sierra high-sulfidation type including Ladera Farellón, Can Can, Coipa Norte,
Checos de Cobre surface which is incised by the upper Oligocene Purén, Purén Sur and Pompeya (Fig. 7). With the exception of Purén
to ~ 18 Ma Asientos pediplain in the eastern Precordillera (Bissig and and Pompeya, mineralization is hosted in the Triassic La Ternera forma-
Riquelme, 2009; Mortimer, 1973). The Oligocene to early Miocene land- tion, consisting of black shales and sandstones, as well as in overlying
scape was incised by the middle- to early-upper Miocene Atacama upper Oligocene (24–22 Ma) dacitic tuffs, volcaniclastic breccias and
pediplain and late Miocene canyons (Bissig and Riquelme, 2009) domes (Cecioni and Dick, 1992; Oviedo et al., 1991). Hydrothermal
which record a progressive tilting of the western Andean slope and up- breccias hosting ore are probably penecontemporaneous with mineral-
lift of the Andean Cordillera since the early Miocene (Riquelme et al., ization. Alunite K/Ar ages range between ca. 20 and 17 Ma (Oviedo et al.,
2007). 1991). A relict, large, upper Oligocene volcanic center, Cerros Bravos
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 341

La Coipa Area, Chile


69°18'0"W 69°16'0"W 69°14'0"W 69°12'0"W

7040000
26°46'0"S

Purén

L. Oligocene/E. Miocene paleosurface


Purén Sur

Coipa Norte

Pompeya
26°48'0"S

Can Can

Ladera Farellón
26°50'0"S

7030000
26°52'0"S

0 0.5 1 2 3 4 5 km

470000 480000

Elevation (m) Slope (Degrees)


High : 4907 Outline of post-mineral
0 - 10 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 Farellón. Unaltered, of advanced argillic alteration and the occurrence of various carbonates
16–14.7 Ma dacitic domes and lavas post-date mineralization (Oviedo including rhodochrosite as alteration minerals (S. Gamonal, pers.
et al., 1991) and locally overlie the mineralized zones at Pompeya commun. 2014). At Purén Sur, some 2.5 km south of Purén, native
(Figs. 7, 8). The Pompeya and the Purén ore bodies were discovered sulfur associated with steam-heated alteration was historically mined.
after 2010, and only Purén has been mined. Pompeya is located ~3 km Steam-heated alteration is also present at Pompeya, Purén and Coipa
NE of Ladera Farellón and is of high-sulfidation type. Purén, 8 km Norte and is generally preserved at 4250 to 4400 m a.s.l. (Figs. 7, 8).
NE of Ladera Farellón features high grade Pb–Zn mineralization Mineralization at Ladera Farellón, Can Can and Coipa Norte is largely
representing an intermediate sulfidation sulfide assemblage, an absence controlled by subvertical fractures when hosted in Triassic shales, but
342 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

16 Ma dacite lava
Pompeya 16 Ma dacite dome

Steam-heated
alteration zone

Late Oligocene/early Miocene erosion surface

Can Can Coipa Norte 16 Ma


Ladera Farellón
dacite dome
D

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 Farellón, 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. to ~ 3800 m a.s.l. (Cecioni and Dick, 1992). The dominant alteration
The latter hosts higher Ag grades, whereas the former is relatively is vuggy–quartz and quartz–alunite in the volcanic rocks and quartz–
Au-rich (Oviedo et al., 1991). The top of the exposed orebodies lies alunite ± dickite and pyrophyllite in the shale and sandstones.
around 4300 m a.s.l., with sulfide oxidation extending downwards There is a close association of alunite vein stockworks with gold
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 343

mineralization (Oviedo et al., 1991). The sulfides observed at depth in- (Peru), Cerro Colorado, Spence and El Salvador (Bissig and Riquelme,
clude pyrite, chalcocite, covellite, bornite, chalcopyrite, tennantite, 2010; Bouzari and Clark, 2002; Gustafson and Hunt, 1975; Quang
tetrahedrite, galena and sphalerite (Oviedo et al., 1991), representative et al., 2003, 2005). The late Eocene climate was probably semi-arid
of precipitation at intermediate to high-sulfidation states. No porphyry- and considerably wetter than the present day since sufficient water is
style mineralization related to epithermal mineralization is known at La necessary for economically significant supergene enrichment and
Coipa. erosion.
The geomorphology of the La Coipa district is dominated by a planar,
slightly west-inclined landscape constructed on the upper Oligocene 6.1. El Hueso
pyroclastic rocks and domes, but covered by the post mineral
16–14.7 Ma dacite lavas and extrusive domes which form the topo- Arguably, nowhere in the Andes is the interrelationship between
graphic highs in the district (Cornejo et al., 1998; Figs. 7, 8). This planar major orogenic episodes and porphyry and epithermal mineralization
surface is located at present day elevations of 4400–4200 m a.s.l., and better demonstrated than in the Potrerillos–El Hueso district (Cornejo
appears to have been the dominant paleosurface at the time of et al., 1993; Marsh et al., 1997; Niemeyer and Munizaga, 2008;
epithermal mineralization as steam-heated mineralization is commonly Thompson et al., 2004). At El Hueso, epithermal mineralization is locat-
exposed on it (Fig. 8). The paleosurface that plausibly controlled ed about 3 km E of the Potrerillos porphyry Cu deposit, in the
epithermal activity can be correlated with the upper Oligocene hangingwall of the SE-verging Potrerillos Mine thrust fault which was
to ≥ 18 Ma Asientos pediplain mapped on the northern side of the active during and immediately after porphyry and epithermal mineral-
Cerros Bravos volcanic complex (Bissig and Riquelme, 2009). The uplift ization (Marsh et al., 1997; Niemeyer and Munizaga, 2008). The district
and subsequent erosion of the Asientos pediplain is probably related to also hosts the Jerónimo carbonate-hosted epithermal gold deposit
the late Oligocene contractile tectonic phase described from the Cordil- (Thompson et al., 2004), located some 1.5 km to the east of El Hueso,
lera Claudio Gay some 50 km E of La Coipa (Mpodozis and Clavero, in the footwall of the Potrerillos Mine Thrust fault.
2002). Gold and silver mineralization at Ladera Farellón and Pompeya Epithermal alteration and mineralization at El Hueso are hosted in
is located near the upper reaches of a NE oriented valley, incised into calcareous silt and sandstones of the Upper Jurassic Asientos formation,
the probable upper Oligocene to lower Miocene paleosurface, but the as well as the uncomformably-overlying andesitic-to-rhyolitic lavas and
timing of valley erosion is currently unknown. volcaniclastic rocks of the Paleocene Hornitos Formation (Marsh, 1997).
Hydothermal activity occurred in two stages. An early mineralization
5.2. La Pepa 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
Within the Maricunga belt, besides la Coipa, only La Pepa contains dominated by sericite and illite, but incorporated only minor advanced
significant high-sulfidation epithermal mineralization. Here it is hosted argillic alteration. A second stage of barren quartz–alunite alteration
in NNW-trending quartz alunite veins that cross-cut porphyry-Au style and silicification occurred at ca. 36.2 Ma (King, 1992; Marsh et al.,
mineralization centered on the Cavancha porphyry (Muntean and 1997; Olson, 1984). The early stage deposited gold, accompanied by
Einaudi, 2001). Similar quartz–alunite veins are widespread in other 1–2% pyrite, stibnite, chalcopyrite, galena and arsenopyrite, all hosted
porphyry Au deposits of the Maricunga belt, but only those at La Pepa in quartz veinlets (Marsh et al., 1997). The principal porphyry Cu–Au
contain gold grades high enough to sustain small-scale historic produc- mineralization centered on the Cobre porphyry was emplaced about
tion (Muntean and Einaudi, 2001). La Pepa is also the only one of these 3 km west of El Hueso and immediately after the epithermal mineraliza-
deposits where alunite precipitation (between 23.5 and 23.25 Ma, tion and alteration at 35.5–35.9 Ma (Marsh et al., 1997). Structural anal-
based on 40Ar/39Ar plateau ages), did not overlap temporally with the ysis of the district suggests that the emplacement depth of porphyry
potassic alteration, here dated at 23.81 ± 0.08 Ma (hydrothermal biotite Cu–Au mineralization at the Cobre porphyry was shallow, as little as
40
Ar/39Ar plateau age, Cavancha porphyry; Muntean and Einaudi, 2001). 1 km below the paleosurface (Niemeyer and Munizaga, 2008). Although
The late Oligocene landscape at La Pepa is obscured by extensive upper a number of Eocene porphyry stocks intrude the Jurassic and Paleocene
Miocene ignimbrite deposits through which La Pepa is exposed in an rocks in the district, no age-equivalent volcanic rocks are recognized
erosional window (Muntean and Einaudi, 2001). (Marsh et al., 1997). In the El Salvador porphyry Cu district, located
some 20 km NW of Potrerillos, hypogene mineralization occurred at
6. High-sulfidation epithermal deposits of the Domeyko fault system 43–41 Ma (Gustafson et al., 2001), but the deposit was exposed to oxi-
dation and supergene modification at ca. 36 Ma (Bissig and Riquelme,
Two noteworthy middle to late Eocene high-sulfidation epithermal 2010; Gustafson and Hunt, 1975; Mote et al., 2001). Thus, uplift and ero-
deposits, El Hueso and El Guanaco, are located in the southern Atacama sion exhumed porphyry style mineraliztion at El Salvador about 6 Ma
Desert. These deposits occur near the Domeyko fault system of the after hypogene mineralization, i.e. at the time of hydrothermal activity
Chilean Precordillera, which also contains the behemothian Escondida, at el Hueso and Potrerillos. The Precordillera (a.k.a. Cordillera Domeyko)
Chuquicamata, and Collahuasi porphyry Cu districts. All of these at the latitude of the present day southern Atacama Desert is located at
deposits were emplaced during, or immediately after the late Eocene the southewestern boundary of the Puna Plateau and probably formed
Incaic orogenic event which has been ascribed to an inferred episode part of the Andean crest from late Eocene until the late Oligocene, before
of slab flattening, crustal thickening, volcanic lull and subsequent east- late Oligocene thrusting in the Cordillera Claudio Gay, some 70 km east
ward shift of the volcanic arc (Mpodozis and Cornejo, 2012; O'Driscoll of the Precordillera, established the main Andean range as known today
et al., 2012). (Bissig and Riquelme, 2010).
Regional uplift and formation of large scale erosional surfaces of The geomorphology of the area around El Hueso is dominated by a
Incaic age are implied from constraints on supergene activity and subplanar erosion surface, now situated at 3900–4000 m elevation
large scale morphotectonic features such as the Precordillera which which has been cut into the Paleocene Hornitos formation (Fig. 9).
formed during the Incaic orogeny in the southern Atacama Desert of This erosion surface is interpreted as a pediplain and its age is uncertain
northern Chile. Uplift in the Eocene to elevations similar to the present but probably represents a somewhat erosion-degraded relic of a
day is also evidenced from stable-isotope paleoaltimetry for the Puna pediplain that formed in response to uplift related to the late Eocene
plateau (Canavan et al., 2014) as well as for the pre-Cordillera at ~26° to early Oligocene Incaic orogeny (Fig. 9). In a broad context, El Hueso
Lat. S (Bissig and Riquelme, 2010). Supergene oxidation and Cu enrich- is located below a topographic high, near a scarp that separates it
ment in Paleocene to middle Eocene porphyry deposits in response to from topographically lower areas to the west and may have been
uplift and erosion, initiated as early as 45 to 36 Ma at Cerro Verde emplaced in a physiographic setting comparable, albeit less well
344 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Cerro El Hueso
El Hueso epi-
thermal 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., Pascua–Lama or El Indio. Given the widespread pres- 7. Late Miocene high-sulfidation epithermal Au deposits of the
ervation of late Eocene to early Oligocene epithermal alteration and western Cordillera of northern Chile and southern Peru
mineralization styles at most about 200–300 m of erosion occurred
since hydrothermal activity in the El Hueso area. Erosion modification High-sulfidation epithermal deposits are rare in the Central Volcanic
of the original Incaic erosion surface probably occurred after 33 Ma as zone, between ~25° Lat S in northern Chile and 15° Lat S in Central Peru,
it affects the granodiorite stock of Cerro El Hueso dated at ca. 33 Ma an area that coincides with the major Chile–Peru orogenic bend. Here,
(Cornejo et al., 1993), but prior to 18 Ma, which is the age of the the subduction angle is at ~ 30° steeper than in the present flat-slab seg-
~100–200 m lower Asientos pediplain cutting into the high level post ments of northern Chile and central and northern Peru (Cahill and
Incaic surface at El Hueso (Bissig and Riquelme, 2009). The base level Isacks, 1992). The current slab geometry resulted from steepening of
of the Asientos pediplain is the Salar de Pedernales which was an Eocene shallow angle slab (Sandeman et al., 1995). The Andes at
established in the late Oligocene (Bissig and Riquelme, 2009). The the Chile–Peru oroclinal bend are metallogenetically diverse and in-
Oligocene to early Miocene landscape elements and underlying shallow clude W, Sn, Ag, Pb, Zn and U deposits in the Inner Arc (partly coinciding
level epithermal mineralization in the El Hueso area are preserved due with the Cordillera Oriental) ranging from Jurassic to Pleistocene age, as
to the arid climate. Erosion since the early Miocene was mostly concen- well as porphyry and skarn type mineralization of Paleocene and
trated in canyons and average denudation rates were low (~14 m/Ma: Eocene ages (Clark et al., 1990; Perelló et al., 2003). However, a few
Riquelme et al., 2008). Miocene epithermal gold deposits occur in the Cordillera Occidental of
northern Chile and southern Peru, all of which are located at local topo-
graphic highs with mineralization exposed at elevations of 4800 to
6.2. El Guanaco 5200 m a.s.l. Significant deposits include the 6.6 Ma Choquelimpie
high-sulfidation deposit in Chile (Groepper et al., 1991) and the Santa
El Guanaco is a second example of a high-sulfidation epithermal de- Rosa and Tucarí high-sulfidation deposits of the Aruntani district in
posit spatially associated with the Domeyko fault system. It is located southern Peru (Barreda et al., 2004; Morche et al., 2008). Hypogene
30 km W of the crest of the Precordillera at 25.1° Lat. S. Mineralization alunites from Santa Rosa and Tucari have been dated at 6.4 and
is hosted in roughly E to ENE-trending quartz replacement veins with 4.6 Ma, respectively (Morche et al., 2008).
a narrow advanced argillic and argillic alteration halo. Sulfides include Mineralization at Choquelimpie is hosted in the core of a stratovolca-
enargite, pyrite and subordinate chalcopyrite. Veins were emplaced no interpreted as broadly cogenetic with epithermal mineralization
in andesite lavas and overlying dacite tuffs of Paleocene age (ca. (Groepper et al., 1991). At Aruntani, multiple episodes of volcanism,
54–60 Ma; Puig et al., 1988). The veins are well-defined, subvertical, represented by dacitic plugs and subvolcanic intrusions as well as an-
tabular structures when hosted in the andesite, but assume an irregular desitic lavas both pre-dated and overlapped with epithermal minerali-
behavior in the overlying, more permeable felsic pyroclastic rocks. zation (Barreda et al., 2004). Thus, the volcanologic setting of these
Supergene oxidation affected the deposit up to 300 m depth and gave deposits is distinct from the deposits farther south in that mineraliza-
rise to a large variety of arsenate minerals, El Guanaco being the type- tion is temporally more closely associated with voluminous volcanism
locality for some of these (Witzke et al., 2006). and does not post-date the age of their host rocks by 5 or more Ma.
At El Guanaco, hydrothermal alunite yields K/Ar ages between ~48 The geodynamic setting of late Miocene high-sulfidation deposits of
and 42 Ma (Puig et al., 1988). No age-equivalent intrusive or volcanic the Cordillera Occidental also differs from the current flat slab segments
rock or porphyry-style mineralization has been documented in the dis- in that magmatism is associated with contraction of the arc due to
trict. El Guanaco is the oldest-known high-sulfidation epithermal de- slab steepening rather than arc-broadening and cessation related to
posit of the Andes, but overlaps in age with the El Salvador and shallowing subuduction angles as observed in the Pampean and
Esperanza porphyry Cu (–Au) districts (Gustafson et al., 2001; Perelló Peruvian flat slab segments (cf. Bissig et al., 2003, 2008; Kay and
et al., 2004), similarly located west of the main Domeyko fault system. Mpodozis, 2001; Sandeman et al., 1995).
Mineralization at El Guanaco is located below Cerro Estrella, a local The landscape evolution documented on the western Andean slope
topographic high at ~ 2900 m a.s.l., and extends to depths of at least indicates that multiple stages of pediment bevelling due to uplift of
2500 m a.s.l. Cerro Estrella rises about 300 m above surrounding relict the Incaic paleosurface occurred between ca. 24 Ma and 8 Ma (Quang
pediment surfaces of unknown age. However, El Guanaco's physio- et al., 2005; Tosdal et al., 1984) followed by a major pulse of Altiplano
graphic setting is consistent with a position below a high-level Incaic uplift and incision of deep Canyons (Thouret et al., 2007). The majority
or pre-Incaic erosion surface being incised by pediments either during of workers agree that most of the Altiplano uplift to average elevations
or after hydrothermal activity. of 3800 m, occurred during the Miocene. Evidence from both
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 345

geomorphology (Hoke et al., 2007; Jordan et al., 2010; Schildgen et al., However, the polymetallic belt also contains a number of epithermal
2007) and paleoaltimetry (based on stable isotope and paleobotany; deposits with important high-sulfidation characteristics; these are
Garzione et al., 2008) suggest that at least 1000 m or more of this eleva- described in the following.
tion 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 9.1. Julcani
pulse.
The Julcani district, mined since colonial times, consists of several-
8. Central to northern Peruvian flat slab segment vein hosted deposits from which Ag and base metals, as well as subordi-
nate Au, have been produced. The mineralization is hosted in coalescing
The Andean segment extending from ~15° Lat S in Peru, to 2° Lat S in domes of andesitic-to-dacitic composition and postdates a large zone of
southern Ecuador, lacks recent arc volcanism, which, as in Chile, has vuggy quartz and quartz alunite alteration in the center of the district, as
been attributed to flat-slab subduction spatially coinciding with subduc- well as quartz–tourmaline–pyrite breccia bodies (Deen et al., 1994;
tion of aseismic ridges and oceanic plateaus (e.g., Gutscher et al., 1999a; Petersen et al., 1977). Veins are zoned along strike from higher
Martinod et al., 2010; Skinner and Clayton, 2013). The high-sulfidation sulfidation state enargite–tennantite bearing assemblages near the cen-
epithermal deposits of this region occur in two distinct belts (Noble ter of the district to lower-sulfidation assemblages containing galena
and McKee, 1999). The southeastern belt, located east of the Cordillera and tennantite–tetrahedrite more distally (Petersen et al., 1977).
Occidental, which forms the Continental divide, and extends from Magmatism largely predated the epithermal mineralization but a
Julcani (~13° Lat S) to the latitude of Antamina (~9.5° Lat S). It coincides 9.7 Ma syn-mineral dyke and small, 7 Ma, post-mineral rhyolite
with the central Peruvian polymetallic province and contains a variety domes have been recognized (Deen et al., 1994; Petersen et al., 1977).
of porphyry-related and carbonate hosted deposits emplaced at differ- Julcani is hosted below a sub-planar, slightly east-inclined, land surface
ent paleodepths (Bissig and Tosdal, 2009; Escalante, 2008; Escalante situated at elevations of 4000–4500 m a.s.l.
et al., 2010; Love et al., 2004). This segment includes the cordilleran
base metal lode deposits (i.e. sulfide rich polymetallic deposits: 9.2. The Marcapunta–Colquijirca district
Bendezú et al., 2003) of Marcapunta–Colquijirca and Cerro de Pasco
which are centered on high-sulfidation epithermal deposits emplaced Marcapunta is a high-sulfidation epithermal Au (–Ag, Cu) deposit.
in reactive host rocks (Bendezú and Fontboté, 2009; Bendezú et al., The district also contains cordilleran base metal mineralization at the
2008). The northwestern metallogenetic belt has its southern terminus Smelter deposit, northward contiguous to Marcapunta, and more distal-
at Pierina at 9.5° Lat S, west of the Cordillera Blanca and is dominated by ly the Colquijirca Ag–Pb–Zn deposit, some 5 km north of Marcapunta.
high-sulfidation epithermal and porphyry Cu–Au deposits but also The San Gregorio deposit is another cordilleran base metal deposit
includes a number of smaller vein hosted Ag–Au deposits such as southerly adjacent to Marcapunta (Bendezú et al., 2008). Vuggy quartz
Quiruvilca (Gustafson et al., 2004; Noble and McKee, 1999). It includes, alteration, phreatomagmatic breccias and disseminated gold minerali-
from south to north, Pierina, Lagunas Norte and Yanacocha, the latter of zation occur in the central Marcapunta dome, whereas stratabound
which contains upwards of 70 Moz contained Au (Longo et al., 2010; base metal and silver rich mineralization at Smelter and Colquijirca,
Teal and Benavides, 2010) and which is by far the largest epithermal extending as far as 5 km north from Marcapunta, are hosted in Eocene
deposit cluster of the Andes. limestones and marls (Bendezú et al., 2008). There is a district-scale
Subduction of the Nazca aseismic ridge in Peru commenced in the lateral zoning from proximal high-sulfidation mineralization containing
middle Miocene at 14–10 Ma (Hampel, 2002) and is the inferred enargite and pyrite to intermediate pyrite, chalcopyrite and tennantite-
cause of flat subduction (e.g., Martinod et al., 2010). However, estimates bearing ore, to a distal pyrite, sphalerite and galena dominated sulfide
on the timing of crustal thickening and uplift of the Cordillera Occiden- assemblage, the later reflecting an intermediate sulfidation state
tal in central and northern Peru vary from the early to the middle Mio- (Bendezú et al., 2008; Vidal and Ligarda, 2004). Alunite associated
cene and do not coincide everywhere with the onset of aseismic ridge with the central high-sulfidation gold mineralization yielded ages of
subduction and flat subduction. Thus, Noble et al. (1990) on the basis 11.9 to 11.1 Ma, whereas that associated with the base metal minerali-
of the age of tuff deposits in deeply incised paleovalleys in northern zation yielded slightly younger ages of 10.8 to 10.5 Ma (Bendezú et al.,
Peru, suggested that uplift and erosion occurred in the early–late 2008). The dacitic dome complex hosting some of the epithermal
Miocene. Crustal thickening inferred from whole rock geochemical mineralization was emplaced at 12.9 to 12.1 Ma (40Ar/39Ar biotite
data of volcanic rocks in the central Peruvian polymetallic province is ages). No syn- or post mineral volcanic or intrusive rocks are exposed
inferred to have occurred around 12–15 Ma (Bissig and Tosdal, 2009). in the district.
In contrast, Montgomery (2012) proposed that a major episode of uplift Colquijirca is located near the eastern margin of a regionally exten-
commenced at about 17 Ma, i.e., before the inferred onset of ridge sub- sive low-relief plain situated at an elevation of 4200 to 4300 m a.s.l.
duction, but coincident with high-sulfidation epithermal mineralization whereas the Cerro Marcapunta summit at 4450 m a.s.l. constitutes the
at Lagunas Norte (Cerpa et al., 2013; Montgomery, 2012). highest feature in the deposit area. Vuggy quartz hosted gold minerali-
zation extends from 4450 m to about 4000 m a.s.l. or 450 m below sur-
9. High-sulfidation epithermal deposits of the central Peruvian face, whereas distal base metal mineralization occurs at depths less than
polymetallic belt ~400 m below surface. Proximal enargite-rich mineralization adjacent
to the dome complex, however, has been drilled to a depth of N600 m
The central Peruvian polymetallic belt contains a number of late below surface and attains thickness of up to 100 m (Bendezú et al.,
Miocene silver and base metal-rich vein deposits that were emplaced 2008; Vidal and Ligarda, 2004).
in the epithermal environment. These include: San Cristóbal (Beuchat
et al., 2004), Morococha (Catchpole et al., 2011) and Uchucchacua 9.3. Cerro de Pasco
(Bussell et al., 1990; Escalante, 2008). At Morococha an evolution from
the porphyry to the epithermal environment occurred concurrently Cerro de Pasco is located about 11 km N of Marcapunta and shares
with erosion, and late Miocene epithermal Ag mineralization probably many similarities to the geologic setting and mineralization style of
occurred only a few 100 m below the paleosurface (Catchpole et al., Colquijirca. The deposit is spatially related to a 15.4 to 15.2 Ma dacitic
2011). All of these epithermal deposits have sulfide assemblages of in- diatreme dome complex which was emplaced at the boundary of Devo-
termediate sulfidation state and veins are largely hosted in Mesozoic nian phyllites to the west and Triassic-to-Jurassic limestones to the east
carbonaceous rocks as well as Triassic volcanosedimentary rocks. (Baumgartner et al., 2008, 2009). Mineralization occurred in two stages
346 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

(Baumgartner et al., 2008; Einaudi, 1977). Stage 1 was focused at the Hydrothermal alunite 40Ar/39Ar ages range from 15.08 ± 0.09 to
eastern margin of the diatreme dome complex and is characterized 13.89 ± 0.13 Ma (n = 19), clustering in two pulses around 15 Ma and
by a low-sulfidation sulfide assemblage consisting of a massive 14.4 Ma (Rainbow, 2009), whereas an age of rare, vug filling porcellane-
quartz–pyrite ± pyrrhotite body with associated distal carbonate re- ous alunite from the oxide zone, yielded a large-error 14.12 ± 1.59 Ma
placement ore containing galena and Fe-rich sphalerite. Alteration relat- plateau age and may be of supergene origin (Rainbow, 2009). This
ed to stage 1 is quartz–sericite. The second mineralizing stage reflects suggests that oxidation closely followed hypogene mineralization. The
high-sulfidation states and has proximal E–W trending enargite–pyrite host rocks of the Huaraz group range in age from 29.3 to 14.8 Ma
veins, hosted in part by the diatreme breccia body, and distal carbonate- (Rainbow et al., 2005; Strusievicz et al., 2000).
hosted Pb–Zn–Ag veins and replacement bodies containing Fe-poor The mineralized part of Pierina is located between about 4000 and
sphalerite (Baumgartner et al., 2008). Alunite is part of the alteration as- 3800 m a.s.l., starting about 100 below the nearest topographic highs.
semblage of stage 2 and gives 40Ar/39Ar ages of 14.5 to 14.2 Ma, whereas A horizontally extensive steam-heated alteration blanket has not been
stage 1 is bracketed between 15.4 and 14.5 Ma (Baumgartner et al., 2009). documented at Pierina, and the presence of any steam-heated alunite
Like the nearby Marcapunta and Colquijirca deposits, Cerro de Pasco remains controversial (Fifarek and Rye, 2005; Rainbow et al., 2005,
is associated with a dome complex which forms a local topographic high 2006). However, paragenetic relationships and stable-isotope geo-
above the regionally-extensive, low-relief surface to the west of it. The chemistry demonstrate that meteoric waters played an increasingly im-
pre-mining land surface was situated at about 4350 m a.s.l. and miner- portant role in deposit formation, from early alteration to subsequent
alization extends about 450 m below this original land surface oxidation, and that these fluids became progressively less isotopically
(Baumgartner et al., 2008). exchanged over time (Rainbow et al., 2006). This suggests that the
water table was progressively lowered during the lifetime of the hydro-
9.4. Quicay thermal system (Fifarek and Rye, 2005; Rainbow et al., 2005, 2006).
In the Pierina–Santo Toribio area, components of middle Miocene
Quicay is a high-sulfidation epithermal center located 14 km W of planar erosional landforms both preceding and broadly contemporane-
Cerro de Pasco. It is being mined by the Peruvian private company ous with mineralization at Pierina can be recognized (Fig. 11). The de-
Chancadora Centauro SAC and little information on its geology and posit formed at the crest of the Cordillera Occidental and Calipuy arc
resources is publically available. According to information obtained facing the Amazonian low-lands to the east, the latter not yet separated
through the INGEMMET website Peru (Rossell et al., 2006) gold miner- hydrographically from the Cordillera Negra as the intervening Cordillera
alization is hosted in a diatreme–dome complex and the highest gold Blanca was only uplifted in the late Miocene (González and Pfiffner,
grades are found in the central vuggy quartz zones (average 3 ppm). 2012; Petford and Atherton, 1992). An erosion surface, marked by an
The ore is oxidized. A mineralization age of 37.5 Ma based on a K/Ar al- angular unconformity, now tilted and dipping at ~ 20° to the ENE, un-
unite age is suggested by Noble and McKee, (1999) which is significant- derlies the andesites below the Pierina deposit. The andesites have
ly older than Cerro de Pasco, despite the similar inferred shallow level of been dated at ca. 21 Ma, suggesting that this angular unconformity re-
formation. Quicay mineralization is centered on a small hill that con- cords uplift and erosion in the late Oligocene to early Miocene, probably
tains outcrops of vuggy residual quartz. It has a summit elevation of representing the Aymará orogenic event (Sébrier et al., 1988).
4350 m, which is roughly 100 m above the same low-relief land surface The upper slope of the Cordillera Negra, immediately south of the
described for Cerro de Pasco. If the age constraint is reliable, it can be Santo Toribio deposit is faceted by four erosional surfaces, one of
inferred that no significant erosion affected the area since the Eocene. which is constrained by the overlying 14.10 ± 1.33–14.99 ± 0.50 Ma
(40Ar/39Ar hornblende plateau ages) lava flows of the unaltered Santo
10. The high-sulfidation epithermal Au (–Cu, Ag) deposits Toribio Formation andesite package. This constrained surface also
of northwestern Peru intersects the extensive area of phyllic alteration surrounding the
Santo Toribio vein system. This shows that hydrothermal activity (at
10.1. Pierina both Santo Toribio and Pierina, dated at ca. 15 to 14.4 Ma) was
penecontemporaneous with both andesite eruption and the incision of
Pierina is located in the Cordillera Negra, Ancash, west of the Cordil- the pediment. Volcanism terminated after 14 Ma in the area. The miner-
lera Blanca. The coeval, dominantly intermediate sulfidation, Santo alization at Pierina (and Lagunas Norte—see below), the cessation of
Toribio Ag-base metal vein systems occur ~ 5 km south of Pierina volcanism and onset of uplift around 14 Ma pre-dated the arrival of
(Rainbow, 2009), and both deposits are located beneath the shoulder the Nazca ridge at the subduction zone by at least 2 Ma (Hampel,
of an erosional surface overlooking the Callejón de Huaylas valley 2002). Uplift in response to crustal thickening in central and northern
(Figs. 10, 11). At Pierina, disseminated mineralization forms a sub- Peru has, instead, been linked to increased mid Miocene plate conver-
horizontal body, and is largely hosted in a lithologically-controlled gence and may not be directly related to initiation of flat subuduction
vuggy quartz alteration zone, focused in a ca. 16.9 ± 0.6 Ma (Montgomery, 2012; Pardo‐Casas and Molnar, 1987).
(40Ar/39Ar, weakly chlorite-altered biotite total gas age) pumice tuff
and an underlying dacitic flow dome complex (Rainbow, 2009); both 10.2. Lagunas Norte
members of the Oligocene to mid-Miocene Huaraz Group (Rainbow
et al., 2005; Strusievicz et al., 2000), the upper succession of the Calipuy Lagunas Norte, La Libertad, is located about 200 km NNW of Pierina
Supergroup subaerial volcanic arc. Hydrothermal breccias and small and 100 km SSE of Yanacocha in the northeastern mineral belt of north-
dacitic domes cut these rocks but appear to pre-date mineralization ern Peru. Mineralization is largely hosted in quartzites with scarce inter-
(Rainbow et al., 2005). Gold–silver mineralization was introduced spersed coal beds of the Lower Cretaceous Chimú Formation and, to a
after initial acidic alteration, and is associated with enargite, pyrite, lesser degree, in overlying dacitic pyroclastic and volcaniclastic rocks
bismuthinite–stibnite, galena and low-Fe sphalerite (Rainbow et al., of the Lagunas Norte Formation (Fig. 12) Montgomery, 2012; Cerpa
2005). However, sulfides have largely been oxidized. Sub-microscopic et al., 2013). Lagunas Norte Formation volcanic units are volumetrically
Au and Ag are now hosted in a goethite–hematite dominated oxide as- minor and are restricted in distribution to the immediate Lagunas Norte
semblage (Rainbow et al., 2005), the formation of which was facilitated deposit area (Fig. 12) Mineralization is centered on at least 2 diatremes
by microbial activity during supergene oxidation. During this process, which are also considered the source of the pyroclasitc rocks overlying
the local reduction of supergene fluids led to the formation of Au- the quartzites (Cerpa et al., 2013; Montgomery, 2012). Hydrothermal
bearing acanthite (Rainbow et al., 2006). Supergene minerals at Pierina alunite 40Ar/39Ar data constrain the age of mineralization at Lagunas
do not include jarosite. Norte to between 17.4 and 16.5 Ma (Cerpa et al., 2013; Montgomery,
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 347

Pierina Area, Peru


77°40'0"W 77°35'0"W 77°30'0"W 77°25'0"W

Pierina

Co
rd
ille

8940000
Santo Toribio
ra
Bl
9°30'0"S

an
ca
no
rm
al
fa
ul
t
Huaraz
Cordillera N

8930000
9°35'0"S

egra
9°40'0"S

8920000
0 2.5 5 7.5 10 km

-450000 -440000 -430000 -420000

Elevation (m) Slope (Degrees)


High : 6280
0 - 10

Low : 1565

Fig. 10. Topographic map flat 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 (González and Pfiffner,
2012; Petford and Atherton, 1992). UTM Zone 18S, WGS84.

2012), whereas a minimum age of 16.5 Ma and maximum age of 2012) were deposited east of Lagunas Norte. These rocks largely pre-
17.2 Ma for volcanic units of the Lagunas Norte Formation have been de- date hydrothermal activity at Lagunas Norte (Fig. 12) but mapping
termined (Montgomery, 2012). Extensive lower Miocene (21.1 Ma to together with geomorphology and geochronology suggests that the
16.4 Ma; Montgomery, 2012) andesitic-to-dacitic volcanic rocks of the Sauco volcanic complex is the result of sector collapse contemporane-
Sauco Volcanic Complex (age- and compositional-equivalent to Huaraz ous with mineralization at Lagunas Norte. At Lagunas Norte a reduction
Group strata of the Pierina district; Strusievicz et al., 2000; Montgomery, of erupted magma volumes and increased SiO2 content over time is
348 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Quebrada Pacchac
Quebrada Huellap (with heap leach facility)

Santo Toribio

Callejón 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) andesite hornblende Ar-Ar plateau age


Surface III (~4040 m) hydrothermal alteration sericite Ar-Ar plateau age
Surface II (~4260 m) sericite Ar-Ar total fusion age
Surface I (~4500 m) 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-sulfidation surface at Lagunas Norte occurred during, and was likely a response
epithermal activity, with restricted, dacitic eruptions overlapping tem- to, the initial stages of an episode of major, mid- to late Miocene
porally and spatially with mineralization (Montgomery, 2012). The vol- (ca. 17–5 Ma) regional uplift (e.g., Montario et al., 2005a, 2005b) and
canic rocks at Lagunas Norte exhibit alteration zonation typical of high- crustal thickening, during which, as much as 2–3 km of surface uplift
sulfidation epithermal deposits with a vuggy quartz core surrounded by may have occurred (Montgomery, 2012).
quartz–alunite and kaolinite–dickite alteration zones. The quartzites, in Hydrothermal activity and ore deposition at Lagunas Norte occurred
contrast, are non-reactive, and display only subtle alteration that in- during the latter stages of development of the Sauco Volcanic Complex,
cludes kaolinite, minor pyrophyllite and traces of alunite (Cerpa et al., centered 4 km east of the deposit, during waning volcanism, and imme-
2013; Montgomery, 2012). Mineralization in the quartzites is fracture- diately post-dated collapse of the volcanic edifice. Regional contractile
controlled, and hypogene ore minerals include digenite, enargite and tectonism and uplift may have ultimately triggered the collapse of this
pyrite, the latter hosting some of the gold. Locally, within coal beds, volcanic edifice, and may also have led to marked decrease in magmatic
low-sulfidation sulfide assemblages are present (Cerpa et al., 2013). activity which set the stage for development of the Lagunas Norte
The ore is oxidized to a depth of over 80 m. ore body (Montgomery, 2012). Inferred emplacement of a shallow
The orebody forms a tabular body between ~4250 to 4000 m a.s.l., intrusion subjacent to Lagunas Norte within this evolving tectono-
although sulfide mineralization extends to greater depths. It occurs magmatic setting, and its subsequent cooling, downward contraction
below relics of the extensive ca. 26–25 Ma subplanar erosional Pampa and associated fluid release, is envisaged to have resulted in early acid
La Julia surface (Montgomery, 2012) into which the ca. 18–16 Ma Río leaching and subsequent ore deposition at Lagunas Norte. Incision of
Chicama surface, consisting of steep-walled, but flat-bottomed valley- the Rio Chicama pediment at the margin of the Lagunas Norte deposit
pediment was cut. The Pampa La Julia surface may be correlative with at this time, would have favored decompression and fluid release, and
the upper Oligocene to lower Miocene angular unconformity recog- subsequent fluid mixing and/or boiling at the site of ore deposition.
nized 200 km to the SSE at Pierina in the Huaraz District (see above), at-
tributed to the regional Aymará orogenic event (Sébrier et al., 1988). 10.3. Yanacocha
The eastern Sauco Volcanic Complex in the Lagunas Norte district
erupted onto the older Pampa La Julia surface, whereas the Lagunas Gold mineralization occurs in about 10 individual centers within
Norte deposit is laterally contiguous with pronounced scarp between broad NE-trending zones of residual quartz, massive quartz, quartz–
this older surface and the younger Rio Chicama valley pediment alunite–pyrophyllite and distal kaolinite alteration (Longo et al., 2010;
(Figs. 13, 14). Active erosion and scarp retreat during hydrothermal ac- Teal and Benavides, 2010). Early pervasive alteration also includes
tivity are constrained in age by truncated a 17.05 Ma alunite veins of at quartz–pyrophyllite featuring the wormy or gusano texture characteris-
the western margin of the Lagunas Norte deposit and 16.75 Ma volcanic tic of the lower portion of the lithocap environment (Sillitoe, 2010;
rocks emplaced on the Rio Chicama surface immediately west of the de- Teal and Benavides, 2010). Gold was introduced in several pulses and
posit (Figs. 13, 14; Montgomery, 2012). Incision of the Rio Chicama is largely associated with permeable pyroclastic rocks, subvertical
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 349

78° 15' W

810000 E
795000 E

805000 E
9130000 N
N
5 km

Unmapped

Lagunas Norte
CSVC

9120000 N
Shulcahuanga dome

9115000 N
Quiruvilca mine
8° S

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


Undifferenated glacial deposits Andesite domes and lavas
Middle Miocene Cretaceous
Las Princesas volcanic domes. 16.3-15.2 Ma Undifferenated sedimentary rocks
Lagunas Norte Formaon, dacite to andesite, Structures
domes lavas and tuffs. 17.3-16.4 Ma
Sauco Volcanic Complex (21.1-17.1 Ma) Reverse fault (observed/inferred)
Dacite to andesite lavas and domes Normal fault (observed/inferred)
(CSVC - Central Sauco Vent complex) Fault (observed/inferred)
Andesite volcaniclasc rocks Syncline
Ancline
Early Andesite Dome/Lava

Fig. 12. Simplified 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 The rocks hosting the epithermal mineralization largely belong to
Benavides, 2010). The highest gold grades are associated with replacive the 14.5–11.2 Yanacocha Volcanics (Longo et al., 2010) which consist
cream-colored chalcedony veins containing barite and fine grained rutile of andesitic-to-dacitic lavas and pyoclastics and associated subvolcanic
and Fe oxides. These are interpreted as representing precipitation under rocks. As at Lagunas Norte, the volcanic pile records a reduction of
intermediate-sulfidation conditions (Teal and Benavides, 2010). Much of erupted magma volumes and increased SiO2 content over time, with
the ore mined to date came from the oxidized portion of the deposit over 90% of the volcanic rocks erupted prior to 11 Ma (Longo et al.,
where gold is associated with supergene Fe oxides. Sulfide assemblages 2010). Hydrothermal alunite ages (13.6 to 8.2 Ma) temporally overlap
associated with precursor hypogene epithermal mineralization include with the Yanacocha Volcanics, but less than 20% of the gold is associated
enargite and covellite and, at depth, chalcopyrite (Teal and Benavides, with hydrothermal activity older than 11 Ma. The bulk of the gold
2010). Late-stage veins of base metal sulfides associated with rhodochro- was introduced after 10.9 Ma and temporally overlaps with the
site have locally been reported from the Cerro Yanacocha deposit and volumetrically-minor Coriwachay dacite domes dated at 10.9 to
reflect higher pH and intermediate-sulfidation state fluids (Teal and 8.4 Ma (Longo et al., 2010). Thus, high-sulfidation epithermal gold min-
Benavides, 2010). Besides high-sulfidation and intermediate-sulfidation eralization clearly post-dated much of the volcanism and occurred in
epithermal mineralization, porphyry style Cu–Au ore has been described multiple pulses over a period of 2–3 Ma. Additionally, hydrothermal bi-
from Kupfertal, an area exposed at ~3800 m a.s.l., 300 m below the sum- otite associated with porphyry Cu–Au mineralization at Kupfertal
mit of Cerro Yanacocha (Gustafson et al., 2004; Teal and Benavides, yielded an 40Ar–39Ar age of 10.7 Ma. Although both porphyry and
2010). high-sulifdation epithermal style mineralization are present and
350 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Lagunas Norte Area, Peru


78°15’0”W

RCh surface
7°55’0”S

18-16 Ma

Sauco volcanic complex


PLJ surface
26-25 Ma

9120000
Lagunas Norte

Q Constructional
surface
PLJ surface
26-25 Ma
8°0’0”S

0 1 2 3 4 5 km

800000 810000

Elevation (m) Slope (Degrees)


High : 4254
0-8

Low : 2275

Fig. 13. Topographic map, flat landscape elements and main physiographic features of the Lagunas Norte area. PLJ: Pampa La Julia 26–25 Ma paleosurface, RCh: Río Chicama 18–16 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 magmatic- in the Lagunas Norte district (Montgomery, 2012) and could conceiv-
hydrothermal settings (cf. La Pepa, Muntean and Einaudi, 2001). ably be age-equivalent. The area hosting mineralization is incised by
The Yanacocha deposits underlie a 3900 to 4000 m a.s.l. subplanar flat bottomed valleys of ~ 3500 m elevation to the west and south
land surface (Fig. 15) evident in digital elevation models, with local to- which themselves have been incised up to 250 m by deep canyons. No
pographic highs, such as Cerro Yanacocha, reaching 4150 m. The high detailed documentation and direct age constraints on erosional surfaces
planation surface has a similar elevation to the Pampa la Julia surface are available. However, steam-heated alteration is locally preserved
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 351

A view looking southwest

Lower Miocene
Sauco volcanic
complex

Lagunas Norte
(early stages of open pit)

~ 16-18 Ma Río Chicama valley-pediment network


~ 25-26 Ma Pampa La Julia pediplain

B Alexa Zone Looking south


17.05 Ma alunite Shulcahuanga dome 16.95 Ma
Lagunas Norte orebody
Lagunas Norte Fm. 4200 m a.s.l. Tres Amigos dome 16.75 Ma

Pampa La Julia surface

Chimú Fm.
4000 m a.s.l. Tres Amigos
lava flows

Chimú Fm.
Río 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 zones). The deposit has been affected to a more significant degree
limited erosion has occurred since the late Miocene. Although glaciation by glaciation than Yanacocha, and it is thought that a significant
affected parts of the district above 3900 m a.s.l., glacial erosion was only part of the mineralization has been eroded (Gustafson et al., 2004).
locally significant and led to the formation of the detrital La Quinua gold Little detailed geological information has been published for La Zanja,
deposit hosted in glacial till (Mallette et al., 2004). operated by Compañia Minera Buenaventura and for Sipán. However,
Tantahuatay as well as Sipán and La Zanja lie beneath a high-elevation
10.4. Tantahuatay, Sipan and La Zanja planar landform similar to- and apparently contiguous with the one at
Yanacocha (Fig. 15). This surface is slightly inclined to the west and
Besides Yanacocha, northern Peru hosts a number of porphyry Sipan and La Zanja are exposed at about 400 m lower elevation than
Cu–Au and high-sulfidation epithermal deposits of late–middle to late Tantahuatay and Yanacocha, the latter being located about 30–40 km
Miocene age (Gustafson et al., 2004; Noble and McKee, 1999; Noble to the E of Sipan and La Zanja.
et al., 2004). Most of these occur within 50 km of Yanacocha (Fig. 15).
The high-sulfidation epithermal deposits include Tantahuatay, Sipan
and La Zanja, whereas Cu ± Au ± Mo porphyry style mineralization 11. The northern Andes
includes El Galeno and Cerro Corona (Gustafson et al., 2004). At
Tantahuatay, mineralization is hosted in an intensely vuggy to, locally, A major break in the basement architecture in southern Ecuador
massive quartz and quartz–pyrophyllite–alunite alteration zone, hosted separates the Northern Andes (3.5° Lat S. to 11° Lat N.) from the Central
in andesitic domes and underlying pyroclastic rocks (Gustafson et al., Andes (Cediel et al., 2003; Gansser, 1973) and is reflected in the differ-
2004). The andesites constitute the local topographic highs at ing metallogeny of the two regions. The western Cordillera of the north-
4050 m a.s.l. Gold is associated with pyrite and enargite and is concen- ern Andes of Ecuador and Colombia is comprised of terranes accreted
trated in areas of secondary permeability (vuggy quartz or fracture during the Cretaceous and Miocene, whereas the basement of the
352 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Yanacocha/Sipan/Tantahuatay/La Zanja Area, Peru


78°55'0"W 78°50'0"W 78°45'0"W 78°40'0"W 78°35'0"W

9260000
Tantahuatay
6°45'0"S

Cerro Corona P. Cu-Au

9250000
6°50'0"S

La Zanja

9240000
Sipan
6°55'0"S

9230000
Yanacocha District
7°0'0"S

9220000
0 1 2 4 6 8 10 km

730000 740000 750000 760000 770000

Elevation (m) Slope (Degrees)


High : 4203
0-8
Porphyry Cu-Au deposit

Low : 1041

Fig. 15. Topographic map, flat 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., 2008) and Buriticá (Lesage, 2011) low-sulfidation epithermal de-
et al., 2003; Ramos, 2009). In the northern Andes, in contrast to the posits. Large high-sulfidation epithermal deposits include Quimsacocha
Central Andes, high-sulfidation epithermal deposits are scarce, although in Ecuador and the California–Vetas district in the eastern Cordillera of
numerous other gold-rich deposits are known. The latter occur along Colombia. The northern Andes are now dominated by humid tropical
the eastern margin of the western Cordillera in northern Ecuador and climate and thus are wetter than the western cordillera of the Central
particularly throughout Colombia (Leal Mejía, 2011; Leal Mejía et al., Andes. Areas at high elevations record 1–2 m annual rainfall, but some
2011; Schuette, 2010) and include the late Miocene gold-only La Colosa of the adjacent low-lying areas record even higher precipitation rates
porphyry deposit (Lodder et al., 2010) as well as the Marmato (Tassinari of up to 10 m/yr (Alvarez Villa et al., 2011).
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 353

The uplift history and landscape evolution of the Northern Andes is (MacDonald et al., 2012). Steam-heated alteration zones of limited thick-
less well established than in the Central Andes and the geomorphology ness have locally been preserved in the highest parts of the deposit. The
is influenced by the terrane architecture. However, extensive planation ore-controlling Rio Falso fault is located along the eastern margin of a
surfaces have been documented in the Ecuadorian Andes (Coltorti and caldera in which the post mineral dacitic domes lie but is a district-
Ollier, 2000) and for the central Cordillera in Colombia (Restrepo- scale feature and does not constitute the caldera margin (Fig. 16).
Moreno et al., 2009). In Ecuador, relict planation surfaces are commonly Mineralization at Quimsacocha is located between 3500 and
exposed at 3500–3800 m a.s.l. The nature of these and the timing of up- 3700 m a.s.l. The paleosurface delimited by the steam-heated alteration
lift are controversial. Coltorti and Ollier (2000) interpret them have zone is located at 3800–3900 m a.s.l. The deposit is located near the
formed as planation surfaces near sea-level as recently as early Pliocene eastern margin of a large plateau at 3800 ± 100 m elevation with
and having been uplifted rapidly since the Pliocene. Conversely, steep margins relative to the adjacent valleys (Fig. 16; Coltorti and
Steinmann et al. (1999) documented the onset of major uplift in the Ollier, 2000). Based on sedimentological evidence and zircon fission-
early–late Miocene on the basis of stratigraphic relationships in track age data for tuff layers (Hungerbuehler et al., 2002; Steinmann
intramontane basins. A major tectonic change and increased uplift and et al., 1999), the area was at sea-level as recently as 15–11 Ma, with up-
contractile deformation in the Andes has been associated with the arriv- lift to almost 4000 m a.s.l. occurring in the Late Miocene, broadly coeval
al of the Carnegie Ridge at the subduction trench but age estimates for with mineralization. Quimsacocha also overlies the subducting Carnegie
this event vary widely between 15 Ma (Spikings et al., 2001) and ridge and no recent volcanism has been documented from the area
2 Ma (e.g., Gutscher et al., 1999b). The Colombian Andes, comprise (Chiaradia et al., 2004; Schuette et al., 2010).
three separate ranges, the Cordilleras Occidental, Central and Oriental,
each of which experienced differing geomorphologic histories. Planar 11.2. California Vetas
surfaces have been described from the Central Cordillera and the East-
ern Cordillera. A high plateau, the Antioqueño Plateau, is located at The California Vetas Mining District is located in the northeastern
the northern limit of the central Cordillera (6° ± 1° Lat. N). It consists Cordillera of Colombia, some 30 km N of the city of Bucaramanga. The
of an extensive upper Oligocene to lower Miocene planar erosion sur- modern climate in the area is tropical and has two pronounced rainy
face which was cut from the NE into an older, uplifted early Eocene seasons per year, with annual rainfall of around 2000 mm. The location
landscape (Restrepo-Moreno et al., 2009; Villagomez and Spikings, corresponds to the southern tip of the triangular Maracaibo block which
2013). The Antioqueño plateau is separated from higher landscape ele- is bounded by the NNW-striking sinistral Santa Marta–Bucaramanga
ments by a relict back-scarp and may have formed originally as a fault and the NE-striking dextral Boconó fault. Mineralization is hosted
pediplain. A renewed late Miocene to early Pliocene uplift pulse and as- by Grenvillia-aged Bucaramanga gneisses as well as upper Triassic
sociated incision of deep canyons is identified on the basis of apatite to lower Jurassic peraluminous granites (Mantilla Figueroa et al.,
U–Th/He data (Villagómez and Spikings, 2013). The magnitude of late 2013). Locally, at the El Cuatro prospect (Fig. 17), small volumes of
Miocene uplift and exhumation increases south of the Antioqueño Pla- metaluminous, coarsely-porphyritic granodiorite dykes of late Miocene
teau and contiguous central Cordillera (Villagómez and Spikings, age (10.9–8.4 Ma: Mantilla F. et al., 2009, 2011; Mantilla Figueroa et al.,
2013) but there the planar nature of the uplifted surface is obscured 2013; Bissig et al., 2014) are associated with porphyry Mo mineraliza-
by recent voluminous volcanic deposits and presumably by more tion (Bissig et al., 2012).
intense landscape dissection in response to uplift. In contrast to the Gold mineralization is mainly vein-hosted and is largely of high-
central Cordillera, the Eastern Cordillera was established in the late sulfidation epithermal type, overprinting earlier porphyry-style Mo
Oligocene when the tectonic regime changed from extensional to con- mineralization (Rodriguez, 2014). Numerous artisanal mines have been
tractile and the frontal thrust separating it from the Llanos Foreland operating since colonial times, but most current resources are contained
was established (Horton et al., 2010). However, based on paleobotanical in the La Bodega/La Mascota and contiguous Angostura deposits. Miner-
evidence, much of the uplift to the present day elevations of around alization was largely controlled by the NE-trending La Baja fault trend,
2600–3600 m a.s.l. probably took place in the late Miocene-to- Angostura focused in a dextral strike-slip sigmoidal loop (Fig. 17;
Pliocene (Gregory-Wodzicky, 2000). The northern extension of the Rodriguez, 2014). Several hydrothermal stages can be distinguished.
eastern Cordillera which hosts the California Vetas Mining District, Early quartz–pyrite ± chalcopyrite veins associated with pervasive
may have experienced uplift pulses in the Paleocene and early Miocene sericitic alteration represent precipitation in the porphyry environment
but, most importantly, in the late Miocene-to-Pleistocene (Villagómez and are associated with low-grade Au mineralization (b1 g/t). These
et al., 2011). are overprinted by several stages of veins and fault-controlled breccias
which introduced the bulk of the gold mineralization (Rodriguez,
11.1. Quimsacocha 2014). Ore minerals of the epithermal stages include covellite, bornite,
chalcopyrite, hubnerite, enargite and pyrite, all of which are associated
Quimsacocha is located in the northern Andean block (sensu Gansser, with native gold and gold–silver tellurides, as well as late Fe-poor sphal-
1973), about 20 km NW of the NE-trending Girón fault which constitutes erite. Epithermal veins and breccias have quartz and alunite gangue that
an important terrane boundary separating the Upper Cretaceous Piñón form banded, colloform and cockade textures (Rodriguez, 2014), such
terrane to the northwest, from the Lower Creteaceous Anaime terrane as are more typically associated with low-sulfidation type deposits
to the southeast (Ramos, 2009; Schuette et al., 2012). The Quimsacocha (Simmons et al., 2005). Conversely, vuggy residual quartz and aerially-
deposit adheres closely to the model for high-sulfidation epithermal de- extensive, pervasive, quartz–alunite–kaolinite alteration typical of
posits (MacDonald et al., 2011). Mineralization is hosted within a large high-sulfidation deposits is largely absent in the California–Vetas district.
vuggy quartz, massive quartz and quartz–alunite ± pyrophyllite alter- Both fluid inclusion and textural evidence indicates that boiling was an
ation zone which is centered on the N-striking Río Falso fault zone. The important ore depositional mechanism, occurring at 200–250° C
ore is confined in pyroclastic strata between subhorizontal andesite (Rodriguez, 2014). Early muscovite alteration associated with quartz–
lava flows of the 9 Ma Quimsacocha Formation (all age dates are pyrite veins at Angostura and La Bodega gives 40Ar/39Ar ages of ca. 4 to
40
Ar/39Ar plateau ages: MacDonald et al., 2011, 2012). The host rocks 3.5 Ma and one alunite sample from Los Laches, at ~ 3500 m a.s.l.,
underwent intense acid leaching resulting in vuggy quartz alteration, some 800 m above the muscovite sample locations, was dated at
prior to the deposition of pyrite, enargite and associated gold and 4.02 ± 0.06 Ma and likely was deposited close to the paleosurface
minor quartz (MacDonald et al., 2011). Alunite associated with the (Fig. 17, Rodriguez, 2014). The age of mineralization inferred from
vuggy quartz and advanced argillic alteration is dated at 7.6 to 7.3 Ma alunite in mineralized veins and breccias ranges from 2.6 to 2 Ma at La
and biotite from a post-mineral dacite dome yielded an age of 6.7 Ma Bodega and Angostura, but alunite the La Plata, San Celestino and El
354 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Quimsacocha Area, Ecuador


79°15’0”W 79°10’0”W

9670000
3°0’0”S

Río Falso fault zone


Quimsacocha
caldera
Quimsacocha Deposit

9660000
3°5’0”S

0 1 2 3 4 5 km

690000 700000

Elevation (m) Slope (Degrees)


High : 4079
0-8

Low : 2925

Fig. 16. Topographic map of the Quimsaochca area. The approximate trace and orientation of the Río Falso fault is indicated.

Cuatro prospects downstream from La Bodega (Fig. 17) yielded ages of The geomorphology in the La Bodega and Angostura area is charac-
3.5 to 3.2 Ma. Post-mineralization alunite associated with sphalerite terized by a deeply-incised, steep-walled valley paralleling the NE-
yielded ages from 1.9 to 1.6 with Ma (Rodriguez, 2014). Paragenetically trending dextral La Baja fault zone (Figs. 18, 19). Angostura is located
late U mineralization has been reported from San Celestino (Polania, at the upper termination of this valley, La Bodega/La Mascota at 2800
1980) 2 km SW of La Bodega, along the same mineralized trend. No to 2350 m, and Angostura between 3500 and 2700 m a.s.l. More
igneous rocks similar in age to the epithermal mineralization have subdued topography characterizes the topographically-high areas
been documented, but stable-isotope data indicate a magmatic source from ~ 3400 to 4000 m a.s.l. (Figs. 18, 19) representing the highest
for the mineralizing fluid (Rodriguez, 2014). parts of the northeastern cordillera of Colombia. This suggests that,
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 355

72°56'0"W 72°54'0"W 72°52'0"W

Fault
lt
ult

fau
fa

la
ra

a
t
tu Veta de Barro

Rio Cucutil
aul

illt
s

cu
go
ja f
ura

Cu
ost

An
Ba
Ang

e ral-
Los Laches 4.02 +/- 0.06 Ma

La
3.4 +/- 0.06 Ma La Picota 2.77 +/- 0.15 Ma

m
o R
La Bodega
3.91 +/- 0.15 Ma

La Mascota PaLa
ez f
Perezosa 2.1- 1.8 Ma (n=2)
ault

2.5 - 1.6 Ma (n=9)


El Cuatro
Pie de Gallo 3.26+/- 0.30 Ma
7°22'0"N

San Celestino
3.23+/- 0.06 Ma

La Plata

1305000
3.43 +/- 0.07 Ma

Violetal Geologic Units


Faults & Contacts
Pliocene
California Hydrothermal breccia/veins Normal fault

Late Miocene Normal fault, inferred

Hydrothermal breccia Thrust fault

Strike-slip, dextral
7°20'0"N

Porphyry Non conformity

Late Cretaceous
Rosablanca Formation Town
California Vetas
Mining District Prospect/Artisanal Mine
Bolivar Tambor Formation

Bucaramanga Late Triassic to Early Jurassic 40Ar/39Ar ages


Diorite to Granodiorite
Antioquia Alunite
Santander
Arauca Leucogranite Sericite
Proterozoic

1300000
Casanare
Boyaca Bucaramanga Gneiss

1125000 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. Modified 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 ero- relief surfaces, in many cases demonstrably pediplains, located at the
sional equilibrium with the present day base-level, as in northern crest of the Andes at the time of mineralization. The low-relief surfaces
Chile and Peru. This high elevation surface is interpreted as a relict were commonly subject to erosion and valley pediments or valleys were
lower to middle Miocene paleosurface. being incised during mineralization. A direct temporal and spatial
Based on apatite fission track data, uplift and valley incision com- relationship of erosion and mineralization has been documented for a
menced around 17 Ma (Van Der Lelij, 2013) but the most important up- number of epithermal districts including the El Indio belt deposits,
lift pulse probably occurred in the Pliocene (Gregory-Wodzicky, 2000; Pierina, Lagunas Norte and California–Vetas.
Shagam et al., 1984; Villagomez et al., 2011). Thus, uplift, erosion and Most high-sulfidation epithermal deposits were emplaced above
epithermal mineralization overlapping in age and the fact that the alu- segments of flat subduction where no volcanism is presently observed.
nite ages along the La Baja trend become younger upstream strongly The few exceptions to the latter occur in the western Cordillera near the
suggests that erosion is directly stimulating epithermal mineralization. Andean orocline in southern Peru and northern Chile. The vast majority
of deposits are 17 Ma or younger, but exceptions include the Eocene El
12. Summary and comparison to low-sulfidation deposits Hueso, and Guanaco deposits in the southern Atacama Desert near the
Domeyko fault system, and possibly the poorly documented Quicay
High-sulfidation epithermal deposits of the Andes occur over a geo- deposit as well as the late Oligocene to early Miocene La Coipa deposit
graphically wide area, yet were emplaced in remarkably predictable for which age constraints are ambiguous. Large volcanic edifices are un-
geologic and geomorphologic settings. Virtually all high-sulfidation common hosts to ore. Host rocks to high-sulfidation systems commonly
epithermal deposits formed during episodes of major uplift in the re- pre-date mineralization by several Ma to as much as 1 Ga, as at
spective Andean segment in which they are hosted. The uplifted land- La Bodega and Angostura in Colombia. However, examples exist
forms hosting epithermal mineralization are mostly extensive low where the ages of ore-hosting volcanic rocks overlap with those of
356 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

California Vetas Mining District, Colombia


72°55’0”W 72°50’0”W
7°25’0”N

Angostura

La Bodega/La Mascota
El Cuatro
San Celestino

La Plata

California
7°20’0”N

Vetas

0 1 2 3 4 5 km

Elevation (m) Slope (Degrees)


High : 4272
High elevation low relief paleosurface
0 - 16

Town
Low : 1358 Prospect

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

alteration and mineralization (e.g., Yanacocha, Lagunas Norte, Aruntani) Andean low-sulfidation epithermal deposits are commonly smaller
but where this temporal overlap exists, the volcanic rocks contempora- and many occur in geologic settings distinct from those of their high-
neous with mineralization are, with exception of Aruntani, of low sulfidation counterparts. Some of the larger examples include the Juras-
volume and volcanism ceased shortly after mineralization. Volcanic sic Fruta del Norte deposit in Ecuador (Henderson, 2009), the Paleocene
rocks of significant volume covering and/or post-dating mineralization El Peñón deposit of northern Chile (Warren et al., 2004, 2007). The Late
are reported from La Pepa and La Coipa. Jurassic to Early Cretaceous deposits of the Deseado and Patagonian
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 357

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 sec-
tor collapse eliminated more than 500 m of rock cover very rapidly, and
is thought to be integral to epithermal ore formation (Blackwell et al.,
California 2014; Carman, 2003; Sillitoe, 1994). While Ladolam has a low-
sulfidation sulfide assemblage and is associated with alkalic magmatism
San Celestino and has, thus, fluid chemistry differing from that of Andean high-
sulfidation epithermal systems, the gold transport and precipitation
El Cuatro processes are still comparable, Au bisulfide 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
La Mascota Au deposits in the Maricunga belt (Muntean and Einaudi, 2001;
La Bodega 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 re-
sponse to uplift events can enhance mineralizing processes (Bissig et al.,
2002a) and may exert a first-order control on exsolution of fluids from
Fig. 19. Photograph looking SW from Angostura, showing the approximate locations of
magmas. This hypothesis is supported by the fact that Eocene and Mio-
prospects. Note the steeply incised La Baja valley and the less rugged higher elevation
terrain. cene porphyry and high-sulfidation epithermal deposits in the Central
Andes were emplaced during contractile deformation, uplift and large-
scale erosion on the western Andean slope, which occurred during the
Massifs (Dietrich et al., 2011; Echavarria et al., 2005; Fernández et al., Eocene Incaic, late Oligocene–early Miocene Aymará and Miocene Que-
2008) are, strictly speaking, located outside the Andes but the Cerro chua orogenic phases. On a more local scale, valley or pediment incision
Bayo low-sulfidation deposit considered the westernmost deposit of during hydrothermal activity can lead to lowering of the water-table
this low sulfidation epithermal province is located in the Chilean near the head of the incising valleys and, by generating local topogra-
Patagonian Andes (Poblete et al., 2014). Mineralization here took phy, may lead to increased lateral groundwater flow and enhanced
place in several pulses between 144 and 113 Ma and partly overlaps mixing of magmatic with meteoric fluids. This process can enhance boil-
in age with the slightly older low-sulfidation epithermal deposits of ing and fluid mixing especially near the shoulders or back-scarps of in-
the Deseado and north Patagonian massifs (Poblete et al., 2014). cising pediments (e.g. El Indio belt: Bissig et al., 2002a, Lagunas Norte:
The low-sulfidation deposits mentioned above occur at elevations Montgomery, 2012) or fault induced steep topographic gradients (po-
below 2400 m a.s.l., are hosted by dominantly rhyolitic volcanic and tentially the case in the Famatina district: Losada-Calderón et al.,
volcaniclastic rocks, lack a spatial or temporal association with porphyry 1994; Pudack et al., 2009). The time scale of valley incision is well within
systems (as defined by Sillitoe, 2010) and occur in extensional tectonic the proposed duration of hydrothermal activity of individual hydrother-
settings without documented contractile deformation and surface uplift mal systems, which is in the order of 10 to 100 k.y. (e.g., Simmons, 2002;
during mineralization. However, there are also examples of epithermal Simmons and Brown, 2006). Erosion can in part influence the locus of
deposits containing low-sulfidation mineralization that occur within mineralization over time. The example of Veladero and Pascua as well
larger magmatic-hydrothermal systems in arc segments also containing as the deposits along the La Baja trend in the California Vetas district il-
porphyry style or high-sulfidation epithermal mineralization. Examples lustrate this relationship (Figs. 20, 21). In both cases, older mineraliza-
of these include Cerro de Pasco (Peru: Baumgartner et al., 2008), tion occurs downstream from younger mineralization. Perhaps the
Marmato and Buriticá (Colombia: Tassinari et al., 2008; Lesage, 2011) best temporal control is available for the La Baja trend where at
which are all mid to late Miocene in age. The low-sulfidation nature of 3.5 Ma, alunite and associated Au mineralization formed to the west
epithermal deposits of some contractile arc settings may be attributed and downstream of contemporaneous higher temperature phyllic alter-
to more reactive, locally-reducing host or basement rock characteristics ation, which formed under 500 m or more of cover rock (Rodriguez,
(e.g., Marmato: Tassinari et al., 2008). Thus, as proposed earlier (Sillitoe 2014; Fig. 20). At a later stage, the La Baja valley incised northeastward
and Hedenquist, 2003) low-sulfidation deposits may be associated with and subsequent epithermal hydrothermal alteration and mineralization
both, overall contractile arc as well as rift settings. Those associated with overprinted the phyllic alteration at La Bodega and Angostura (Fig. 20).
rift settings and rhyolitic magmatism evidently are more likely to be In the case of Veladero and Pascua, no constraints for phyllic alteration
preserved over longer time intervals than epithermal deposits in overprinted by high-sulfidation epithermal mineralization are available.
contractile settings. However, mineralization at the Filo Federico zone at Veladero is at
11–10.3 Ma distinctly older than 9.5 Ma mineralization at the Penelope
13. Controls of geomorphic processes and climate on mineralization ore zone and 9.1–8.1 Ma mineralization at Pascua which occur upstream
from Filo Federico (Fig. 21). At the time of mineralization at Pascua,
Epithermal deposits are emplaced at shallow crustal levels, in the Veladero was already subject to oxidation, as suggested by the jarosite
case of high-sulfidation epithermal deposits at depths commonly less ages of Veladero, roughly contemporaneous with hypogene alunite for-
than 500 m (see above). Such shallow hydrothermal environments mation at Pascua (Deyell et al., 2005b; Holley, 2012).
are directly influenced by surface processes and climatic conditions. Andean high-sulfidation epithermal deposits occur over a wide
Precious metals in the epithermal environment are typically precipitat- range of climatic zones from humid tropical to hyper-arid climate. The
ed by processes of boiling or fluid mixing (Simmons et al., 2005). The Miocene climate of the Central Andes was probably slightly more
depth of boiling in the epithermal environment is controlled by the humid than at present but probably not fundamentally different as
hydrostatic pressure which depends on the elevation of the water South America was located at similar latitudes as at present (Clarke,
table. A lowering of the water table during hydrothermal activity such 2006). Thus, the availability of meteoric fluids during hydrothermal ac-
as would occur during catastrophic erosion or volcanic sector collapse tivity does not seem to be the principal controlling factor for mineraliza-
would enhance boiling and can be important for the efficiency of pre- tion. This is in agreement with the growing number of stable-isotope
cious metal deposition (Bissig et al., 2002a; Simmons, 1991). This can studies, all of which propose a dominantly magmatic source for the
358 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

La Bodega Angostura
10 Ma Paleosurface
NE
A

Los Laches
SW

La Bodega

Perezosa
San Celestino

La Mascota
California

El Cuatro
La Plata
3,500

ce
rfa
su
ay
Elevation (m)

td
3,000

en
es
Pr
2,500

Mo Mo
Cu?
2,000 Cu?
0 500 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)
B
3,500
4 - 3.25 Ma 4 Ma
e
f ac ce
ur rfa
al
e os
os
u Au
p le
pa
Elevation (m)

3,000 a
M a
5 M
3. 2 5
~ 3.
~ Au
3.9 Ma Cu
2,500 3.25 Ma Au
3.25 Ma 3.4 Ma
Cu
3.5 Ma Au Au ?
Au Ag Ag
2,000 Ag
0 500 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
C Distance (m)

3,500 ~ 2.5 - 2 Ma 2.7 Ma


Advanced argillic
Sericite/illite quartz
Potassic, K-feldspar, biotite
3,000
Elevation (m)

Porphyries
Ar/Ar on alunite age Alunite < 2 Ma 2.1 Ma
2.2 Ma Au
Ar/Ar on sericite
2.6-2.3 Ma 1.8 Ma Ag
Re/Os molybdenite
2,500 U/Pb zircon (Cu)
Au
U Au 1.6 Ma
Ag
Ag 1.9 Ma (Cu)
2,000
Zn
0 500 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)

Fig. 20. Schematic relationships between erosion, tectonics and mineralization along the La Baja trend, California–Vetas Mining District, Colombia. Based on data from Rodriguez (2014);
Refer to Figs. 17 and 18 for context. A) snap-shot of the landscape configuration in relation to hydrothermal systems at ~10 Ma, at the time of porphyry intrusion and Mo (–Cu) miner-
alization. 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
700–1000 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 2.5–2 Ma, during the main stage of epithermal Au (–Ag, Cu) mineralization at La Bodega,
La Mascota and Angostura, erosion is further advanced and overprints 3.5–3.9 Ma sericite. Note, Los Laches has not been affected by significant 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) 327–364 359

A NW SE
Pascua
9.1-8 Ma Penelope Río de las Taguas
9.5 Ma N-S valley
5000
4750 Lama Central Filo Federico
barren alteration 11.1-10.3 Ma
4500
13.3 Ma
4250
4000

Frontera-Deidad surface Steam-heated alteration Present day Río Turbio profile


Azufreras-Torta surface
16-~12.5 Ma volcaniclastic rocks
Los Ríos surface
Mineralized zones Future mineralized zones
Jarosite formation

B
5000
4750
4500
4250
4000 Valley profile at 9.5 Ma

Present day Río Turbio profile

C
5000
4750
4500
4250
4000 Valley profile at 10.5 Ma

Present day Río Turbio profile

Fig. 21. Schematic relationship between erosion, tectonics and mineralization at Veladero and Pascua–Lama, Argentina, Chile. The cartoon is based on data from Bissig et al. (2002a),
Charchaflié et al. (2007) and Holley (2012). A) Present-day configuration of landscape elements, faults, mineralized zones. B) Snap-shot of the landscape configuration 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 configuration at
10.5 Ma, during hypogene mineralization at Filo Federico.

precious metal-bearing fluids (e.g., Cerpa et al., 2013; Deyell et al., 2004; Most high-sulfidation epithermal deposits were emplaced between
Rainbow et al., 2005; Rye, 2005). However, the hydrothermal systems 17 and 6 Ma, irrespective of the climatic zone in which they are located.
overall are influenced by the climate. Thus, the position of the water There is no clearcut overall correlation of younger age with better pres-
table is influenced by the climate and the extent of near surface ervation within this age range. The location and age of currently ex-
steam-heated alteration formed in the vadose zone varies. Up to several posed Miocene high-sulfidation epithermal deposits is, thus, not only
100 m thick steam-heated blankets overlying and overprinting mineral- a function of the preservation potential but also due to tectono-
ization are observed in the El Indio belt and La Coipa. These are magmatic factors. However, the Eocene deposits of El Guanaco and El
interpreted as evidence for dry climate at the time of mineralization Hueso are located in one of the most arid climatic zone of the planet,
and general water table lowering during hydrothermal activity whereas the youngest deposits La Bodega and Angostura in Colombia
(e.g., Bissig et al., 2002a; Holley, 2012), but short term water table fluc- are located in the wettest climatic zone in which high-sulfidation
tuations due to episodic incursion of magmatic fluid or periods of more epithermal deposits are known in the Andes. This indicates that a
humid climate has been inferred from hypogene alunite or barite longer-term control on preservation potential does exist. The fact that
overprinting jarosite (Chouinard et al., 2005; Deyell et al., 2005b; no high-sulfidation epithermal deposits older than Eocene are known,
Holley, 2012). In contrast to deposits of northern Chile, in the more together with the fact that they generally were emplaced near surface
humid Miocene climates of Peru, Ecuador or Colombia where the during uplift indicates that the preservation of high-sulfidation
water-table was probably much closer to the land-surface, steam- epithermal deposits older than the Eocene is unlikely except in the dri-
heated alteration blankets tend to be less well developed or even lack- est climatic zones. Oxidation of primary sulfide assemblages is an im-
ing (McDonald et al., 2011; Rainbow et al., 2005; Rodriguez, 2014; portant factor for the economic viability of high-sulfidation epithermal
Teal and Benavides, 2010). Although this difference may be due to deposits, as it may liberate encapsulated gold. Oxidation occurs where
better preservation under dry climatic conditions, the persistence of sulfides are exposed to atmospheric oxygen and water and is favored
high-elevation paleosurfaces in northern Peru and Ecuador that pre- by decreasing water table as well as sufficient rock permeability. Some
date hydrothermal activity suggests that steam-heated blankets should examples of high-sulfidation epithermal deposits of Northern Peru as
have been forming but were probably never as extensive in those areas well as the Pampean flat slab are oxidized to considerable depth
as in dryer climatic zones, and could therefore have been removed by (e.g., Yanacocha, Lagunas Norte, Pierina, Veladero), whereas others
modest erosion. (e.g., Pascua–Lama, Cerro de Pasco) have undergone considerably less
360 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

extensive supergene oxidation. In the case of Pascua–Lama and Although low-density magmatic vapor may be responsible for early
Veladero as well as the La Coipa district, oxidized and non-oxidized vuggy quartz alteration, later high-sulfidation epithermal mineraliza-
deposits occur within b 15 km from each other, which indicates that tion hosted in vuggy quartz likely precipitated from contracted mag-
the degree of oxidation cannot simply ascribed to climatic factors. matic vapor exsolved from a deeper porphyry intrusion (Heinrich
Rock-permeability or physiographic setting as well as the post- et al., 2004; Pudack et al., 2009). This evolution from shallow toward
hydrothermal history influence the oxidation and hence economic deeper locus of fluid exsolution is consistent with waning magmatism
viability of epithermal deposits as well. and increasing SiO2 content of magmas; as is evident for the majority
of Andean high-sulfidation epithermal districts. Large-scale tectonic
14. Igneous rocks, volcanology and magmatic fluids related changes such as the onset of flat subduction, increased plate coupling,
to high-sulfidation epithermal deposits contractile deformation and uplift result in decreasing volumes of
eruptive magmatism and increasing depth of porphyry emplacement
Although magmas are the primary contributors of metals and vola- and may favor high-sulfidation epithermal Au and associated deeper
tiles in porphyry systems (e.g., Audétat and Simon, 2012), extrusive porphyry Cu over porphyry Cu–Au mineralization.
magmatism is commonly waning, of low volume or absent during the
formation of high-sulfidation epithermal Au–Ag deposits in the Andes.
Conversely, Au-rich porphyry deposits in the Maricunga belt and 15. Conclusions
in northern Peru are typically emplaced in volcanic centers or associated
with shallowly emplaced magmatic bodies (e.g. Cerro Casale: Muntean High-sulfidation epithermal deposits of the Andes formed in pre-
and Einaudi, 2001). Murakami et al. (2010) showed that Cu/Au ratios in- dictable geological and geomorphological settings. They are located at
crease with greater depth of emplacement of porphyry systems and high elevation and largely near the crest of the Andean cordillera in seg-
suggested that high-sulfidation epithermal Au mineralization occurs ments where volcanism is currently absent or subdued and where sub-
where significant amounts of Au has been physically separated from duction angles are shallow. All were emplaced during major periods of
Cu, a process which is apparently favored where the depth of intrusion contractile deformation and uplift and most coincided with the terminal
of the magma providing the magmatic volatiles is N 3 km (Murakami stages of local arc magmatism. The vast majority of deposits are be-
et al., 2010). A compelling explanation for this is that precious metals tween ca. 17 and 5 Ma old. Although the tectonic setting in which
are better transported in high-density magmatic vapors released from these deposits form makes them prone to erosion, their restricted age
magmas emplaced at more than 3–4 km depth than in low-density va- range is not only a function of preservation potential but also attributed
pors exsolved at depths of less than 2 km (Heinrich, 2005, 2007; Fig. 22). to favorable tectonomagmatic settings at the time of hydrothermal ac-
Fluid evolution at depth can lead to porphyry Cu mineralization but tivity. Eocene high-sulfidation epithermal deposits are only known
allows for efficient Au transport in contracted magmatic vapor. If from the driest regions of the Atacama Desert but they plausibly formed
condensed magmatic vapors are efficiently transported and focused, a under similar tectonomagmatic conditions as the Miocene ones. Con-
high-sulfidation epithermal Au deposit may form several km above. versely, the youngest deposits, La Bodega and Angostura, Colombia,

A B C vvvv

vvvv v v Au vvvv
Au v v v
v v
v Au
1 km v v v v v
v v
Au Cu

2 km

3 km
Cu
Cu
Mo Mo

4 km 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 Au–Cu hosted in a volcanic edifice such as observed in the Maricunga belt (e.g. Cerro Casale). Depth of intrusion and fluid exsolution is b2 km, magmatic
vapor cannot dissolve significant Au and temperature gradient to surrounding rocks is steep, leading to bulk co-precipitation of Cu and Au. B) High-sulfidation epithermal deposit forms
N3–4 km above porphyry intrusion. At that depth, magmatic vapor has a higher density and is, after contraction, capable of transporting significant Au as bisulfide complexes at low fluid
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 fluid release from the magma and the structural pathways permitting efficient separation of magmatic vapor derived fluids from the
porphyry Cu environment (this scenario reflects the situation at La Bodega and Angostura, Colombia as well as El Indio and Pascua–Lama). C) A scenario where multiple pulses of
both, porphyry Cu–Au and high-sulfidation epithermal mineralization occurred. This scenario reflects the situation at Yanacocha. Volcanism in the form of flow-domes and pyroclastic
flow deposits occurred intermittedly as well.
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 361

were emplaced in the late Pliocene in an area with pluvial tropical degassing during formation of epithermal Au–Ag and Cordilleran polymetallic ores.
Mineral. Deposita 43, 777–789.
climate with low preservation potential. Beuchat, S., Moritz, R., Pettke, T., 2004. Fluid evolution in the W–Cu–Zn–Pb San Cristobal
High-sulfidation epithermal Au–Ag deposits can be emplaced in a vein, Peru: fluid inclusion and stable isotope evidence. Chem. Geol. 210, 201–224.
range of host rocks, and volcanic rocks both pre- or syn-hydrothermal Bissig, T., Riquelme, R., 2009. Contrasting landscape evolution and development of super-
gene enrichment in the El Salvador porphyry Cu and Potrerillos–El Hueso Cu–Au dis-
activities, are not a pre-requisite for epithermal mineralization. Howev- tricts, Northern Chile. In: Titley, S. (Ed.), Society of Economic Geologists Special
er, a magma from which the mineralizing magmatic fluids are derived Publication No. 14: Supergene Environments. Processes and Products, pp. 59–68.
must have present, albeit fluid exsolution typically occurred N3–4 km Bissig, T., Riquelme, R., 2010. Andean uplift and climate evolution in the southern
Atacama Desert deduced from geomorphology and supergene alunite-group
below the surface at the time of hydrothermal activity. minerals. Earth Planet. Sci. Lett. 299, 447–457.
Significant erosion during hydrothermal activity is documented for a Bissig, T., Tosdal, R.M., 2009. Petrogenetic and Metallogenetic Relationships in the Eastern
number of districts (e.g., Lagunas Norte, Pierina, El Indio belt, California Cordillera Occidental of Central Peru. J. Geol. 117, 499–518.
Bissig, T., Lee, J.K.W., Clark, A.H., Heather, K.B., 2001. The cenozoic history of volcanism
Vetas Mining District) and mineralization ages generally become youn-
and hydrothermal alteration in the central Andean flat-slab region: New 40Ar–39Ar
ger upstream in both, the El Indio belt and the California Vetas Mining constraints from the El Indio–Pascua Au (–Ag, Cu) belt, 29° 20 ′–30° 30 ′ S. Int.
District. Fluid boiling and mixing with meteoric water are enhanced Geol. Rev. 43, 312–340.
near topographic breaks such as the heads of incising valleys or back- 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 Indio–Pascua Au–Ag–Cu
scarps of pediments, ultimately stimulating mineralization in those belt, Chile and Argentina. Econ. Geol. Bull. Soc. 97, 971–996.
locations. Bissig, T., Clark, A.H., Lee, J.K., 2002b. Cerro de Vidrio rhyolitic dome: evidence for Late
Pliocene volcanism in the central Andean flat-slab region, Lama–Veladero district,
29 20′ S, San Juan Province, Argentina. J. S. Am. Earth Sci. 15, 571–576.
Bissig, T., Clark, A.H., Lee, J.K.W., von Quadt, A., 2003. Petrogenetic and metallogenetic
Acknowledgments responses to Miocene slab flattening: new constraints from the El Indio–Pascua
Au–Ag–Cu belt, Chile/Argentina. Mineral. Deposita 38, 844–862.
This paper reviews and summarizes many years of work on Andean Bissig, T., Ullrich, T.D., Tosdal, R.M., Friedman, R., Ebert, S., 2008. The time–space distribu-
tion of Eocene to Miocene magmatism in the central Peruvian polymetallic province
epithermal systems, much of it carried out at Queen's University. The
and its metallogenetic implications. J. S. Am. Earth Sci. 26, 16–35.
authors would like to express their special thanks to Barrick Gold Bissig, T., Mantilla, F.L.C., Rodriguez, M.A., Raley, C.A., Hart, C.J.R., 2012. The California–
Corp. and former chief geologist Jay Hodgson for the continued support Vetas District, Eastern Cordillera, Santander, Colombia: Late Miocene Porphyry and
Epithermal Mineralization Hosted in Proterozoic Gneisses and Late Triassic–Early
and funding of TB's, AM's and AR's PhD theses at Queen's University.
Jurassic Intrusions, XVI Congreso Peruano de Geología & SEG2010 Conference, Lima,
Barrick also funded projects at MDRU and Universidad Católica del Peru, (Poster 11).
Norte, Chile in which the senior author continues to be involved. With- Bissig, T., Mantilla Figueroa, L.C., Hart, C.J., 2014. Petrochemistry of igneous rocks of
out Barrick's support, our understanding of Andean high-sulfidation the California–Vetas mining district, Santander, Colombia: Implications for
northern Andean tectonics and porphyry Cu (–Mo, Au) metallogeny. Lithos
epithermal deposits would be far less advanced. 200, 355–367.
We also acknowledge the contributions and student support from Blackwell, J.L., Cooke, D.R., McPhie, J., Simpson, K.A., 2014. Lithofacies Associations and
NSERC in the form of grants to AHC and to Kurt Kyser, as well as from Evolution of the Volcanic Host Succession to the Minifie Ore Zone: Ladolam Gold
Deposit, Lihir Island, Papua New Guinea. Econ. Geol. 109, 1137–1160.
many other companies who supported research at Queen's University Bouzari, F., Clark, A.H., 2002. Anatomy, evolution, and metallogenic significance of the su-
and MDRU, including (but not limited to) IamGold, Ventana Gold, Eco pergene orebody of the Cerro Colorado porphyry copper deposit, I region, northern
Oro minerals and Kinross. Chile. Econ. Geol. 97, 1701–1740.
Bussell, M.A., Alpers, C.N., Petersen, U., Shepherd, T.J., Bermudez, C., Baxter, A.N., 1990. The
Sara Jenkins is acknowledged for help with the figures and we thank Ag–Mn–Pb–Zn vein, replacement, and skarn deposits of Uchucchacua, Peru; studies
the OGR editors Franco Piranjo and Tim Horscroft for the invitation of of structure, mineralogy, metal zoning, Sr isotopes, and fluid inclusions. Econ. Geol.
this review article. Comments provided by reviewers Stuart Simmons 85, 1348–1383.
Cahill, T., Isacks, B.L., 1992. Seismicity and shape of the subducted Nazca plate. J. Geophys.
and Dick Tosdal helped improve clarity of the paper.
Res. 97, 17503–17517 (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
References isotope paleoaltimetry of hydrated volcanic glass. Geology 42, 447–450.
Carman, G.D., 2003. Geology, mineralization, and hydrothermal evolution of the Ladolam
Aguilar, G., Riquelme, R., Martinod, J., Darrozes, J., Maire, E., 2011. Variability in erosion gold deposit, Lihir Island, Papua New Guinea. In: Simmons, S.F., Graham, I. (Eds.),
rates related to the state of landscape transience in the semi-arid Chilean Andes. Volcanic. Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of Processes
Earth Surf. Process. Landf. 36, 1736–1748. Within the Earth, Special Publication-Society of Economic Geologists 10, pp. 247–284.
Aguilar, G., Riquelme, R., Martinod, J., Darrozes, J., 2013. Role of climate and tectonics in Catchpole, H., Kouzmanov, K., Fontbote, L., Guillong, M., Heinrich, C.A., 2011. Fluid evolu-
the geomorphologic evolution of the Semiarid Chilean Andes between 27–32° S. tion in zoned cordilleran polymetallic veins; insights from microthermometry and
Andean Geol. 40, 79–101. LA–ICP-MS of fluid inclusions. Chem. Geol. 281, 293–304.
Álvarez‐Villa, O.D., Vélez, J.I., Poveda, G., 2011. Improved long‐term mean annual rainfall Cecioni, A.J., Dick, L.A., 1992. Geología del yacimiento epitermal de oro y plata Can Can,
fields for Colombia. Int. J. Climatol. 31, 2194–2212. Franja de Maricunga, precordillera de Copiapo, Chile. Andean Geol. 19, 3–17.
Arribas, A., 1995. Characteristics of high-sulfidation epithermal deposits, and their Cediel, F., Shaw, R.P., Cáceres, C., Bartolini, C., 2003. Tectonic assembly of the Northern
relation to magmatic fluid. Mineral. Assoc. Can. Short Course Ser. 23, 419–454. Andean block. The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats.
Audétat, A., Simon, A.C., 2012. Magmatic controls on porphyry copper deposits. In: Basin Form. Plateau Tectonics 79, 815–848.
Hedenquist, J.W., Harris, M., Camus, F. (Eds.), Geology and Genesis of Major Copper Cerpa, L.M., Bissig, T., Kyser, K., McEwan, C., Macassi, A., Rios, H.W., 2013. Lithologic con-
Deposits and Districts of the World: A Tribute to Richard H. Sillitoe. Society of trols on mineralization at the Lagunas Norte high sulfidation epithermal gold deposit,
Economic Geologists Special Publication 16, pp. 573–618. northern Peru. Mineral. Deposita 48, 653–673.
Barreda, J., Loayza, D., Juarez, P., Torres, R., 2004. Depósitos epitermales de alta sulfuración Charchaflie, D., Tosdal, R.M., Mortensen, J.K., 2007. Geologic framework of the Veladero
en el distrito minero Aruntani, Moquegua, XII Congreso Peruano de Geología, Lima, high-sulfidation epithermal deposit area, Cordillera Frontal, Argentina. Econ. Geol.
Peru. pp. 605–608. Bull. Soc. 102, 171–192.
Baumgartner, R., Fontbote, L., Vennemann, T., 2008. Mineral zoning and geochemistry of Charrier, R., Baeza, O., Elgueta, S., Flynn, J.J., Gans, P., Kay, S.M., Muñoz, N., Wyss, A.R.,
epithermal polymetallic Zn–Pb–Ag–Cu–Bi mineralization at Cerro de Pasco, Peru. Zurita, E., 2002. Evidence for Cenozoic extensional basin development and tectonic
Econ. Geol. 103, 493–537. inversion south of the flat-slab segment, southern Central Andes, Chile (33°–36°S.L.).
Baumgartner, R., Fontbote, L., Spikings, R., Ovtcharova, M., Schaltegger, U., Schneider, J., J. S. Am. Earth Sci. 15, 117–139.
Page, L., Gutjahr, M., 2009. Bracketing the age of magmatic-hydrothermal activity at Chiaradia, M., Fontbote, L., Beate, B., 2004. Cenozoic continental arc magmatism and
the Cerro de Pasco epithermal polymetallic deposit, central Peru; a U–Pb and associated mineralization in Ecuador. Mineral. Deposita 39, 204–222.
40
Ar/39Ar study. Econ. Geol. 104, 479–504. Chouinard, A., Williams-Jones, A.E., Leonardson, R.W., Hodgson, C.J., Silva, P., Tellez, C.,
Bendezú, R., Fontboté, L., 2009. Cordilleran Epithermal Cu–Zn–Pb–(Au–Ag) Mineraliza- Vega, J., Rojas, F., 2005. Geology and genesis of the multistage high-sulfidation
tion in the Colquijirca District, Central Peru: Deposit-Scale Mineralogical Patterns. epithermal Pascua Au–Ag–Cu deposit, Chile and Argentina. Econ. Geol. 100, 463–490.
Econ. Geol. 104, 905–944. Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., France, L.J., McBride, S.L., Woodman, P.L.,
Bendezú, R., Fontboté, L., Cosca, M., 2003. Relative age of Cordilleran base metal lode and Wasteneys, H.A., Sandeman, H.A., Douglas, A., Archibald, D.A., 1990. Geologic and
replacement deposits, and high sulfidation Au–(Ag) epithermal mineralization in the geochronologic constraints on the metallogenic evolution of the Andes of southeast-
Colquijirca mining district, central Peru. Mineral. Deposita 38, 683–694. ern Peru. Econ. Geol. 85, 1520–1583.
Bendezú, R., Page, L., Spikings, R., Pecskay, Z., Fontboté, L., 2008. New 40Ar/39Ar alunite Clarke, J.D.A., 2006. Antiquity of aridity in the Chilean Atacama Desert. Geomorphology
ages from the Colquijirca district, Peru: evidence of a long period of magmatic SO2 73, 101–114.
362 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Coltorti, M., Ollier, C., 2000. Geomorphic and tectonic evolution of the Ecuadorian Andes. Bethke, P.M. (Eds.), Geology and Geochemistry of Epithermal Systems. Society of Eco-
Geomorphology 32, 1–19. nomic Geologists, Reviews in Economic Geology No.2, 129–168.
Cooke, D.R., Hollings, P., Walshe, J.L., 2005. Giant porphyry deposits: characteristics, Heald, P., Foley, N.K., Hayba, D.O., 1987. Comparative anatomy of volcanic-hosted
distribution, and tectonic controls. Econ. Geol. 100, 801–818. epithermal deposits; acid-sulfate and adularia-sericite types. Econ. Geol. 82, 1–26.
Cornejo, P., Mpodozis, C., Ramírez, C., Tomlinson, A., 1993. Estudio geológico de la Región de Heather, K.B., Leach, T., Clark, A.H., 2003a. Geology of the El Indio Au–Cu–Ag Deposit: The
Potrerillos y El Salvador (26–27 Lat. S). Servicio Nacional de Geología y Minería- Final Chapter? 10° Congreso Geológico Chileno. Universidad de Concepción,
Corporación del Cobre, informe registrado IR-93–01, 12 cuadrángulos escala 1:50.000, Concepción, Chile (CD-ROM).
258 p. Heather, K.B., Bissig, T., Staff of Exploraciones Barrick Chile Ltda, 2003b. Regional Geology
Cornejo, P., Mpodozis, C., Tomlinson, A.J., 1998. Hoja Salar de Maricunga, Región de of the El Indio Mineral District, North-central Chile. 10° Congreso Geológico Chileno.
Atacama. Servicio Nacional de Geología y Minería, Mapas Geológicos N° 7, 1 mapa Universidad de Concepción, Concepción, Chile (CD-ROM).
escala 1:100.000. Santiago. Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of hydro-
Deckart, K., Clark, A., Cuadra, P., Fanning, M., 2013. Refinement of the time–space thermal ore deposits. Nature 370, 519–527.
evolution of the giant Mio-Pliocene Río Blanco–Los Bronces porphyry Cu–Mo cluster, Hedenquist, J.W., Taran, Y.A., 2013. Modeling the Formation of Advanced Argillic
Central Chile: new U–Pb (SHRIMP II) and Re–Os geochronology and 40Ar/39Ar Lithocaps: Volcanic Vapor Condensation Above Porphyry Intrusions. Econ. Geol.
thermochronology data. Mineral. Deposita 48, 57–79. 108, 1523–1540.
Deen, J.A., Rye, R.O., Munoz, J.L., Drexler, J.W., 1994. The magmatic hydrothermal system Hedenquist, J.W., Arribas Jr., A., Reynolds, T.J., 1998. Evolution of an intrusion-centered
at Julcani, Peru: evidence from fluid inclusions and hydrogen and oxygen isotopes. hydrothermal system; Far Southeast-Lepanto porphyry and epithermal Cu–Au
Econ. Geol. 89, 1924–1938. deposits, Philippines. Econ. Geol. Bull. Soc. 93, 373–404.
Deyell, C., Bissig, T., Rye, R., 2004. Isotopic evidence for magmatic-dominated epithermal Hedenquist, J.W., Arribas, R.A., Gonzalez-Urien, E., 2000. Exploration for epithermal gold
processes in the El Indio–Pascua Au–Cu–Ag Belt and relationship to geomorphologic deposits. Rev. Econ. Geol. 13, 245–277.
setting. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discov- Heinrich, C.A., 2005. The physical and chemical evolution of low-salinity magmatic fluids
eries, Concepts, and Updates. Society of Economic Geologists Special Publication 11, at the porphyry to epithermal transition: a thermodynamic study. Mineral. Deposita
pp. 55–74. 39, 864–889.
Deyell, C.L., Rye, R.O., Landis, G.P., Bissig, T., 2005a. Alunite and the role of magmatic fluids Heinrich, C.A., 2007. Fluid–fluid interactions in magmatic-hydrothermal ore formation.
in the Tambo high-sulfidation deposit, El Indio–Pascua belt, Chile. Chem. Geol. 215, Rev. Mineral. Geochem. 65, 363–387.
185–218. Heinrich, C.A., Driesner, T., Stefánsson, A., Seward, T.M., 2004. Magmatic vapor contraction
Deyell, C.L., Leonardson, R., Rye, R.O., Thompson, J.F.H., Bissig, T., Cooke, D.R., 2005b. and the transport of gold from the porphyry environment to epithermal ore deposits.
Alunite in the Pascua–Lama high-sulfidation deposit: Constraints on alteration and Geology 32, 761–764.
ore deposition using stable isotope geochemistry. Econ. Geol. 100, 131–148. Henderson, R.D., 2009. Fruta del Norte Project Ecuador NI 43-101 Technical Report. p. 135.
Dietrich, A., Gutierrez, R., Nelson, E., Layer, P., 2011. Geology of the epithermal Ag–Au Hoke, G.D., Isacks, B.L., Jordan, T.E., Blanco, N., Tomlinson, A.J., Ramezani, J., 2007. Geomor-
Huevos Verdes vein system and San José district, Deseado massif, Patagonia, phic evidence for post-10 Ma uplift of the western flank of the Central Andes 18° 30′–
Argentina. Mineral. Deposita 1–17. 22°S. Tectonics 26, TC5021. http://dx.doi.org/10.1029/2006TC002082.
Echavarría, L.E., Schalamuk, I.B., Etcheverry, R.O., 2005. Geologic and tectonic setting of Holley, E.A., 2012. The Veladero High-sulfidation Epithermal Au–Ag deposit, Argentina:
Deseado Massif epithermal deposits, Argentina, based on El Dorado–Monserrat. J. S. Volcanic Stratigraphy, Alteration, Mineralization, and Quartz Paragenesis. Colorado
Am. Earth Sci. 19, 415–432. School of Mines, Golden, Colorado, p. 226 (Unpublished PhD thesis).
Einaudi, M.T., 1977. Environment of ore deposition at Cerro de Pasco, Peru. Econ. Geol. 72, Horton, B.K., Parra, M., Saylor, J.E., Nie, J., Mora, A., Torres, V., Strecker, M.R., 2010. Resolv-
893–924. ing uplift of the northern Andes using detrital zircon age signatures. GSA Today 20,
Escalante, A.D., 2008. Patterns of Distal Alteration Zonation Around Antamina Cu–Zn 4–10.
Skarn and Uchucchacua Ag-base Metal Vein Deposits, Peru: Mineralogical, Chemical Hungerbuehler, D., Steinmann, M., Winkler, W., Seward, D., Egueez, A., Peterson, D.E.,
and Isotopic Evidence for Fluid Composition, and Infiltration, and Implications for Helg, U., Hammer, C., 2002. Neogene stratigraphy and Andean geodynamics of south-
Mineral Exploration. The University of British Columbia, Vancouver, Canada, p. 788 ern Ecuador. Earth Sci. Rev. 57, 75–124.
(Unpublished PhD Thesis). Jannas, R.R., Beane, R.E., Ahler, B.A., Brosnahan, D.R., 1990. Gold and copper mineralization
Escalante, A., Dipple, G.M., Barker, S.L.L., Tosdal, R., 2010. Defining trace element alteration at the El Indio deposit, Chile. J. Geochem. Explor. 36, 233–266.
halos to skarn deposits hosted in heterogeneous carbonate rocks; case study from the Jannas, R.R., Bowers, T.S., Petersen, U., Beane, R.E., 1999. High-sulfidation deposit types in
Cu–Zn Antamina skarn deposit, Peru. J. Geochem. Explor. 105, 117–136. the El Indio District, Chile. Spec. Publ. Soc. Econ. Geol. 7, 219–266.
Felder, F., Ortíz, G., Campos, C., Monsalve, I., Silva, A., 2005. Angostura Project: a high Jordan, T.E., Nester, P.L., Blanco, N., Hoke, G.D., Davila, F., Tomlinson, A.J., 2010. Uplift of the
sulfidation gold-silver deposit located in the Santander Complex of north eastern Co- Altiplano–Puna Plateau; a view from the west. Tectonics 29. http://dx.doi.org/10.
lombia. Proceedings Congreso Internacional de Prospectores y Exploradores (Pro- 1029/2010TC002661.
Explo). Instituto de Ingenieros de Minas del Perú, Lima p. 15. Kay, S.M., Mpodozis, C., 2001. Central Andean Ore deposits linked to evolving shallow
Fernández, R.R., Blesa, A., Moreira, P., Echeveste, H., Mykietiuk, K., Andrada de Palomera, P., subduction systems and thickening crust. GSA Today 11, 4–9.
Tessone, M., 2008. Los depósitos de oro y plata vinculados al magmatismo jurásico de Kay, S.M., Mpodozis, C., Coira, B., 1999. Neogene magmatism, tectonism, and mineral de-
la Patagonia: revisión y perspectivas para la exploración. Rev. Asoc. Geol. Argent. 63, posits of the Central Andes (22° to 33° S latitude). In: Skinner, B.K. (Ed.), Geology and
665–681. Ore Deposits of the Central Andes. Soceity of Economic Geologists Special Publication
Fifarek, R.H., Rye, R.O., 2005. Stable isotope geochemistry of the Pierina high-sulfidation 7, pp. 27–59.
Au–Ag deposit, Peru; influence of hydrodynamics on SO2–H2S sulfur isotopic King, A.R., 1992. Magmatism, Structure and Mineralization in the Maricunga Belt. Imperial
exchange in magmatic-steam and steam-heated environments. Chem. Geol. 215, College London (University of London), N. Chile.
253–279. Leal-Mejía, H., 2011. Phanerozoic Gold Metallogeny in the Colombian Andes — A Tectono-
Gansser, A., 1973. Facts and theories on the Andes (Twenty-sixth William Smith lecture). magmatic Approach. Universitat de Barcelona, p. 1000.
J. Geol. Soc. Lond. 129, 93–131. Leal-Mejía, H., Melgarejo, I., Draper, J., Shaw, R., 2011. Phanerozoic gold metallogeny in
Garzione, C.N., Hoke, G.D., Libarkin, J.C., Withers, S., MacFaddon, B., Eiler, J., Ghosh, P., the Colombian Andes, Proceedings Let's talk Ore Deposits. SGA Biennial Meeting,
Mulch, A., 2008. Rise of the Andes. Science 320, 1304–1307. Antofagasta, Chile.
Gonzalez, L., Pfiffner, O.A., 2012. Morphologic evolution of the Central Andes of Peru. Int. Lesage, G., 2011. Geochronology, Petrography, Geochemical Constraints, and Fluid
J. Earth Sci. 101, 307–321. Characterization of the Buriticá Gold Deposit, Antioquia Department, Colombia,
Gregory-Wodzicki, K.M., 2000. Uplift history of the Central and Northern Andes: A Masters Abstracts International.
review. Bull. Geol. Soc. Am. 112, 1091–1105. Lodder, C., Padilla, R., Shaw, R., Garzon, T., Palacio, E., Jahoda, R., 2010. Discovery history of
Groepper, H., Calvo, M., Crespo, H., Bisso, C.R., Cuadra, W.A., Dunkerley, P.M., Aguirre, E., the La Colosa gold porphyry deposit, Cajamarca, Colombia. In: Goldfarb, R.J., Marsh, E.
1991. The epithermal gold–silver deposit of Choquelimpie, northern Chile. Econ. E., Monecke, T. (Eds.), The Challenge of Finding New Mineral Resources: Global
Geol. Bull. Soc. 86, 1206–1221. Metallogeny, Innovative Exploration, and New Discoveries. Society of Economic
Gustafson, L.B., Hunt, J.P., 1975. The porphyry copper deposit at El Salvador, Chile. Econ. Geologists special publication vol. I, pp. 19–28 (15).
Geol. 70, 857–912. Longo, A.A., Dilles, J.H., Grunder, A.L., Duncan, R., 2010. Evolution of calc-alkaline volca-
Gustafson, L.B., Orquera, W., McWilliams, M., Castro, M., Olivares, O., Rojas, G., Maluenda, nism and associated hydrothermal gold deposits at Yanacocha, Peru. Econ. Geol.
J., Mendez, M., 2001. Multiple Centers of Mineralization in the Indio Muerto District, 105, 1191–1241.
El Salvador, Chile. Econ. Geol. 96, 325–350. Love, D.A., Clark, A.H., Glover, J.K., 2004. The lithologic, stratigraphic, and structural setting
Gustafson, L.B., Vidal, C., Pinto, R., Noble, D., 2004. Porphyry-epithermal transition, of the giant Antamina copper–zinc skarn deposit, Ancash, Peru. Econ. Geol. 99,
Cajamarca region, northern Peru. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), 887–916.
Andean Metallogeny: New Discoveries, Concepts, and Updates. Society of Economic Lozada Calderón, A.J., McBride, S.L., Bloom, M.S., 1994. The geology and 40Ar/39Ar geochro-
Geologists Special Publication 11, pp. 279–299. nology of magmatic activity and related mineralization in the Nevados del Famatina
Gutscher, M.A., Olivet, J.L., Aslanian, D., Eissen, J.P., Maury, R., 1999a. The ‘lost Inca mining district, La Rioja province, Argentina. J. S. Am. Earth Sci. 7, 9–24.
Plateau’: Cause of flat subduction beneath Peru? Earth Planet. Sci. Lett. 171, 335–341. MacDonald, P.J., Bissig, T., Hart, C.J., Barreno, J., Viera, F., 2011. The hydrothermal evolution
Gutscher, M.A., Malavieille, J., Lallemand, S., Collot, J.Y., 1999b. Tectonic segmentation of of the Quimsacocha high sulfidation Au–Ag–Cu deposit, Azuay Province, Ecuador,
the North Andean margin: impact of the Carnegie Ridge collision. Earth Planet. Sci. 11th Biennial Conference Society for Geology Applied to Ore Deposits, Antofagasta,
Lett. 168, 255–270. Chile.
Hampel, A., 2002. The migration history of the Nazca Ridge along the Peruvian active MacDonald, P.J., Bissig, T., Hart, C.J., Barreno, J., Viera, F., Mantilla, G., Rogers, J., 2012.
margin: a re-evaluation. Earth Planet. Sci. Lett. 203, 665–679. Hydrothermal evolution of the Rio Falso Fault, Quimsacocha high sulfidation
Hayba, D.O., Bethke, P.M., Heald, P., Foley, N.K., 1985. The geological, mineralogical and Au–Ag–Cu District, Azuay Province, Ecuador. XVI Congreso Peruano de Geología
geochemical characteristics of volcanic-hosted epithermal deposits. In: Berger, B.R., & SEG2010 Conference, Lima, Peru, abstract and oral presentation.
T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364 363

Mallette, P., Rojas, R., Guttierrez, A.R., 2004. Geology, mineralization, and genesis of the La Ossandon, C.G., Freraut, C.R., Gustafson, L.B., Lindsay, D.D., Zentilli, M., 2001. Geology of
Quinua gold deposit, Yanacocha district, northern Peru. In: Sillitoe, R.H., Perello, J., the Chuquicamata Mine: A Progress Report. Econ. Geol. 96, 249–270.
Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and Updates. Oviedo, L., Fuster, N., Tschischow, N., Ribba, L., Zuccone, A., Grez, E., Aguilar, A., 1991.
Society of Economic Geologists Special Publication 11, pp. 301–312. General geology of La Coipa precious metal deposit, Atacama, Chile. Econ. Geol.
Mantilla Figueroa, L.C., Bissig, T., Valencia, V., Hart, C.J.R., 2013. The magmatic history of Bull. Soc. 86, 1287–1300.
the Vetas–California mining district, Santander Massif, Eastern Cordillera, Colombia. Pardo‐Casas, F., Molnar, P., 1987. Relative motion of the Nazca (Farallon) and South
J. S. Am. Earth Sci. 45, 235–249. American plates since Late Cretaceous time. Tectonics 6, 233–248.
Mantilla, L.C., Valencia, V.A., Barra, F., Pinto, J., Colegial, J., 2009. GEocronología U–Pb de los Perello, J., Carlotto, V., Zarate, A., Ramos, P., Posso, H., Neyra, C., Caballero, A., Fuster, N.,
cuerpos porfíriticos del distrito auríofero de Vetas–California (Dpto de Santander, Muhr, R., 2003. Porphyry-style alteration and mineralization of the Middle Eocene
Colombia). Bol. Geol. 31, 31–43. to Early Oligocene Andahuaylas–Yauri Belt, Cuzco Region, Peru. Econ. Geol. 98,
Mantilla, F.L.C., Mendoza, F.H., Bissig, T., Craig, H., 2011. Nuevas evidencias sobre el 1575–1605.
magmatismo Miocenico en el distrito minero de Vetas–California (Macizo de Perelló, J., Brockway, H., Martini, R., 2004. Discovery and geology of the Esperanza por-
Santander, Cordillera Oriental, Colombia). Bol. Geol. 33, 43–58. phyry copper–gold deposit. Antofagasta Region, northern Chile: S. In: Sillitoe, R.H.,
Marsh, T.M., 1997. Geochronology, Thermochronology, and Isotope Systematics of the Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and
Cu–Au and Au–Ag Deposits of the Potrerillos District, Atacama Region, Chile. Stanford Updates. Society of Economic Geologists Special Publication 11, pp. 167–186.
University, p. 341, (Unpubl. PhD thesis). Petersen, U., Noble, D., Arenas, M., Goodell, P., 1977. Geology of the Julcani mining district,
Marsh, T.M., Einaudi, M.T., McWilliams, M., 1997. 40Ar/39Ar geochronology of Cu–Au and Peru. Econ. Geol. 72, 931–949.
Au–Ag mineralization in the Potrerillos District, Chile. Econ. Geol. 92, 784–806. Petford, N., Atherton, M.P., 1992. Granitoid emplacement and deformation along a major
Martin, M., Clavero, J., Mpodozis, C., Cutiño, L., 1995. Estudio geológico regional de la crustal lineament: the Cordillera Blanca, Peru. Tectonophysics 205, 171–185.
franja El Indio. Cordillera de Coquimbo: Informe registrado IR-95-6 Servicio Nacional Poblete, J.A., Bissig, T., Mortensen, J.K., Gabites, J., Friedman, R., Rodriguez, M., 2014.
de Geología y Minería, Chile, and Compañía Minera San José. The Cerro Bayo District, Chilean Patagonia: Late Jurassic to Cretaceous Magmatism
Martinod, J., Husson, L., Roperch, P., Guillaume, B., Espurt, N., 2010. Horizontal subduction and Protracted History of Epithermal Ag–Au Mineralization. Econ. Geol. 109,
zones, convergence velocity and the building of the Andes. Earth Planet. Sci. Lett. 299, 487–502.
299–309. Polania, J.H., 1980. Die Uranvorkommen von California bei Bucaramanga (Kolumbien).
Masterman, G.J., Cooke, D.R., Berry, R.F., Clark, A.H., Archibald, D.A., Mathur, R., Walshe, J.L., University of Stuttgart, Germany, p. 152 (Unpubl. PhD thesis).
Duran, M., 2004. 40Ar/39Ar and Re–Os geochronology of porphyry copper–molybdenum Pudack, C., Halter, W.E., Heinrich, C.A., Pettke, T., 2009. Evolution of magmatic vapor to
deposits and related copper–silver veins in the Collahuasi District, northern Chile. Econ. gold-rich epithermal liquid: the porphyry to epithermal transition at Nevados de
Geol. Bull. Soc. 99, 673–690. Famatina, Northwest Argentina. Econ. Geol. 104, 449–477.
Moncada, D., Mutchler, S., Nieto, A., Reynolds, T.J., Rimstidt, J.D., Bodnar, R.J., 2012. Puig, A., Diaz, S., Cuitino, L., 1988. Sistemas hidrotermales asociados a calderas en el arco
Mineral textures and fluid inclusion petrography of the epithermal Ag–Au volcanico paleogeno de la region de Antofagasta, Chile; distritos El Guanaco, Cachinal
deposits at Guanajuato, Mexico: Application to exploration. J. Geochem. Explor. de La Sierra y El Soldado. Rev. Geol. Chile 15, 57–82.
114, 20–35. Quang, C.X., Clark, A.H., Lee, J.K., Guillén, J., 2003. 40Ar–39Ar Ages of Hypogene and Super-
Montario, M., Garver, J., Reiners, P., Ramage, J., Brandon, M., 2005a. Thermochronological gene Mineralization in the Cerro Verde–Santa Rosa Porphyry Cu–Mo Cluster,
evidence for kilometre-scale incision of the northern Peruvian Andes. Geol. Soc. Am. Arequipa, Peru. Econ. Geol. 98, 1683–1696.
Abstr. Programs 37, 553. Quang, C.X., Clark, A.H., Lee, J.K., Hawkes, N., 2005. Response of supergene processes to
Montario, M., Garver, J.I., Reiners, P., Ramage, J.M., 2005b. Timing of canyon incision of the episodic Cenozoic uplift, pediment erosion, and ignimbrite eruption in the porphyry
Río Pativilca in response to uplift of the Andes in northern Perú. Geol. Soc. Am. Abstr. copper province of southern Peru. Econ. Geol. 100, 87–114.
Programs 37, 76. Rainbow, A., 2009. Genesis and Evolution of the Pierina High-sulphidation Epithermal
Montgomery, A.T., 2012. Mmetallogenetic Controls on Miocene High-sulphidation Au–Ag Deposit, Ancash, Perú. Queen's University, Kingston, On, Canada, p. 277
Epithermal Gold Mineralization, Alto Chicama District, La Libertad, Northern Perú. (Unpublished PhD thesis).
Queen's University, Kingston, Ontario, Canada, p. 381 (Unpublished PhD thesis). Rainbow, A., Clark, A.H., Kyser, T.K., Gaboury, F., Hodgson, C.J., 2005. The Pierina
Morche, W., Velasco, C., Loayza, D., Clark, A., 2008. Late Miocene High-sulfidation epithermal Au–Ag deposit, Ancash, Peru; paragenetic relationships, alunite textures,
Epithermal Gold Deposits of the Aruntani District, Southern Peru, Congreso Geologico and stable isotope geochemistry. Chem. Geol. 215, 235–252.
del Peru. Sociedad Geologica del Peru, Lima, Peru. Rainbow, A., Kyser, T.K., Clark, A.H., 2006. Isotopic evidence for microbial activity during
Mortimer, C., 1973. The Cenozoic history of the southern Atacama Desert, Chile. J. Geol. supergene oxidation of a high-sulfidation epithermal Au–Ag deposit. Geology 34,
Soc. Lond. 129, 505–526. 269–272.
Mote, T.I., Becker, T.A., Renne, P., Brimhall, G.H., 2001. Chronology of exotic mineralization Ramos, V.A., 2008. The basement of the Central Andes; the Arequipa and related terranes.
at El Salvador, Chile, by 40Ar/39Ar dating of copper wad and supergene alunite. Econ. Annu. Rev. Earth Planet. Sci. 36, 289–324.
Geol. Bull. Soc. 96, 351–366. Ramos, V.A., 2009. Anatomy and global context of the Andes: Main geologic features and
Mpodozis, C., Clavero, J., 2002. Tertiary evolution of the southwestern edge of the Puna the Andean orogenic cycle. Geol. Soc. Am. Mem. 204, 31–65.
Plateau: Cordillera Claudio Gay (26°′27° S), northern Chile. International Symposium Restrepo-Moreno, S.A., Foster, D.A., Stockli, D.F., Parra-Sánchez, L.N., 2009. Long-term ero-
on Andean Geodynamics (ISAG), 5th,Toulouse, France, 2002, Extended Abstracts, sion and exhumation of the “Altiplano Antioqueño”, Northern Andes (Colombia)
pp. 445–448. from apatite (U–Th)/He thermochronology. Earth Planet. Sci. Lett. 278, 1–12.
Mpodozis, C., Cornejo, P., 2012. Cenozoic Tectonics and Porphyry Copper Systems of the Riquelme, R., Herail, G., Martinod, J., Charrier, R., Darrozes, J., 2007. Late Cenozoic geomor-
Chilean Andes. In: Hedenquist, J.W., Harris, M., Camus, F. (Eds.), Geology and Genesis phologic signal of Andean forearc deformation and tilting associated with the uplift
of Major Copper Deposits and Districts of the World: A Tribute to Richard H. Sillitoe. and climate changes of the southern Atacama Desert (26° S–28° S). Geomorphology
Society of Economic Geologists Special Publication 16, pp. 329–360. 86, 283–306.
Mpodozis, C., Kay, S.M., 2003. Neogene tectonics, ages and mineralization along the tran- Riquelme, R., Darrozes, J., Maire, E., Herail, G., Soula, J.-C., 2008. Long-term denudation
sition zone between the El Indio and Maricunga mineral belts (Argentina and Chile rates from the Central Andes (Chile) estimated from a Digital Elevation Model
28–29°). 10° Congreso Geológico Chileno. Universidad de Concepción, Concepción, using the Black Top Hat function and Inverse Distance Weighting: implications for
Chile (CD-ROM). the Neogene climate of the Atacama Desert. Andean Geol. 35, 105.
Muntean, J.L., Einaudi, M.T., 2001. Porphyry-Epithermal Transition: Maricunga Belt, Rodriguez, M.A.L., 2014. Geology, Alteration, Mineralization and Hydrothermal Evolution
Northern Chile. Econ. Geol. 96, 743–772. of the La Bodega–La Mascota Deposits, California–Vetas Mining District, Eastern
Murakami, H., Seo, J.H., Heinrich, C.A., 2010. The relation between Cu/Au ratio and Cordillera of Colombia, Northern Andes. The University of British Columbia, p. 325,
formation depth of porphyry-style Cu–Au ± Mo deposits. Mineral. Deposita 45, (Unpubl. MSc thesis).
11–21. Rodriguez, M.P., Aguilar, G., Urresty, C., Charrier, R., 2014. Neogene Landscape Evolution in
Niemeyer, H., Munizaga, R., 2008. Structural control of the emplacement of the Potrerillos the Andes of North-central Chile Between 28 and 32°S: Interplay Between Tectonic
porphyry copper, central Andes of Chile. J. S. Am. Earth Sci. 26, 261–270. and Erosional Processes: Geological Society. Special Publications, London, p. 399
Noble, D.C., McKee, E.H., 1999. The Miocene metallogenic belt of central and northern http://dx.doi.org/10.1144/SP399.15.
Peru. In: Skinner, B.K. (Ed.), Geology and Ore Deposits of the Central Andes. Society Rosenbaum, G., Giles, D., Saxon, M., Betts, P.G., Weinberg, R.F., Duboz, C., 2005. Subduction
of Economic Geologists Special Publication 7, pp. 155–193. of the Nazca Ridge and the Inca Plateau: Insights into the formation of ore deposits in
Noble, D.C., Silberman, M.L., 1984. Evolución volcánica e hidrotermal y cronología de K-Ar Peru. Earth Planet. Sci. Lett. 239, 18–32.
del distrito minero de Julcani. Sociedad Geológica del Perú, Volumen Jubilar, Perú p. Rossell, W., De La Cruz, O., Romero, D., 2006. Visita técnica a la Mina Quicay–
1–35. Pacoyán. Instituto Geologico Minero y Metalurgico (Ingemmet), Peru (www.
Noble, D.C., McKee, E.H., Mourier, T., Mégard, F., 1990. Cenozoic stratigraphy, magmatic slideshare.net. Technical powerpoint Presentation available for download, July
activity, compressive deformation, and uplift in northern Peru. Geol. Soc. Am. Bull. 3rd, 2014).
102, 1105–1113. Rye, R.O., 1993. The evolution of magmatic fluids in the epithermal environment; the
Noble, D.C., Vidal, C.E., Perelló, J., Rodriguez, P.O., 2004. Space–Time Relationships of Some stable isotope perspective. Econ. Geol. 88, 733–752.
Porphyry Cu–Au, Epithermal Au, and other magmatic-related mineral deposits in Rye, R.O., 2005. A review of the stable-isotope geochemistry of sulfate minerals in
Northern Peru. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: selected igneous environments and related hydrothermal systems. Chem. Geol.
New Discoveries, Concepts, and Updates. Society of Economic Geologists Special 215, 5–36.
Publication 11, pp. 313–319. Rye, R.O., Bethke, P.M., Wasserman, M.D., 1992. The stable isotope geochemistry of acid
O'Driscoll, L.J., Richards, M.A., Humphreys, E.D., 2012. Nazca–South America interactions sulfate alteration. Econ. Geol. 87, 225–262.
and the late Eocene–late Oligocene flat‐slab episode in the central Andes. Tectonics 31. Sandeman, H.A., Clark, A.H., Farrar, E., 1995. An integrated tectono-magmatic model for
Olson, S.F., 1984. Geology of the Portrerillos District, Atacama, Chile. Stanford University, the evolution of the southern Peruvian Andes (13–20 S) since 55 Ma. Int. Geol. Rev.
(Unpubl. PhD thesis). 37, 1039–1073.
364 T. Bissig et al. / Ore Geology Reviews 65 (2015) 327–364

Scher, S., Williams-Jones, A.E., Williams-Jones, G., 2013. Fumarolic activity, acid-sulfate Taran, Y.A., Bernard, A., Gavilanes, J.-C., Africano, F., 2000. Native gold in mineral precipi-
alteration, and high sulfidation epithermal precious metal mineralization in the tates from high-temperature volcanic gases of Colima Volcano, Mexico. Appl.
crater of Kawah Ijen volcano, Java, Indonesia. Econ. Geol. 108, 1099–1118. Geochem. 15, 337–346.
Schildgen, T.F., Hodges, K.V., Whipple, K.X., Reiners, P.W., Pringle, M.S., 2007. Uplift of the Tassinari, C.C.G., Pinzon, F.D., Buena Ventura, J., 2008. Age and sources of gold mineraliza-
western margin of the Andean Plateau revealed from canyon incision history, tion in the Marmato mining district, NW Colombia: a Miocene–Pliocene epizonal
southern Peru. Geology 35, 523–526. gold deposit. Ore Geol. Rev. 33, 505–518.
Schuette, P., Chiaradia, M., Beate, B., 2010. Geodynamic controls on Tertiary arc magmatism Teal, L., Benavides, A., 2010. History and geologic overview of the Yanacocha mining
in Ecuador; constraints from U/Pb zircon geochronology of Oligocene–Miocene district, Cajamarca, Peru. Econ. Geol. 105, 1173–1190.
intrusions and regional age distribution trends. Tectonophysics 489, 159–176. Thompson, J.F.H., Gale, V.G., Tosdal, R.M., Wright, W.A., 2004. Characteristics and
Schuette, P., Chiaradia, M., Barra, F., Villagomez, D., Beate, B., 2012. Metallogenic features formation of the Jeronimo carbonate-replacement gold deposit, Potrerillos district,
of Miocene porphyry Cu and porphyry-related mineral deposits in Ecuador revealed Chile. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discov-
by Re/Os, 40Ar/39Ar, and U/Pb geochronology. Mineral. Deposita 47, 383–410. eries, Concepts, and Updates. Society of Economic Geologists Special Publication 11,
Sébrier, M., Lavenu, A., Fornari, M., Soulas, J.-P., 1988. Tectonics and uplift in central Andes pp. 75–95.
(Perú, Bolivia and northern Chile) from Eocene to present. Géodynamique 3, 85–106. Thouret, J.C., Wörner, G., Gunnell, Y., Singer, B., Zhang, X., Souriot, T., 2007. Geochronologic
Shagam, R., Banks, P.O., Kohn, B.P., Dasch, L.E., Vargas, R., Pimentel, N., Rodríguez, G.I., and stratigraphic constraints on canyon incision and Miocene uplift of the Central
1984. Tectonic Implications of CretaceousePliocene fission-track ages from rocks of Andes in Peru. Earth Planet. Sci. Lett. 263, 151–166.
the circum-Maracaibo Basin region of western Venezuela and eastern Colombia. Tosdal, R.M., Clark, A.H., Farrar, E., 1984. Cenozoic polyphase landscape and tectonic
Geol. Soc. Am. Mem. 162, 385–412. evolution of the Cordillera Occidental, southernmost Peru. Geol. Soc. Am. Bull. 95,
Sillitoe, R.H., 1994. Erosion and collapse of volcanoes: Causes of telescoping in intrusion- 1318–1332.
centered ore deposits. Geology 22, 945–948. Tosdal, R.M., Dilles, J.H., Cooke, D.R., 2009. From source to sinks in auriferous magmatic-
Sillitoe, R.H., 1995. Exploration of porphyry copper lithocaps. Proc. Pac. Rim Congr. 95, hydrothermal porphyry and epithermal deposits. Elements 5, 289–295.
527–532. Van Der Lelij, R., 2013. Reconstructing North-western Gondwana With Implications for the
Sillitoe, R.H., 2008. Special Paper: Major Gold Deposits and Belts of the North and South Evolution of the Iapetus and Rheic Oceans: A Geochronological, Thermochronological
American Cordillera: Distribution, Tectonomagmatic Settings, and Metallogenic and Geochemical Study. University of Geneva, Switzerland, p. 221 (Unpubl. PhD thesis).
Considerations. Econ. Geol. 103, 663–687. Vargas, C.A., Mann, P., 2013. Tearing and Breaking Off of Subducted Slabs as the Result of
Sillitoe, R.H., 2010. Porphyry copper systems. Econ. Geol. 105, 3–41. Collision of the Panama Arc‐Indenter with Northwestern South America. Bull.
Sillitoe, R.H., Hedenquist, J.W., 2003. Linkages between volcanotectonic settings, ore-fluid Seismol. Soc. Am. 103, 2025–2046.
compositions, and epithermal precious metal deposits. In: Simmons, S.F., Graham, I. Vidal, C., Ligarda, R., 2004. Enargite–gold deposits at Marcapunta. Colquijirca mining dis-
(Eds.), Volcanic, Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of trict, central Peru: Mineralogic and geochemical zoning in subvolcanic, limestone-
Processes Within the Earth. Special Publication–Society of Economic Geologists 10, replacement deposits of high-sulfidation epithermal type. In: Sillitoe, R.H., Perello, J.,
pp. 315–343. Vidal, C.E. (Eds.), Andean Metallogeny: New Discoveries, Concepts, and Updates.
Sillitoe, R.H., Mortimer, C., Clark, A.H., 1968. A Chronology of Landform Evolution and Society of Economic Geologists Special Publication 11, pp. 231–241.
Supergene Mineral Alteration, Southern Atacama Desert. Institution of Mining and Vila, T., Sillitoe, R.H., 1991. Gold-rich porphyry systems in the Maricunga Belt, northern
Metallurgy Transactions, Chile, pp. 166–169 (sec. B 77). Chile. Econ. Geol. Bull. Soc. 86, 1238–1260.
Sillitoe, R.H., McKee, E.H., Vila, T., 1991. Reconnaissance K–Ar geochronology of the Villagómez, D., Spikings, R., 2013. Thermochronology and tectonics of the Central and
Maricunga gold–silver belt, northern Chile. Econ. Geol. 86, 1261–1270. Western Cordilleras of Colombia: Early Cretaceous–Tertiary evolution of the
Sillitoe, R.H., Tolman, J., Van Kerkvoort, G., 2013. Geology of the caspiche porphyry gold– Northern Andes. Lithos 160, 228–249.
copper deposit, maricunga belt, Northern Chile. Econ. Geol. 108, 585–604. Villagomez, D., Spikings, R., Mora, A., Guzman, G., Ojeda, G., Cortes, E., van der Lelij, R.,
Simmons, S.F., 1991. Hydrologic implications of alteration and fluid inclusion studies in 2011. Vertical tectonics at a continental crust-oceanic plateau plate boundary zone;
the Fresnillo District, Mexico; evidence for a brine reservoir and a descending fission-track thermochronology of the Sierra Nevada de Santa Marta, Colombia.
water table during the formation of hydrothermal Ag–Pb–Zn orebodies. Econ. Geol. Tectonics 30.
86, 1579–1601. Warren, I., Zuluaga, J.I., Robbins, C.H., Wulftange, W.H., Simmons, S.F., 2004. Geology and
Simmons, S.F., 2002. The paleosurface and paleowater table and their relationship to geochemistry of epithermal Au–Ag mineralization in the El Peñón district, northern
epithermal mineralization. Soc. Econ. Geol. Newsl. 51, 28–29. Chile. In: Sillitoe, R.H., Perello, J., Vidal, C.E. (Eds.), Andean Metallogeny: New Discov-
Simmons, S.F., Brown, K.L., 2006. Gold in magmatic hydrothermal solutions and the rapid eries, Concepts, and Updates. Society of Economic Geologists Special Publication 11,
formation of a giant ore deposit. Science 314, 288–291. pp. 113–139.
Simmons, S.F., White, N.C., John, D.A., 2005. Geological characteristics of epithermal Warren, I., Simmons, S.F., Mauk, J.L., 2007. Whole-rock geochemical techniques for evalu-
precious and base metal deposits. In: Hedenquist, J.W., Thompson, J.F.H., Goldfarb, ating hydrothermal alteration, mass changes, and compositional gradients associated
R.J., Richards, J.P. (Eds.), Economic Geology; one hundredth anniversary volume, with epithermal Au–Ag mineralization. Econ. Geol. 102 (5), 923–948.
1905–2005. Society of Economic Geologists, pp. 455–522. Winocur, D.A., Litvak, V.D., Ramos, V.A., 2014. Magmatic and tectonic evolution of the
Skinner, S.M., Clayton, R.W., 2013. The lack of correlation between flat slabs and bathy- Oligocene Valle del Cura basin, Main Andes of Argentina and Chile: evidence for
metric impactors in South America. Earth Planet. Sci. Lett. 371, 1–5. generalized extension. Geol. Soc. Lond. Spec. Publ. 399. http://dx.doi.org/10.1144/
Spikings, R.A., Winkler, W., Seward, D., Handler, R., 2001. Along-strike variations in the SP399.2.
thermal and tectonic response of the continental Ecuadorian Andes to the collision Witzke, T., Kolitsch, U., Krause, W., Wiechowski, A., Medenbach, O., Kampf, A.R., Steele, I.M.,
with heterogeneous oceanic crust. Earth Planet. Sci. Lett. 186, 57–73. Favreau, G., 2006. Guanacoite, Cu2Mg2 (Mg0.5Cu0.5)(OH)4 (H2O)4(AsO4), a new arse-
Steinmann, M., Hungerbuehler, D., Seward, D., Winkler, W., 1999. Neogene tectonic evo- nate mineral species from the El Guanaco Mine, near Taltal, Chile; description and
lution and exhumation of the southern Ecuadorian Andes; a combined stratigraphy crystal structure. Eur. J. Mineral. 18, 813–821.
and fission-track approach. Tectonophysics 307, 255–276. Yáñez, G.A., Ranero, C.R., Huene, R., Díaz, J., 2001. Magnetic anomaly interpretation across
Stoffregen, R.E., 1987. Genesis of acid–sulfate alteration and Au–Cu–Ag mineralization at the southern central Andes (32–34 S): The role of the Juan Fernández Ridge in the
Summitville, Colorado. Econ. Geol. 82, 1575–1591. late Tertiary evolution of the margin. J. Geophys. Res. Solid Earth 106, 6325–6345
Strusievicz, O.R., Clark, A.H., Lee, J.K.L., Farrar, E., Slauenwhite, M., Hodgson, C.J., 2000. (1978–2012).
Metallogenetic relationships of the Huaraz, Ancash, segment of the precious-base Zappettini, E., 2008. Geología de frontera Argentina–Chile. http://sig.segemar.gov.ar/
metal sub-province of northern Peru. Abstr. Programs Geol. Soc. Am. 32, 504. (accessed March 2014).

You might also like