1.igneous Geochronology

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

The time-space distribution of Eocene to Miocene magmatism in the Central


Peruvian high plain and its metallogenetic implications

Introduction
Geologists have long tried to recognize patterns in the temporal and spatial distribution of ore
deposits and thus identify metallogenetic belts where exploration efforts promise high rates of
success. Identifying factors that lead to the observed distribution of mineral resources continues
to be important as the ever-changing focus of exploration, new discoveries and new tools for
exploration and science force geologists to adapt their ideas and concepts. Thus, within this
context, reliable geochronologic data are an important component for understanding of
metallogenetic relationships.
In this chapter we present an extensive 40Ar/39Ar geochronologic database, which contains 44
new dates in addition to 30 previously dated intrusive and volcanic rocks of the Central Peruvian
Cordillera Occidental and adjacent areas to the east (Fig. 1). These data are complemented by
four additional U/Pb zircon ages from the transect; an additional three more U-Pb ages on zircon
were determined for Antamina, and are referred to herein.

The study region hosts numerous important polymetallic deposits of different types which
have been assigned to two roughly orogen parallel NW striking belts of approximately 700 km
strike length and up to 70 km width overall (e.g., Noble and McKee, 1999; Petersen, 1965). The
central portion of these belts between ~10.2 and 12.5° Lat. S is the focus of this study (Fig. 1)
and contains predominantly carbonate rock-hosted deposits. A direct spatial and, by inference,
genetic relationship of the mineralization to shallow level intrusions can be demonstrated or at

1
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

least confidently postulated for the majority of the region’s deposits. However, numerous other
shallow level intrusions lack evidence for associated economic metal enrichment, despite the
similar carbonate and siliciclastic sedimentary host-rock units throughout the central Peruvian
cordillera. We discuss the spatial and temporal distribution of magmatism and its metallogenetic
potential in a more restricted area and thus less generalized way than the review of Noble and
McKee (1999).
Published K-Ar geochronological studies (Noble and McKee, 1999; Soler and Bonhomme,
1988a,b) on the Cerro de Pasco-Churín transect (approximately 10.4-11° lat. S) contained in the
northern parts of our study area have identified a previously unrecognized Oligocene
metallogenetic episode comprising the Milpo and Atacocha skarn systems. Noble and McKee
(1999) and Noble et al. (2004) compiled geochronological data spanning an area from the
Castrovirreina district (13.5° S) to the wider Yanacocha area (6° S) and found that the overall
ages of the polymetallic deposits range from late Eocene to late Miocene, but the bulk of the
mineralization was emplaced in the middle and late Miocene. However, the available
geochronological data were largely obtained by the K-Ar method and in many cases not mineral
separates, but whole-rock samples were dated. Furthermore, as also stated by Noble and McKee
(1999, p. 176), the number of dated intrusions and mineralized deposits in the Central Peruvian
metallogenetic belts remains scarce, compared to the large number of igneous rocks and
polymetallic prospects still undated. This is particularly applicable to the areas between Domo
de Yauli and the Huacravilca intrusion (herein referred to as “south of Domo de Yauli”; Figs. 1,
2), but also applies to numerous intrusions and domes between Cerro de Pasco and La Oroya
(herein referred to as “north of Domo de Yauli”; Figs. 1, 3). Domo de Yauli is situated on an
important ENE striking cross-strike structural discontinuity (Benavides, 1999; Love et al., 2004)
and is therefore considered an appropriate subdivision of the area discussed herein. We focused
on the areas from the continental divide to approximately 60 km east of it, but did not study and
therefore do not discuss in detail the main Cenozoic volcanic arcs located to the west of the
Cordillera Occidental. Many of our new data cover the area south of Domo de Yauli (Fig. 2) and
correspond to intrusions where no previously published age constraints are available. New data
were also obtained for a number of intrusions north of Domo de Yauli (Fig. 3), which are
complemented by published data. In addition, we confirmed and refined the age constraints for
the previously dated Milpo-Atacocha, Chungar and Uchucchacua districts as well as those for
some apparently unmineralized domes. Age constraints are now available for the majority of
late Eocene to late Miocene intrusions in the segment between about 10.2° and 12° Lat. S.

The carbonate rock hosted polymetallic ore deposits of Central Perú


Central Perú has a long history of mining, spanning from pre-Incaic cultures to the present
day. Place-names such as Antamina (anta = copper) or Colquijirca (colqui = silver) indicate that
copper and silver have been produced in these districts already in pre colonial times. The types
of deposits in the study area range from relatively deep-seated skarn (Milpo-Atacocha: Gunnesch
and Baumann, 1984; Johnson, 1955; Soler, 1986) to shallow cordilleran base metal lode and
high-sulfidation replacement deposits (e.g. Colquijirca and Cerro de Pasco: Baumgartner et al.,
2003; Bendezú et al., 2003; Vidal and Ligarda, 2004). Besides these carbonate rock-hosted
deposits, epithermal mineralization hosted by syngenetic volcanic rocks is known from
Carhuacayán and Quicay (Fig. 3), whereas porphyry related Cu mineralization is known at
Morococha (Fig., 2). For some deposits it is unclear to which intrusive rocks they are associated,
as in the case of the vein-hosted Ag (-Mn-Pb-Zn) deposit of Uchucchacua (Bussell et al., 1990;

2
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Escalante et al., 2004; Petersen et al., 2004) or the Santander skarn deposit (e.g. Zimmerninck,
1983). Nevertheless, for most deposits it can be readily discerned which of the intrusive rocks
are genetically or at least spatially and temporally related to the mineralization. It should be
noted, however, that a syngenetic or diagenetic origin contemporaneous with deposition of the

3
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Jurassic or Cretaceous sedimentary rocks have been proposed for a number of deposits. Thus,
Rivera (2002) suggests such an origin for early stages of the mineralization at Cerro de Pasco,
whereas Gunnesch and Baumann (1984) did not exclude a syngenetic origin for the ores at
Atacocha and Milpo. South of Domo de Yauli, Cedillo and Tejada (1988) proposed that the
stratiform lead-zinc deposit of Cercapuquio (Fig. 2) is syngenetic with the Upper Jurassic strata
hosting the ore. Considering our own field observations, regional relationships and more recent
studies (e.g., Baumgartner et al., 2003), we believe, however, that a syngenetic origin for
mineralization in these districts is unlikely. Stratiform and stratabound ores in the San Cristóbal
district (Beuchat, 2003) or at Azulcocha (Muñoz, 1994) are assumed to be distal features of
intrusion related magmatism, although in the case of Cercapuquio it is not clear to which
intrusion the mineralization would be associated. At Milpo, field observations of alteration
assemblages within and outside the associated intrusion leave little doubt that it represents an

4
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Table 1. Ore deposits and major prospects of the region of interest herein. Deposits in italics
represent abandoned but historically producing deposits or currently subeconomic prospects.
Deposit or district Type Selected References
Uchucchacua Ag-Mn-Pb-Zn vein, Bussell et al., 1990; Petersen
carbonate rock et al., 2004; Escalante et al.,
replacement and skarn 2004
Colquijirca Base metal lode and Bendezú et al., 2003; Vidal
carbonate rock and Ligarda, 2004
replacement high-
sulfidation Au-Ag.
Milpo-Atacocha Zn-Pb-Cu (-Au) Skarn Johnson, 1955 ; Gunnesch
hosted by Jurassic and Baumann, 1984
limestone
Cerro de Pasco Base metal carbonate-rock Baumgartner et al., 2003;
replacement Baumgartner et al., 2004
Chungar Polymetallic Skarn Soler and Bonhomme, 1988a
Huarón Zn-Pb-Cu-Ag veins Thouvenin, 1983
Carhuacayán Epithermal Au-Ag-Zn-Pb Noble and McKee, 1999
veins and stockwork
Santander Polymetallic skarn without Zimmerninck, 1985
direct association with
intrusion
Yauliyacu-Casapalca ?
Morococha Porphyry Cu and Eyzaguirre et al., 1975;
associated base-metal Beuchat, 2003
veins
Domo de Yauli W–Cu–Zn–Pb veins, Beuchat et al., 2003
(San Cristóbal) hosted largely by
paleozoic volcanic rocks
and metapelites

Rey Salomon Cu-skarn ?


Mario Epithermal (?) ?
Azulcocha-Chuquipita-Janunhuasi Zn-Cu-Mn (-Au) skarn and Muñoz, 1994
manto type mineralization
Yauricocha Pb-Zn skarn and carbonate Alvarez et al., 1988; Jurado
replacement deposit. et al., 2004.
Carlin-type (?) Au-Ag
veins (Purisima
Concepción), distal
polymetallic veins.
Cercapuquio Stratiform Pb-Zn Cedillo and Tejada, 1988
mineralization

5
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

intrusion-related skarn system. For a comprehensive summary and collection of references we


refer to Noble and McKee (1999), but the major mineral deposits of the area discussed herein are
summarized in Table 1 (see also Figs. 2, 3).

Samples and analytical procedures


Igneous rocks have been sampled throughout central Perú. A range of rock compositions
that represent the complete geographic range of the Neogene igneous province east of the main
volcanic arcs were sampled. Regional geological maps published by INGEMETT at a scale of
1:100,000 were used as a field guide. Both apparently barren and mineralized intrusions were
sampled. The freshest possible specimens from the Neogene igneous rocks were taken. After
examination under a petrographic microscope, 45 samples from 42 intrusive bodies or volcanic
domes were chosen for Ar-Ar geochronology. Among these, a statistically significant plateau
age was obtained for 41 samples, and all except two of the rocks dated are of Cenozoic age
(Tables 2 and 3). The Ar-Ar database is complemented by seven new U/Pb zircon analyses. See
the appendices for analytical data.

Ar-Ar geochronology
After crushing the rocks with a steel mortar, approximately 10 mg of biotite and hornblende
of grain sizes between 0.25 to 0.5 mm were handpicked and subsequently washed in ethanol.
Whole-rock samples were chosen in only one sample (2PYB524: fine-grained basalt) whereas
plagioclase and nepheline were dated from one sample each. A hand magnet was passed over the
samples to remove magnetic minerals and metallic crusher fragments. The samples were washed
in deionized water, rinsed and then air-dried at room temperature. The samples were wrapped in
aluminum foil with similar-aged samples and with neutron flux monitors (Fish Canyon Tuff
sanidine, 28.02 Ma (Renne et al., 1998)). The samples were irradiated at the McMaster Nuclear
Reactor in Hamilton, Ontario, for 44 MWH, with a neutron flux of approximately 3x1016
neutrons/cm2. Analyses (n=54) of 18 neutron flux monitor positions produced uncertainties of
<0.5% in the J value.
The samples were analyzed at the Noble Gas Laboratory, Pacific Centre for Isotopic and
Geochemical Research, University of British Columbia, Vancouver, BC, Canada. The separates
were step-heated at incrementally higher powers in the defocused beam of a 10W CO2 laser
(New Wave Research MIR10) until fused. The gas evolved from each step was analyzed by a
VG5400 mass spectrometer equipped with an ion-counting electron multiplier. All
measurements were corrected for total system blank, mass spectrometer sensitivity, mass
discrimination, radioactive decay during and subsequent to irradiation, as well as interfering Ar
from atmospheric contamination and the irradiation of Ca, Cl and K (Isotope production ratios:
(40Ar/39Ar)K=0.0302, (37Ar/39Ar)Ca=1416.4306, (36Ar/39Ar)Ca=0.3952, Ca/K=1.83
(37ArCa/39ArK). The plateau and correlation ages were calculated using Isoplot ver.3.09
(Ludwig, 2003). Errors are quoted at the 2-sigma (95% confidence) level and are propagated
from all sources except mass spectrometer sensitivity and age of the flux monitor.

U/Pb geochronology
Zircon was separated from rock samples using conventional crushing, grinding, and Wilfley
table techniques, followed by final concentration using heavy liquids and magnetic separation.
Mineral fractions for analysis were selected based on grain morphology, quality, size and

6
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Table 2. New Ar-Ar age constraints from the region between Domo de Yauli and Huacravilca. See Fig. 2 for locations. Coordinates are given
in UTM zone 18, PSAD 56. Full dataset is presented in the appendix.
Assoc. Plateau Correlation Integrated Comment
39
Sample Rock /mineral Location Coord.-UTM min. Plateau age Ar/steps age age
Cretaceous
2PYB529 alkali gabbro/neph Chicchce 470.944/8640.569 115.32 ± 0.55 74.2%/ 10 of 15 steps 116.7 ± 1.8 114.54 ± 0.6
Eocene
Cerro 40.3%/ 6 of 16 steps Ar loss
2PCB600 Rhyolite sill/bi Maraypaquina 412.554/8689.785 40.14 ± 0.61 38.3 ± 2.65 31.5 ± 1.62
2PYB518 granodiorite/bi Huacravilca 456.672/8622.151 39.34 ± 0.28 90.5%/ 7 of 11 steps 39.43 ± 0.85 39.25 ± 0.31
Oligocene
2PYB540 porph.diorite/hbl Near Chaucha 429.611/8652.552 33.09 ± 0.43 63.7%/ 6 of 11 steps 33.03 ± 1.22 31.87 ± 0.54
2PYB532 dacite/bi Yanacancha 451.578/8654.256 32.01 ± 0.20 95.5%/ 8 of 10 steps 32.13 ± 0.34 31.9 ± 0.22
2PYB531 andesite flow/hb W of Yanacancha 454.903/8650.156 31.2 ± 0.5 100%/ 8 steps 31.55 ± 0.73 31.2 ± 0.5
Miocene (Aquitanian and Burdigalian)
2PCB594 gt-bearing rhyolite/bi SW of Canchayllo 424.055/8693.006 18.42 ± 0.15 31.8%/ 3 of 12 steps 18.19 ± 0.76 18.22 ± 0.13 Plag, qz incl.
2PYB539a Tonalite/bi W of Vitis 409.590/8648.150 17.02 ± 0.11 57.6%/ 5 of 12 steps 17.00 ± 0.24 16.8 ± 0.12 Minor Ar loss
2PYB539b Monzodiorite/bi W of Vitis 411.395/8647.743 16.66 ± 0.13 98.9%/ 10 of 11 steps 16.37 ± 0.34 16.54 ± 0.18
2PYB544 Granodiorite/bi Grán Bretaña 421.755/8665.461 16.20 ± 0.20 73.7%/ 6 of 12 steps 16.75 ± 0.24 15.89 ± 0.18
Miocene (Langhian and Seravallian)
2PCB608 gt-bearing dacite/bi Laguna Tunshu 395.335/8684.496 14.31 ± 0.10 59.5%/ 4 of 10 steps 14.46 ± 0.96 14.55 ± 0.18 Excess 40Ar ?
2PCT-56 diorite/bi Laguna Tunshu 395.818/8683.517 14.24 ± 0.09 42.6%/ 3 of 15 steps 13.99 ± 0.69 13.98 ± 0.36 Minor Ar loss?
2PCT-57 Granodiorite/bi Laguna Tunshu 395.491/8684.246 13.95 ± 0.12 93.1%/ 11 of 12 steps 13.47 ± 0.28 14.03 ± 0.57
2PCB607 Diorite/bi Laguna Tunshu 394.775/8685.187 13.67 ± 0.11 82.7%/ 7 of 10 steps 13.93 ± 0.11 13.45 ± 0.10
2PCB602 Diorite/bi Laguna Vicecocha 406.414/8686.610 13.67 ± 0.13 48.6%/ 3 of 7 steps 13.73 ± 1.64 13.45 ± 0.18
2PYB524 Basalt/WR Rio de la Virgen 464.155/8629.957 14.54 ± 0.49 35%/ 4 of 10 steps 15.5 ± 1.3 15.92 ± 0.42 inherited 40Ar
2PYB525 Dacite/bi Rio de la Virgen 463.998/8630.068 13.85 ± 0.15 69.7%/ 5 of 12 steps 14.26 ± 2.55 13.86 ± 0.14
2PCB609 Granodiorite/bi Cerro Portachuelo 398.103/8672.938 12.41 ± 0.13 55.5%/ 6 of 12 steps 12.41 ± 0.59 12.25 ± 0.13
± 70%/ 5 of 11 steps
2PCB611 Tonalite/bi Mina Rey Salomon 390.600/8679.614 Skarn 13.00 ± 0.13 13.07 ± 0.27 12.9 ± 0.11
Miocene (Tortonian and Messinian)
2PYB503 Qz-monzonite/bi Yauricocha 422.798/8637.969 Skarn N/A 10 steps 9 to 7Ma 6.21 ± 0.31 7.56 ± 0.09 Disturbed
2PYB505 Qz-monzonite/bi Yauricocha (Exito) 423.734/8635.257 Skarn 7.47 ± 0.06 78.8%/ 6 of 10 steps 7.3 ± 0.35 7.52 ± 0.10
2PYB512 Dacite dome/bi Huasicancha 470.768/8635.223 5.87 ± 0.08 57.4%/ 3 of 14 steps 5.59 ± 1.02 5.28 ± 0.08 inclusions
2PYB514 Dacite dome/bi Huasicancha 471.103/8634.445 5.40 ± 0.25 59.7%/ 4 of 14 steps 5.38 ± 1.52 5.51 ± 1.32 High atm.

7
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Table 3. New Ar-Ar age constraints from the region between Domo de Yauli and Cerro de Pasco. See Fig. 3 for locations.
Coordinates are given in UTM zone 18, PSAD 56. Full dataset is presented in the appendix.
Assoc. Plateau age Plateau Correlation Integrated Comment
39
Sample Rock /mineral Location Coord.-UTM Min (Ma) Ar/steps age (Ma) age (Ma)
Cretaceous
3PSB633 Hb diorite/hb Calhuacocha 330.545/8778.712 77.8 ± 1.3 95%/ 6 of 11 steps 81.2 ±3.1 77.22 ± 0.88
Eocene
2PPB562 Diorite stock /plag Cerro Señal Raco* 348.992/8808.030 34.6 ± 1.3 100%/ 7 of 7 steps 34.1 ± 3.0 35.21 ± 1.97
3PPB710 Andesite/hb Huangur (Quicay) 352.654/8819.271 33.5 ± 1.5 95.8%/ 8 of 10 steps 32.1 ± 1.9 34.58 ± 3.13
Early Oligocene (Rupelian)
3PYB715 Porph. diorite/hb Ticlio/Señal* 369.607/8715.577 31.6 ± 1.3 80.5%/ 5 of 8 steps 33.1 ± 2.9 25.92 ± 1.6
2PMB582 K-alt. diorite/bi Milpo stock* 368.305/8827.975 Skarn 29.59 ± 0.20 98.6%/ 14 of 16 steps 29.64 ± 0.26 29.69 ± 0.22
G.-diorite
2PMB585 porphyry/bi Socorro (Milpo)* 367.851/8829.003 N/A 31.02 ± 5.33 30.72 ± 0.72
Late Oligocene (Chattian)
3PUA558 Dacite porphyry/bi Uchucchacua* 316.643/8825.516 Skarn? 25.28 ± 0.44 98.1/11 of 13 steps 25.55 ± 0.72 25.16 ± 0.25
Porphyritic 41.2%/ 3 of 11 steps
2PUB560 dacite/bi E of Uchucchacua* 325.320/8826.184 24.49 ± 0.40 25.73 ± 0.89 24.25 ± 0.23
Early Miocene
2PTB612 Rhyolite/bi Mina Santa Sabina 408.160/7839.589 21.00 ± 0.21 47.7%/ 3 of 11 steps 22.37 ± 0.52 20.92 ± 0.16
2PTB613 Granite/bi Cerro Santa Ana 414.536/8743.007 21.04 ± 0.20 68.8%/ 5 of 12 steps 21.19 ± 0.32 20.59 ± 0.19
403.577/8751.765 61.8%/ 6 of 14 steps Min. 40Ar-
2PTB615 Dacite porphyry/bi Soccochuccho 20.54 ± 0.25 20.02 ± 2.29 20.31 ± 0.2 loss
Middle Miocene
3PSB636 Diorite/bi Marcapomacocha 355.490/8741.209 14.96 ± 0.30 98.4%/ 7 of 10 steps 15.20 ± 0.39 14.7 ± 0.21
3PSB624 Andesite porph./bi Huacracancha 348.407/8761.788 14.55 ± 0.26 99.9%/ 9 of 10 steps 14.37 ± 0.39 14.41 ± 0.19
Dacite porph. 85.8%/ 8 of 12 steps
2PCE249 dike/bi W of Colquijirca* 358.806/8808.684 14.13 ± 0.24 14.41 ± 0.4 13.89 ± 0.35
2PUE246 Dacite porphyry/bi Iskaycruz 308.308/8813.241 Skarn? 13.49 ± 0.30 91.8%/ 8 of 10 steps 13.53 ± 0.35 13.20 ± 0.20
3PSB631 Granite/bi Chungar* 332.366/8770.240 Skarn? 12.88 ± 0.36 38.7%/ 5 of 10 steps 13.11 ± 0.64 13.64 ± 0.22 ± excess Ar
Late Miocene
3PYB724 Monzonite/bi Señal Carrizal 356.599/8709.914 10.92 ± 0.4 90.5%/ 6 of 9 steps 11.16 ± 0.98 10.92 ± 0.23
3PSB630 Granodiorite/bi Calhuacocha* 330.262/8778.340 9.74 ± 0.24 99.16%/ 7of 8 steps 9.83 ± 0.26 9.34 ± 0.22
3PSB617 Dacite dome/bi Carhuacayán* 363.459/8760.442 Epith.? 8.2 ± 0.18 99.9%/ 12 of 13 steps 8.18 ± 0.23 8.11 ± 0.14
2PUB553 Rhyolite dike/bi NE Uchucchacua 318.360/8828.950 Skarn? N/A Most steps near 7Ma 7.3 ± 0.47 7.98 ± 0.11 Age ~ 7 Ma
2PPB589 Dacite porphyry/bi Alpamarca 333.743/8806.494 Manto 5.75 ± 0.09 44.1%/ 4 of 10 steps 5.53 ± 0.47 6.07 ± 0.1 ~ 6-7 Ma

8
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

magnetic susceptibility. All zircon fractions were abraded, using the technique of Krogh (1982),
prior to dissolution to minimize the effects of post-crystallization Pb-loss. All mineral
separations, geochemical separations and mass spectrometry were done in the Pacific Centre for
Isotopic and Geochemical Research in the Department of Earth and Ocean Sciences, University
of British Columbia. Samples were dissolved in concentrated HF and HNO3 in the presence of a
mixed 233-235U-205Pb tracer. Separation and purification of Pb and U employed ion exchange
column techniques modified slightly from those described by Parrish et al. (1987). Pb and U for
each fraction were eluted sequentially into a PFA beaker and loaded together on a single Re
filament using a phosphoric acid-silica gel emitter. Isotopic ratios were measured using a
modified single collector VG-54R thermal ionization mass spectrometer equipped with a Daly
photomultiplier. Measurements were done in peak-switching mode on the Daly detector. U and
Pb analytical blanks were in the range of 1 pg and 1-3 pg, respectively, during the course of this
study. U fractionation was determined directly on individual runs using the 233-235U tracer, and
Pb isotopic ratios were corrected for a fractionation of 0.32%/amu for Faraday and Daly runs,
respectively, based on replicate analyses of the NBS-982 Pb standard. All analytical errors were
numerically propagated through the entire age calculation using the technique of Roddick
(1987). Concordia intercept ages and associated errors were calculated using a modified version
the York-II regression model (wherein the York-II errors are multiplied by the MSWD) and the
algorithm of Ludwig (1980). All age errors are quoted at 2s (Ma) level and errors for isotopic
ratios are quoted at 1s (%).

Geologic evolution of Central Perú


A comprehensive review of the geologic and tectonic evolution of central Perú was presented
by Benavides (1999) and much of the following description is summarized from there.
Geological features related to the Andean geodynamic cycle dominate the central Peruvian
Cordillera Occidental and adjacent areas. The Andean cycle is characterized by a subduction
setting fundamentally similar to the present day, that is the oceanic Nazca Plate and predecessor,
the Farallón plate, subducting beneath the South American continent. Subduction initiated with
the onset of spreading of the mid-Atlantic ridge in the Triassic and persists to the present. A
Mariana-type subduction with an extensional back arc domain located east of an island arc
dominated the Mesozoic. Shallow marine sedimentary sequences, composed of limestone and
lesser siliciclastic rocks, dominate the stratigraphy of the units deposited in the back arc basin
during this time. An alkali-gabbro intruding Lower Cretaceous rocks of the Goyllarisquizga
group and Chulec-Pariatambo Formation was dated in this study at 115.3 ± 0.6 Ma (nepheline
plateau age: Table 2). This date provides a minimum age for these largely siliciclastic rocks, but
also indicates the presence of an extensional back-arc regime up to at least the Aptian. The
acceleration of spreading in the South Atlantic soon thereafter led to an increase in the rate of
convergence. Based on evidence from the magmatic rock record, the character of arc magmatism
changed from voluminous volcanism with the products partially deposited in a subaqueous
setting (Casma Group) up to the Early Cretaceous, to plutonism represented by the 100-60 Ma
coastal batholith (Cobbing, 1973). It is worth mentioning, however, that there are numerous
basaltic dikes, sills and laccoliths intruding Casapalca Formation red beds in the area of Chungar.
These basalts are considered calc-alkaline as they are hornblende and plagioclase phyric and lack
petrographic evidence for alkaline character. Hornblende from one of these rocks yielded a
plateau age of 77.8 ± 1.3 Ma (Table 3), which coincides in age with the coastal batholith.

9
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Up to seven Paleocene and early Pliocene contractional deformation events have been
proposed to be responsible for crustal thickening and uplift in the Andes (Benavides, 1999;
Sandeman et al., 1995; Sébrier and Soler, 1991). The late Paleocene Incaic I (59-55 Ma)
orogenic pulse formed a fold and thrust belt with the most intense deformation between the rigid
block represented by the coastal batholith and the Marañón arch tectonic high constituting the
western border of the South American continent at that time. A subsequent Incaic II (~42 Ma)
deformation event overprinted the previous fold and thrust belt; the current Cordillera Occidental
represents the backbone of the Incaic I and II orogen. Most of the folds and thrust faults in the
study area have their origin in the Incaic deformation pulses. Deformation, subsequent uplift and
erosion were most intense in the Cordillera Occidental and decreased in intensity towards the
east.
Evidence for subsequent uplift and deformation events has been mainly described from
southern Perú (Sandeman et al., 1995; Sébrier and Soler, 1991; Tosdal et al., 1984). Sébrier and
Soler (1991) described five short lived deformation pulses, Inca III and Quechua I-IV at 26, 17,
10, 7 and 2 Ma, separated by periods of relative tectonic quiescence. Important mid-Miocene
deformation has been constrained to between ~14.5 and 5.2 Ma by Farrar and Noble (1976) for
areas south of Cerro de Pasco.

The Paleogene and Neogene volcanic arcs of Central Perú


The timing of the Cenozoic arc volcanism of central Perú is still relatively poorly understood,
due to the general lack of detailed stratigraphic mapping supported by extensive geochronology
within the volcanic sequences to the west of the continental divide. The current understanding of
the Cenozoic arc volcanism of central Perú points to a concentration of volcanic activity in times
of reduced compressional stress. However, much of the detailed work on the Cenozoic volcanic
arcs in Perú has been carried out in northern (e.g. Noble et al., 1990) or southern Perú (e.g.,
Noble et al., 1979; Sandeman et al., 1995; Sébrier and Soler, 1991) and is extrapolated into the
areas of central Perú. The largest volumes of volcanic rocks were deposited from the Cordillera
Occidental to the west and thus west of the area discussed herein, although notable exceptions
are mapped (e.g., the early Oligocene Astabamba Formation near Hunacayo: (Mégard, 1968; see
below).
Volcanic activity in the western cordillera of central Perú, and areas adjacent to the west of
the study area, is represented by the Calipuy Group volcanic rocks, a widespread sequence that
covers the post Incaic II erosional surface in the Cordillera Negra west of the Cordillera Blanca
and was assigned a late Eocene age (Noble et al., 1999a). However, significant portions of the
volcanic sequence traditionally assigned to the Calipuy Group have been subsequently assigned
to a newly defined Huaráz Group of upper Oligocene to middle Miocene age (Strusievicz, 2000).
The Calipuy and Huaráz Group volcanic rocks are included in the Calipuy Supergroup of Love
et al. (2004).
The available information indicates that voluminous arc-volcanism may have lasted until the
middle Miocene in central Perú, but appears to have been concentrated west of the continental
divide. As pointed out below, volcanic rocks traditionally assigned to the Calipuy group, now
assigned to the late Oligocene, also crop out within the study area.

10
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Timing of magmatism south of Domo de Yauli


The segment between Domo de Yauli (Lat., 11.68° S; Long., 76.05° W) and Huacravilca
(Lat., 12.48° S; Long., 75.41° W; Fig. 2) contains, similar to the northern segment (see below),
isolated shallow intrusive centers and volcanic domes. However, only one polymetallic deposit,
Yauricocha was in production (2005), in marked contrast with numerous producing and historic
mines from Domo de Yauli north to Antamina. In contrast to the area north of Domo de Yauli,
where a number of age constraints have been available prior to this study, only one K-Ar age, for
the Yauricocha stock (7.5 Ma; K-Ar biotite: Giletti and Day, 1968; recalculated: Noble and
McKee, 1999) has previously been published for the southern segment of the study area. Our
new data are summarized in Table 2 and Fig. 2.

Late Eocene and early Oligocene intrusive activity


The central portion of this transect contains a number of late Eocene, and thus older than
expected, intrusive centers. The Huacravilca granodiorite intrusion yielded an undisturbed biotite
plateau age of 39.34 ± 0.28 Ma. This intrusion features only minor skarn alteration with some
magnetite at its margin. Seventy kilometers farther northwest, a coarse porphyritic rhyolite sill
intruding strata of the Casapalca Formation (location: Cerro Maraypaquina) yielded a biotite
plateau age of 40.14 ± 0.61 Ma. The argon release pattern indicates argon loss in the low-T
heating steps, but the high-T plateau age is considered reliable. A similar, slightly argillically
altered rhyolite sill cropping out near a small and undated prospect (Mario project) some 60 km
SE may be assigned to the same intrusive phase. No conclusive evidence for mineralization
associated to Eocene intrusions is known. However, the stratiform Pb-Zn deposit of Cercapuquio
is located approximately 5 km north of the Huacravilca intrusion and, thus may represent distal
associated mineralization. All early to mid-Eocene intrusive rocks are calc-alkaline and dacitic
to rhyolitic (SiO2 = 68 to 72 wt.%) in composition, but lack petrographic evidence for
peraluminous character.
In the late Eocene to early Oligocene, the style and composition of igneous activity changed
with the intrusion of diorite stocks near Yauricocha. The largest of these intrusions is a coarse-
grained hornblende and plagioclase porphyritic diorite of Cerro Caja Real (Chaucha), which
yielded a hornblende plateau age of 33.09 ± 0.43 Ma, whereas a smaller diorite intrusion
approximately 8 km farther west was dated by U/Pb on zircon and yielded a concordant age of
36.07 ± 0.05 Ma. In the earliest Oligocene, comparatively large volumes of andesitic to dacitic
lava flows, ignimbrites and tuffs of the Astabamba Formation erupted approximately 20 km east
from the late Eocene intrusive centers mentioned above. Two samples from this volcanic unit
were dated: hornblende from an andesite flow yielded an age of 31.2 +/- 0.5 Ma whereas a dacite
dome yielded a biotite plateau age of 32.01 +/- 0.2 Ma. Both samples yielded undisturbed age
spectra and sensibly identical isotope correlation ages.
Evidence for significant polymetallic mineralization associated with either the Astabamba
Formation or the diorite stocks is lacking.

Miocene magmatism
The late Oligocene and earliest Miocene are characterized by an apparent volcanic lull, since
no volcanic or intrusive rocks were dated between 31 and 18.5 Ma. Magmatic activity resumed
with the intrusion of rhyolite sills, domes and cryptodomes a few kilometers south of the town of
Canchayllo (Fig. 2, 4). Biotite phenocrysts from these rhyolites yielded a plateau age of 18.42 ±
0.15 Ma, the overall age spectrum of this sample is somewhat disturbed, possibly due to

11
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

inclusions of other minerals, but all heating steps yielded ages between 19 and 17 Ma and the
plateau age, even though it only contains 32 % of the 39Ar released in three steps is considered
meaningful. The rhyolites at Canchayllo contain magmatic garnet which is interpreted as
evidence for peraluminous character and considerable crustal contamination of the magma.
Igneous activity after emplacement of the rhyolites at Canchayllo shifted west, as
demonstrated by the large tonalite and granodiorite intrusions near the town of Vitis, about 12
km NW of Yauricocha. Two reliable biotite plateau ages were obtained from this intrusive
complex: 17.02 ± 0.11 Ma from a tonalite and 16.66 ± 0.13 Ma from a monzodiorite sample.
The large intrusions generated a contact metamorphic halo of considerable size, consisting
mostly of grey marbles. No polymetallic mineralization has been observed associated with these
intrusions.

Intrusive centers younger than 17 Ma are mostly distributed in the western portions of the
area. Age constraints have been obtained from the Chuquipita intrusion (biotite: 16.20 +/- 0.20
Ma) where modest historic mine workings at Chuquipita, Jatunhuasi and Azulcocha (that is,
Gran Bretaña) exploited stratabound and skarn Mn-Zn (-As, Au, Cu) (Muñoz, 1994). Along
strike of the main Andean structural grain, about 20 to 30 km northwest near Cerro Tunshu,
seven intrusive rocks were dated and all yielded ages between 14.31 ± 0.10 Ma and 12.41 ± 0.13
Ma. The rocks range from dioritic to granodioritic in composition and one of these samples,
taken from a granodiorite porphyry dike contains magmatic garnet indicating peraluminous
composition. Calc-silicate alteration, albeit without evidence for skarn, is widespread. However,
at Mina Rey Salomon significant, but subeconomic copper skarn lies along the contact of a large
intrusive complex. Tonalite from this complex yielded a reliable biotite plateau age of 13.0 ±
0.13 Ma. Granodiorite from the same intrusive complex was dated at 12.41 ± 0.13 Ma and
represents the youngest intrusive rock of the wider Cerro Tunshu area.
Basalt in spatial association with dacite domes has been observed near Río de la Virgen in
the southeastern corner of the study area (Figs. 2, 5). An age of 14.54 ± 0.49 Ma was obtained
for the basalt. The date corresponds to a whole rock analysis which exhibits evidence for excess
argon, and should therefore be considered a maximum estimate. To complement the whole-rock

12
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

analysis, a plagioclase separate was dated but yielded only a poor plateau age of 17.9 ± 4.6 Ma.
One dacite dome spatially associated to the basalt was dated at 13.85 ± 0.15 Ma (biotite plateau)
and may be considered contemporaneous with the basalt, leading to a hypothesis that the
volcanism of this area may be bimodal.

Magmatism resumed in the late Miocene with the emplacement of the Yauricocha and coeval
Exito quartz-monzonite to granodiorite intrusions in the Cordillera Occidental. An undisturbed
7.47 ± 0.06 Ma biotite age was obtained from the Exito intrusion, whereas a sample from the
equivalent Yauricocha intrusion yielded a total gas age of 7.56 ± 0.09 Ma, but no interpretable
plateau. Both ages agree with the published K-Ar biotite age of 7.5 Ma (Giletti and Day, 1968;
recalculated by Noble and McKee 1999).
The youngest currently known magmatic event in the region is represented by dacite domes
of the Herú Formation (Mégard, 1968), which extruded at the eastern limit of the study area near
Chongos Alto. Two samples from different domes were analyzed and yielded somewhat
disturbed age spectra, probably due to inclusions in the magmatic biotite. Nevertheless,
statistically significant biotite plateau ages of 5.87 ± 0.08 Ma and 5.4 ± 0.25 Ma were obtained,
indicating the cessation of volcanism in the latest Miocene.

Timing of magmatism north of Domo de Yauli


Owing to the relatively large number of polymetallic deposits and prospects of the area north
of Domo de Yauli, many of the intrusive and volcanic rocks have been dated in previous studies
(summarized by Noble and McKee 1999; Table 4). The evolution of the magmatic arc is
therefore well established.

Late Eocene and early Oligocene intrusive activity


A number of domes of dacitic and andesitic composition underlie prominent topographic
features rising above the flat plains to the west of Cerro de Pasco. The oldest magmatic activity
was dated indirectly at Quicay where hypogene alunite yielded a K-Ar age of 37.5 Ma (A.
Alvarez and D.C. Noble, unpublished age commented in Noble and McKee, 1999). A mid-
Eocene K-Ar whole-rock age of 38.5 ± 1 Ma is also reported by Soler and Bonhomme (1988b)
for the nearby quartz-monzonite stock at Huangoc; this rock appears unrelated to mineralization.
During this study, we obtained two late Eocene ages for unmineralized domes. A 40Ar-39Ar
plagioclase plateau age of 34.6 ± 1.3 Ma was determined from a xenolith-rich diorite stock of
Cerro Señal Raco. This stock is unaltered, and lacks hornblende and biotite phenocrysts. Soler

13
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Table 4. Published age constraints for intrusions and ore deposits in the study area

Rock/mineral Age (Ma) Method Reference Our Ar-Ar age/comment


Granodiorite porph./WR 15.2 ± 0.4 K-Ar Soler and Bonhomme, 1988a not dated in this study
Granite/biotite 13.33 ± 0.3 K-Ar Soler and Bonhomme, 1988a 12.88 ± 0.36/ difference explained by excess40Ar
Our total gas age within error, but disturbed Ar- release
Granodiorite porph./plagioclase 29.8 ± 1.4 K-Ar Soler and Bonhomme, 1988a pattern
Granodiorite porph./biotite 30.9 ± 0.4 K-Ar Soler and Bonhomme, 1988a not dated in this study
Granodiorite porph. /biotite 31.1 ± 0.4 K-Ar Soler and Bonhomme, 1988a not dated in this study
Granodiorite porph./WR 29.3 ± 0.5 K-Ar Soler and Bonhomme, 1988a not dated in this study
Granodiorite/biotite 10.0 ± 0.3 K-Ar Soler and Bonhomme, 1988a 9.74 ± 0.24/ ages within error of each other date
Diorite stock/WR 35.2 ± 0.1 K-Ar Soler and Bonhomme, 1988b 34.6 ± 1.3/ages within error
Quartz Monzonite/WR 38.5 ± 1.0 K-Ar Soler and Bonhomme, 1988b not dated in this study
40
Dacite dome/biotite 12.43 ± 0.06 Ar-39Ar Bendezú et al., 2003 not dated in this study
40
Au-Ag mineralization/alunite 11.6 to 11.3 Ar-39Ar Bendezú et al., 2003 not dated in this study
40
Base metal mineralization/alunite 10.8-10.5 Ar-39Ar Bendezú et al., 2003 not dated in this study
Diorite/biotite 14.58 ± 0.48 K-Ar Farrar and Noble 1976 not dated in this study
Dacite ignimbrite/biotite 5.2 ± 0.2 K-Ar Farrar and Noble 1976 not dated in this study
Hypogene alunite 37.5 K-Ar Alvarez, A. and Noble, D.C., unpub.
Biotite rhyolite dike 9.32 ± 0.57 K-Ar (Romani, 1982) not dated in this study
Quartz monzonite/biotite 7.5 K-Ar, recalc Giletti and Day, 1968 7.47 ± 0.06/ coincides well
Quartz monzonite/zircon 8.8-9.1 U-Pb Beuchat, 2003 not dated in this study
Quartz monzonite 8.2-8.3 K-Ar Eyzaguirre et al., 1975 not dated in this study
Granodiorite intrusion/zircon 6.6 +1/-3.6 U-Pb Beuchat, 2003 not dated in this study
Huayllay ignimbrite 5.4 ± 0.3 K-Ar Unpublished, Noble and McKee, 1999 not dated in this study
Pre-mineral dike/sanidine 24.5 K-Ar Noble and McKee, 1999 25.28 ± 0.44/considered equivalent

14
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

and Bonhomme (1988a) report a similar K-Ar whole-rock age of 35.2 ± 0.1 Ma. At Huangur,
approximately 5 km east of Quicay, a flow-textured andesite yielded a hornblende plateau age of
33.5 ± 1.5 Ma, which agrees within the 2s confidence interval of the age for the Cerro Señal
Raco diorite.
Magmatic activity in the early Oligocene migrated to the east. A number of intrusions have
been dated as early Oligocene in the wider Milpo-Atacocha district. Near the Milpo and
Atacocha mines, numerous small intrusive bodies of quartz, plagioclase and biotite porphyritic
granodiorite stocks intrude steeply dipping Jurassic limestone. However, most of these intrusive
bodies lack a significant alteration halo within the limestone, as exemplified by the Socorro
Stock approximately 500 m north of the Milpo skarn deposit (Fig. 6). Soler and Bonhomme
(1988a) report a 29.8 ± 1.4 Ma K-Ar plagioclase age for this intrusion. Biotite attempted during
this study revealed a disturbed argon release pattern and no interpretable Ar-Ar plateau age was
obtained. However, the total gas age of 30.72 ± 0.72 Ma lies within the 2 s confidence level of
the published K-Ar age. The skarn at Milpo is genetically related to an altered diorite to
granodiorite stock lacking quartz phenocrysts. A potassically-altered sample containing
magmatic biotite yielded an age of 29.59 ± 0.20 Ma (biotite plateau age, matching the correlation
age of 29.64 ± 0.26 Ma). Crystallization and potassic alteration of this rock is considered to be
closely related temporally. Skarn, thus, may have post-dated the barren granodiorite stocks of
the district.

Soler and Bonhomme (1988a, b) report a number of additional age constraints for intrusions
in the area. The Sunkullo and Mariac intrusions to the southeast of Milpo have ages of about 31
Ma and thus may be considered contemporaneous with the Socorro stock. The Atacocha-San
Gerardo stock in contrast yielded a slightly younger K-Ar whole rock age of 29.3 ± 0.5 Ma.
Only one sample from outside the Uchucchacua to Milpo transect yielded an Oligocene age.
It corresponds to a hornblende and plagioclase porphyritic diorite which intruded Casapalca
Formation red beds at Ticlio west of Domo de Yauli that yielded a hornblende plateau age of
31.6 ± 1.3 Ma (Fig. 2).

15
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Late Oligocene magmatic activity was restricted to the wider Uchucchacua area. A
potassically altered dacite porphyry from the lower mine levels yielded a biotite plateau date of
25.28 ± 0.44 Ma, which is similar to an Ar-Ar sanidine date of 24.5 Ma (unpublished age
reported by Noble and McKee, 1999). The latter corresponds to a dike and was considered to
immediately predate the mineralization by Noble and McKee (1999), but may actually
correspond to an early phase of skarn mineralization as suggested by our new dated sample.
Dacitic and andesitic flows and volcaniclastic rocks were deposited to the east of Uchucchacua
near the road to Cerro de Pasco. One sample from an autobrecciated dacitic flow was dated at
24.49 ± 0.40 Ma and thus coincides in age with those from the Uchucchacua mine. The volcanic
rocks east of Uchucchacua were assigned to the early Eocene Calipuy Group on the regional
maps (Cobbing, 1973), an assignation that may require re-evaluation due to the new age
constraints provided herein.
After an apparent lull of magmatism until 21 Ma, igneous activity resumed locally when a
few small intrusive and volcanic centers were emplaced 15 to 30 km SE of the town of Junín,
approximately 50 to 70 km southeast of the region where magmatic activity was concentrated
previously. Three latest Oligocene to earliest Miocene rocks have been dated during this study;
all ages are based on biotite plateaus. A small rhyolite dome at Santa Sabina near the road to
Tarma was dated at 21.00 ± 0.21 Ma A similar age of 21.04 ± 0.20 Ma was obtained from a
granite some 6 km to the northeast, and a slightly younger age of 20.54 ± 0.25 Ma represents the
emplacement of the Soccochuccho dacite porphyry, located 15 km southeast of Junín. No
known mineralized rock is associated with these igneous rocks except at Soccochuccho where
minor copper oxide mineralization has been previously mined.

Miocene Magmatic activity


After the short early Miocene magmatic episode, igneous activity ceased until about 16 Ma
when it resumed in isolated and scattered areas over the entire length of the segment north of
Domo de Yauli. Magmatism was concentrated along the Cordillera Occidental and extended up
to 50 km west of the latest Oligocene igneous activity. The major mineral districts of Colquijirca
and Cerro de Pasco are related to this middle Miocene magmatic activity. Igneous rocks dated
from within the areas contrast with other contemporaneous intrusions in that they are found
about 40-50 km east of the continental divide. Magmatic activity in the wider Cerro de Pasco
area initiated with the intrusion of the 15.2 ± 0.4 Ma (K-Ar, WR, Soler and Bonhomme, 1988a)
granodiorite intrusion at Yanamate, and a 14.13 ± 0.24 Ma (Ar-Ar biotite) dacite porphyry dike
west of Colquijirca. Dacite from a diatreme-dome complex at Colquijirca yielded biotite ages
between 12.7 and 12.4 Ma (Bendezú et al., 2003), which predated two pulses of mineralization
dated at 11.6-11.3 and 10.9-10.5 Ma (Bendezú et al., 2003, 2004). At Cerro de Pasco, Bendezú
et al (2004) report that the subvolcanic complex is broadly contemporaneous with that at
Colquijirca, despite a slightly older K-Ar age of about 14.5 Ma reported by Silberman and Noble
(1977). Lead-Zn ores were deposited between 10.9 and 10.5 Ma, in the same time range as those
at Colquijirca (Bendezú et al, 2004). Middle Miocene magmatism in the Uchucchacua area is
manifested by a rhyolite dike, which generated some skarn mineralization to the northeast of
Uchucchacua. A concordant U/Pb zircon age of 13.63 ± 0.11 Ma was obtained for this rhyolite
dike. A dacite porphyry located on the northern end of the Iskaycruz skarn deposit, which is
located west of the main axis of the Cordillera Occidental, yielded a similar age of 13.49 ± 0.30
Ma (biotite plateau); the relationship of this dome to the massive sulfide deposit is uncertain.
Small-mineralized centers, which have undergone historic production, are also located south of

16
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Cerro de Pasco. These include the Río Pallanga carbonate replacement deposit (14.6 ± 0.5 Ma,
Farrar and Noble, 1976: date corresponds to a K-Ar biotite age of a spatially associated dacitic
dome) and the Chungar skarn deposit (12.88 ± 0.36 Ma, biotite plateau age of a granite).
Some apparently barren mid-Miocene intrusions complement the data summarized above.
The large diorite intrusion at Marcapomacocha and a small diorite stock at Huacracancha west of
Carhuacayán yielded biotite plateau ages of 14.96 ± 0.30 Ma and 14.55 ± 0.26 Ma respectively.
A granodioritic intrusion at Ticlio was dated at 14.11 ± 0.04 Ma (U-Pb, zircon; Beuchat, 2003),
whereas west of Domo de Yauli, near Cerro Señal Carrizal northwest of the Rosaura mine, a
monzonite intrusion apparently unrelated to mineralization yielded an age of 10.92 ± 0.4 Ma.

Magmatism younger than 10 Ma generated a number of mineralized centers, most


importantly the Toromocho porphyry at Ticlio dated at 9.11 ± 0.1 Ma (U-Pb zircon, Beuchat,
2003). The historically mined epithermal deposit of Carhuacayán 15 km S of Huayllay was
emplaced at approximately 8 Ma as indicated by a 7.8 ± 0.3 Ma K-Ar alunite age (Noble and
McKee, 1999) and an associated dacite dome dated at 8.2 ± 0.18 Ma (biotite plateau) in this
study. A rhyolite dike northeast of Uchucchacua yielded a somewhat disturbed age spectrum
with a biotite isotope correlation age of 7.3 ± 0.47 Ma. This dike intruded red beds of the
Casapalca Formation, and is petrographically similar to a mid-Miocene feldspar phyric rhyolite
dike but only the latter generated some skarn mineralization (Fig. 7). One of these dikes has a
reported K-Ar age of 9.32 ± 0.57 Ma (Romani, 1982). The 6.8 ± 1 Ma Anamaray intrusion, north

17
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

of Uchucchacua, generated minor skarn. The Uchucchacua area, thus, hosts felsic igneous rocks
of late Oligocene, middle Miocene and late Miocene age, all of which have generated some
skarn type mineralization, but it still remains unclear, which of these events is responsible for the
bulk of the mineralization in the main Ag-Mn-Pb-Zn vein systems. The youngest intrusive
events with related mineralization include the dacite porphyry dome complex of Alpamarca
(5.75 ± 0.09 Ma biotite plateau), located 20 km due west of Colquijirca, the Chumpe intrusion at
San Cristóbal, Domo de Yauli (6.6 +1/-3.6 U-Pb zircon age: Beuchat, 2003; 5.4 ± 0.3 Ma
commercial K-Ar date reported by Noble and McKee 1999). Magmatism apparently ceased after
the eruption of the Huayllay ignimbrite, for which a K-Ar date of ~5.2 Ma was reported by
Farrar and Noble (1976).

Discussion
Magmatic activity was not uniformly distributed in time and space and important differences
and temporal (Fig. 8) in spatial (Fig. 9) distribution of volcanic and intrusive rocks between the
segments separated by Domo de Yauli are evident. Temporal variations in the magmatic activity
may be assigned to variations in plate convergence (e.g. Sébrier and Soler, 1991), but may also
be a function of changing subduction angle, as demonstrated for northern Chile (e.g. Bissig et al.,
2003; Kay and Mpodozis, 2001; Kay et al., 1999). Areas or trends across the orogen where
magmatism was concentrated over extended periods of time, on the other hand, may be
explained by structural weaknesses in the upper plate or inhomogeneities within the slab, such as
subducted fracture zones. For instance, an across the strike of the Incaic folds aligned belt of
intermitted intrusive and volcanic activity between Uchucchacua, Cerro de Pasco and Milpo-
Atacocha is conspicuous and may be assigned to a first-order crustal control. Unfortunately,
insufficient information on the evolution of Paleogene and Neogene volcanic arcs of Central
Perú is available to permit an exhaustive and comprehensive discussion of the geodynamic
evolution. Nevertheless, we present ideas and possible interpretations for the evolution of the
arc-magmatism based on the temporal and spatial distribution of the igneous activity east of the
main Neogene volcanic arc, but understand that our interpretations may have to be adapted once
more data become available.

Late Eocene and Oligocene magmatic activity


The oldest Cenozoic intrusions have been dated south of Domo de Yauli where isolated felsic
magmas intruded at ca. 40 Ma in a roughly orogen parallel belt approx 15 km east of the main
Cordillera Occidental continental divide (Fig. 9). Magmatic activity in the northern segment
initiated slightly later and only occurs in the high plains west of Cerro de Pasco. It is represented
by the Quicay dome and the stocks of Cerro Señal Raco, Huangoc and Huangur. The Eocene
magmatism of the Cerro de Pasco transect is spatially more restricted compared to the
magmatism south of Domo de Yauli.
Voluminous late Eocene volcanism is represented by the Calipuy Group in the Cordillera
Negra (Noble et al., 1999b). Calipuy Group volcanic rocks are mapped in extensive areas
adjacent to the Cordillera Occidental, however the only scarce age constraints for these volcanic
rocks have to be born in mind when interpreting the evolution of the magmatic arc. Nevertheless,
it is apparent that Eocene Calipuy Group volcanism has its counterpart to the east of the

18
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Cordillera Occidental as represented by domes (mostly west of Cerro de Pasco) and felsic
intrusions (south of Domo de Yauli).

19
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Oligocene magmatic activity


A temporary, albeit locally limited, increase of magma volumes produced in the early
Oligocene is indicated by several dacitic intrusions in the wider Milpo-Atacocha area, ca. 50-60
km east of the continental divide, and by the intrusion of diorite at Ticlio. Diorite intrusions at
Chaucha and the voluminous volcanic Astabamba Formation, in the southern segment of our
study area, also were emplaced in the late Eocene and early Oligocene. Compared to the early
and middle Eocene, magmatism south of Domo de Yauli appears to have shifted to more mafic
compositions in the early Oligocene.

Current data suggest a general middle Oligocene magmatic lull from approximately 29.3 to
25.5 Ma in Central Perú. In contrast to the areas north of the Domo de Yauli, there is no
evidence in the southern portion of the study area for igneous activity for a more extended period
between the early Oligocene (31 Ma) and the early Miocene (~18.5 Ma).
An early to middle Oligocene magmatic quiescence is also observed in southern Perú
(Sébrier and Soler, 1991), suggesting a regional geodynamic control on the reduction of magma

20
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

output. Indeed, paleomagnetic data indicate slow rates of convergence between the Nazca and
South American plates in the middle Oligocene (Pardo-Casas and Molnar, 1987), which may
explain the observed reduction of magma production in southern (Sébrier and Soler, 1991) and
central Perú.
Magmatism resumed to the east of Uchucchacua in the Late Oligocene, around 25 Ma. It
appears that this renewed magmatic activity was, again, restricted to the Uchucchacua - Cerro de
Pasco - Milpo transect, an interpretation, however, that may be an artifact of the scarcity of
geochronological constraints of volcanic rocks between Domo de Yauli and Uchucchacua. A
dramatic shift in the locus of magmatic activity is observed in the latest Oligocene. Small
volumes of felsic rocks dated around 21 Ma occur in a restricted area about 25 km NE of La
Oroya, or approximately 70 km SE of the transect where magmatism was active previously.
The renewed magmatic activity near Uchucchacua around 25 Ma may be attributed to a
change in direction of subduction of the Nazca plate from a northeasterly to a more easterly
motion and an increase of the rate of convergence (Pardo-Casas and Molnar, 1987; Pilger, 1981).
In the northern Chilean Andes, this reconfiguration resulted in a more orthogonal convergence
vector (Pilger, 1981) and in some regions a marked increase in erupted volumes of volcanic
rocks (e.g., Bissig et al., 2001). The same change of the direction in relative plate motion
resulted in an increase of the obliquity of the subduction beneath the central Peruvian Andes, a
convergent geometry, which could explain the generally reduced volcanism and the complete
lack thereof south of Domo de Yauli in the late Oligocene. However, the reduced magma output
in the late Oligocene and early Miocene and in particular the difference between the areas north
and south of Domo de Yauli may point more convincingly to a flat subduction configuration
which was more pronounced in the southern extents of the study area.

Early Miocene magmatic activity


Magmatism south of Domo de Yauli resumed with the intrusion of the 18.5 Ma rhyolite sill
near Canchayllo, about 35 km east of the continental divide. This rock has a peraluminous
composition, and thus differs from most other Neogene rocks of the area. It represents the first
magmatic product after a period of quiescence and may reflect partial melting, or extensive
assimilation, of crustal material prior to the renewed onset of arc-magmatism. Thereafter,
magmatism moved west, closer to the continental divide at about 17 Ma. Magmatic activity
flared-up around 15 Ma over the entire strike-length of the Andes within the study area. This
general evolution, from a magmatic lull over large areas and over an extended period of time, to
relatively wide-spread magmatism that, moreover, moved towards the trench in the earliest
middle Miocene may indicate a transition from a flat subduction regime to a more normal
subduction setting. The arguments for such a geodynamic change are most convincing for the
area south of Domo de Yauli, where large middle Miocene intrusions form part of the Cordillera
Occidental, but where ~14 Ma bimodal volcanism locally occurs in the back arc domain. Such
bimodal or basaltic volcanism would have been facilitated by extension induced by slab
sweetening. Basaltic rocks of Eocene, Miocene, and Pliocene age have been reported from the
Huacravilca area south of the region discussed herein (Noble et al., 1999a), and provide further
evidence for middle Miocene extension in the Huancayo-Huancavelica area.
The size of middle Miocene intrusions along the Cordillera Occidental decreases markedly
north of Domo de Yauli and bimodal middle Miocene volcanism has not been observed
anywhere else in the study area. North of Domo de Yauli, instead, magmatic activity is recorded

21
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

for the two largest mineralized districts of the study area, that is Cerro de Pasco and Colquijirca,
in the late middle Miocene, about 40-50 km east of the continental divide.

Late Miocene magmatic activity


After a short episode of comparatively intense middle Miocene magmatism, magma
production ceased in the areas south of Domo de Yauli, where no igneous rocks of ages between
12.4 and 7.5 Ma have been recognized. In contrast, from Domo de Yauli to the north
magmatism appears to have been more continuous but at reduced intensity, as several small
intrusions and volcanic domes yielded ages between 12 and 9 Ma. Intrusions of this interval are
located close to the continental divide and include Cerro Carrizal at Rosaura, the Toromocho
porphyry at Morococha, both located west of Domo de Yauli, as well as a granodiorite at
Calhuacocha some 60 km farther north near the somewhat older Chungar granite intrusion.
However, mineralization ages of 10.5-10.9 Ma were obtained for Colquijirca and Cerro de Pasco
(Bendezú et al., 2003, 2004), representing magmatic driven hydrothermal activity that locally
occurred significantly east of the Cordillera Occidental.
Late Miocene magmatism was important farther north than the area studied herein. The
Cordillera Blanca Batholith intruded between 13-3 Ma (some 100 km NW of Cerro de Pasco;
(Petford and Atherton, 1992), but in a position west of the Incaic fold and thrust belt.
In the southern portion of the study area, magmatism resumed with the intrusion of the
Yauricocha and Exito stocks at 7.5 Ma. At approximately the same time, dacite domes, dated at
8.2 Ma, were emplaced at Carhuacayán and felsic dikes generated skarn in the Uchucchacua
district. Indirect evidence for magmatism at this time comes from alteration minerals at Huarón
(Thouvenin, 1983 in Noble and McKee, 1999). Both Huarón and Carhuacayán are located
approximately 20 km east of the continental divide, indicating that magmatism has locally
shifted to the east at this time.
The youngest recorded magmatic events south of Domo de Yauli are represented by a
number of small dacite domes dated at 5.9 Ma, almost 50 km E of the continental divide. At the
same time, magmatism north of Domo de Yauli is represented by dacitic to dioritic domes at
Alpamarca, as well as the dacitic ignimbrites of Bosque de Piedra. Magmatism occurred
throughout the region in a position some tens of km east of the continental divide before it
ceased at about 5 Ma.
The magmatic evolution of the late Miocene may be interpreted as a function of the flat
subduction regime of central and northern Perú currently thought to be responsible for the
general lack of volcanism. Plate reconstruction models (Hampel, 2002) predict the onset of the
subduction of the Nazca ridge to 11.2 Ma at 11° Lat. S which approximately corresponds to the
latitude of Cerro de Pasco. Flat subduction is commonly attributed to increased buoyancy of the
slab due to the subduction of aseismic ridges such as the Nazca ridge (Gutscher et al., 2000; van
Hunen et al., 2002; Yañez et al., 2001), but can not explain the observed patterns of magmatism
(Kay and Mpodozis, 2002) in all cases.
The general decrease after 11-10 Ma and, in the latest Miocene, and the eastward shift of
magmatism, may be interpreted as evidence for slab flattening due to the subduction of the
Nazca Ridge that apparently did not affect areas farther north to the same degree, since
magmatism was important in the Cordillera Blanca through the late Miocene. Differences in
distribution of magmatism between areas south and north of Domo de Yauli are best explained
by interplay of upper-plate structural control and the southward more flat subduction geometry.

22
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Metallogenetic implications of observed magmatism


Epithermal mineralization at Quicay represents the only currently producing mine related to
Eocene magmatic activity. This mineralization is thought to coincide in time with the eruption of
the voluminous Calipuy Group volcanic rocks farther west and represents an isolated
metallogenetic event. Presumably these magmatic centers are an extension of the much more
continuous metallogenic province in the Cordillera Andahuaylas south of Cusco (Perelló et al.,
2003).
Early Oligocene skarn observed at Milpo and Atacocha is associated with the last pulses of
the late Eocene to early Oligocene magmatic episode before the onset of a magmatic lull in the
middle Oligocene. Reconstructed plate convergence rates and vectors (Pilger 1981) suggest that
the timing of mineralization coincided with a near stall of the subduction process. Such a
situation allows heating the slab to reach temperatures, which permit production of highly
oxidized, slab derived, supercritical fluids or melts (Mungall, 2002). High-oxidation states of
magmas are favorable for porphyry and skarn because copper and other metals behave
incompatible in under these magmatic compositions (e.g. Cline and Bodnar, 1991). We propose
that the Milpo and Atacocha-San Gerardo skarn represent a brief metallogenetic event. Short-
lived incursion of slab derived melts or supercritical fluids into the mafic underplates in the
lower crust may have generated highly oxidized hybrid melts which ultimately are related to the
mineralization.
Late Oligocene mineralization probably took place at Uchucchacua in the late Oligocene and
coincides with an increased plate convergence rate (Pardo-Casas and Molnar, 1987), but also a
possible flat-subduction configuration.
The most prolific regional metallogenetic episode initiated in the middle Miocene, as has
been pointed out previously (Noble and McKee, 1999). The episode also coincided with a
generally increased and more widely distributed magma production. In the southern segment,
where the geologic relationships suggest a slab steepening and westward migration of magmatic
activity, restricted mineralization in the abandoned mines of the Azulcocha district and at Rey
Salomón formed at this time. Relatively small, abandoned middle Miocene deposits, including
Río Pallanga and Chungar, formed along the Cordillera Occidental farther north. Some intrusive
complexes south of Domo de Yauli are quite large, being up to 10 by 5 km in outcrop. The
modest size of mineralization at Rey Salomón and the absence thereof in the context of other
intrusions may be attributed to the apparently deep level of exposures in the large igneous
complexes. A contrary argument is that mineral deposits farther north along the entire
continental divide are of modest size as well, despite the presumably lesser degree of exhumation
which has so far failed to expose large intrusive bodies in those areas. We therefore propose the
presence of a first order metallogenetic control on the middle Miocene mineralization.
Comparing the historically mined deposits along the continental divide, the largest middle
Miocene deposits formed in the Cerro de Pasco and Colquijirca districts at a distance of 40 to 50
km east of the Cordillera Occidental. These two major mineral districts occupy a position
significantly east of where the bulk of the middle Miocene magmatism farther south was
concentrated and also lie in the broad east-west trending metallogenetic belt where at least three
previous episodes of mineralization can be discerned. Cerro de Pasco and Colquijirca thus
formed in a setting that is unrepresented in the middle Miocene magmatic arc farther south. This
clustering of ore deposits of different ages in the Uchucchacua-Cerro de Pasco transect strongly
argues for an upper plate, crustal scale structural and (or) metallogenetic control.

23
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

After about 11 Ma, renewed flattening of the subducted slab caused a broadening as well as a
decrease of intensity of the magmatic arc, which ended in the cessation of arc magmatism at
approximately 5 Ma. Flat subduction, similar to a stalling slab as described above, can
potentially provide the conditions necessary to generate a supercritical fluid in the subducted slab
(Mungall, 2002). Mineralization at Yauricocha, Carhuacayán, Huarón and most importantly at
Morococha and San Cristóbal and other deposits near Domo de Yauli can be related to this
renewed flat subduction regime. Similarly, the giant Cu-Zn skarn deposit of Antamina, some 100
km NW of Cerro de Pasco and in a position approximately 20 km east of the Cordillera Blanca
Batholith was emplaced at 10.7-10.1 Ma (Love et al., 2004; this study). World-class
mineralization in a flat subduction environment is well known and described from the Chilean
flat-slab segment, where metallogenetic relationships are more straightforward to decipher
(Bissig et al., 2003; Kay et al., 1999; Reich et al., 2003).

Conclusions
Scattered igneous activity occurred in the Central Peruvian Cordillera Occidental and the
adjacent high-plains between late Eocene (~40.2 Ma) and late Miocene (5.2 Ma), but is
characterized by regional and temporal variations in intensity. Late Eocene and early Oligocene
widespread albeit volumetrically limited magmatism was followed by a magmatic lull that lasted
from 29.3 to 25.3 Ma north of Domo de Yauli, but persisted until the early Miocene farther
south. The reduced magmatic output in the Oligocene correlates with a period of slow
convergence between the Nazca plate and South American continent and, after about 25 Ma, a
more oblique subduction and a possible flat slab configuration.
Magmatism increased in intensity in the middle Miocene. South of Domo de Yauli, felsic to
intermediate rocks intruded near the axis of the Cordillera Occidental and local basaltic-dacitic
bimodal volcanism took place 30 km into the back-arc domain, a pattern that is interpreted as
reflecting the steepening of the subducted slab following a possible late Oligocene episode of flat
subduction. North of Domo de Yauli, the magmatic pattern differs in that intrusions near the
continental divide are small middle Miocene felsic to intermediate magmatism, represented by
those in the Cerro de Pasco and Colquijirca area some 40-50 km east of the Cordillera
Occidental. The different patterns may reflect slight differences in subduction angle or upper-
plate structural control on magma ascent.
Magma output rates decreased in the late Miocene and magmatism locally expanded to the
east, a pattern that may be interpreted as a renewed onset of flat subduction beneath central Perú.
Plate reconstruction models predict the onset of Nazca Ridge subduction at 11°S at 11.2 Ma, a
timing that agrees well with the onset of flat subduction thereafter.
Several episodes of intrusion related ore deposits are recognized in the study area. The
metallogenetically most prolific area is a broad easterly striking belt between Uchucchacua and
Milpo-Atacocha where late Eocene (Quicay), early Oligocene (Milpo-Atacocha), late Oligocene
(Uchucchacua), middle Miocene (Iskaycruz(?), Cerro de Pasco, Colquijirca) and late Miocene
(Uchucchacua(?), Alpamarca) mineralization events are concentrated. Economic mineralization
of Eocene and Oligocene age has not been dated anywhere else in the study area, but the
abandoned and yet undated Cercapuquio mine and the Mario prospect could conceivably be of
this age.
Middle Miocene mineralization of generally modest size is found along the entire length of
the Cordillera Occidental within the study area, whereas late Miocene mineralization is

24
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

concentrated at Domo de Yauli, but includes the operating mines of Yauricocha to the south and
Huarón to the north as well.
We conclude that the Uchucchacua to Milpo transect represents a first order metallotect
where mineralization occurred in several episodes since the late Eocene. However middle and
late Miocene deposits of generally more modest size formed throughout central Perú.
Exploration efforts thus would promise best results around intrusions of ages between 14 and 7
Ma. Intrusions of other ages appear to have had the potential for generating mineralization in the
Uchucchacua-Milpo transect only.

References cited
Baumgartner, R., Fontboté, L., and Bendezú, R. R., 2003, Low temperature, late Zn-Pb-(Bi-Ag-
Cu) mineralization and related acid alteration replacing carbonate rocks at Cerro de Pasco,
Central Peru: Society for Geology Applied to Mineral Deposits, 7th Biennial Meeting,
Athens, Greece, 2003, p. 441-444.
Benavides, V., 1999, Orogenic evolution of the Peruvian Andes: the Andean cycle, in Skinner,
B. J., ed., Geology and ore deposits of the central Andes: Society of Economic Geologists
Special Publication No. 7: Littleton, Colorado, Society of Economic Geologists, p. 61-107.
Bendezú, R., Fontboté, L., and Cosca, M., 2003, Relative age of Cordilleran base metal lode and
replacement deposits and high sulfidation Au-(Ag) epithermal mineralization in the
Colquijirca mining district, central Peru: Mineralium Deposita, v. 38, p. 683-694.
Bendezú, R., Baumgartner, R., Fontboté, L., Page, L., Pecskay, Z., and Spikings, 2004, The
Cerro de Pasco-Colquijirca ‘super-district”, Peru” ~2 m.y. of pulsed high-sulphidation
hydrothermal activity: SEG 2004 Predictive Mineral Discovery Under Cover, p. 340-342.
Beuchat, S., 2003, Geochronological, structural, isotope and fluid inclusion constraints of the
polymetallic Domo de Yauli district, Peru: PhD thesis, Terre et environment, 41, ISBN 2-
940153-40-X.
Bissig, T., Clark, A. H., Lee, J. K. W., and von Quadt, A., 2003, Petrogenetic and
Metallogenetic responses to Miocene slab flattening: new constraints from the El Indio-
Pascua Au-Ag-Cu belt, Chile/Argentina: Mineralium Deposita, v. 38, p. 844-862.
Bussell, M. A., Alpers, C. N., Petersen, U., Shepherd, T. J., Bermudez, C., and Baxter, A. N.,
1990, The Ag-Mn-Pb-Zn vein, replacement and skarn deposits of Uchucchacua, Peru: studies
of structure, mineralogy, metal zoning, Sr isotopes and fluid inclusions: Economic Geology,
v. 85, p. 1348-1383.
Cedillo, E., and Tejada, J., 1988, Yacimientos estratoligados de plomo y zinc en la formación
Chaucha del Jurasico Superior; Cercapuquio, Junín: Boletin de la Sociedad Geológica del
Perú, v. 78, p. 35-43.
Cline, J. S., and Bodnar, R. J., 1991, Can economic mineralization be generated by a typical
calc-alkaline melt?: Journal of Geophysical Research, v. 96, p. 8113-8126.
Cobbing, E. J., 1973, Geología de los cuadrángulos de Barranca, Ambar, Oyón, Huacho, Huaral
y Canta: Lima, 172 p.
Escalante, A., Dipple, G., Ebert, S., Tosdal, R. M., and Lipten, E., 2004, Distal alteration and
veins around Antamina and Condorcocha skarn systems, northern Peru: XII Congreso
Peruano de Geología, Lima, Peru, 2004, p. 650-651.
Farrar, E., and Noble, D. C., 1976, Timing of late Tertiary deformation in the Andes of Peru:
Geological Society of America Bulletin, v. 87, p. 1247-1250

25
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Giletti, B. W., and Day, H. W., 1968, Potassium-argon ages of igneous rocks of Peru: Nature, v.
220, p. 570-571.
Gunnesch, K. A., and Baumann, A., 1984, The Atacocha district, central Peru: some
metallogenetic aspects, in Wauschkuhn, A., Kluth, C., and Zimmermann, R. A., eds.,
Syngenesis and Epigenesis in the Formation of Mineral Deposits: Heidelberg, Germany,
Springer-Verlag, p. 448-456.
Gutscher, M.-A., Maury, R., Eissen, J.-P., and Bourdon, E., 2000, Can slab melting be caused
by flat subduction?: Geology, v. 28, p. 535-538.
Hampel, A., 2002, The migration history of the Nazca Ridge along the Peruvian active margin: a
re-evaluation: Earth and Planetary Science Letters, v. 203, p. 665-679.
Johnson, R. F., 1955, Geology of the Atacocha mine, department of Pasco, Peru: Economic
Geology, v. 50, p. 249-270.
Kay, S. M., and Mpodozis, C., 2001, Central Andean ore deposits linked to evolving shallow
subduction sysstems and thickening crust: GSA Today, v. 11, p. 4-9.
Kay, S. M., and Mpodozis, C., 2002, Magmatism as probe to the Neogene shallowing of the
Nazca plate beneath the modern Chilean flat-slab: Journal of South American Earth Sciences,
v. 15, p. 39-57.
Kay, S. M., Mpodozis, C., and Coira, C., 1999, Neogene magmatism, tectonism and mineral
deposits of the Central Andes (22° to 33° S Latitude), in Skinner, B. J., ed., Geology and Ore
Deposits of the Central Andes: Society of Economic Geologists Special Publication No. 7:
Littleton, Colorado, Society of Economic Geologists, p. 27-59.
Krogh, T.E, 1982, Improved accuracy of U-Pb zircon ages by the creation of more concordant
systems using an air abrasion technique. Geochimica et Cosmochimica Acta, v. 46, p.
637-649.
Love, D. A., Clark, A. H., and Glover, K. J., 2004, The lithologic, stratigraphic and structural
setting of the giant Antamina copper-zinc skarn deposit, Ancash, Peru: Economic Geology,
v. 99, p. 887-916.
Ludwig, K.R., 1980, Calculation of uncertainties of U-Pb isotopic data, Earth and Planetary
Science Letters, v. 46, p. 212-220.
Ludwig, K.R 2003. Isoplot 3.00 A Geochronological Toolkit for Microsoft Excel: Berkeley
Geochronology Center, Special: Publication No. 4
Mégard, F., 1968, Geología del cuadrángulo de Huancayo: Lima, Peru, 123 p.
Mungall, E. J., 2002, Roasting the mantle: slab melting and the genesis of major Au and Au-rich
Cu deposits: Geology, v. 30, p. 915-918.
Muñoz, C., 1994, Geologische, mineralogische und metallogenetische Untersuchungen des
Jatunhuasi-Azulcocha-Chuquipita Gebietes, mit besonderer Berücksichtigung der Zn-As-
(Au)-Lagerstätte Azulcocha, Zentralperu, Ruprecht-Karls-Universität, Heidelberg, 380 p.
Noble, D. C., Farrar, E., and J., C. E., 1979, The Nazca Group of south-central Peru: age,
source, and regional volcanic and tectonic significance: Earth and Planetary Science Letters,
v. 45, p. 80-86.
Noble, D. C., and McKee, E. H., 1999, The Miocene metallogenic belt of central and northern
Peru, in Skinner, B. J., ed., Geology and Ore Deposits of the Central Andes: Society of
Economic Geologists Special Publication No 7: Littleton, Colorado, Society of Economic
Geologists, p. 155–193.
Noble, D. C., Mckee, E. H., Mourier, T., and Mégard, F., 1990, Cenozoic stratigraphy, magmatic
activity, compressive deformation and uplift in northern Peru: Geological Society of America
Bulletin, v. 102, p. 1105-1113.

26
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Noble, D. C., Vidal, C. E., Perelló, J., and Rodríguez, P., 2004, Space-time relationships of
some porphyry Cu-Au, epithermal Au and other magmatic related mineral deposits in
northern Peru, in Sillitoe, R. H., Perelló, J., and Vidal, C. E., eds., Andean Metallogeny: New
Discoveries, Concepts, and Updates, Society of Economic Geologists Special Publication
No. 11: Littleton, Colorado, p. 313-317.
Noble, D. C., Wise, J. M., and Vidal, C. E., 1999a, Episodes of Cenozoic extension in the
Andean orogen of Peru and their relation to compression, magmatic activity and
mineralization, in Benavides, V., and Rosas, S., eds., Volumen Jubilar No. 5 “75 Aniversario
Sociedad Geológica del Peru”: Lima, Peru, Sociedad Geológica del Perú, p. 45-66.
Noble, D. C., Wise, J. M., Vidal, C. E., and Heizler, M. T., 1999b, Age and deformational
history of the “Calipuy Group” in the Cordillera Negra, northern Peru, in Benavides, V., and
Rosas, S., eds., Volumen Jubilar No. 5 “75 Aniversario Sociedad Geológica del Peru”: Lima,
Perú, Sociedad Geológica del Peru, p. 219-226.
Pardo-Casas, F., and Molnar, P., 1987, Relative motion of the Nazca (Farallon) and South
American plates since Late Cretaceous time: Tectonics, v. 6, p. 233-248.
Parrish, R., Roddick, J.C., Loveridge, W.D., and Sullivan, R.W., 1987, Uranium-lead analytical
techniques at the geochronology laboratory, Geological Survey of Canada. In Radiogenic
Age and Isotopic Studies, Report 1, Geological Survey of Canada, Paper 87-2, p. 3-7.
Perelló, J., Carlotto, V., Zárate, A, Ramos, P, Posso, H, Neyra, C., Caballero, A., Fuster, N., and
Muhr, R., 2003, Porphyry-style alteration and mineralization of the middle Eocene to early
Oligocene Andahuaylas-Yauri Belt, Cuzco Region, Peru: Economic Geology, v. 98, p. 1575-
1606.
Petersen, U., 1965, Regional Geology and major ore deposits of central Perú: Economic
Geology, v. 30, p. 407-476.
Petersen, U., Mayta, O., Gamarra, L., Vidal, C. E., and Sabastizagal, A., 2004, Uchucchacua: a
major silver producer in South America, in Sillitoe, R. H., Perelló, J., and Vidal, C. E., eds.,
Andean Metallogeny, New Discoveries, Concepts and Updates, Society of Economic
Geologists Special Publication No. 11, p. 243-257.
Petford, N., and Atherton, M. P., 1992, Granitoid emplacement and deformation along a major
crustal lineament: the Cordillera Blanca, Peru: Tectonophysics, v. 205, p. 171 - 185.
Pilger, R. H., 1981, Plate reconstructions, aseismic ridges, and low-angle subduction beneath the
Andes: Geological Society of America Bulletin, v. 92, p. 338-456.
Reich, M., Parada, M. A., Palacios, C., Dietrich, A., Schultz, F., and Lehmann, B., 2003,
Adakite-like signature of Late Miocene intrusions at the Los Pelambres giant porphyry
copper deposit in the Andes of central Chile: metallogenic implications: Mineralium
Deposita, v. 38, p. 876 – 885.
Renne, P.R., C.Swisher, C.C., III, Deino, A.L., Karner, D.B., Owens, T. and DePaolo, D.J.,
1998, Intercalibration of standards, absolute ages and uncertainties in 40Ar/39Ar dating:
Chemical Geology, v. 145, no. 1-2, p. 117-152.
Roddick, J.C., 1987, Generalized numerical error analysis with application to geochronology
and thermodynamics. Geochimica et Cosmochimica Acta, v. 51, p. 2129-2135.
Romani, M., 1982, Géologie de la region minière Uchcchacua-Hacienda Otuto, Pérou: Ph.D.
thesis, L’Universite scientifique et medicale de Grenoble. 176 p.
Sandeman, H. A., Clark, A. H., and Farrar, E., 1995, An integrated tectono-magmatic model for
the evolution of the southern Peruvian Andes (13-20° S) since 55 Ma: International Geology
Review, v. 37, p. 1039-1073.

27
Final report: Sources and Exhausts in Polymetallic Carbonate Rock-hosted Ore Deposits

Sébrier, M., and Soler, P., 1991, Tectonics and magmatism in the Peruvian Andes from Late
Oligocene time to the present,, in Harmon, R. S., and Rapela, C. W., eds., Andean
Magmatism and its Tectonic Setting, Geological Society of America Special Paper 265,
Geological Society of America, p. 259-279.
Silberman, M. L., and Noble, D. C., 1977, Age and igneous activity and mineralization, Cerro de
Pasco, central Perú: Economic Geology, v. 72, p. 925-930.
Soler, P., 1986, La Province polymétalique des Andes du Pérou central Chronique de la
Recherche Minière, v. 482, p. 39-54.
Stacey, J.S. and Kramer, J.D., 1975. Approximation of terrestrial lead isotope evolution by a
two-stage model: Earth and Planetary Science Letters, v. 26, p. 207-221.
Thouvenin, J.-M., 1983, Les minéralisations polymetalliques à Zn-Pb-Cu-Ag de Huarón, Pérou
Central : Mineralographie des minéralisés et petrographie des altérations de épontes, École
Supérieure des Mines de Paris, 223 p.
Tosdal, R. M., Clark, A. H., and Farrar, E., 1984, Cenozoic polyphase landscape and tectonic
evolution of the Cordillera Occidental, southernmost Peru: Geological Society of America
Bulletin, v. 95, p. 1318-1332.
van Hunen, J., van den Berg, A. P., and Vlaar, N. J., 2002, On the role of subducting oceanic
plateaus in the development of shallow flat subduction: Tectonophysics, v. 352, p. 317– 333.
Vidal, C. E., and Ligarda, R., 2004, Enargite-gold deposits at Marcapunta, Colquijirca mining
district, central Peru: mineralogic and geochemical zoning in subvolcanic limestone-
replacement deposits of high-sulfidation epithermal type, in Sillitoe, R. H., Perelló, J., and
Vidal, C. E., eds., Andean Metallogeny: New Discoveries, Concepts, and Updates, Society of
Economic Geologists Special Publication No. 11, p. 231-241.
Yañez, G. A., Ranero, C. R., von Huene, R., and Diaz, J., 2001, Magnetic anomaly
interpretation across the southern central Andes (32°-34° S): The role of the Juan Fernandez
Ridge in the late Tertiary evolution of the margin: Journal of Geophysical Research, v. 1206,
p. 6325-6345.
Zimmerninck, W. G., 1983, Investigaciones mineralógicas y petrológicas en el depósito
Felicidad: Boletin de la Sociedad Geológica del Perú, v. 70, p. 51-60.

28

You might also like