You are on page 1of 28

Caylloma epithermal district Accepted, Econ. Geology, Nov.

2004

Structural and stratigraphic controls on and paragenetic evolution of the Caylloma


epithermal district, southern Perú

Leandro Echavarria** and Eric Nelson*


Department of Geology and Geological Engineering, Colorado School of Mines
Jorge Chavez and Leopoldo Escobedo
Mauricio Hochschild & Cia. Ltda., S.A.C.
Alex Iriondo ‡
U.S. Geological Survey, Denver

* Corresponding author: enelson@mines.edu


** Present address:
‡ Present address: Centro de Geociencias, Universidad Nacional Autónoma de México
Campus Juriquilla, C.P. 76230 Juriquilla, Querétaro, México

Abstract
The silver- and base metal-rich Caylloma epithermal district in the Tertiary volcanic belt of southern Perú has
been worked intermittently since the Incaic period and has produced over 100 million ounces of Ag.
Intermediate sulfidation mineralization is present in veins hosted by Miocene andesitic volcanic and
volcaniclastic rocks, with minor ore in underlying folded Jurassic sedimentary basement. New Ar/Ar dates
give a host rock age of 20.30±0.11 Ma (andesitic volcanic matrix), a mineralization age of 18.35±0.17 Ma
(adularia in vein wall rock), and post-mineralization ages of 11.8±0.8 Ma and 12.25±0.07 Ma (sanidine and
biotite from a rhyolite dome). Gangue minerals include quartz, calcite, rhodonite, rhodochrosite, pyrite and
minor adularia, illite, barite, and helvite. Ore minerals include sphalerite, galena, chalcopyrite, and
tetrahedrite. Hydrothermal alteration is pervasive in lava flows, but weak and narrow in volcaniclastic rocks.
The three types of alteration include: quartz-adularia (+pyrite±illite), quartz-illite (+pyrite) and propylitic
(chlorite+calcite±illite). Banded veins show four stages of mineral precipitation: early (sugary quartz,
chalcedony, pyrite), manganese (ore), quartz (+sulfides), and late (calcite+quartz). Cyclic bands in the
manganese stage (early sulfides, coarse to medium-grained quartz, late rhodonite+calcite+chalcedony) are a
few mm to 5 cm thick, and repeat to form ore bands up to 1 meter thick.
Veins occupy dextral normal faults (020° to 050°E, 45° to70° SE) and extension fractures (060° to
090°E, 70° to 90° SE), and are 1 to 25 meters wide and several kilometers long. The veins display complex
and multiepisodic filling with text ures characteristic of open space precipitation such as crustiform banding,
symmetric banding, vugs, breccias, and cockade and comb textures. We present a structural model in which
the principal veins formed in sub-parallel NE-striking dextral-normal faults and related extension fractures
within a NW-striking structural corridor bounded by sinistral regional faults; second and third order veins
formed by bookshelf sliding related to movement on higher-order, dominantly strike-slip faults. Slickenlines
have low rakes (20°-50° SW), and fault -kinematic analysis suggests the extension axis was oriented 0°/330°
and the shortening axis 55°/ 234°, in agreement with the strain field commonly proposed for the Early -Middle
Miocene of southern Peru.
Ore grade is discontinuous, with high-grade ore shoots tens to hundreds of meters long and about 300
meters of known vertical extent. Ore shoots in fault-hosted veins are narrow subvertical zones a few tens of
meters wide (measured in the plane of the vein), separated by lower-grade ore zones. Ore shoots in extension
fractures are vertically shorter but horizontally more continuous. Some ore shoots occur as half-cymoids
which form along the main vein only in the hangingwall of syn-mineral transverse faults. Mineral
composition and metal content are zoned horizontally and vertically; silver, base metals, and calcite and
manganese minerals increase to the northeast and downward, and sulfide content increases gradually with
depth. Shallower levels of the hydrothermal system are exposed to the west and southwest in the district.
Evidence from sulfur and lead isotopes, and the high salinity of the fluids from fluid inclusion
analysis, support the contribution of magmatic fluids into the hydrothermal system through channels of
enhanced structural permeability. The cyclic nature of the mineral precipitation with abrupt changes in
mineralogy, textures, and fluid composition leads us to postulate that cyclic injections of magma -derived

1
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

hydrothermal fluids occurred. The depth of formation of economic mineralization is estimated to be about 650
meters below the paleo-water table, with a temperature of formation of ~270°C.

Introduction
The Andes of Peru are characterized by numerous ore deposits within the Tertiary volcanic arc (Noble et al.,
1989, 1999, Noble and Vidal, 1994). The Miocene portion of this belt is one of the most important
metallogenic provinces, hosting numerous epithermal deposits (Noble and McKee, 1999). In the Arequipa
Department these precious metal-rich deposits consist of veins hosted by Miocene volcanic rocks. Major
examples include Orcopampa (Gibson et al., 1990, 1995), Arcata (Candiotti et al., 1990; Echavarria et al.,
2003), Ares (Candiotti and Guerrero, 2002), Shila (Cassard et al., 2000), and Caylloma (Fig. 1).
The silver- and base metal-rich Caylloma epithermal district, located about 150 km north of
Arequipa city (Fig. 1) at 4,500 meters above sea level, has been worked intermittently since the Incaic period.
Although no data are available for the first period of workings carried out by the Incas before the arrival of
the Europeans, only small-scale extraction occurred. During the Spanish colonial period, between 1626 to
1824, the upper part of the principal veins (San Cristóbal, San Pedro, El Toro, Elisa, Bateas, Trinidad, and La
Leona, Fig. 2) were worked intensively, and an estimated 490,000 t (metric ton) of high grade ore averaging
114 oz/t of Ag were extracted (Stephan, 1974). After many years of inactivity, German, English, and Chilean
companies attempted to work the Caylloma mines during the period 1889 to 1926. Finally, in 1933 Compania
Minera de Caylloma started mining and is still working today. During its extensive history, the Caylloma
district has produced over 4 million tons of ore from which more than 100 million ounces of Ag have been
extracted. Present workings exploit the San Cristóbal, Animas and La Plata veins; daily production is 600 t
with an average Ag grade of 13 oz/t.
Several studies discuss the metallogenesis and geological evolution of the region (e.g., Fornari and
Vilca Neyra, 1979; McKee and Noble, 1982, 1989; Noble et al., 1974, 1984, 1989, 1999; Peterson et al.,
1983; Sebrier and Soler, 1991; Mercier et al., 1992, Noble and Vidal, 1994). However, there are few detailed
studies of the Caylloma dis trict. Stephan (1974) discusses district mineralogy, and Silberman et al. (1985)
and McKee and Noble (1989) provide geochronological data on the age of mineralization. Despite its long
history and economic importance, there are no published studies of the geological and structural setting of the
Caylloma district, or of the controls on mineral distribution and hydrothermal fluid evolution.
This paper presents the geological setting and age of mineralization and related rocks of the
Caylloma d istrict, as well as the structural and stratigraphic controls on the veins, and their paragenesis and
metal distribution. This work is based on new data from detailed field mapping, isotopic age determinations,
3D modeling of the most important veins, and mineralogical, petrographic, and fluid inclusions studies.

Regional Geology
The Neogene volcanic arc of the Central Peruvian Andes developed over a thick continental crust
composed of multistage deformed Paleozoic and Mesozoic rocks. The volcanic arc in southern Peru is
characterized by large, locally superimposed calderas (Fig. 1). The rock sequence contains calc-alkaline
andesitic to rhyolitic flows, ignimbrites, laharic deposits, and volcanic domes. Volcanic and tectonic activity
was episodic (McKee and Noble, 1989; Sebrier and Soler, 1991; Mercier et al., 1992; Noble et al., 1999),
reflecting periods of quiescence with mild extension followed by short periods of compressional deformation.
Magmatic and hydrothermal activity seem to be related to extensional or transtensional tectonism (Noble et
al., 1999).
The study area contains two calderas, Chonta and Caylloma (Fig. 1). However, both are younger
than mineralization of the Caylloma district. The Chonta caldera contains thick rhyolitic welded ash flow
tuffs, and several post-caldera rocks including dacite and andesite flows and dacite domes. Several epithermal
deposits, including Sukuitambo, San Miguel, Chonta, and Corimina are suggested to be genetically related to
the Chonta caldera (Noble, 1981a, Peterson et al., 1983, Noble et al., 1989). The Caylloma caldera is
characterized by two or three episodes of collapse and pyroclastic activity (Noble, 1981b), followed by
resurgence and effusion of phenocryst-poor andesite. Portions of the caldera are partia lly covered by
extensive flat-lying andesitic flows associated with young strato-volcanoes and assigned to the Plio-
Pleistocene Barroso Group.
Two main fracture systems are recognized in the area, striking NW and NE. The NW-striking system
shows sinistral displacements and is better developed than the dextral NE-striking system. The NE-striking
Caylloma veins are present within a structural corridor bounded by two major NW-striking lineaments

2
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

interpreted as faults, the Trinidad and Santiago faults. These NE- and NW-striking features are interpreted to
be conjugate fractures formed in response to a regional stress field with NNW tension axis and ENE
compression axis, in agreement with the stress field commonly proposed for the Early-Middle Miocene of
Southern Perú (Cassard et al., 2000). A major set of lineaments striking N is superimposed on the ones
described above, and is interpreted to have controlled the emplacement of the post-mineral Caylloma caldera .

District Geology
The oldest rocks exposed in the area are Jurassic sedimentary rocks of the Yura Group (Fig. 2 and 3),
which consist of intercalated sandstone and marine black shale and sandstone. In outcrop, these strata are
strongly deformed into frequently recumbent kink folds with straight limbs and narrow hinges. However,
strain in the Yura Group is locally weaker at depth where only open folds have been recognized within the
mine.
A thick sequence of andesitic lavas and volcaniclastic rocks unconformably overlies the Yura Group
(Fig. 3) and is the main host of the mineralized veins. The volcanic sequence consists of 20 to 100 meter
thick andesitic lava flows interlayered with thinner beds of volcaniclastic rocks. Lavas vary from coarse-
grained, phenocryst-rich to fine-grained matrix-rich andesite. The volcaniclastic rocks generally form thin
beds, 2 to 5 meters in thickness, between andesite flows. They consist primarily of reworked andesitic
material and fresh-water limestone. South of San Cristóbal vein (Fig. 2), weakly welded ignimbrites,
composed of matrix-supported breccia of rhyolitic to dacitic composition, are widespread and form gently
dipping packages tens of meters thick.
Sedimentary and volcanic rocks are intruded by post-mineral fault-related rhyolitic lava domes and
dikes, characterized by coarse quartz and sanidine phenocrysts, spherulites and lithophysae, and a well-
developed lamination. Finally, Plio-Pleis tocene lava flows of the Barroso Group unconformable overly older
rocks and form plateaus (Fig. 2).
The district is crosscut by NE- and lesser developed NW-striking oblique-normal faults (Fig. 2).
Most faults in both sets host mineralized veins. NE-striking faults dip 45° to 70° southeast and have an
important dextral component with slickenline rakes between 20° and 50° SW (Fig. 4). The movement along
NE-striking faults produced a downdropping of blocks to the south and tilting of the volcanic sequence about
20° to the north (Fig. 5). It was not possible to measure the net displacements along the faults due to lack of
markers; however, displacements between tens and hundreds of meters were estimated. Post-mineral high-
angle normal faults striking N to NW offset the veins several to tens of meters.

Caylloma District Mineralization


The veins at Caylloma are a classic example of the intermediate sulfidation subtype (Hedenquist et
al., 2000; Sillitoe and Hedenquist, 2003), or Ag-rich system (Nolan, 1933). The district contains veins 1 to 25
meters wide and several kilometers long, hosted principally by a thick M iocene sequence of andesite lavas
and ash flow tuffs (Fig. 2). Grade is discontinuous, with high-grade ore zones tens to hundreds of meters long
and with about 300 meters of known vertical extension. In the following sections we discuss the structural and
stratigraphic control on veins and ore-grade distribution, hydrothermal alteration, mineral paragenesis, and
metal content and distribution.

Structural setting and control on mineralization


Veins in the Caylloma district are hosted by sub-parallel oblique normal faults and extension
fractures (Fig. 6). Sense of movement along the faults was determined based on normal separation of strata,
splay geometry, and slickenlines. Most faults that host ore strike between 020° and 050°, and dip 45° to 70° to
the SE (Fig. 4). NW-striking faults were also identified, although most do not host economic mineralization.
Northeast-striking faults have dextral-normal movement with a dominant strike-parallel component and host
some of the richest veins of the district including San Cristóbal, Animas, Paralela -Santa Rosa, Elisa-Apostoles
2-Jerusalem, La Peruana-Santo Domingo (Fig. 2). On the other hand, veins hosted by NW-striking faults are
smaller and less common. They strike between 330° and 295° and dip with high angles both NE and SW, and
have a normal-sinistral movement with a dip-slip component greater than the NE-striking veins.
Extension fractures related to the faults strike between 060° and 090° and dip between 70° and 90°
mostly to the S (Fig. 4). Extension fractures are observed in both the hangingwall and footwall of the faults,
but those in the hangingwall are more common. These are mainly dilatational structures with minor sinistral

3
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

movement, commonly form thick and high grade veins, though short and discontinuous, and include the El
Toro, San Pedro, Bateas, Santa Cata, Trinidad, and Carolina among the most notable (Fig. 6).
Although slickenlines within and on the walls of the fault-veins are uncommon, several were
measured on the main veins (Fig. 4). Shear sense interpreted from slickenlines agrees with other kinematic
indicators such as mineralized extension (T) fractures (Fig. 6). Therefore we believe that the slickenlines
formed synchronously with the veins. Slickenlines on NE-striking fault planes have low-angle rakes between
20° to 50° SW (Fig. 4). Fault -kinematic analysis of the data suggests that the extension axis was sub-
horizontal and trended 330°, and the shortening axis plunged 55° and trended 234° (Fig. 4). Extension
fractures form close to the λ2 -λ3 plane during fault generation.
Vein textures include crustiform banding, breccias, and cockade textures, suggesting that vein
opening and filling was episodic, with several episodes of fault movement related to brecciation and
mineralization. In some NE-striking veins with breccia on the vein margins and high-angle slickenlines, a late
reactivation of faults is interpreted to have produced a normal-sinistral movement.
We present a structural model (Fig. 6c) to explain the various vein types, orientations, and
kinematics. In this model, the second- and third-order veins formed by bookshelf sliding related to movement
on higher-order, dominantly strike-slip faults. Splays on any order fault usually formed extension veins.

Ore grade distribution


Silver grade within the veins is irregular, with the economic mineralization hosted in discontinuous
first- or second-order ore shoots. Fault-hosted veins are more than 3 km in length, and generally have a semi-
continuous ore zone located near the central portion of the vein, with grade decreasing toward the vein tips. In
general, ore shoots in fault-hosted veins are narrow, subvertical zones a few tens of meters wide (measured in
the plane of the vein) with up to three hundred meters of vertical extent. These zones are separated by low-
grade zones which may be economic (Fig. 7). In contrast, the distribution of ore within extension fractures is
somewhat different, with shorter, more horizontally continuous ore shoots (Fig. 8) that are up to 600 meters in
length. The area of intersection between faults and extension fractures only rarely hosts high-grade ore. The
position of ore shoots in fault-hosted veins may be related to the hangingwall of crosscutting faults (see half
cymoids below).
Half cymoids: Ore shoots in some veins in the district are related to transverse syn-mineralization
faults. The most important of these ore shoots, due to the high grade and volume of ore, are present in the San
Cristóbal vein (Fig. 9), which is the main vein of the district. It has been exp loited since the Spanish period
and is still worked at present. The vein is hosted by a NE-striking, SE-dipping oblique (dextral) normal fault
that is more than 3.5 km long and locally over 25 m wide. Several ore shoots located along the vein have
more than 350 meters of vertical extent. Several high-grade ore shoots are truncated by transverse NW-
striking, NE-dipping normal faults. However, alteration and mineralization along the transverse faults indicate
that these faults moved during vein formation. These ore shoots have a characteristic shape, here called half-
cymoid, and form only in the hangingwall of the transverse fault. In the half-cymoid the vein has several wide
splits and numerous narrower veins that extend up to 150 m away from the transverse fault (Fig. 9). There are
usually two main subparallel splits, of which the footwall split is wider, higher grade, dips at a lower angle,
and ends at the transverse fault. In the footwall of the transverse fault only one single vein with poor Ag grade
is usually present (Fig. 10). Similar examples of half-cymoids have been described in the Breckenridge
district, Colorado (Lovering and Goddard, 1950, Figs. 31 and 32).
Perhaps the best example of a half cymoid along the San Cristóbal vein is in the hangingwall of the
Jesus Maria transverse fault (Fig. 9). The Jesus Maria fault strikes about 290° and dips 70°NE, and its syn-
mineralized character is recognized from tectonic breccias cemented by vein material with crustiform
banding. Although the kinematics of this fault has not been established, slickenlines measured on other NW-
striking normal faults show a dip-slip movement. In the footwall of the Jesus Maria fault, the San Cristóbal is
a single vein about 6 meters wide striking 030° and dipping 57° to 60° SE, with low Ag content. However, in
the hangingwall of the Jesus Maria fault the vein is complex with several splits and mineralized fractures.
Two main splits are recognized; the widest and highest Ag grade vein is the footwall split that dips about 50°
SE, constituting one of the most important ore shoots of the San Cristóbal vein. The hangingwall split is
narrower, with lower Ag content (Fig. 10) and slightly higher dips, averaging 56°SE. The host rock between
both splits is strongly altered and is cut by numerous veins and veinlets. About 150 m NE of the Jesus Maria
fault the splits join and the San Cristóbal vein continues as a single vein.
Another example of a half cymoid in the district is present along the La Plata vein (Fig. 2), where the
controlling transverse normal fault is the NW-striking Antimonio Bajo fault-vein. In this case, the

4
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

northernmost footwall split of La Plata vein is mineralized, whereas the hangingwall split is only recognized
as a fault.

Stratigraphic control on mineralization


As mentioned before, the Caylloma district veins are partly hosted by folded Jurassic marine
sedimentary rocks (basement) and by unconformably overlying Miocene volcanic rocks. Within the
sedimentary basement veins are generally narrow, tortuous and branched, with only low grades or small zones
of high-grade ore. However, when they reach the volcanic rocks veins widen, and are more continuous with
markedly higher grades. In some veins (e.g., Apostoles 2 vein; Fig. 7) the Jurassic-Miocene unconformity
marks the base of the economic mineralization. Because the region was affected by block tectonics, the
topographic level of the unconformity is quite variable across the district. For example, the unconformity is
present at level 9 (4650 m.a.s.l.) in the Apostoles 2 workings, but is below level 13 (4450 m.a.s.l.) in San
Cristóbal vein and is below level 14 (4400 m.a.s.l.) in Bateas vein, where it has not been recognized. The
same stratigraphic control described above was identified in other epitermal vein districts of the area. For
examp le, Gibson et al. (1990) note that in the Orcopampa district veins are narrow and low grade when they
are hosted by Mesozoic sedimentary rocks.

Hydrothermal alteration
Lava flows and volcaniclastic rocks of the Caylloma district are hydrothermally altered. In general,
alteration is more widely distributed within the lava flows, with only narrow and weak alteration zones within
the volcaniclastic rocks. The most common minerals present in the alteration zones are quartz, pyrite,
adularia, and illite that constitute three different types of alteration: quartz-adularia (+pyrite±illite), quartz-
illite (+pyrite) and propylitic (chlorite+calcite±illite).
Quartz-adularia alteration is restricted to vein margins, and its width is directly proportional to vein
thickness, varying between a few centimeters to several meters. Quartz is present as a replacement of the
volcanic matrix, and also as irregular veinlets, whereas adularia is almost completely restricted to the veinlets.
Pyrite is disseminated within quartz in the veinlets, and also grows over mafic mineral phenocrysts. Illite is
present within plagioclase phenocrysts and in the volcanic matrix.
In the upper levels the quartz-adularia zone is absent, and narrow quartz-illite halos grade outward to
propylitic alteration characterized by chloritic alteration of mafic minerals, and illite alteration of plagioclase.
With depth, quartz-adularia alteration becomes more abundant, grading outward to quartz-illite and finally to
propylitic alteration.
District-wide propylitic alteration predated the mineralization, whereas quartz-adularia and quartz-illite
alterations are closely related to the mineralization and probably formed simultaneously.

Mineralogy and paragenesis


A complex paragenetic sequence has been identified (Fig. 12) from detailed underground mapping
and microscopic studies of samples from the Bateas, San Cristóbal, and Animas veins. Four main stages of
mineral precipitation have been recognized (early stage, manganese stage, quartz stage, and late stage), and
two of them have complex histories with substages.
Veins display complex textures characteristic of episodic, open-space precipitation such as
crustiform banding, symmetric banding, vugs, breccias, and cockade and comb textures. In general, the more
complexly banded veins have higher ore grade. Breccias consist of angular host rock or vein clasts up to 50
cm in diameter cemented by vein material that exhibits cockade texture. Symmetric crustiform banding is the
most typical texture, with the early stages of deposition arranged toward the wall rocks and the younger stages
in the center of the veins (Fig. 11). Principal gangue minerals are quartz, manganese-rich silicates (dominated
by rhodonite, but including bustamite and johansenite), rhodochrosite, pyrite, and calcite. Variable amounts of
adularia, sericite, barite, and helvite also are present. Common ore minerals include sphalerite, galena,
chalcopyrite, and tetrahedrite (freibergite). Subsidiary ore minerals include polybasite, stephanite, argentite,
native silver, pyrargyrite, miargyrite, chalcocite, native gold, boulangerite, stibnite, alabandite, and other Pb
and Ag sulfosalts (Stephan, 1974). In general, silver ore minerals form thin bands, with minor or no sulfide
other than pyrite. The quartz stage is an exception, with some coarse-grained sulfides, mainly sphalerite,
disseminated in the gangue. In general, the order in which the sulfide and sulfosalt minerals precipitated is
essentially the same throughout the deposit, with early sphalerite and probably some pyrite, followed by
galena. Chalcopyrite and Ag sulfosalts are among the last ore minerals to precipitate. Chalcopyrite displays

5
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

disease texture and replaces sphalerite. Some pyrite encloses galena and sphalerite, indicating that it also
precipitated after base metal sulfides. Early sulfides are usually coarser than subsequent sulfides.
Early stage: The early stage consists of sugary quartz and gray quartz, chalcedony, and disseminated
pyrite with common colloform-banded texture. Pyrite is coarse grained in white quartz, but is finely
disseminated in gray quartz. Colloform banding presents a slight color variation between white, pinkish, and
gray. This stage is also related to wallrock silicification and quartz-adularia alteration, which consists of
quartz veinlets with massive, comb, and recrystallization textures. Some veinlets contain rhombic adularia,
which crystallized before quartz. Coarse-grained pyrite and illite are abundantly disseminated in the silicified
host rocks. The early stage forms a discontinuous band up to 15 centimeters thick in contact with the
wallrocks in Bateas (Fig. 11), San Cristóbal, and Animas veins, but can be absent in other veins where it is
represented only by wallrock silicification. Early stage bands do not host economic mineralization and Ag
grade, in general, is less than 1 oz/t (mine staff, pers. commun., 2002).
Manganese stage: The manganese stage is the most complex stage due to the variety of minerals
present and the cyclic character of formation, and also contains the most sulfides and constitutes the best ore.
Crustiform banding is the prevailing texture of the manganese stage, although breccias, veinlets, and cockades
also are locally developed. The Bateas (Fig. 11), San Pedro, El Toro, and Carolina veins are characterized by
symmetric banding, whereas in the San Cristóbal vein, breccia is more common. The vein width of the
manganese stage varies from a fe w centimeters to several meters.
The manganese stage is divided into three substages (Fig. 12). The earliest substage consists of
abundant calcite, rhodonite, and quartz, with variable amounts of rhodochrosite, adularia, and helvite arranged
in cyclic bands a few to tens of centimeters thick. Calcite and rhodonite are early minerals that commonly
cement clasts of wallrock or early-stage minerals . Rhodochrosite follows and replaces rhodonite, whereas
quartz is the last phase to precipitate. Adularia, when present, follows rhodonite and precedes quartz. Lemon-
yellow helvite forms thin bands, 1 to 2 mm thick, intercalated in thin bands with adularia, quartz and
rhodonite. Sulfides are uncommon in the earliest substage; they are very scarce within calcite or manganese
minerals, and they are usually found disseminated within fine-grained quartz, but are totally absent within
coarse-grained comb quartz.
The second substage is sulfide-rich and consists of bands of ore minerals from 1 to 20 cm thick.
Earlier bands are coarser with higher sphalerite content and galena, whereas later bands are rich in fine-
grained chalcopyrite and tetrahedrite. The ore mineral precipitation sequence, as discussed above, has early
sphalerite replaced by chalcopyrite and tetrahedrite. Sulfide bands contain, and are enveloped by, abundant
quartz and lesser amounts of rhodonite and carbonates.
The last substage consists of quartz, rhodonite, and calcite bands with scarce disseminated sulfides.
A mineralogical evolution with quartz and rhodonite becoming less abundant and calcite more abundant with
time was identified by Stephan (1974). Veinlets of quartz and rhodonite from the last substage commonly
crosscut earlier substages, and quartz-rhodonite breccia clasts from the first substage are cemented by
minerals of the last substage.
Microstages within the manganese stage: Each of the manganese substages is formed by a
succession of small-scale bands (microstages) that commonly follow the same order of precipitation. Figure
13 schematically shows bands representative of a complete cycle of precipitation during substage 2
(microstages S1-S6), which consist of 2 sub-cycles, each formed by 3 microstages . The microstages are
repetitive and are a few mm to cm thick. The early microstage (S1) consists of coarse-grained, brown to
amber sphalerite, followed by finer-grained pyrite and chalcopyrite plus tetrahedrite. Next is coarse-grained
crystalline quartz (S2). S3 consists of narrow crustiform bands of fine-grained rhodonite plus rhodochrosite,
commonly intercalated with calcite and/or recrystallized quartz from amorphous silica; locally the S2 quartz
is absent and the S3 rhodonite bands are in contact with the S1 sulfide minerals. Isolated crystals of rhombic
adularia are present in contact with the rhodonite. S4 marks a renewed subcycle of disseminated fine-grained
sulfides, where a transition from early sphalerite to late chalcopyrite commonly was identified. Sulfides are
enveloped by milky, fine-grained quartz. S4 evolved to clearer and coarser-grained quartz in the S5
microstage, with a notable lack of disseminated sulfides. Finally, S6 is characterized by rhodonite which can
form a massive fine-grained band (S6a in Fig. 13) or a chaotic network of millimetric acicular crystals (S6b in
Fig. 13).
Quartz stage: The quartz stage consists of two substages that follow the manganese stage (Fig. 12).
The first substage is composed of medium-grained sulfides disseminated within quartz. The sulfides are
principally pyrite, sphalerite, galena, and minor tetrahedrite that precipitated prior to quartz in euhedral to
subhedral grains forming clusters or irregular veinlets, which are then enveloped by quartz. Quartz is massive

6
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

or rarely banded, fine to medium grained, milky or gray due to fine disseminated sulfides, and locally replaces
bladed calcite. Minor rhombic adularia is present, and always precedes quartz.
The last substage consists of coarse-grained, clear, amethyst quartz in veinlets with comb texture that
crosscut the first substage. The lack of sulfides is typical of this substage.
The quartz stage is developed towards the center of some veins and in contact with the manganese
stage from which gradational contacts were observed. It is also present as cement in breccia with clasts of the
manganese stage. The quartz stage was recognized in the Bateas and Animas veins as an irregular and
discontinuous band more than 50 cm in width (Fig. 11), whereas in the San Cristóbal vein it is present as
veinlets crosscutting the manganese stage.
Late stage: The late stage is the latest stage of mineral deposition recognized in the district. It
consists of coarse-grained, white, and massive calcite followed by milky to clear medium-grained quartz
filling vugs and open spaces. This stage is present in the center of the Bateas vein in bands up to 80 cm thick
(Fig. 11).

Mineral zonation
Variations in mineral composition and metal content, primarily within the manganese stage which
hosts ore, show a horizontal and vertical zonation, both at the district and the vein scale.
District-scale horizontal zonation: Surface mapping and sampling show changes in mineral and
chemical composition across the district (Table 1). In a NW to SE section across the veins, Ag and base metal
content diminish to the SE. Northern veins (north of San Cristóbal vein, Fig. 2) are characterized by
crustiform banding with abundant calcite, manganese minerals and quartz, with high sulfide content at the
surface. Southern veins (south of San Cristóbal vein, Fig. 2) are characterized by minor calcite and manganese
minerals, massive fine-grained quartz with stibnite and barite and almost no Ag and base metal sulfides at the
surface.
A section along the veins from the NE to the SW shows a similar pattern of zonation. The eastern
portion of the veins at surface has crustiform banding characterized by rhodonite-rhodochrosite, quartz, and
abundant base metal sulfides. To the west, especially at Vilafro Hill (Fig. 2), the Ag and base metal minerals
disappear, and the veins contain only fine- to medium-grained white quartz.
Bateas vein vertical zonation: The Bateas vein, as well as most veins in the district, strike NE and
dip SE. Vein thickness varies between 30 cm and 3 m, and is mostly controlled by changes in strike and
probably dip of the vein. The Bateas vein is hosted by a silicified andesite flow and contains symmetric bands
of all the mineral stages identified in the district (Fig. 11). These bands are cut by quartz-rhombic adularia
veinlets. The early stage is represented by a 15 cm wide band of fine-grained white to gray colloform-banded
quartz in contact with the wallrock. The three substages of the manganese mineral stage are well developed
and form about 80% of the vein volume. The intermediate sulfide substage is a 30 cm band of crustiform
sulfides intercalated with quartz and rhodonite. Early bands are sphalerite-galena rich and late bands are
chalcopyrite-pyrite rich (Fig. 11). The center of the vein is occupied by a thick and discontinuous band of
white quartz with disseminated sulfides of the quartz stage followed by calcite with scarce quartz of the late
stage.
The Bateas vein is unique in the district due to a remarkable vertical zonation in which a tetrahedrite-
rich horizon with high Ag grade was identified (Fig. 8). The vertical zonation was recognized within the
manganese stage and is characterized by three approximately horizontal zones with transitional boundaries
(Fig. 8; Tables 1 and 2). The mineral composition and metal content vary significantly within a vertical range
of 200 m. The sulfide content increases gradually with depth. The upper zone is present mostly above level
12A (4475 m.a.s.l.), and is characterized by banded quartz-rhodonite with minor calcite; Ag is contained in
pyrargyrite and miargyrite disseminated within quartz. Scarce base metal sulfides and tetrahedrite are present
in minor sulfide bands. The intermediate zone is present mostly between levels 12A and 13A (4425 m.a.s.l.)
and has a vertical range between 30 and 50 meters for more than 400 meters horizontally; it is characterized
by more abundant and finely banded manganese minerals (rhodonite-rhodochrosite). Silver is contained in
coarse-grained tetrahedrite in bands several mm to 1 cm thick. The tetrahedrite is locally intergrown with
sphalerite, galena and quartz. In the upper zone the sulfides are mostly disseminated within quartz, resulting
in an unclear differentiation of the three substages of the manganese stage, whereas in the intermediate zone a
30 cm-thick band of massive sulfides makes up the intermediate substage and separates the almost barren
early and late substages. In the lower zone, below level 13A, the massive sulfide band is still recognizable,
although it consists of mainly base metal sulfides (sphalerite, galena, chalcopyrite and pyrite) with minor
tetrahedrite in quartz gangue. Rhodonite and rhodochrosite are still abundant, though unrelated to the ore.

7
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

La Plata vein vertical zonation: La Plata vein, located in the south of the Caylloma district, strikes
about 050° and dips 50° to the SE. Due to the high topographic relie f it is exposed over a vertical distance of
100 m and a horizontal distance of 250 m. The topographically high outcrops consist of fine- to medium-
grained, gray, massive to weakly banded quartz with disseminated pyrite, or of wallrock breccias cemented by
the described quartz. Chip and channel samples show low Ag grades of less than 3 oz/t (Fig. 14). At depth, as
well as in the topographically low outcrops, the vein widens to 2 m and the mineralogical and metal
compositions are completely different. In these areas the vein consists of the typical manganese stage
crustiform bands of alternating quartz, rhodonite-rhodochrosite with disseminations and bands of sulfides.
Also, the vein varies from quartz-dominant in the upper levels to rhodonite-dominant at depth. Within the
manganese stage the Ag grade is 8 to 25 o z/t.

Metal content and distribution


Caylloma is a Ag-rich epithermal system with subordinate Au, Zn, Pb, and Cu. Economic grade is
almost exclusively present in the manganese stage. Mineral zonation is also accompanied by metal zonation,
producing a large variation in metal content and metal ratios both vertically and horizontally, across and along
the veins.
Three vertical zones are recognized; upper, intermediate, and lower. The upper zone of the San
Cristóbal vein, down to about 150 meters below surface, is characterized by a low Ag/Au ratio (<100). The
low value was interpreted by Stephan (1974) as being produced by oxidation and Au enrichment of the upper
portions of the veins. Below the upper (oxidized) zone the Ag/Au ratio increases to 500 to >1000 within the
manganese stage. This intermediate zone, which is about 200 m in vertical extent, is the focus of most mining.
The average Ag content is about 16 oz/t, the average Au content is 0.3 to 0.7 g/t, and the base metal content
(Zn+Pb+Cu) is less than 3%.
In the intermediate zone of the Bateas vein (Table 2), the average Ag grade is 75 oz/t, with ore shoots
reaching average grade of 100 oz/t, and maximum values over 1000 oz/t. The base metal content in this zone
is >18%. In the lower zone of the Bateas vein the Ag grade decreases to values near 6 oz/t, and the total base
metal content is ~6%, with increased ratios of Zn/Ag, Pb/Ag, and Cu/Ag.
A section across the district shows a metal zonation from base metal-rich veins in the north (e.g., San
Pedro, Paralela, and El Toro veins), through Ag-rich veins in the central zone (e.g., San Cristóbal and Bateas
veins), to Sb-rich veins in the south (e.g., Antimonio and Corona Antiminio veins). The base metal
distribution shows a zonation within the base metal-rich northern zone. Zinc and Cu are more abundant to the
north, whereas Pb and Ag contents increase toward the south.
A 3 km-long section along the San Cristóbal vein also reveals a metal zonation, with a relatively high
base metal content in the NE (~3%), Ag content increasing in the central portion, and finally Sb appears
towards the SW extreme, in miargyrite, pyrargyrite, and other sulfosalts (Stephan, 1974).

Age Determination of Host Rock and Mineralization


Age control on volcanic rocks and mineralization of the Caylloma district is based on four new
40
Ar/ 39 Ar ages determined at the USGS laboratory in Denver, Colorado. The samples include adularia in the
altered host rock, the matrix (nearly 100% fine plagioclase) of an andesite flow, and sanidine and biotite in
the post-mineralization Cuchilladas rhyolite dome. Samples were prepared at the USGS Ar/Ar laboratory
using standard methods of mineral separation including heavy liquids, magnetics, and hand picking. Due to
the coarse-grained nature of the adularia , biotite and sanidine, these concentrates had few or no impurities.
The new 40 Ar/39Ar dates are compared with published (Peterson et al., 1983; Silberman et al., 1985; McKee
and Noble, 1989) and unpublished (Noble, 1981a, 1981b) K-Ar dates of mineralization and volcanic rocks of
the area (Fig. 15).
The age of the rocks that host the mineralization was obtained from a sample collected from a thick
sequence of andesitic flows that are intercalated with ash flow tuffs. The sequence corresponds approximately
to the center portion of the Lower Miocene stratigraphic sequence. The volcanic matrix separate yielded an
age of 20.30±0.11 Ma.
The age of mineralization, obtained from coarse-grained adularia in the wall rock of the Soledad
vein, is 18.35±0.17 Ma. Because published and unpublished K-Ar ages (Noble, 1981a, Silberman et al., 1985,
McKee and Noble, 1989) are younger, we interpret the age difference to reflect argon loss during argon
extraction for the K-Ar analysis .

8
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

The age of post-mineralization volcanism was obtained from felsic volcanic domes and dikes in the
Caylloma district (Fig. 2) which yielded ages of 11.80±0.80 Ma and 12.25±0.07 Ma. Thus these volcanic
rocks were emplaced several million years after the mineralization and are interpreted to be related to the
Chonta caldera (Figs. 1, 3, and 15). The main collapse event of the Chonta caldera is dated at 11.4±0.4 Ma
based on the age of a thick ash flow tuff (Peterson et al., 1983). Post-collapse dacite and andesite flows and
domes are dated at 11.6±0.4 and 11.3±0.3 Ma, respectively (Peterson et al., 1983). Several epithermal
deposits dated by Peterson et al. (1983) between 11.7±0.5 and 10.5±0.3 Ma were suggested to be related to
the Chonta caldera (Fig. 15). The last volcanic event of the area is the formation of the Caylloma caldera that
occurred between 4.4±0.1 and 2.4±0.07 Ma, (Noble et al., 1981b) with several episodes of ash flow tuff
eruption and caldera collapse that are superimposed on the Lower and middle Miocene volcanism.

Fluid Inclusions
Flu id inclusions were studied in sphalerite, quartz, and calcite in samples from the manganese, quartz
and late stages of the Bateas vein. Fluid inclusions in coarse-grained sphalerite of the manganese stage are
abundant, range from 10 to 60 micrometers in diameter, and are tubular or negative-crystal shaped. Most
inclusions in sphalerite are secondary, whereas primary inclusions are rare and isolated. Primary fluid
inclusions in medium- to coarse-grained quartz of the manganese and quartz stages are abundant and occur
along growth lines. In general, quartz crystals have inclusion-free clear cores and dark rims with many
irregular shaped inclusions with a radial geometry. Fluid inclusions in calcite of the late stage are rectangular,
between 5 and 40 micrometers in size, and mostly are present along the calcite cleavage planes.
Thermometric data were obtained using a FLUID INC. adapted U.S.G.S. Gas-Flow Heating/Freezing
System on an Olympus BX51 microscope. Data were obtained on individual planes or groups of inclusions
(fluid inclusions assemblages, FIA; Goldstein and Reynolds, 1994), mainly of primary origin, which had
consistent liquid-to-vapor ratios and yielded consistent data. All the studied inclusions are two-phase liquid-
rich inclusions, with homogenization always to the liquid phase.
FIA measured in sphalerite yielded a homogenization temperature (Th) range between 310° and 240°
C, with consistent high salinity on the primary FIA with ice melting point temperature (Tm) around -21.2° C,
which corresponds to a salinity of 23.5 wt % NaCl equivalent (Fig. 16). Secondary FIA in sphalerite show
the same range of Th, though the Tm are significantly lower, between -8.1° and -2.4° C.
Primary FIA in quartz for microstages S2, S4, and S5 of the manganese stage yielded a Th range
between 225° and 308° and the Tm between -0.3° to -4.1°, indicating low salinities, between 0.5 and 6.6 wt %
NaCl equivalent (Fig. 16). The S2 microstage is characterized by coarse-grained crystalline quartz with many
primary fluid inclusions, mainly on growth lines towards the rim of the crystals. The inclusions are large,
about 15 micrometers, with a liquid-vapor proportion of about 80%. The cores of the S2 quartz yielded the
highest Th found in the system for quartz crystals, 275° C for primary FIA and 310° for secondary FIA, with
Tm of -2° and -4.1° C, respectively. However, the rims of the same crystals yielded Th of 220° to 230°, and a
salinity of 6.6 wt.% NaCl equivalent (Fig. 16), indicating a drop in temp erature, as well as an increase in the
salinity.
In the S4 microstage, two types of quartz were identified, an early medium-grained, clear quartz with
few fluid inclusions and no disseminated sulfides, and a late fine-grained quartz with abundant small fluid
inclusions and fine-grained disseminated sulfides (Fig. 13). The fluid inclusions are irregular or rectangular,
between 5 to 10 micrometers in diameter and contain a liquid proportion of about 85%. There is a slight
decrease in Th from the early to the late quartz, and an increase in salinity (Fig. 17).
Finally, the S5 microstage consists of medium-grained massive quartz, with abundant primary fluid
inclusions. The inclusions are irregularly shaped, about 15 micrometers in diameter and contain a liquid
proportion of 80%. The FIA measured in this microstage also show a decrease in Th, from 275° to 255° C,
and an increase in salinity from 0.7 wt.% to 4.1 wt.%, from the early quartz to the late quartz (S5a to S5b in
Fig. 17).
Medium- to coarse-grained massive quartz with disseminated sulfides of the quartz stage contains
numerous small primary inclusions with a high liquid proportion (~85%). Studied FIA yielded Th between
235° and 265° C, and Tm between -4.4° and -5° C (Fig. 16). Fluid inclusions studied in the late stage calcite
yielded Th between 235° and 275° C, and Tm between -0.6° and -1° C (Fig. 16).

9
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

Sulfur Isotopes
Sulfur isotopes are commonly used to constrain the source of sulfur, and to infer the source of
metals, in the hydrothermal system; sulfide pairs can also be used as a geothermo meter. We conducted nine
analyses of δ34 S in pyrite, galena and sphalerite from three samples from the Bateas and San Cristóbal veins
(Table 4).
The d 34 S values for the entire set of analyzed sulfides have a tight composition range between -3.2
and 1.1 per mil. Values of d 34 S of the H2 S in isotopic equilibrium with the sulfides where calculated using a
temperature of 275° C and the equation of Ohmoto and Rye (1979). The d 34 SH2S calculated is between -2.9
and 2.6 per mil.
The pyrite-galena and sphalerite-galena sulfide pairs gave temperatures between 380° and more than
500° C, which are more than a 100° C higher than the temperatures obtained from the fluid inclusion study.
The disparity between temperatures obtained from sulfur isotope geothermo metry and those obtained from
fluid inclusions has been noted by other authors (e.g., Rye and Ohmoto, 1974, Field and Fifarek, 1985) and
can be the result of isotopic non-equilibrium during sulfide precipitation. Indeed, the mineral paragenesis
shows that the sulfides are seldom in equilibrium, being more common the replacement of one sulfide by
another.

Discussion

Tectonic setting
Intermediate sulfidation epithermal veins of the Caylloma d istrict were formed at 18.86±0.1 Ma (Fig.
15) within a continental calc-alkaline volcanic arc. Although this date is about 1.4 Myr younger than the dated
host rocks, it is suggested that volcanic rocks and mineralization could be related to the same magmatic event.
Hydrothermal activity that produces epithermal deposits commonly begins a short time after the timing of
associated silicic to intermediate volcanism. In high sulfidation deposits , hydrothermal activity closely
follows the emplacement of volcanic host rocks by 0.1 to 0.5 Myr, such as the case for Goldfield, Nevada
(Ashley and Silberman, 1976), Julcani, Peru (Noble and Silberman, 1984), El Indio, Chile (Sillitoe, 1991),
and Rodalquilar, Spain (Arribas et al., 1995). However, other epithermal deposits, including Caylloma and
Orcopampa, Peru (Gibson et al., 1995), Pachuca-Real del Monte, Mexico (McKee et al., 1992), Guanajuato,
Mexico (Gross, 1975), San Dimas, Mexico (Enriquez and Rivera, 2001), Comstock, Nevada (Vikre et al.,
1988), Creede, Colorado (Bethke et al., 1976), and Baia Mare area, Romania (Lang et al., 1994) show a time
gap of 0.5 to more than 3 Myr between the emplacement of the youngest volcanic rocks and the
mineralization. Such a period of volcanic inactivity might reflect the presence of buried intrusive rocks,
slightly younger than the youngest exposed volcanic sequence (McKee et al., 1992) that triggered the
hydrothermal activity.
Early Miocene volcanism in southern Peru is related to other well-known epithermal deposits, such
as those in the Orcopampa district (Gibson et al., 1995), which represents the oldest Neogene epithermal
deposits known in southern Peru. The renewal of widespread volcanism and mineralization during the Early
Miocene, after a period of quiescence during the Oligocene (Petersen, 1958, Noble et al., 1974, McKee and
Noble, 1982, 1989, Megard et al., 1984), may be related to an increase in the Pacific plate absolute rotation
(Clague and Jarra rd, 1973), which resulted in renewed or more rapid subduction beneath the Central Andes
(Pilger, 1983, 1984, Pardo Casas and Molnar, 1987, Sebrier and Soler, 1991) that ulitmately generated the
volcanism (Noble et al., 1974).
The northeast-trending oblique normal faults that host ore in the Caylloma district were formed
during a period of extension, which spans the period between the late Oligocene Aymara compressional event
(ca. 26 Ma; Sebrier et al., 1988, Machare et al., 1986, Sebrier and Soler, 1991) and the Miocene Quechua
compressional phase (ca. <15 Ma; Farrar and Noble 1976, McKee and Noble, 1989, Megard et al., 1984,
Sebrier and Soler, 1991).

Structural setting
Veins of the Caylloma district formed during an episode of extension, as interpreted from vein
textures (crustiform banding, breccias, cockade texture, symmetric banding), the great thickness of some open
space fill veins, and the fact that the veins are hosted by oblique normal faults and related extension fractures.
We propose a structural model in which the principal veins formed in sub-parallel NE-striking dextral normal
faults and related extension fractures within a NW-striking structural corridor bounded by sinistral regional

10
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

faults (Fig. 6). Several orders of faults, with related extensional fractures and block rotations have been
identified. Movement along first-order NW-striking sinistral faults produced counterclockwise rotation and
the development of second order NE-striking dextral normal faults and third-order extension fractures that
host mineralization. Movement along the second-order NE-striking faults in turn produced clockwise block
rotations and a sinistral component of movement along the third-order extension fractures, with the
development of fourth-order extension fractures as splays off the third order veins (Fig. 6).
Kinematic analysis of structures hosting mineralization in the Caylloma district (Fig. 4) shows that
the plunge and trend of the extension and shortening axe s responsible for their formation were 0° 330° and
55° 234°, respectively. Although the orientation of extension during the time span between the short-lived
Miocene compressional events is not well-known, the data presented here are in agreement with kinematic
indicators from the Lower Miocene from other sites of the high Andes of southern Peru (e.g., Gibson et al.,
1995).
The position of some ore shoots is controlled by syn-mineral movement on NW-striking transverse
normal faults (Fig. 9). This movement caused localized extension in the hangingwall of the transverse fault
and produced fractures, bends, and open spaces enhancing the hydrothermal fluid flow and promoting mineral
precipitation, leading to high-grade and thick ore shoots called half cymoids.

Mineral and metal zonation


Mineral and metal zonation show a similar behavior across the district (from north to south), along
the principal veins (from the NE to the SW), and from deep to shallow levels of the mine (especially at Bateas
vein; Table 1). The northern veins of the district, the NE extreme of the major veins (e.g., San Cristóbal, El
Toro, Animas, San Pedro) and the deeper portions of the veins (e.g., Bateas) contain rhodonite-
rhodochrosite+calcite±quartz with relatively high base metal content. The central zone, both vertical and
horizontal, is Ag-rich with rhodonite-rhodochrosite and quartz as gangue minerals. Finally, the southern veins
(e.g., La Plata, Antimonio), the SW extreme of the veins (e.g., San Cristóbal), and the upper portion of the
veins, dominantly contain quartz with Sb in stibnite or sulfosalts. Shallower levels of the hydrothermal system
are thus exposed to the west and southwest in the district, whereas deeper base metal-rich levels are exposed
at the surface to the east and north, indicating a tilting of the mineralization zones to the west-southwest. This
tilting is the result of a post-mineral regional compressive tectonic event that occurred probably between 12
and 6 Ma (Farrar and Noble 1976, McKee and Noble, 1989, Swanson et al., 1993) and is assigned to the
Quechua 2 phase of deformation (Megard et al., 1984). A west dip between 10° and 20° of the Orcopampa ore
bands is also related to the Quechua 2 compressive phase (Gibson et al, 1995).
Although mineral and metal zonation in longitudinal sections, such as those shown for Bateas and La
Plata veins (Figs. 9 and 10, Table 2), are near horizontal or slightly inclined to the west, most ore-shoots have
nearly 90° rakes. This suggests that the hydrothermal fluid flow which generated the mineralization may have
had a strong vertical component. The 90° rakes of the first and/or second-order ore shoots are in agreement
with the dominantly strike -slip sense of movement of the faults that host the mineralization. Strike-slip faults
produce jogs with steeply dipping axe s, and good vertical connectivity (Cox et al., 2001) that control the fluid
flow generating narrow and near-90° rake ore shoots. The hydrothermal fluids thus ascended through
enhanced vertical structural permeability, and precipitated the ore and gangue minerals arranged in sub-
horizontal zones.

Fluid and metal source


Fluid inclusion analysis shows a wide range of salinities, from 0.5% to 7.5 wt% NaCl equivalent in
quartz and between 4 and 23 wt% NaCl equivalent in sphalerite. Consistently higher salinities in sphalerite
than in co-existing quartz also has been described in other epithermal deposits that are either base metal rich
or precious metal rich with subordinate amounts of base metals, such as Fresnillo, Coneto de Comonfort,
Durango, Real de Guadalupe, Guerrero, and La Guitarra, all in Mexico (Simmons, 1991; Albinson et al.,
2001).
The high salinity fluid recognized in sphalerite-hosted fluid inclusions can be interpreted as the
parental fluid and could be derived from either a magmatic source, from interaction with evaporitic deposits,
or from the influence of formation waters. However, evaporites are not known in the stratigraphic sequence
related to the emplacement of the Caylloma mineralization, and saline formation waters are unlikely in
terranes with no evaporites or sea water circulation (Albinson et al., 2001). In this type of environment with
no circulation of sea water and lack of evaporites at depth, the geothermal waters in volcanic-hosted systems

11
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

contain only up to 2 wt% NaCl equivalent (Simmons, 1995). Therefore, the presence of high salinity fluids in
the Caylloma district is more likely related to brines exsolved from a magmatic source.
The d 34 SH2S values of the analyzed sulfides, and by inference, the hydrothermal fluids, were between
-2.9 and 2.6 per mil, consistent with a magmatic source for the sulfur, as the composition of orthomagmatic
total sulfur in igneous rocks is about -3 to 3 per mil. However, the assumption of magmatic source for sulfides
with δ34 S near 0 per mil is a hazardous generalization when applied to hydrothermal environments, and it is
valid only assuming that the H2 S concentration is equal to that of total sulfur in the system (Field and Fifarek,
1985).
Albinson et al. (2001) compiled information for several epithermal deposits of Mexico and found
that the higher salinity fluids were present in Ag- and base metal-rich deposits with Ag/Au ratios greater than
100, comparable with the Caylloma deposit. Based on geochemical indicators such as salinities, gas
signatures , and isotopic composition they also proposed a magmatic source for the ore-forming fluids in those
deposits.
Tosdal et al. (1995) carried out lead isotope studies on samples of the Caylloma, Orcopampa, Shila
and Arcata epithermal districts of southern Peru. Although they focused on the Orcopampa mineralization,
and only studied a limited suite of samples from the other sites (only three samples from Caylloma), they
concluded that the metal source was the same for all the studied locations. They proposed that metal-bearing
hydrothermal fluids originated in the Miocene plutons, and then obtained more radiogenic lead through fluid-
rock interaction, especially from the lower Paleozoic basement.

Fluid evolution and mineral precipitation


Veins show a complex paragenetic evolution; however, most of the ore is located only in the
manganese stage, whereas the other stages are mostly barren. The manganese stage shows a cyclic evolution
with six identified microstages arranged in a cycle composed of two subcycles (Fig. 13). Each cycle is a few
mm to 5 cm thick, and repeats successively to form ore bands of up to 1 meter thick. Each subcycle is formed
by early sulfides, followed by coarse or medium-grained quartz, and finally ending with bands of
rhodonite+calcite+amorphous silica. The sulfides commonly form thin bands, where the early bands are
coarser-grained sulfides and the later bands are finer-grained sulfides. There is also a compositional variation
of sulfides in the bands; the early sulfides are mostly sphalerite and galena, evolving to chalcopyrite and silver
sulfosalts in the later bands.
The transitional but rapid passage between different near-horizontal mineral and metal zones
suggests that precipitation was caused by abrupt changes in fluid conditions during ascent, whereas the
alternating occurrence of quartz-rich and rhodonite-rich bands may indicate a cyclic change in fluid
composition. We present below the evidences that favor boiling and/or renewed fluid injection as the causes
that led to the observed mineral zonation and banding.
Mineral associations, such as rhodonite+chalcedony+calcite followed by adularia , present in some
bands of the manganese stage suggest that their precipitation may have been caused by boiling. Rhodonite is a
very common vein mineral in the Caylloma district; it appears as bands of massive fine-grained or needle-like
crystals, sometimes replaced by rhodochrosite. Rhodonite is often intergrown with minor amounts of fine-
grained bladed or massive calcite and recrystallized amorphous silica, and is followed by adularia. According
to Gammons and Seward (1996), the solubility of rhodonite is retrograde, and thus cannot precipitate during a
simple episode of cooling. Rather, its precipitation is more likely caused by a sudden increase in pH (from
fluid boiling), or a decrease in chloride concentration (from fluid dilution) (Leroy et al., 2000). Calcite is
another retrograde mineral, and in epithermal veins calcite precipitation is most likely caused by the loss of
CO2 due to boiling, and the subsequent generation of CO3 2- from the dissociation of HCO3 - (Henley, 1985,
Reed and Spycher, 1985). Furthermore, in many active geothermal systems bladed calcite is commonly
restricted to the boiling zone (Browne, 1978, Simmons and Christenson, 1994). Amorphous silica
precipitation is often interpreted as a result of rapid cooling due to decompressional boiling (Fournier, 1985,
Drummond and Ohmoto, 1985, Saunders and Schoenly, 1995). Finally, adularia is commonly found in
epithermal veins, and is also related to boiling, resulting in increase in the fluid pH due to loss of CO2 to the
vapor phase (Browne and Ellis, 1970).
Evidence for boiling from fluid inclusions in Caylloma samples is not conclusive. Coexisting planes
of fluid-rich and gas-rich inclusions were not identified. Also, Th decreases and salinity increases, which
exceed 50° C and 2.5 wt % NaCl equivalent, respectively, are observed in the same quartz crystal from the
core to the rim (S2) or from early to late quartz within the same microstage (S5). These changes can be
interpreted either as the result of loss of liquid to the vapor phase due to non-equilibrium boiling (local boiling

12
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

towards dryness; Simmons and Browne, 1997), or due to injection of more saline fluid into the hydrothermal
system.
Although we are in favor of sporadic boiling to explain the presence of rhodonite+amorphous
silica+calcite bands after a drop in temperature and an increase in salinity, there is no relationship between
these bands (where boiling might have occurred) and the ore minerals. The microstages with evidence of
boiling lack sulfides, and the microstages where the sulfides are present do not show any evidence of boiling.
The fluid inclusions in quartz show a slightly higher salinity (4 to 7 wt% NaCl equivalent) for the microstages
that carry the sulfide minerals than the microstages immediately following sulfide precipitation, such as S2a
and S5a, with average salinities of ~2.7 wt% NaCl equivalent. The fluid inclusions in the quartz stage, which
carries disseminated sulfides, also show salinities (7 to 8 wt% NaCl equivalent) higher than the quartz without
disseminated sulfides, whereas the fluid inclusions in calcite from the latest quartz-calcite stage show that
they were precipitated from a dilute fluid (1.4 wt% NaCl equivalent).

Deposit model
The high salinity of the fluids and the lead isotope evidence support the contribution of magmatic
fluids into a hydrothermal system through channels of enhanced structural permeability. Although boiling
may have occurred at some time during the system evolution, it is not related to ore deposition. The slightly
higher salinity of quartz-hosted FI in stages with sulfides present, together with the high salinity of sphalerite-
hosted FI, suggest that new injection of magmatic fluid may have caused ore mineral precipitation. The cyclic
nature of the mineralization with abrupt changes in mineralogy, textures, and fluid composition leads us to
postulate that cyclic injections of magmatic -derived hydrothermal fluids occurred, as argued by Simmons
(1991). These injections may have been related either to complex history of magma crystallization and fluid
exsolution, and/or to repeated pulses of deformation and related enhancement of structural permeability. A
feedback mechanism may have operated in which exsolution of magmatic fluid drove fluid pressure cycling
leading to faulting events during high fluid pressure and low effective normal stress conditions. Faulting
events thus allowed fluid migration and precipitation of individual mineral bands.
If we consider a boiling fluid at hydrostatic pressure at 270° C and with 2.7 wt.% NaCl equivalent as
responsible for most of the gangue minerals precipitation, the depth of formation for the interval with the
economic mineralization was about 640 meters below the paleowater table .

Conclusions
Intermediate sulfidation epithermal veins of the Caylloma district formed at 18.86±0.1 Ma within a
continental calc-alkaline volcanic arc. Northeast-striking veins formed in dextral normal faults and associated
second- and third-order structures within a NW-trending structural corridor, and ore shoots are localized in
fault jogs and in the hangingwall of transverse faults where structural permeability was enhanced. Banding in
veins formed from mineral precipitation during cyclic injection of magma-related hydrothermal fluids at
temperatures of ~270°C and depths of ~640m below the paleowater table. Post-mineral tectonism resulted in
tilting of district-wide mineral and metal zones with shallower levels now are exposed to south and southwest.

References
Albinson T., Norman D. I., Cole D., and Chomiak B., 2001. Controls on formation of low-sulfidation epithermal deposits
in Mexico: constraints from fluid inclusion and stable isotope data. In: Albinson T. and Nelson C. E (eds.). New mines
and discoveries in Mexico and Central America. SEG Special Publication 8: 1-32.

Ashley, R.P and Silberman M. L., 1976. Direct dating of mineralization at Goldfield, Nevada, by potassium-argon and
fission-track methods. Economic Geology, 71: 904-924.

Arribas, A. Jr., Cunningham C. G., Rytuba, J. J., Rye, R. O., Kelly, W. C., Podwysocky M. H., McKee E. H., and Tosdal,
R. M., 1995. Geology, geochronology, fluid inclusions, and isotope geochemistry of the Rodalquilar gold alunite deposit,
Spain. Economic Geology, 90: 795-822.

Bethke P. M., Barton P. B. Jr., Lanphere M. A., and Stevens T. A., 1976. Environment of ore deposition in the Creede
mining district, San Juan Mountains, Colorado: II. Age of mineralization. Economic Geology, 71: 904-924.

Browne, P. R. L., 1978. Hydrothermal alteration in active geothermal fields. Annual Reviews in Earth and Planetary
Sciences, 6: 229-250.

13
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

Browne, P.R.L. and Ellis, A.J., 1970. The Ohaaki Broadlands Hydrothermal Area, New Zealand: Mineralogy and Related
Geochemistry. American Journal of Science, 269:97-131.

Candiotti de lo Rios H. and Guerrero T., 2002. Low sulphidation epithermal gold-silver veins in the Ares deposit,
southern Peru. XI Congreso Peruano de Geología. Nuevos descubrimientos en el Perú, Bolivia y Chile Symposium.
Sociedad Geológica del Perú, Abstracts.

Candiotti de los Rios H, Noble D. C., and McKee E. H., 1990. Geologic setting and epithermal veins of the Arcata
District, southern Peru. Economic Geology, 85: 1473-1490.

Cassard D., Chauvet A., Bailly L., Llosa F., Rosas J., Marcoux E., and Lerouge C., 2000. Structural control and K/Ar
dating of the Au-Ag epithermal veins in the Shila Cordillera, southern Peru. Comptes Rendus de l’Academie des Sciences,
Serie II. Sciences de la Terre et des Planetes, 330, 1: 23-30.

Clague D. A. and Jarrard R. D., 1973. Tertiary Pacific plate motion deduced from the Hawaiian-Emperor chain. GSA
Bulletin, 84: 1135-11???.

Corbett, G. J. and Leach, T. M., 1998, Southwest Pacific gold-copper systems: Structure, alteration and mineralization:
Special Publication 6, Society of Economic Geologists, 238p.

Cox S. F., Knackstedt M. A., and Braun J., 2001. Principles of structural control on permeability and fluid flow in
hydrothermal systems. In: Richards J. P. and Tosdal R. M. (eds.). Structural controls on ore genesis. Reviews in Economic
Geology, 14: 1-24.

Drummond S. E. and Ohmoto H., 1985. Chemical evolution and mineral deposition in boiling hydrothermal solutions.
Economic Geology, 80: 126-147.

Echavarria L., Yagua T., Nelson E., and Benavides J., 2003. Sistema epitermal de Arcata, sur de Perú. III Congreso
Internacional ProExplo 2003, Instituto de Ingenieros de Minas del Perú. CD-Rom: 17p.

Enriquez E. and Rivera R., 2001. Timing of magmatic and hydrothermal activity in the San Dimas district, Durango,
Mexico. In: Albinson T. and Nelson C. E (eds.). New mines and discoveries in Mexico and Central America. SEG Special
Publication 8: 3-38.

Farrar E, and Noble D. C., 1976. Timing of late Tertiary deformation in the Andes of southern Peru. Geological Society of
America Bulletin, 87: 1247-1250.

Field, C. W., and Fifarek, R. H., 1985. Light stable isotope systematics in the epithermal environment. Reviews in
Economic Geology 2: 99-128.

Fornari, M. and Vilca Neyra, C., 1979. Mineralización argentífera asociada al volcanismo Cenozoico en la faja Puquio-
Cailloma. Boletín de la Sociedad Geológica del Perú, 60: 101-128.

Fournier R. O., 1985. The behavior of silica in hydrothermal solutions. In: Berger B. R. and Bethke P. M. (eds.). Reviews
in Economic Geology 2. Geology and Geochemistry of Epithermal Systems: 45-61.

Gammons C. H. and Seward T. H., 1996. Stability of manganese (II) chloride complexes from 25° to 300°C. Geochimica
et Cosmochimica Acta, 60: 4295-4311.

Gibson P. C., Noble D. C., and Larson L. T., 1990. Multistage evolution of the Calera epithermal vein system, Orcopampa
District, Southern Peru: First results. Economic Geology, 85: 1504-1519.

Gibson P.C., McKee E. H., Noble D. C., and Swanson K. E., 1995. Timing and interrelation of magmatic, tectonic, and
hydrothermal activity at the Orcopampa district, southern Peru. Economic Geology, 90: 2317-2325.

Goldstein R. H. and Reynolds T. J., 1994. Systematics of fluid inclusions in diagenetic minerals. Society for Sedimentary
Geology, Short Course 31: 199p.

Gross W. H., 1975. New ore discovery and source of silver-gold veins, Guanajuato, Mexico. Economic Geology, 70:
1175-1189.

14
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

Hedenquist J. W., Arribas A., and Gonzalez Urien E., 2000. Exploration for epithermal gold deposits. Hagemann S. G.
and Brown P. E. (eds.). Reviews in Economic Geology , 13: 245-277.

Henley R. W., Truesdell A. H., and Barton P. B., 1984. Fluid-mineral equilibria in hydrothermal systems. Reviews in
Economic Geology, 1. 267 pp.

Henley R. W., 1985. The geothermal framework of epithermal deposits. In: Berger B. R. and Bethke P. M. (eds.). Reviews
in Economic Geology 2. Geology and Geochemistry of Epithermal Systems: 1-24.

Lang B, Edelstein O., Steinitz G., Kovacs M., and Halga S., 1994. Ar-Ar dating of adularia –a tool in understanding
genetic relations between volcanism and mineralization: Baia Mare area (Gutii M ountains), northwestern Romania.
Economic Geology, 89: 174-180.

Leroy J. L., Hube D., and Marcoux E., 2000. Episodic deposition of Mn minerals in cockade breccia structures in three
low-sulfidation epithermal deposits: a mineral stratigraphy and fluid-inclusion approach. The Canadian Mineralogist, 38:
1126-1136.

Lovering T.S. and Goddard E. N., 1950. Geology and Ore Deposits of the Front Range Colorado. U. S. Geological
Survey Professional Paper 223, 319p.

Machare J., Sebrier M., Huaman D., and Mercier J. L., 1986. Tectónica Cenozoica de la margen continental peruana.
Boletín de la Sociedad Geológica del Perú, 76:45-77.

McKee E. H. and Noble D. C., 1982. Miocene volcanism and deformation in the western Cordillera and high plateaus of
south-central Peru. Geological Society of America Bulletin, 93: 657-662.

McKee E. H. and Noble D. C., 1989. Cenozoic tectonic events, magmatic pulses, and base- and precious metal
mineralization in the Central Andes. In: Ericksen G.E., Cañas Pinochet M., and Reinemund J.A. (eds.), Geology of the
Andes and its relation to hydrocarbon and mineral resources. Circum-Pacific council for Energy and mineral resources
Earth Science Series, 11: 189-194.

McKee E. H., Dreier J. E., and Noble D. C., 1992. Early Miocene hydrothermal activity at Pachuca-Real del Monte,
Mexico: an example of space-time association of volcanism and epithermal Ag-Au mineralization. Economic Geology,
87: 1635-1637.

Megard F., Noble D. C., McKee E. H., and Bellon H., 1984. Multiple pulses of Neogene compressive deformation in the
Ayacucho intermontane basin, Andes of central Peru. Geological Society of America Bulletin, 95: 1108-1117.

Mercier J. L., Sebrier M., Lavenu A., Cabrera J., Bellier O., Dumont J., and Machare J., 1992. Changes in the tectonic
regime above a subduction zone of Andean type: the Andes of Peru and Bolivia during the Plio-Peistocene. Journal of
Geophysical Research, 97, N° B8, 11945-11982.

Noble D. C., 1981a. Preliminary report on four radiometric ages from the Cailloma-Sucuitambo region, Peru.
Unpublished report Mauricio Hochschild Company.

Noble D. C., 1981b. Age of the Cailloma caldera. Unpublished report Mauricio Hochschild Company.

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. SEG Special Publication 7: 155-193.

Noble D. C. and Silberman M. L., 1984. Evolución volcánica e hidrotermal y cronología de K/Ar del distrito minero de
Julcani. Volumen Jubilar LX aniversario, Homenaje al Dr. Georg Petersen. Sociedad Geológica del Perú, 5: 35p.

Noble D. C. and Vidal C. E., 1994. Gold in Peru. Society of Economic Geologists newsletter, 17: 1, 7-13.

Noble D. C., Eyzaguirre V. R., and McKee E. H., 1989. Precious-metal mineralization of Cenozoic age in the Central
Andes of Peru. In: Ericksen G.E., Cañas Pinochet M., and Reinemund J.A. (eds.), Geology of the Andes and its relation to
hydrocarbon and mineral resources. Circum-Pacific council for Energy and mineral resources Earth Science Series, 11:
207-212.

15
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

Noble D. C., McKee. E. H., Eyzaguirre V. R., and Marocco R., 1984. Age and regional tectonic and metallogenic
implications of igneous activity and mineralization in the Andahuaylas-Yauri belt of southern Peru. Economic Geology,
79: 172-176.

Noble D. C., Wise J. M., and Vidal C. E., 1999. Episodes of Cenozoic extension in the Andean orogen of Peru and their
relation to compression, magmatic activity and mineralization. In: Machare J., Benavides V., and Rosas S. (eds.),Volumen
Jubilar N° 5 “75 Aniversario Sociedad Geológica del Perú”: 45-66.

Noble D. C., McKee E. H., Farrar E., and Petersen U., 1974. Episodic Cenozoic volcanism and tectonism in the Andes of
Peru. Earth and Planetary Science Letters, 21: 213-220.

Nolan T. B., 1933. Epithermal precious-metal deposits of the western states. Lindgren Volume, American Institute of
Mining and Metallurgical Engineerings, New York: 623-640.

Ohmoto, H., and Rye, R. O., 1979. Isotopes of sulfur and carbon. In: Barnes H.L. (ed.). Geochemistry of hydrothermal ore
deposits, second edition: 509-567. J. Wiley and Sons, New York.

Pardo Casas F. and Molnar P., 1987. Relative Motion of the Nazca (Farallón) and South America plates since late
Cretaceous time. Tectonics, 6: 233-248.

Petersen, U., 1958. Structure and uplift of the Andes of Peru, Bolivia, Chile, and adjacent Argentina. Boletín de la
Sociedad Geológica del Perú, 33:57-129.

Peterson P. S., Noble D. C., McKee E. H., and Eyzaguirre V. R., 1983. A resurgent mineralized caldera in southern Peru:
preliminary report. EOS (American Geophysical Union Transactions), 64: 884p.

Pilger R. H. Jr., 1983. Kinematics of the South American subduction zone from global plate reconstruction In: Cabre R.
(ed.). Geodynamics of the eastern pacific region; Caribbean and Scotia arcs. Geodynamics, 9: 113-125.

Pilger R. H. Jr., 1984. Cenozoic plate kinematics, subduction and magmatism; South American Andes. Journal of the
Geological Society of London, 141: 793-802.

Reed, M. H., and Spycher, N. F., 1985. Boiling, cooling and oxidation in epithermal system: a numerical modelling
aproach. In: Berger, B. R. and Bethke P. M. (eds.), Reviews in Economic Geology 2, Geology and Geochemistry of
Epithermal Systems: 249-272.

Rye R. O. and Ohmoto H., 1974. Sulfur and carbon isotopes and ore genesis: a review. Economic Geology, 69: 26-842.

Saunders J. A. and Schoenly P. A., 1995. Boiling, colloid nucleation and aggregation, and the genesis of bonanza Au-Ag
ores of the Sleeper deposit, Nevada. Mineralium Deposita, 30: 199-210.

Sebrier M. and Soler P., 1991. Tectonics and magmatism in the peruvian Andes from late Oligocene time to the present.
In: Harmon R. S. y Rapela C. W. (eds.), Andean magmatism and its tectonic setting. Geological Society of America
Special Paper 265: 259-278.

Sebrier M., Lavenu A., Fornari M., and Soulas J. P., 1988. Tectonics and uplift in Central Andes, Peru, Bolivia and
northern Chile, from Eocene to present. Geodynamique, 3: 739-780.

Silberman M. L., McKee E. H., and Noble D. C., 1985. Age of mineralization at the Cailloma and Orcopampa silver
districts, southern Peru. Isochron/West, 43: 17-18.

Sillitoe, R. H., 1991. Gold metallogeny of Chile –an introduction. Economic Geology, 86: 1187-1205.

Sillitoe R., and Hedenquist J., 2003. Linkages between volcanotectonic settings, ore-fluid compositions, and epithermal
precious-metal deposits. In: Simmons, S.F. and Graham, I. (eds.) SEG Special Publication 10: 315-343.

Simmons, S. F., 1991. Hydrologic implications of alteration and fluid inclusion studies in the Fresnillo District, Mexico;
evidence for a brine reservoir and a descending water table during the formation of hydrothermal Ag-Pb-Zn orebodies:
Economic Geology, 86: 1579-1601.

Simmons, S. F., 1995. Magmatic contributions to low-sulfidation epithermal deposits. Mineralogical Association of
Canada Short Course, 23: 455-477.

16
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

Simmons S. F. and Browne P. R. L., 1997. Saline fluid inclusions in sphalerite from the Broadlands-Ohaaki geothermal
system: A coincidental trapping of fluids boiled towards dryness. Economic Geology, 92: 485-489.

Simmons S. F. and Christenson B. W., 1994. Origins of calcite in a boiling geothermal system. American Journal of
Science, 294: 361-400.

Stephan S., 1974. Genese der subvolkanischen Silbererzgangue von Caylloma in Peru. Geologisches Jahrbuch, Reihe D,
Heft 8. 143 p.

Swanson K. E., Noble D. C., McKee E. H., and Gibson P. C., 1993. Collapse calderas and other Neogene volcanic and
hydrothermal features of the Chila Cordillera and adjacent areas, southern Peru. Abstracts. Geological Society of America
Abstracts with Programs, 25: 154.

Tosdal R. M., Gibson P. C., and Noble D. C., 1995. Metal source for Miocene precious-metal veins of the Orcopampa,
Shila, Cailloma and Arcata mining districts, southern Peru. In: Volumen Jubilar Alberto Benavides. Sociedad Geológica
del Perú: 311-326.

Vikre P. G., McKee E. H., and Silberman M. L., 1988. Chronology of Miocene hydrothermal activity and igneous events
in the western Virginia Range, Washoe, Storey and Lyon Counties, Nevada. Economic Geology, 83: 864-874.

Acknowledgements
The authors would like to acknowledge Mauricio Hochschild & Cia. Ltda., S.A.C. for financial and
logistical support for this project. Jorge Benavides is thanked for his arranging logistical support and for
geological discussions in the field. Jeff Hedenquist and Murray Hitzman gave thoughtful reviews of the
manuscript.

Figure Captions
Figure 1. Simplified location map of the Caylloma district, showing the principal Neogene calderas (gray) and
epithermal deposits of the region (after Noble et al., 1989). Calderas: 1. Ccarhuarasco, 2. Pampa Galeras, 3.
Parinacocha, 4. Tumiri, 5. Teton, 5. San Martin, 7. Esquillay, 8. Chonta, 9. Caylloma, 10. Coropuna. Bold
dashed line = political (department) boundaries.
Figure 2. Geologic map of the Caylloma district, showing main veins and related silicification zones.
Figure 3. Schematic stratigraphic column of the Caylloma district showing main sequences of deposition,
isotopic ages , mineralization, and dominant lithology.
Figure 4. Lower hemisphere, equal area stereonet diagrams. A. Plot of fault-veins and extension veins, arrow
show direction of movement of hangingwall on slickenline measurements. Note, some extension veins have a
component of shear and thus display slickenlines. B. Pole plot of the same data. 1, 2, and 3 represent modeled
extension (300°/4°), intermediate (063°/35°) and shortening (234°/55°) strain axes, respectively. Extension
veins plot in the 2-3 plane.
Figure 5. Schematic cross section across the Caylloma district, showing high-angle south dipping fault-veins
that tilted and displaced basement blocks. Section line shown on Fig. 2.
Figure 6. Caylloma district structural model. A. Map of district veins. The veins are restricted between two
first-order northwest-striking faults with interpreted sinistral movement. B. Blow up showing details of the
Paralela vein system. C. Structural model. Veins are hosted by second-order dextral-normal faults and third-
order extension fractures dipping to the south. Dextral block rotation produced minor sinistral movement
along the third-order extension fractures and the development of fourth-order extension fractures. Rotations of
both second and third order blocks are interpreted to be due to bookshelf sliding.
Figure 7. Ag grade contours plotted on long section of the Apostoles 2 vein. Narrow and discontinuous ore
shoots have high rake angles. The unconformity between basement and volcanic rocks represents, in most
cases, the base of economic mineralization. The ore shoots open up and are better developed within the
volcanic rocks.
Figure 8. Long section of Bateas vein showing Ag grade contours and mine working. A strong vertical
zonation is present with a ruby silver-rich upper zone, a tetrahedrite-rich high grade intermediate zone, and a
base metal-rich lower zone.
Figure 9. Detailed outcrop map of a portion of the San Cristóbal vein. In the hangingwall of the transverse
Jesus Maria fault the San Cristóbal vein consists of a half cymoid with a thick footwall split.

17
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

Figure 10. Silver grade contours plotted on long section of San Cristóbal vein. The half cymoid, in the
hangingwall of the Jesus Maria fault, displays the richest and largest (most continuous) ore shoot along the
vein.
Figure 11. Photograph mosaic and drawing of Bateas vein cross section at level 14, hosted by a silicified
porphyritic andesite. All the depositional stages described are present, with a symmetric disposition where the
younger bands are arranged towards the center of the vein. The manganese stage occupies more than 75% of
the vein, where the three principal substages are clearly identified. The intermediate substage shows a clear
zonation with an outer sphalerite-galena band that passes inward to a chalcopyrite-tetrahedrite band.
Figure 12. Paragenetic stages and substages identified with characteristic minerals present. Line weight
indicates relative abundance.

Figure 13. Detailed schematic section and photomicrograph of the manganese substage two at Bateas vein,
level 14. A complete cycle of deposition from S1 to S6 is represented. S1a: coarse-grained sphalerite, S1b:
chalcopyrite+tetrahedrite, S2: coarse-grained quartz, S3: fine-grained rhodonite+amorphous
silica+rhodochrosite, S4: quartz with disseminated sulfides, S5: quartz, S6a: fine-grained rhodonite, S6b:
acicular rhodonite. Photomicrograph close up shows details of the S4 and S5 microstages, with fine-grained
sulfides disseminated within quartz in the S4 microstage, and sulfide-free quartz in the S5 microstage. sph:
sphalerite, cpy: chalcopyrite, qz: medium- to coarse-grained quartz, qzf: fine-grained quartz, rhn: acicular
rhodonite, rh: massive fine-grained rhodonite.
Figure 14. Long section of La Plata vein showing schematically the distribution of paragenetic stages, with an
upper portion of low-grade quartz-pyrite and a lower portion of manganese stage with higher Ag content.
Numbers show Ag grade in oz/t.
Figure 15. Compilation of new 40 Ar/ 39 Ar and existing K-Ar isotopic ages obtained from volcanic rocks and
mineral deposits and prospects in the region. K-Ar data compiled from Noble (1981a), Noble (1981b),
Peterson et al. (1983), Silberman et al. (1985), McKee and Noble (1989), and Noble et al. (1989). Three
episodes of volcanic activity are recognized: the older one is related to the Caylloma volcanism and
mineralization, the intermediate one is related to the Chonta caldera formation, and the younger is one related
to post-mineralization volcanism of the Caylloma caldera.
Figure 16. Homogenization temperature versus salinity plot of fluid inclusion assemblages measured on
sphalerite, quartz, and calcite in samples from the manganese, quartz and late stages in the Bateas vein. Inset:
Homogenization temperature versus salinity plot of selected primary fluid inclusion assemblages from the
manganese stage, showing an increase in salinity and drop in temperature from early to late quartz of the S2
and S5 microstages. Dash lines show fluid path during equilibrium boiling (Henley et al., 1984).
Figure 17. Homogenization temperature and salinity measured in quartz crystals for each microstage of the
Mn stage. No data are available for the fine-grained rhodonite microstage. Two subcycles are recognized,
each one characterized by early deposition of sulfides and late rhodonite after an interpreted boiling process.

Table 1. Schematic zonation of gangue and ore minerals, and metals, across the Caylloma district, and
horizontally and vertically along the veins.

Section across district: NW Center SE


Section along veins: NE Center SW
Vertical zones: Lower Intermediate Upper
Dominant gangue minerals: Rhodonite-rhodochrosite- Rhodonite- Quartz
calcite rhodochrosite-quartz
Dominant ore minerals: Sphalerite-galena-pyrite- Tetrahedrite-polybasite Pyrargyrite-
chalcopyrite-stromeyerite miargyrite-stibnite
Metal zonation: Cu-Zn? Pb Ag Sb

Table 2. Average metal contents within the different levels of the Bateas vein.

Level Ag (g/t) Au (g/t) Zn (%) Pb (%) Cu (%)

18
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

Upper 510 0.85 0.8 0.6 0.1


Intermediate 2330 0.4 9.8 6.4 2.5
Lower 200 No data 2.2 2.2 0.9

Table 3. New Ar-Ar isotopic ages from the Caylloma area.


Location Rock type Material Age Error
(UTM PSAD 56) analyzed
N8319806-E193894 Andesite lava Volcanic matrix 20.30 0.11
N8323406-E192548 Rhyolitic dome Sanidine 11.80 0.07
N8323406-E192548 Rhyolitic dome Biotite 12.25 0.07
N8320857-E195116 Volcaniclastic Adularia 18.35 0.17
host rock

Table 4. Isotopic analyses from pyrite, galena and sphalerite from the Bateas and San Cristóbal (S.C.) veins.
Isotopic ratios are reported in standard notation relative to CDT, with 0.03‰ of analytical error. Calculation
of δ34 SH2S is based on fractionation factors in Ohmoto and Rye (1979). Coarse-grained sulfides were separated
by hand picking and analyzed at the Colorado School of Mines Stable Isotope Laboratory. Samples were
placed in individual tin capsules and combusted in a Eurovector elemental analyzer for continuous-flow
analysis of sulfur isotopic composition in a GV Instruments IsoPrime isotope ratio mass spectrometer.

Sample/vein Mineral δ34 S Mineral δ34 SH2S Equation


‰ (measured) ‰ (calculated)
B3 (Bateas) Galena -2.31 -0.21 103 lna=-0.63(106 /T2 )
B3 (Bateas) Pyrite -1.35 -2.68 103 lna=0.40(106 /T2 )
B3 (Bateas) Sphalerite -2.00 -2.33 103 lna=0.10(106 /T2 )
B4 (Bateas) Galena -3.24 -1.14 103 lna=-0.63(106 /T2 )
B4 (Bateas) Pyrite -1.57 -2.9 103 lna=0.40(106 /T2 )
B4 (Bateas) Sphalerite -1.65 -1.98 103 lna=0.10(106 /T2 )
SC3 (S.C.) Galena 0.48 2.58 103 lna=-0.63(106 /T2 )
SC3 (S.C.) Pyrite 0.86 -0.47 103 lna=0.40(106 /T2 )
SC3 (S.C.) Sphalerite 1.12 0.79 103 lna=0.10(106 /T2 )

19
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

20
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

21
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

22
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

23
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

24
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

25
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

26
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

27
Caylloma epithermal district Accepted, Econ. Geology, Nov. 2004

28

You might also like