You are on page 1of 95

Copper Skarn Deposits, Per

Copper Skarn Deposits, Per

To J. Isabel

Copper Skarn Deposits, Per


Ore-Deposits II

Student: ESPIRITU PUJAY, Junior


School of Geology
Teaching: Flores Coaguila, Saturnino
Eighth Semester
Universidad Nacional Daniel Alcides Carrin

Preface
The Antamina mine is located at approximately 9 32' S and 77 03' W, 270 km
north of Lima and 130 km from the Pacific coast, in Ancash Department in north-central
Peru. It lies in the eastern part of the Cordillera Occidental, east of the Cordillera Blanca
and west of the Ro Maran valley. The skarn is exposed between approximately 4,200
and 4,800 m a.s.l., at the head of a southwest-draining glacial valley, but prior to mining
much of the orebody was covered by Lago Antamina, a glacial tarn (Fig. 2). The history of
exploration at Antamina and the general geology of the deposit are summarized by
Redwood (1998, 1999). OConnor (2000) reviewed the geologic and geophysical
approaches that delineated the ore and outlined the development of the mine,
metallurgical testing, and resource calculations. Both authors generally supported
previous geologic descriptions and interpretations (e.g., Petersen, 1965). The Antamina
deposit formed at 9.86 to 10.18 Ma (40Ar/39Ar step-heating data of Love et al., 2003)
around a small monzogranitic porphyry intrusion. It is hosted by Upper Cretaceous
carbonate strata within the Maran thrust and fold belt, formed by the late Eocene Incaic
orogeny (Noble et al., 1979; Mgard, 1984). The western Andes of Peru were the site of
episodic arc magmatism from the Late Triassic to the late Miocene (Cobbing et al., 1981).
Copper mineralization was known at Antamina (anta: copper in Quechua) in pre-Colonial
times, but only modest amounts of Pb and Ag are known to have been produced in the
district prior to 2001 (Redwood, 1999). The skarn contains proven and probable reserves
of 561 Mt with an average grade of 1.24 percent Cu, 1.03 percent Zn, 13.71 g/t Ag,
and0.029 percent Mo (calculated at a 0.7% Cu equiv cutoff grade).
THE COPPER SKARN deposit at Coroccohuayco lies 7 km southeast of the Tintaya open
pit mine in the Tintaya district, Cusco Department, Peru (14 57' 21" S, 7115'17" W; Fig.
1). This deposit is one of several in the Andahuaylas-Yauri copper belt, a late EoceneOligocene belt of igneous rock which extends for 350 km from Andahuaylas in the north to
southeast of Yauri (Noble et al., 1984; Perell et al., 2003). Copper deposits in this belt are
hosted by Eocene-Oligocene igneous and Cretaceous sedimentary rocks and include the
Cu (+Au) porphyry and skarn deposits at Katanga, Charcas, and the Las Bambas and
Tintaya districts. Perell et al. (2003) provided a range of K-Ar ages from biotite and
hornblende of 30 to 40 m.y. for intrusive rocks of the Andahuaylas-Yauri batholith. This
general age range is contemporary with igneous activity of the Eocene to early Oligocene
volcanic arc in central Peru (Noble et al., 1984) and magmatic activity related to large
porphyry copper deposits in northern Chile (e.g., Chuquicamata, La Escondida; Clark et al.,
1990). A thorough regional review of the tectonic context, structure, stratigraphy,
petrography, and mineralization of the Andahuaylas-Yauri belt has
been summarized by Perell et al. (2003) and will not be repeated here. Other publications
have outlined regional

stratigraphy, age constraints, and tectonic and structural aspects of the belt (Noble et al.,
1984; De La Cruz B., 1995; Benavides-Caceres, 1999; Jaillard et al., 2000).

The Student

Contents
Preface.............................................................................................................. .5
Antamina Copper-Zinc Skarn Deposit, Ancash................................................8
Regional Geologic Setting................................................................................................ 10
Stratigraphic relationships............................................................................................ 10
Tectonic relationships................................................................................................... 13
Stratigraphic variations across the cross-strike structural discontinuity:.....................13
Igneous activity along the cross-strike structural discontinuity:...................................16
Local Geologic Setting..................................................................................................... 17
Sedimentary host rocks................................................................................................ 17
Alteration of sedimentary rocks adjacent to the skarn.................................................19
Jumasha Formation:...................................................................................................... 20
Celendn Formation:..................................................................................................... 21
Structural geology of the mine area............................................................................. 24
Local folds and faults:................................................................................................... 24
Structural control of mineralization:............................................................................. 28
Discussion........................................................................................................................ 29
Regional stratigraphic and tectonic relationships.........................................................29
A model for the geologic history of the Antamina region.............................................30
Regional effects on local structure and mineralization.................................................32
Implications for local exploration and ore genesis.......................................................35
Metallogenic and geotectonic implications...................................................................35
Conclusions..................................................................................................................... 37
References.......................................................................................................................... 38
Copper Skarn Deposit at Coroccohuayco, Tintaya District......Error! Marcador no
definido.42
Deposit Setting................................................................................................................ 42
Stratigraphy................................................................................................................. 42
Structural geology........................................................................................................ 43
Igneous rocks............................................................................................................... 43
Deposit Alteration............................................................................................................ 48
Igneous rocks............................................................................................................... 48
Hornfels........................................................................................................................ 49
7

Marble and skarn.......................................................................................................... 50


Retrograde and low-temperature alteration.................................................................51
Mineralization.................................................................................................................. 52
Igneous rocks............................................................................................................... 52
Skarn............................................................................................................................ 52
Ore petrography........................................................................................................... 52
Magnetite..................................................................................................................... 54
Igneous Rock Geochemistry............................................................................................ 54
Mineral Chemistry............................................................................................................ 54
Garnet.......................................................................................................................... 55
Pyroxene...................................................................................................................... 56
Sulfide and precious metal mineralization chemistry...................................................57
Fluid Inclusions................................................................................................................ 59
Discussion........................................................................................................................ 60
Role of igneous rocks................................................................................................... 60
Skarn zonation.............................................................................................................. 61
Skarn-forming conditions............................................................................................. 62
Mineralization............................................................................................................... 63
Magnetite..................................................................................................................... 63
Conclusions..................................................................................................................... 64
References.......................................................................................................................... 66
Projects and mining prospects

Antamina Copper-Zinc Skarn Deposit,


Ancash
8

The Antamina deposit


The Antamina deposit comprises endoskarn and exoskarn, with subordinate breccia
bodies that cut both skarn and intrusion within the perimeter of skarn. The mineralized
skarn is dominated by grandite garnet, which grades from brown to green nearer the host
limestone (Petersen, 1965). Thus the deposit conforms to the oxidized calcic clan of
Einaudi et al. (1981). The geology of the deposit and its immediate surroundings is
summarized in Figure 3a on three northwestsoutheast drill-hole cross sections, spaced 50
m apart, through the middle of the orebody. The simplified surface geology map in Figure
3b was constructed by projecting the lithologic boundaries defined on these and thirty
other cross sections.
The area delimited by the mineralized skarn front is approximately 1.18 km 2 overall.
The skarn zone straddles the original intrusive contact and surrounds a circa 0.24 km 2 core
of porphyry that forms a crudely parallelogram-shaped prism with a vertical axis (Fig. 3b).
The outer boundary of the mineralized skarn is more elongated northeast-southwest than
this core because it expands around faults and dikes extending to the east, northeast, and
southwest. It is therefore roughly elliptical in plan and has a northwest-southeast width of
up to 1,000 m, and a northeast-southwest length of more than 2,500 m. The long axis
parallels the Antamina valley and is perpendicular to the regional structural grain of the
deformed carbonate host rocks. As recognized by Petersen (1965) the outer limit of skarn
is generally subvertical. The skarn narrows with depth as the core of porphyry widens (Fig.
3a), but there is no significant change in the Cu and Zn grades to depths of at least 400 m
below the original valley floor.
Except at high elevations in the eastern part of the deposit, the exoskarn is almost
everywhere mineralized. In detail, the individual skarn facies and breccia bodies are
complexly shaped and discontinuous, but the deposit can be simplified as comprising an
inner shell of endoskarn, stockwork, breccia, and brown garnet exoskarn that contains the
copper molybdenum ore and an outer shell of green garnet exoskarn comprising the
copper-zinc ore. Molybdenite is disseminated in irregular zones within and at the margin
of the intrusion. Chalcopyrite is the dominant copper mineral except at shallow depths in
the southwestern part of the deposit, where bornite predominates in wollastonitic
exoskarn, which forms an enclave in green garnet Cu-Zn exoskarn. Hydrothermal breccia is
common at or near the endoskarn-exoskarn contacts along the northwest and southeast
sides of the deposit. Breccia also cuts the porphyry core as anastomosing sheets and pipes
that are commonly enveloped by fine-grained maroon garnet endoskarn (Fig. 3). The
breccia zones contain minor chlorite and were originally described by Petersen (1965) as
chlorite skarn. The breccias are generally poorly sorted and comprise angular to rounded
fragments of the skarn and metallic minerals supported in a sand-sized matrix of similar
composition. As in skarn, sphalerite and molybdenite do not commonly occur together in
the breccias; sphalerite is found in breccias in or near exoskarn, but Molybdenite occurs in
breccias in endoskarn. The common metallic
minerals in the breccias are pyrite, chalcopyrite, and magnetite, which are mostly
comminuted but also occur as veins, massive bodies, and large fragments. This mineral
assemblage is rare in exoskarn but forms widespread stockworks and sheeted vein swarms
in fine-grained maroon garnet endoskarn, in many places grading into crackle, mosaic, and
matrix-dominated breccia. We estimate that approximately onethird of the ore at
Antamina may have formed during this late brecciation, veining, and endoskarn-forming
stage. Widely spaced, late calcite-tetrahedrite sphalerite galena veinlets are common
and cut all skarn and breccia types, although they are also locally dismembered in breccia,

probably because of settling. Scarce realgar veinlets occur in calc-hornfels above and
peripheral to the skarn.

FIG. 1. Location map of the Antamina deposit and general geology of part of Ancash and La Libertad Departments,
Peru. Compiled and modified after Egeler and De Booy (1956), Cosso (1964), Wilson and Reyes (1964), Cosso and
Jan (1967), Wilson et al. (1967, 1995), Myers (1976, 1980), Reyes (1980), Snchez (1995), Allende (1996), Cobbing
et al. (1996), Jacay (1996), Snchez et al. (1998), INGEMMET (1999), and Strusievicz et al. (2000). The Tapacocha
axis delineates the western edge of the Cretaceous shelf (Myers, 1974, 1975). Other deposits (Pierina and Pasto
Bueno) and prospects (Magistral) mentioned in the text are shown, as are the areas illustrated in Figs. 5 and 6. MTFB
= Maran thrust and fold belt.

10

FIG. 2. Premine physiography of the Antamina area, with place names referred to in the text, illustrating the clustering of Ag-bearing Pb-Zn ve

Regional Geologic Setting


The Upper Cretaceous strata enclosing the Antamina deposit are part of a
metallogenically important Albian and Upper Cretaceous package of carbonate rocks, the
Machay Group, that hosts many ore deposits in the polymetallic skarn and carbonatereplacement belt of central Peru (Soler et al., 1986). These rocks formed during the later of
two Permian to Paleocene episodes of basin development that deposited a succession of
alternating siliciclastic and carbonate facies in western South America (Sempere et al.,
2002). This Late Jurassic to Paleogene subsidence formed the West Peruvian trough, which
separated the magmatic arcs to the west from the eastern geanticline now represented by
the Maran metamorphic complex (Benavides, 1956; Wilson, 1963; Atherton et al., 1983;
Mgard, 1987). The Cretaceous sedimentary rocks that crop out in the Antamina area
accumulated in the shallow-water portion of this trough, the Yauli shelf (Szekely, 1967).
The Tapacocha axis, now a north-northwest trending high strain zone, separates the
11

western, deeper-water portion of the trough from the shelf sedimentary rocks (Fig. 1;
Myers, 1974, 1975).
Stratigraphic relationships
The host Machay Group (Figs. 1 and 4) includes all Albian to mid-Campanian
carbonate rocks south of 9 S and east of the Tapacocha axis in north-central Peru. Szekely
(1967) introduced the Machay Group in central Peru and Samam-Boggio (1980) applied it
throughout Peru. These rocks also have been referred to informally as the middle
Cretaceous limestone series (Harrison, 1940), the upper Cretaceous and Albian
carbonate series (Mgard, 1984), and the upper carbonate sequence (Manrique, 1998).
The group is underlain by a predominantly siliciclastic sequence comprising Upper Jurassic
marine black shales of the Chicama Group and Lower Cretaceous (Berriasian to Aptian)
continental to shelf sandstones, shales, and minor limestones of the Goyllarisquisga Group
(Fig. 4). It is widely overlain, conformably or slightly unconformably, by red beds (Wilson,
1963), which are mainly Campanian to Paleocene but as old as Santonian in central Peru
(Jaillard, 1987). Not preserved in the immediate Antamina area, these nonmarine, coarse
clastic rocks have been variously described as the Pocabamba Formation, 25 km southeast
of Antamina at La Unin (Wilson, 1963, after McLaughlin, 1924), the Chota Formation, 20
km north of Antamina (Benavides, 1956: after Broggi, 1942), and the Casapalca Group, 25
km southwest of Antamina in the Cordillera Huayhuash (Coney, 1971, after McLaughlin,
1924).
The Machay Group contains two transgressive sequences separated by a
disconformity ascribed to late-middle Albian uplift and erosion related to the Mochica
orogeny (Mgard, 1984). In the lower part of the group, the successive Pariahuanca,
Chulec, and Pariatambo Formations (Fig. 4) record a transition from near-shore, calcareous
sandstone and massive, shelly limestone, through thin-bedded limestone and marl, to
deep-water, thin-bedded, bituminous, dark gray marl and limestone (Benavides, 1956,
1999; Wilson, 1963; Jaillard, 1987).

12

FIG. 3. Simplified cross sections (a) and projected surface geology map (b) of the Antamina deposit, illustrating the
crudely elliptical, vertical zones of endoskarn and exoskarn developed between a core of largely skarn-free porphyry
and the limestone host rocks. Simplified surface geology map is based on drill-hole geology projected to surface and
on surface mapping by D.A.L. and J.K.G., combined with that by L. Hathaway (Inmet) and M. Wunder (Noranda).
Prominent northweststriking Incaic folds and the left-stepping, transverse Valley fault (VF) are indicated. The location

13

of the proposed Valley lateral ramp (VLR), inferred to be responsible for the apparent dextral offset of the Antamina
anticline (AA), is also shown.

Following the late-middle Albian hiatus, carbonate sedimentation on the platform resumed
with the deposition of the shallow-water, upper Albian to upper Turonian Jumasha
Formation (Jaillard, 1987), originally defined by McLaughlin (1924) in central Peru. This
formation is overlain by the muddier, deeper-water Celendn Formation (Benavides, 1956),
largely Coniacian to Santonian in age (Jaillard, 1987) but attaining the mid-Campanian in
northern Peru (Mourier et al., 1988). The upper part of the group, comprising the Jumasha
and Celendn Formations (Fig. 4), thus represents a second major transgressive sequence.
The lower to middle Albian carbonate strata (the first transgressive sequence) are similar
in lithology and thickness in both northern and central Peru (Benavides, 1956; Jaillard,
1987). However, north of approximately 9 S, the overlying upper Albian to midCampanian carbonate rocks are much thicker, more fossiliferous, and lithologically more
variable than in central Peru, and the Jumasha Formation interval is divided into five
formations (Benavides, 1956). Jaillard (1987) provides correlations between these
stratigraphic sections in northern and central Peru.

FIG. 4. Inferred stratigraphic column in the mine area compared with that for the Jumasha and Celendn Formations,
compiled from observations (Benavides, 1956) on the measured sections in the Ro Puchca valley, approximately 20

14

km north of Antamina. The contact between the Jumasha and Celendn Formations coincides with the boundary
between the Turonian and Coniacian stages and is indicated with a solid line; unknown stage boundary locations have
been approximated and indicated with dashed lines and question marks. The stratigraphic interval interpreted to host
the skarn is shown. Inset shows an outline of the regional stratigraphic section.

Tectonic relationships
Published descriptions of the structural setting of the Antamina deposit (Bodenlos
and Ericksen, 1955; Terrones, 1958; Petersen, 1965; Redwood, 1999) have focused on the
Maran thrust and fold belt, which developed circa 30 m.y. before mineralization, with
scant consideration of the regional tectonic environment that existed in the mid-Miocene.
We argue, however, that Antamina lies athwart a large-scale,cross-strike (northeastsouthwest) structural discontinuity (Wheeler, 1978) that was tectonically active at the time
of skarn formation and hence has metallogenic significance. Love et al. (2001) termed the
structure the Querococha arch because its southwestern limit at the margin of the Callejon
de Huaylas lies close to Laguna Querococha (Figs. 5 and 6). Its influence on the abundance
of Neogene intrusions, the regional strikes in the Maran thrust and fold belt, and the
stratigraphic relationships of underlying Mississippian to Lower Jurassic strata (Fig. 4) are
described below. The overall strike of the Maran thrust and fold belt changes, and the
common plunge directions reverse, across the proposed northeast-trending cross-strike
structural discontinuity (Fig. 5). These thin-skinned Eocene structures, attributed to the
Incaic orogeny, are the dominant tectonic elements in the region, although two regional
unconformities in the Cretaceous succession in central Peru, one at the base of the
Jumasha Formation and the other below the Casapalca Group and its equivalents,
represent earlier emergence during, respectively, the Mochica and Peruvian tectonic
phases (Noble et al., 1979; Mgard, 1984). The Maran thrust and fold belt extends from
5 S to 12 30' S, generally striking north-northwestsouth-southeast, parallel to the
present plate boundary, and comprises structures that predominantly verge northeast
(Mgard, 1984; Fig. 5).
However, southeast of the cross-strike structural discontinuity, folds and thrust
faults strike north-northwest, whereas to the northwest of it, they strike northerly (Figs. 1,
5, and 6). Moreover, fold plunges are reversed across this zone: to the southeast, most
major anticlines and synclines plunge to the south-southeast, whereas to the northwest,
they plunge north (Fig. 5). The locus of changes in strike and plunge extends northeast
from Laguna Querococha, but about 5 km southwest of Antamina it steps 8 km to the
north before continuing northeastward (Fig. 5). Faults with the same overall northeast
strike as the cross-strike structural discontinuity control some present-day drainages, such
as the northeast trending Ro Puchca valley, 20 km north of Antamina, which is discordant
to the overall north to north-northwest grain of the terrain (Fig. 5).
The deflection in the Antamina area is one of several that articulate the Maran
thrust and fold belt. Northerly strikes continue to Llamalln, 50 km to the north of
Antamina in the eastern part of the belt (Fig. 1). Still farther north, the overall northnorthwest regional strike of the Maran thrust and folds belt resumes. A comparable
sharp deflection to northsouth strikes occurs at the northern end of the Cordillera Blanca
(Fig. 1), 175 km to the north-northwest of Antamina, where the Casma-Pasto Bueno fault
zone intersects the regional north-northwest strikes (Rivera, 1996). Benavides (1999)
identified many segments in the fabric of the Maran thrust and fold belt, including the
two described above, although he proffered a different mechanism for their formation, as
discussed below.
15

Stratigraphic variations across the cross-strike structural discontinuity:


The contact relationships of the upper Paleozoic and Mesozoic strata (Fig. 4) to the
lower Paleozoic Maran metamorphic complex vary in accordance with the segmentation
of the thrust and fold belt. Figure 6 shows the relationships along the western margin of
the Maran complex throughout an area more extensive than that shown in Figure 5.
Mississippian to Lower Jurassic strata that normally separate the pre-Ordovician
metamorphic rocks from the Cretaceous sedimentary rocks are absent near the proposed
cross-strike structural discontinuity. In the north-striking segment of the fold belt
immediately north of the Antamina area, the eastern limit of the Mississippian to
Cretaceous succession is a subhorizontal unconformity that strikes north overall but has
an irregular surface trace owing to the incised topography (Figs. 5 and 6).

16

FIG. 5. Structural geology of the Maran thrust and fold belt in the vicinity of Antamina, illustrating the marked
change in the orientations of thrust faults and fold axes across a northeast-trending zone through Antamina (after
Egeler and de Booy, 1956; Wilson et al., 1967, 1995; Cobbing et al., 1996; Jacay, 1996; and Strusievicz et al., 2000).
The northeast-southwest trending loci of these changes in structural attitude is indicated by the heavy dashed line,
which is offset in a left-stepping sense in the vicinity of Antamina. This area is the location of the proposed crossstrike structural discontinuity discussed in the text. Inset shows the location of the map area relative to the Huari (19i), Singa (19-j), Requay (20-i) and La Unin (20-j) quadrangles. No attempt has been made to establish continuity
between the map units of Wilson et al. (1967, 1995) north of 9 30' S and those of Cobbing et al. (1996) to the south.
The undated volcanic rocks of Pampa Junn have been assigned to the Calipuy Supergroup (see text for discussion).

17

FIG. 6. Simplified geology of approximately 8,000 km2 of the Maran thrust and fold belt in the region around
Antamina; the dominantly upper Miocene Cordillera Blanca batholith is removed (cf. Figs. 1 and 5), reflecting the
geology at the time of intrusion and mineralization in the late Miocene. The northeast-southwest trending cross-strike
structural discontinuity that passes through Antamina is delimited by heavy dashed lines. Mississippian to Lower
Jurassic sedimentary rocks are absent beneath the Cretaceous Goyllarisquisga Group northeast of Antamina along
the
cross-strike structural discontinuity but are present north and southeast of the cross-strike structural
discontinuity, except where cut out by faulting. An unusual abundance of igneous bodies intrudes the Maran Belt
along the cross-strike structural discontinuity, compared to transects to the north and south. After Egeler and de
Booy (1956), Wilson et al. (1967, 1995), Cobbing et al. (1996), Jacay (1996), and Strusievicz et al. (2000).

18

19

In the north-northweststriking segments of the belt, however, the eastern extent of the
Mesozoic rocks is generally delimited by north-northweststriking faults (e.g., southeast of
Antamina, and northwest of Llamalln; Fig. 6). Southeast of Antamina, the southwest
margin of the Maran metamorphic complex is largely defined by a series of major
northeast-verging reverse faults that involved basement, but the Mississippian to Lower
Jurassic strata are preserved between the pre-Ordovician metamorphic rocks and the
Cretaceous sedimentary rocks (Fig. 6). However, in the north-northweststriking segment
northwest of Llamalln, the Maran complex is backthrust over the Cretaceous rocks and
the thickness of the Mississippian to Lower Jurassic strata is only locally apparent (Fig. 6).
Northeast of Antamina, Mississippian to Lower Jurassic strata (Fig. 4) are absent
along the proposed cross-strike structural discontinuity, and the clastic rocks of the Lower
Cretaceous Goyllarisquisga Group lie unconformably on the pre-Ordovician Maran
metamorphic complex (Figs. 5 and 6). In contrast, north and southeast of the cross-strike
structural discontinuity, a relatively thick sequence of Mississippian to Lower Jurassic
strata separates these units (Fig. 6). Sandstones and shales of the Mississippian Ambo
Group locally unconformably overlie the Maran complex near its western limit. Similarly,
continental sedimentary rocks and alkaline to subalkaline volcanic rocks of the Lower
Permian Mitu Group commonly overlie the Ambo Group and also locally lie unconformably
on the Maran complex along its western edge but are absent northeast of Antamina.
Both north and south-southeast of the cross-strike structural discontinuity, these
siliciclastic sedimentary rocks are overlain by Upper Triassic to Lower Jurassic carbonate
rocks and shales (Pucar Group). The absence of these three groups directly northeast of
Antamina records either a sub-Cretaceous erosional unconformity or nondeposition from
Mississippian to Late Jurassic times. The cross-strike structural discontinuity
therefore also coincided with a topographic high or arch, at least in the Middle to Late
Jurassic, but possibly persisting throughout the Mississippian to Jurassic interval.
The cross-strike structural discontinuities probably also influenced the distribution of
the uppermost Cretaceous to Paleocene red beds that unconformably overlie the
Cretaceous strata (Fig. 4). These are absent near the proposed cross-strike structural
discontinuity through the Antamina area, and they also thin significantly near the more
northerly Casma-Pasto Bueno deflection, but they attain a considerable thickness between
these two transverse zones as well as to the south of the arch (Fig. 1).
Igneous activity along the cross-strike structural discontinuity:
Between the Cordillera Blanca and the Maran metamorphic complex, the 1:100,000
quadrangle maps of Cobbing et al. (1996) and Wilson et al. (1967, 1995) record a greater
abundance of Tertiary hypabyssal and extrusive rocks along the proposed cross-strike
structural discontinuity than to the northwest or southeast (Fig. 6). In addition, west of the
thrust and fold belt, volcanic rocks of the Calipuy Supergroup (Strusievicz et al., 2000) are
abundant northwest and southeast of this cross-strike structural discontinuity (e.g., in the
Nevado Huantsn area and the Cordillera Huayhuash; Fig.6). These rocks are interpreted
to be lower-middle Miocene because hypabyssal intrusions associated with similar volcanic
rocks elsewhere in the region (Huaraz Group of the Calipuy Supergroup; Fig. 5) have been
shown by 40Ar/39Ar step-heating geochronology to have persisted to 14.2 Ma (Strusievicz
et al., 2000; Love et al., 2001). However, the 115 km 2 middle Miocene granodioritic
Carhuish pluton crops out on the axis of the proposed cross-strike structural discontinuity
(Fig. 5), representing deeper-seated rocks of broadly equivalent age (13.7 Ma U/Pb zircon
date on the main phase, Mukasa, 1984; 16.5 Ma K/Ar date on the marginal phase, Cobbing
20

et al., 1981). The volcanic rocks of Pampa Junn (Egeler and de Booy, 1956) northeast of
the Carhuish pluton (Fig. 5) have not been dated, so their significance with regard to the
transverse structure is uncertain. However, we propose that the cross-strike structural
discontinuity focused the distribution of igneous activity as it diminished through the midMiocene, prior to intrusion and mineralization at Antamina. In Figure 6 the arch is depicted
with a width of approximately 20 km to incorporate the left-stepping locus of changes in
Incaic strikes and plunges (Fig. 5), the absence of Mississippian to Lower Jurassic rocks
beneath the sub-Cretaceous unconformity at the northeast end of the arch, the diameter
of the Carhuish pluton at its southwestern margin, and the array of Miocene intrusions.
Local Geologic Setting
Sedimentary host rocks
The contact of the Jumasha and Celendn Formations within the Machay Group has not
previously been defined in the Antamina mine area, owing to the intense skarn and
hornfels development and to the paucity of biostratigraphic markers. The cliff-forming
strata surrounding the Antamina deposit were mapped initially by Bodenlos and Ericksen
(1955) as marine limestones of the Jumasha Formation. J.J. Wilson (unpublished report to
Cerro de Pasco Corp., Lima, 1959, in Petersen, 1965) observed that the beds on the
northwest slopes of Quebrada Antamina were stratigraphically higher than those at similar
elevations on the southeast side, but he did not locate the interformational contact.
Petersen (1965) described the carbonate units as the Machay Formation, and they were
subsequently reassigned to the Jumasha Formation by Cobbing et al. (1996). The skarnhosting limestone strata at Antamina are herein subdivided into two sequences, the upper
markedly more shaly than the lower, and assigned to the Celendn and Jumasha
Formations, respectively. These sequences are described and compared with a section of
the relevant stratigraphic interval constructed from two sections measured by Benavides
(1956) at Uchupata in the Ro Puchca valley, 20 km north of Antamina (Fig. 4). A measured
section has not been established for the Antamina minesite because of the structural
complexity of the area (described below) and the absence of marker beds in the exposed
host rocks. The inferred stratigraphic interval occupied by the skarn is also indicated in
Figure 4. The upper sequence of thin-bedded rocks that prior to mining underlay the ridge
crests flanking Quebrada Antamina (Fig. 2) is widely altered to hornfels adjacent to the
Antamina intrusive center and grades into shaly limestone with increasing distance from it
(Fig. 7). The lower sequence

21

FIG. 7. The two distinct host rock types of the Antamina skarn system: an upper sequence of thin-bedded, silty
limestones, here largely converted to calc-hornfelses, assigned to the Celendn Formation (Ce), and a lower sequence
of thick-bedded, relatively pure limestones that form marbles, interpreted as the Jumasha Formation (J). Locations
from which the photographs were taken are indicated in Fig. 2. Photographs taken in 1997 through 1999; this ridge
has now been largely removed by mine development. (a) The southeast side of Quebrada Antamina, looking
southeast. The contact (long black dashes) between the Jumasha and Celendn Formations is placed at the top of the
uppermost massive thick-bedded limestone (pale gray band) deformed by the Antamina anticline. Note the irregular
upper, southeastern contact of the skarn (long white dashes), here largely confined to the Jumasha Formation. (b)
Looking northeast at the head of Quebrada Antamina showing the distinct control of bedding in the Celendn
Formation on skarn development.

of calcitic marble contains only subordinate intercalated diopside-rich units and grades
outward into
predominantly thickbedded, relatively pure limestone (Fig. 7). The lower
sequence (Figs. 4, 7a, and 8a) generally contains more than 75 percent calcite and
comprises calcitic, variably bioclastic limestones that range in grain size from mudstone to
wackestone. This sequence is generally thick-bedded and massive and displays karstic
weathering (Fig. 8b). Toward the top of this sequence, medium to thick beds (2050 cm) of
impure, silty limestone are interbedded with the pure limestone

22

FIG. 8. The Jumasha and Celendn Formations. (a) Cliff-forming thick beds of massive limestone on the southeast flank
of Quebrada Callapo (see Fig. 2 for location). (b) Massive limestone of the Jumasha Formation, showing fluted
weathering (south of Yanacancha area, Fig. 2). (c) Upper: massive (i.e., unlaminated and with no preferred orientation
of fossils) pelecypod carbonate wackestone of the Jumasha Formation (DDH CMA-039, 155 m). Lower: sheared
fossiliferous carbonate wackestone, equivalent to upper piece, not marmorized, but with a strong preferred
orientation (DDH CMA-039, 160 m). (d) More recessive weathering, vegetated, thin-bedded limestones and marls of
the Celendn Formation cropping out on the southeast slopes of Quebrada Ayash near its intersection with Quebrada
Tucush (see Fig. 2 for location). Photograph taken in 1997; the lower half of this area is now obscured by construction
of the tailings dam. (e) Nodular limestone of the Celendn Formation, consisting of light gray, coalescing, carbonaterich nodules separated by wisps of dark gray calcareous siltstone (Fortuna Mine area, see Fig. 2; hammer for scale).
(f) Upper: Celendn Formation nodular limestone (DDH CMAC3, 221 m). Lower: lenticular bedding interpreted as
sheared nodular limestone of the Celendn Formation (DDH CMA-C5, 87.5 m).

(Figs. 4 and 7a). This sequence hosts ore at the surface and at depth in the southwest part
of the deposit, but only in the subsurface in the northeast. Few whole fossils are preserved
(Fig. 8c) and no ammonites were observed in the mine area, precluding biostratigraphic
correlations. However, we assign this sequence to the Jumasha Formation on lithologic
grounds. Sheared fossiliferous carbonate wackestone occurs locally in this sequence (Fig.
23

8c). In north-central Peru, Benavides (1956) described the Jumasha Formation as


dominated by massive, thick-bedded, light orange-brown to yellowish-brown and gray,
fossil-poor dolostones and limestones that weather dark yellowish-brown to brownish-gray.
The formation has been further described as comprising topographically prominent, cliffforming, light-gray limestones and yellowish dolostones that are characteristically
bioclastic (Wilson, 1963).
In the Uchupata section (Fig. 4), the upper 437 m is limestone, and the lower 353 m
is dolostone. Unlike
the overlying Celendn Formation, this formation only rarely contains calcareous siltstones,
although it incorporates marly limestone beds near its top (Jaillard, 1987). The upper limit
of the Jumasha Formation was defined lithostratigraphically where medium- or thickbedded limestones pass upward into thin-bedded marls and limestones (Wilson, 1963).
The upper sequence in the mine area (Figs. 4, 7, and 8d) comprises thin- to thick-bedded
impure limestone-marl that varies in carbonate/silicate ratio from relatively calcite-rich
muddy limestone (generally 5075% carbonate) to calcareous siltstone (less than 50%
carbonate). Many units contain lightgray, calcitic nodules, composing 10 to 90 percent of
the rock, enclosed by dark, silty calcareous mudstone (Fig. 8e, f). No limestone beds with
siliceous nodules have been observed. The nodular texture is interpreted to be diagenetic.
Sheared limestone with lenticular bedding exposed at the northeast head of Quebrada
Antamina represents deformed nodular limestone in which the nodules have been
flattened into lenses during folding and faulting (Fig. 8f). The upper sequence lacks
identifiable fossils but is assigned to the Celendn Formation on lithologic grounds. In the
Uchupata section, 20 km north of the mine, the Celendn Formation comprises very soft,
friable, fossil-poor, light greenish-gray, nodular, moderately silty marls and calcareous
shales (Fig. 4; Benavides, 1956). This formation is generally medium-bedded (0.30.8 m)
and variably dolomitic, and it ranges from fine-grained to pelletal (Wilson, 1963).
The contact between the Jumasha and Celendn Formations in the mine area is
conformable and generally gradational throughout several meters and is marked by
upward-increasing siltiness and decreasing bed thickness. This contact is interpreted to be
the top of the uppermost thick-bedded limestone or marble unit on the basis of lithology
and bedding characteristics. This stratigraphic position is illustrated in Figure 7a for
outcrops on the southeast wall of the Antamina valley. The contact is folded by the
anticlines exposed on the valley walls adjacent to the southwestern part of the deposit,
and it dips northeast in the subsurface in the northeast part of the deposit, as do the
overlying strata exposed in that area. The thin-bedded calc-hornfelses that predominate
on the ridge crests flanking Quebrada Antamina grade laterally away from skarn into the
marly, variably nodular sequence assigned to the Celendn Formation, but they do not
contain the shales characteristic of its upper part. The relatively pure calcitic marble of the
Jumasha Formation that hosts ore at surface in the southwest part of the deposit and at
depth in the northeast sector does not contain the thick dolostones characteristic of the
lower part of that formation, and no magnesian skarn has been found. On this basis, we
conclude that the skarn developed in the upper part of the Jumasha Formation and the
lower part of the Celendn Formation (Fig. 4). However, we estimate that approximately
three-quarters of the mineralized exoskarn, and all of the wollastonite exoskarn, developed
in the Jumasha Formation, and that a similar proportion of the contiguous and genetically
related endoskarn was formed in the stock adjacent to that formation. It is not known
whether magnesian skarn developed in dolostone below the limit of exploration and
development drilling.

24

Alteration of sedimentary rocks adjacent to the skarn


The outer contact of coarse-grained mineralized exoskarn is abrupt, convoluted, and
uninfluenced by bedding in the marbles formed from the limestones of the Jumasha
Formation (Fig. 7a), but it is more gradational and stratigraphically controlled in the finegrained calc-silicate rocks developed in the interbedded limestones and marls of the
overlying Celendn Formation (Fig. 7b). In the Jumasha Formation, the mineralized skarn is
juxtaposed with marbles containing local horizons rich in calc-silicate minerals that may be
ascribed to thermal metamorphism. In contrast, the calc-silicate-bearing rocks developed
in the marly Celendn Formation include rock types that are interpreted as the products of
either metamorphism or metasomatism. The latter do not host sulfide minerals and differ
radically in texture and mineralogy from the exoskarn ore. Their wide distribution around
the upper part of the orebody constitutes a significant exploration target.
Jumasha Formation:
On approaching skarn, dark gray Jumasha limestone is converted to coarsely
crystalline, gray to white calcitic marble (Table 1, Fig. 9), forming an aureole ranging from
tens to hundreds of meters wide. Within the aureole, local development of sparse diopside,
wollastonite, or scapolite porphyroblasts (Fig. 9a) or slight differential erosion and color
variation (Fig. 9b) in calcite marble record minor compositional variations in the limestone.
Rare, fine-grained, light green, diopsidic layers in calcite marble (Fig. 9c), the porcellanite
of Terrones (1958), are interpreted as calc-hornfels, representing thermally
metamorphosed, medium to thick beds of dolomitic, muddy, fine-grained limestone near
the top of the formation. Within the marble aureole, medium-grained, mottled, or banded
gray marble predominates, locally grading inward to medium- to coarse-grained, pure
white marble, reflecting either degraphitization or metamorphism of organic matter to
clear vitrinite. Rare, medium- to coarse-grained, thin, buff garnet layers in both white and
mottled facies of marble mimic the forms of folded silicate-rich laminae (Fig. 9d) and
appear to have replaced them. This type of garnet development is interpreted as a
bimetasomatic reaction skarn. In gray marble, scapolite is locally developed in some
darker-gray bands (Fig. 9e) but does not persist into white marble or skarn (Table 1).
Scapolite therefore forms a discontinuous halo up to tens of meters wide and commonly
separated from

25

the skarn front by tens of meters. White marble develops in patches (Fig. 9d and f) and is
concentrated along stylolites and fractures (Fig. 9f). The latter probably represent fluid
pathways, suggesting that development of white marble involved the local channeling of
fluids. Such pathways were not consistently used by later skarn-related fluids and are
locally cut by fractures with wollastonite-rich selvages (Fig. 9f). Immediately southwest of
the skarn, the Jumasha Formation has a distinct planar banded fabric, locally a spaced
cleavage, that dips northeast at a high angle to bedding (Fig. 9a). To the northeast, this
fabric steepens and becomes vertical adjacent to the skarn and along the strike of a major
fold axis. The consistent strike, systematically changing dip, and relationship to folds
visible in the adjacent limestones indicate that this is an upward-fanning cleavage
associated with local folding. In the adjacent skarn, alternating andradite-rich and
sphalerite-rich bands, generally 5 to 50 cm wide, are subparallel to this fabric, suggesting
that at least some preexisting structures influenced mineralization. This banding may
correspond with the coarse-grained garnet-defined fabric observed by Terrones (1958).
Celendn Formation:
In the Celendn Formation, on the ridges around Lago Antamina, the skarn has a
halo of mineralogically diverse fine-grained calc-silicate rocks with subconchoidal to
conchoidal fractures. With increasing distance from the skarn, these rocks grade into
interbedded limestone and marl that reacted differentially to thermal metamorphism
around the porphyry intrusion and orebody, thus highlighting the bedding (Fig. 7b).
Dark gray massive units change in color, mineralogy, and grain size proximal to
monzogranitic dikes. Although we have not mapped the various facies of these rocks in
detail, we recognize the development of three distinct zones: a distal very fine grained,
light-brown phlogopitic facies; an intermediate fine-grained, light-gray, tremolitic facies;
26

and a proximal medium-grained, light-green diopsidic facies (Table 1, Fig. 10a, b). The
boundaries between these facies are commonly sharp and smooth but are locally irregular
where controlled by stockwork fractures (Fig. 10a, b). The outer limit of the distal brown
facies is typically hundreds of meters from the boundary of sulfide-bearing skarn, and the
progression from brown through gray to green facies occurs over distances ranging from
tens of meters adjacent to the main intrusion to tens of centimeters (Fig 10b) adjacent to
dikes extending beyond the main intrusion. Although local fluid channeling and probably
metasomatism occurred at the boundaries of the facies, the systematic mineralogic
zonation (Table 1) most likely reflects increasing temperature. The first appearance of light
brown calc-hornfels corresponds with the phlogopite-in reaction, Dol + Kfs = Phl + Cal,
which occurs in the range 350 to 465C under geologically reasonable conditions (i.e., P =
1,000 bars and XCO2 = 0.1 to 0.9: Tracy and Frost, 1991). The gray calc-hornfels contains
tremolite with or without phlogopite, whereas brown calc-hornfels is tremolite free (Table
1). The transition from brown to gray calc-hornfels therefore is interpreted to represent a
phlogopite-out reaction that coincides with the formation of tremolite, probably through
the reaction Phl + Cal + Qtz = Tr + Kfs (Tracy and Frost, 1991; Table 1). The lightgreen
calc-hornfels records the first appearance of massive diopside (Table 1), although veinlets
and small blebs of this mineral locally occur farther from the intrusive contacts. This zone
probably formed through the reaction Tr + Qtz + Cal = Di, which under similar pressure
and XCO2 occurs at temperatures of between 405 and 495C (Tracy and Frost, 1991). The
local narrowness of the gray calc-hornfels facies and the absence of tremolite in the
phlogopite facies suggest that the metamorphism took place at high XCO2. Under such
conditions
the diopside-in and phlogopite-in reactions are separated by only circa 10C. Also, the
development of tremolite through the elimination of phlogopite is promoted under these
conditions, but its formation within the field of phlogopite stability through the reaction
Dol + Qtz = Tr + Cal is inhibited.
In contrast to these calc-hornfelses that formed under essentially thermal
metamorphic conditions, calc-silicate development through metasomatism is revealed by
development of skarnoid in some beds of the Celendn Formation, extending tens of
meters beyond the mineralized skarn front. Skarnoid is a descriptive term for calc-silicate
rocks that are relatively fine grained, calcium rich, and iron poor, and that reflect, at least
in part, an aluminosilicate component in the protolith (Zharikov, 1970). It is genetically
intermediate between a purely metamorphic hornfels and a purely metasomatic, fluidcontrolled skarn, which is typically coarser grained and does not as closely reflect the
composition or texture of the immediately surrounding rocks (Einaudi, 2000). In some beds
in the Celendn Formation, diopsidic nodules stand out in relief against the calcite-bearing
calc-silicate matrix (Fig. 10c, d). The nodules, locally concentrically banded, range from
>10 cm to <1 cm in diameter and from sparse to abundant, independent of their size (Fig.
10c, d). They are interpreted as products of the preferential metasomatism of calcitic
nodules (cf. Fig. 8e, f).

27

FIG. 9. Marmorized Jumasha Formation. (a) Subtle bedding, defined by porphyroblast abundance (outlined) in calcitic
marble, dips moderately southwest and is overprinted by a northeast-dipping planar fabric (looking northwest, portal
of bulk sample adit, 450 m southwest of west shore of Lago Antamina; hammer for scale). (b) Subtle bedding, defined
by slight color difference and differential erosion (outlined), reflecting cryptic difference in porphyroblast abundance
in calcitic marble (looking northwest, 350 m southwest of west shore of Lago Antamina; 20 cm notebook in
foreground for scale). (c) Medium bedded layer of light green, fine-grained, diopsidic calc-hornfels defines bedding
within coarse-grained calcitic marble near the top of the formation (looking northeast, 300 m southwest of the
western shore of Lago Antamina; 15 cm ruler for scale). (d) Upper: mottled gray calcitic Jumasha Formation marble
with tightly folded, boudinaged, dark-gray silty laminae. Lower: mottled white and gray marble with tightly folded
buff garnet-rich layer, similar in shape to the silty layer in the upper core (DDH CMA-086, 218 m). (e) Banded gray
Jumasha Formation marble showing development of approximately 10 percent black scapolite crystals within darker
layers (DDH CMA-136, 232 m). (f) Upper: white marble in gray Jumasha Formation marble both as patches throughout
and locally around veinlets and stylolites (DDH CMA-136, 228 m). Lower: veinlet-controlled white marble in gray
Jumasha Formation marble cut by irregular veinlets with chalky-white wollastonitic selvages
(DDH CMA-236, 441 m). Scale bars in centimeters.

28

FIG. 10. Hornfels, skarnoid, and wrigglite developed in the Celendn Formation. (a) Irregular stockwork of sealed
fractures controlling development of light-gray tremolitic calc-hornfels in light-brown phlogopitic calc-hornfels (on the
southeast ridge crest, center of Fig. 12c; 15 cm pencil for scale). (b) Calc-hornfels facies with distal, light-brown, very
fine grained phlogopitic calc-hornfels at upper left, light-gray tremolitic in the center, and proximal, light-green
medium-grained diopsidic calc-hornfels at lower right. These facies are unusually closely spaced because they are on
the margin of a narrow dike, approximately 350 m southeast of the main intrusion (same location as [a]; 55 mm lens
cap for scale on the far left). (c) Sparse, large, medium-green diopsidic nodules in pale gray calc-hornfels (base of
cliffs, northeast of Lago Antamina; 10 cm knife for scale). (d) Very abundant coalescing diopsidic nodules with calcitebearing calc-hornfels matrix (base of cliffs, east of Lago Antamina; 15 cm pencil for scale). (e) Irregular concentric

29

banding, interpreted by Bodenlos and Ericksen (1955) to be of algal origin, but reinterpreted as wrigglite texture of
metasomatic replacement origin (base of cliffs, east of Lago Antamina; 15 cm pencil for scale). (f) Banding developed
in calc-hornfels and apparently controlled by northeast striking fractures (on the ridge crest southeast of Antamina
valley, just east of the east end of Fig. 12c; 55 mm lens cap for scale).

30

Both close to the skarn front and tens to hundreds of meters from it, sulfide-free
concentrically layered structures occur in some massive (nonnodular) beds of altered
Celendn limestone (Fig. 10e). These were interpreted as algal structures by Bodenlos and
Erickson (1955). However, they do not have the requisite three-dimensional geometry, and
no unambiguous algal structures have been seen in unaltered limestones in the area. Also,
at the transition from brown to gray facies in calc-hornfels, rhythmic banding is developed
adjacent to through-going, northeast-striking fractures (Fig. 10f). Similar laminated
features occur in ore-bearing wollastonite skarn. Because of their similarity to banded
features described in other skarns (e.g., Knopf, 1908; Eskola, 1951), we interpret the
layering as metasomatic wrigglite (Kwak and Askins, 1981). From these relationships, we
conclude that minor metasomatism occurred beyond the extent of metallic mineralization
in the Celendn Formation.
Structural geology of the mine area
The Upper Cretaceous strata of the Machay Group that host the skarn were
intensely deformed in the Maran thrust and fold belt. Southwest of the mine, the
Jumasha Formation is thrust over the younger Celendn Formation, and the Lower
Cretaceous formations of the Goyllarisquisga Group are thrust both over the Celendn
Formation and over each other in reverse stratigraphic order (Fig. 11).

Local folds and faults:


The Antamina deposit is hosted by a relatively flat-lying section of the Jumasha and
Celendn Formations, is deformed by open folds, and is cut by numerous small thrust-fault
ramps and bedding-parallel thrust faults (Fig. 11), all interpreted to have developed during
the late Eocene Incaic orogeny. Prior to open-pit development, the steep slopes around the
Antamina valley provided clear, three-dimensional images of the local structural
relationships (Fig. 12), summarized in an isometric block diagram in Figure 13. The
vergence of the thrust faults are interpreted from cutoff angles, and it is assumed that the
strata are all upright, unless obviously overturned. The local thrusting and related folding
are described below in sequence from southwest to northeast. Southwest of Antamina, a
northeast-verging thrust, herein named the Yaquirsh-Buque Punta thrust (YBPT in Figs. 11,
12, and 13), has caused structural repetition of the Jumasha Formation. A few hundred
meters to the east of this fault, the Jumasha Formation is thrust over the younger Celendn
Formation on the northeast-verging Antamina thrust fault (mine terminology; AT in Figs.
11, 12, and 13). The imbricate thrust tack is developed in massive limestones of the
Jumasha Formation (Fig. 8a) in the next transcurrent valley to the northwest of Antamina
(Quebrada Callapo in Fig. 2). This is part of a duplex in which the Antamina thrust is the
sole and one of the unnamed thrusts west of the Yaquirsh-Buque Punta thrust is the roof,
although the latter has been extensively eroded. These imbricate slices do not all continue
south from Quebrada Callapo into Quebrada Antamina because some fault surfaces
merge.
On the northwest slope of the Antamina valley, only the Jumasha Formation occurs
between the Antamina thrust and the Yaquirsh-Buque Punta thrust, whereas the Jumasha
and Celendn Formations in conformable stratigraphic sequence separate these two
thrusts on the opposing southeast slope (Figs. 11, 12, and 13). The immediately
31

underlying thrust slice, carried on the northeast-verging Fortuna thrust (mine terminology:
FT), occurs only on the northwest slope of the valley, east of and below the Antamina
thrust. In this area, the Fortuna thrust repeats the Celendn Formation and has a small
recumbent anticline in its hanging wall (Figs. 11, 12a, and 13). However, neither this thrust
nor its hanging-wall anticline can be traced across the valley to its southeast side (Fig.
12b, c), either because the thrust surfaces merged or because the branch line of the two
thrusts plunges north and therefore climbs above the erosion level to the south. The
Antamina thrust cuts the shallowly dipping Fortuna thrust (Figs. 11, 12a, 13) as well as
other minor, bedding-parallel thrusts in the Celendn Formation.
Most of the folds and thrust faults at Antamina are northeast verging, as is
characteristic of the Maran thrust and fold belt (Mgard, 1984), but two southwestverging thrustsoccur southeast of Lago Antamina (Figs. 11, 12c, and 13). These thrusts are
interpreted as backthrusts related to the underlying northeast-verging Oscarina thrust
(mine terminology: OT), which is the lowest in the immediate mine area. This thrust can be
traced along the crest of the southwest slope of the Tucush valley (Fig. 2), east of
Antamina. It also repeats the Celendn Formation and deforms it into a hanging- wall
anticline (Figs. 11; 12a, b, d; and 13). Many other minor thrust faults are recorded by
ramps with bedding cutoffs for distances of tens of meters, especially in the Celendn
Formation, but are not illustrated here. Prominent northwest-striking, open, upright
anticlines are exposed above the deposit on the northwest and southeast walls of the
valley (Figs. 7a; 12a, c; and 13). They become broader and more open upward but the
intervening synclines are tight. Such concentric folds are common elsewhere in the
Maran thrust and fold belt (e.g., Coney, 1971).
The largest of the parallel folds visible on the flanks of the Antamina valley has
been widely referred to as the Antamina anticline, the axial plane of which has an
apparent dextral offset of approximately 500 m in a northeast-southwest direction across
the valley. This offset has been variously ascribed to local bends in the strike of the folds
(Bodenlos and Ericksen, 1955), dextral offset on a postulated northeast-striking Valley
fault (Terrones, 1958), transverse normal faulting (Petersen, 1965,) and cross-folding
(McKee et al., 1979). However, in this study we show that the offset is best explained by a
lateral ramp model. In this model (Fig. 14), the anticlines represent fault-bend folds related
to hanging-wall cutoffs folded over the top of northwest-striking frontal ramps that are
offset because a northeast-striking, 500-m-long transfer fault or lateral ramp separates the
frontal ramps in an apparent dextral sense. Thispostulated structure, the Valley lateral
ramp, would account for the absence of corresponding offsets of other structures to the
southwest and northeast (Fig. 11).
Minor diking and skarn alteration extend beyond the main intrusion and orebody to
the southwest down Quebrada Antamina and to the northeast and east at the head of the
valley. The vertical dikes that extend east from the main porphyry mass to the Rosita de
Oro area (Figs. 11 and 12c) offset two minor forethrusts and backthrusts, indicating that
there was minor, apparently vertical, faulting between the time of thrust faulting and
folding and that of intrusion and mineralization. The orientation of these dikes changes
considerably along strike; near the main stock they are nearly vertical, strike east, and cut
strata (Fig 12c), whereas farther east beyond the ridge crest they become sills and follow
either strata or bedding-parallel thrust flats, strike south, and dip moderately to the west
(Fig. 12b). The dikes become progressively more altered to endoskarn toward the main
intrusion, and although they transect zoned exoskarn they do not appear to have intruded
it. A similarly variable orientation is shown by a dike that extends northeast from the
northern shore of Lago Antamina; it locally crosses strata at a steep angle but also follows
flat and moderately dipping strata (Fig. 12a). The parallel features at opposite ends of the
32

main intrusion and orebody suggest an offset, left-stepping, northeast-southwest fracture


zone that is longer than the 500 m lateral ramp, the postulated Valley fault (Figs. 3b and
11). A well developed northeast-striking, widely spaced, nearly vertical fracture set occurs
in the sedimentary host rocks throughout the area (e.g., on the southwest flank of Cerro
Racpe; Fig. 15a), and we interpret it as regional a-c jointing related to folding. A similarly
oriented fracture set is intensely developed near the deposit, especially along the
southeast side of the Antamina valley (Fig. 15b, c).
In many places, calc-hornfels is more intensely developed and widespread where
these fractures are closely spaced. All of the northeast-striking fractures may have been
related to Eocene folding and thrust faulting, but they were better developed in the mine
area because of flexure and tearing of strata above a transfer fault or lateral ramp
(discussed below). Following intrusion, they provided access for fluids to the
hornfelsbearing strata. In the Antamina district, therefore, the Jumasha and Celendn
Formations have been thrust-faulted, folded, and juxtaposed into a thick, complex thrust
stack. The total thickness
of rock that overlay the site of mineralization in the late Miocene cannot be estimated
because the local thickness of overlying uppermost Cretaceous and Paleocene red beds is
unknown and because it is unclear if the subaerial volcanic rocks and associated
sedimentary rocks of the Eocene to middle Miocene Calipuy Supergroup (Strusievicz et al.,
2000) extended into this area.

33

FIG. 11. Local geology of the Antamina area (modified after Cobbing et al.,1996; Glover, 1997, unpublished report to
Compana Minera Antamina S.A., Lima; Palomino, 1997, unpublished report to Compana Minera Antamina S.A., Lima;
and Love and Clark, 1998, unpublished report to Compana Minera Antamina S.A., Lima). The physiography of this
area is shown in Fig. 2. The areas depicted in Fig. 13 are indicated by the three diagonal rectangles. AT = Antamina
thrust, FT = Fortuna thrust, VF = left-stepping Valley fault, YBPT = Yaquirsh-Buque Punta thrust.

34

FIG. 12. Photomosaics illustrating the structure and stratigraphy of the Antamina mine area. Locations from which the
photographs were taken are indicated in Fig. 2. The views in (a) and (b) look northwest, whereas those in (c) and (d)
look south. Photographs taken in 1997 through 1999; much of (a), (b), and (c) has now been removed by mine
development, and the skyline of (d) has been modified. (a) The northwest side of the Antamina valley; the irregular
upper or northwestern contact of the skarn (long white dashes) cuts off the stratigraphic contact (long black dashes)
of the Jumasha (J) and Celendn (Ce) Formations. (b) The southeast flank of the ridge extending northeast of Cerro
Buque Punta (i.e., the opposite side of
the ridge illustrated in [c]), viewed from the Yanacancha area. (c) The southeast side of the Antamina valley; the
irregular upper, southeastern contact of the skarn (long white dashes) cuts off the stratigraphic contact (long black
dashes) of the Jumasha (J) and Celendn (Ce) Formations. (d) The northwest flank of the ridge extending northeast of

35

Cerro Yaquirsh (i.e., the opposite side of the ridge illustrated in [a]). Faults and fold axes are shown by solid black
lines, the stratigraphic contact between the Jumasha (J) and Celendn (Ce) Formations by long black dashes, the
contact of the skarn by long white dashes, intrusive contacts by solid white lines, intrusions by crosses. AT =
Antamina thrust, Ce = Celendn Formation, FT = Fortuna thrust, J = Jumasha Formation, OT = Oscarina thrust, VF =
surface expression of the northeastern segment of the Valley fault, YBPT = Yaquirsh-Buque Punta thrust.

36

FIG. 14. Schematic diagram illustrating the lateral ramp model for the structural setting of the Antamina deposit.
Looking south so that the face of the lateral ramp is exposed, showing offset anticlines produced in the hanging wall
of a thrust fault by two frontal ramps separated by a lateral ramp. (a) Geometry of the fault prior to movement. (b)
Geometry of the footwall (hanging-wall block removed), showing the southwest-dipping frontal ramps linked by a
northeast-striking lateral ramp or transfer fault. (c) After minor thrust movement, offset ramp-cutoff anticlines
produced in the hanging wall of the thrust fault, with more intense northeast-striking fracturing above the lateral
ramp, represented by dashed lines. (d) Hanging- wall cutoff anticlines separated from fault-bend anticlines by
additional thrust movement. Extensive fracturing developed in the thrust sheet where it flexed over the lateral ramp.

Structural control of mineralization:


As in most intrusion related skarn deposits, the most obvious feature controlling
mineralization at Antamina is the contact between the main stock and the host limestones
(Terrones, 1958; Petersen, 1965; Redwood, 1999). However, the southeast and northwest
intrusive contacts are themselves parallel to, and along strike from, the two segments of
the proposed northeast-striking Valley fault. The hydrothermal breccia sheets that are
common at or near the endoskarn-exoskarn contact also have this orientation. The
irregular zones of breccia and endoskarn within the intrusion are interpreted to strike
predominantly north-south (Fig. 3b) and may have been controlled by crosscutting
structures. The western end of the main body of the intrusion also generally strikes northsouth, as does the skarn front in that area (Fig. 3b). The parallelogram shape, in plan, of
the main body of the intrusion and the surrounding skarn is complicated by the network of
anastomosing dikes with
envelopes of fine-grained garnet skarn extending to higher
elevations to the east of Lago Antamina (Fig. 12c). At least one of these dikes intrudes a
normal fault on which movement occurred between the time of thrust faulting and folding
and that of intrusion and mineralization (Fig. 12c). Several other dikes with skarn
envelopes also extend beyond the main mass of porphyry and most are controlled by
Incaic structures. Along the southern edge of the porphyry and skarn, minor peripheral
rhyodacitic dikes and sills and their skarn envelopes locally follow bedding and thrust
faults. Several minor Pb-Zn- Ag veins (Fig. 2), some of which have been mined on a small
scale, occur at or near the contacts between these dikes and the host limestone,
describing a circa 3 3 km area of dispersed hydrothermal activity centered on the
Antamina skarn (Bodenlos and Ericksen, 1955).
Discussion
Regional stratigraphic and tectonic relationships
The postulated Valley lateral ramp at Antamina may have been controlled by underlying,
northeast-striking basement faults. Lateral ramps are thought to be largely the result of
the interaction of thrust sheets and old basement fracture systems (Pohn, 2000). Both
Mgard (1987) and Benavides (1999) proposed that the segmentation and articulation of
the Maran thrust and fold belt probably reflect control by basement structures, although
they differ in their interpretations of the mechanism. Benavides (1999) attributed the
segmentation

37

of the Maran thrust and fold belt to major


northeaststriking, dextral basement faults,
without specifying when the faults were active.
Further, Bussell and Pitcher (1985) suggested
that a well developed set of northeast-striking,
en echelon, dextral faults in the CretaceousPaleogene Coastal batholith may have
controlled some of the contacts of early
Paleocene intrusive ring complexes (Fig. 16).
However, no significant tear faulting parallel to
the northeast-trending deflections is apparent
in the deformed Mesozoic cover strata,
suggesting that the basement faults have not
been active since the Jurassic, and that any
dextral offsets are apparent, not real. Mgard
(1987), in contrast, proposed that en echelon
sinistral growth faults in the basement were
the common
boundary of the miogeosyncline and Yauli shelf
in the eastern part of the West Peruvian
trough, and controlled the attitude of the
present thrust and fold belt. The effects of the
cross-strike
structural
discontinuities
on
sedimentation have varied through time and
with rock type. A compilation of documented
stratigraphic columns (Fig. 16b) shows that
from Huancayo in the south to 50 km north
of Antamina the Cretaceous strata vary in
aggregate thickness from approximately 500
to 1,500 m and thicken overall to the north.
Generally, only the easternmost sections have
been used in this compilation, minimizing the
effect of thickness variations perpendicular to
the margin, and thus emphasizing along-strike
variations that may be related to segmentation
of the margin. It is evident that the Cretaceous
strata exhibit no clear relationship between
thickness and proximity to a crossstrike structural discontinuity.
FIG. 15. Northeast-striking, nearly vertical fracturing in the eastern and southeastern part of the Antamina mine area.
Locations from which the photographs were taken are indicated in Fig. 2. Photographs taken in 1998; much of (b) and
(c) has now been removed by mine development. (a) Looking eastnortheast at the southwest flank of Cerro Racpe
where all bedding dips southwest, toward the point of view. Widely spaced northeast-striking fractures are expressed
as lineaments that converge in the distance and are perpendicular to the strike of the Celendn (Ce) and upper
Jumasha (J) Formations on the lower slope, but are more obvious as vertical fractures in the lower Jumasha Formation
on the steep upper slope. Long black dashes indicate the stratigraphic contact between the Celendn and Jumasha
Formations. (b) Looking north-northeast along the ridge crest on the southeast side of the Antamina valley in the
immediate vicinity of the mine. The more intensely developed northeast-striking fracture set in the Celendn
Formation can be seen as closely spaced vertical fractures in the cliff face. (c) Looking northeast along the ridge
bounding the southeast side of the Antamina valley, near its head, in the immediate mine area. The closely spaced
northeaststriking fractures in the Celendn Formation strike away from the point of view across the steep slope in the
lower-right foreground. The fractures are less strongly developed along strike to the northeast on Cerro Aparina
(upper right).

38

A model for the geologic history of the Antamina region


The geologic history of the Antamina region from the formation of the West Peruvian
trough in the Jurassic to the

FIG. 16. Segmentation and articulation of the Maran thrust and fold belt and its effect on Cretaceous
sedimentation. (a) Simplified geology of the western Andes of central and north-central Peru, showing the major
anticlinal axes, faults and segment boundaries in the Maran thrust and fold belt (Benavides, 1999), the major
plutonic centers of the Coastal batholith and Cordillera Blanca batholith (Pitcher et al., 1985), and the numbered
locations of the sections used in (b). The major producing mines, Yanacocha, Pierina, Antamina and Cerro de Pasco
are also shown. (b) Longitudinal fence diagram of the Cretaceous stratigraphic section of central and north-central
Peru along the eastern margin of the Maran belt. Sections B-14, B-15, B-16, and B-19 from Benavides (1956),
sections 6, 9, 11, 14, 17, 20, and 22 from Wilson (1963), and section D from Manrique (1998). Antamina is located
between sections 6 and B-19.

Miocene intrusion and formation of the orebody is summarized schematically in Figure 17.
In this model, the marginparallel West Peruvian trough formed during the Middle or Late
39

Jurassic by extension on en echelon, northwest-striking normal faults separated by


northeast-striking transform faults (Fig. 17a, b; Mgard, 1987). The normal faults on the
western margin of the basement are thus right-stepping, but the segments experienced no
relative displacement and each northeast-striking transform fault experienced only
sinistral movement. This distribution of growth faults in the Jurassic would result in
promontories and reentrants in
the margin of the
West Peruvian trough, similar to
those of the larger-scale early
Paleozoic eastern margin of
North America (Thomas, 1977).
The
Querococha
arch
coincided with an intermittent
topographic high that developed
at least in the Middle Jurassic, or
even
throughout
the
Mississippian to Middle Jurassic
interval,
and
which
also
influenced the distribution of
Cretaceous
carbonate
rocks,
Paleocene red beds, and Miocene
igneous rocks. It is apparent that
the
development
of
the
northeaststriking
basement
structures
predated
JurassicCretaceous sedimentation. The
Querococha
arch
apparently
influenced the distribution of
Mississippian to Lower Jurassic
rocks northeast of Antamina,
resulting
in
either
local
nondeposition
in
the
Mississippian to Early Jurassic or
a Middle to Late
Jurassic
erosional unconformity (Fig. 17b).
Additional minor extension at
any angle discordant to the
northeast-trending faults could have resulted in reactivation of the original ly sinistral faults
as north-side-down normal faults

FIG. 17. Regional-scale schematic diagrams, looking east, illustrating the proposed structural evolution of the
Antamina area. (a) In the Middle or early Late Jurassic, the West Peruvian trough started to develop through formation
of an en echelon pattern of left-lateral growth faults on the western edge of the Maran metamorphic complex. The
Mississippian to Lower Jurassic sedimentary rocks that overlay the metamorphic complex are removed from the
diagram for clarity (after Mgard, 1987). (b) Also in the Middle or Late Jurassic, minor extension at an angle to the
transform faults segmented the continental basement, producing on the promontories structural highs parallel to the
original offsets, such as the Querococha arch, along which the Mississippian to Lower Jurassic sedimentary rocks were
eroded. In the Late Jurassic the Chicama Group was deposited in the western, deeper-water part of the basin, not
illustrated. (c) Cretaceous clastic and carbonate (Goyllarisquisga and Machay Groups) sedimentation on the Yauli
shelf was not strongly controlled by the segmentation although carbonate facies may have extended farther seaward

40

adjacent to promontories. (d) Latest Cretaceous to Paleocene deposition of red beds (checkerboard patterned) was
controlled by the segmentation, with depocenters in the reentrants. (e) In the Eocene Incaic orogeny, major strike
deflections and en echelon fold patterns that conform to the segmentation of the basement developed in the
Maran thrust and fold belt. Reentrants became structural salients, and promontories became structural recesses.
(f) In the Miocene, additional uplift of the Querococha arch reactivated basement structures, which allowed intrusions
to extend farther inland along the arch and to reach shallow levels. A = Antamina mine, CPB = Casma- Pasto Bueno
zone, Q = Querococha arch.

sinistral faults as north-side-down normal faults and hence erosion of preexisting


sedimentary rocks from the transverse highs (Fig. 17b). The effect of the Querococha arch
in the Late Jurassic cannot be deduced because the Upper Jurassic Chicama Group is not
exposed in the eastern part of the Maran thrust and fold belt.
In the Cretaceous, the Querococha arch and other transverse structures that
segment the Maran thrust and fold belt had little apparent effect on the thickness of
clastic (Goyllarisquisga Group) and carbonate (Machay Group) sedimentary rocks on the
Yauli shelf (Figs. 16b and 17c), but it appears to have affected the distribution of the
carbonate rocks. The Machay Group does not extend as far west to the north of the arch
as it does on the arch and to the south of it. From the arch north the westernmost limit of
outcrop of the Machay Group therefore crosses the belt in a northeastsouthwest direction
(Fig. 1). Thomas (1977) noted a similar relationship between the extent of carbonate facies
and the locations of structural salients and recesses in the Appalachians. The distribution
of latest Cretaceous and Paleocene red beds has also been influenced by the transverse
structures
(Fig. 17d). The red beds pinch out toward both the Querococha arch and the Casma-Pasto
Bueno zone, and they attain their greatest thickness between these arches in a structural
salient, which would have been a reentrant and
depocenter in the margin of the West Peruvian trough at the time of sedimentation.
Shortening and variations in the regional attitudes of folds and thrust faults
generated in the Eocene Incaic orogeny are represented in Figure 17e. The folds and
thrusts in the Antamina region constitute an articulated structural recess in the margin of
the craton, which probably formed through deformation around a basement promontory,
the northwestern edge of which is now delineated by the Querococha arch. Because the
orientation of folds and thrust faults in thinskinned tectonic belts generally reflects the
underlying ramps rather than the translation direction of deformation (Pohn, 2000), the
strike and articulation of the Maran thrust and fold belt mimic the geometry of the
basement, despite variations in the Cenozoic plate convergence direction and rate (PardoCasas and Molnar, 1987; Somoza, 1998; Norabuena et al., 1999). Old transform faults in
the margin of the West Peruvian trough did not experience extensive later strike-slip
movement because the maximum shortening direction was not parallel to the orientation
of movement during rifting.
We propose that the sinuous configuration of the mountain belt generally
reproduces the original zig-zag margin, although the articulation is pronounced on the
eastern side of the belt, and shortening in the interior of the belt had a smoothing effect
on the segmentation. Thus, the regional strike at the western extent of the Cretaceous
carbonate strata gradually changes throughout a distance of approximately 175 km along
strike, from northerly near Antamina to north-northwesterly at the northwest end of the
Cordillera Blanca (Fig. 1). As Thomas (1977) concluded in the context of eastern North
America, the Maran thrust and fold belt has formed a best-fit curve around old
promontories and reentrants.

41

The persistence of Miocene igneous rocks farther from the main axis of the
magmatic arc into the foreland thrust and fold belt along a proposed basement transverse
structure (Fig. 17f) is consistent with observations in other belts. Igneous intrusions have
been documented directly over, and elongated parallel to, three of the four lateral ramps
in the Appalachians analyzed in detail by Pohn (2000). Furthermore, the difference in
exposure level of the Miocene igneous rocks on and adjacent to the southwest end of the
arch implies that post-Eocene uplift enhanced the plunge reversals of the Eocene folds
across the arch (Love et al., 2001). Thus, faults parallel to the arch may have
accommodated its uplift and furnished the structural anisotropies that providedconduits
for magma ascent in the late Miocene.
Regional effects on local structure and mineralization
At Antamina, the northeast-trending fracture set, the Valley fault and the Valley lateral
ramp are parallel to the regional cross-strike structural discontinuity, the Querococha arch,
and may have been controlled by similarly oriented underlying basement structures. At
the local scale, the Valley lateral ramp and the Antamina intrusion have been localized by
the left-stepping jog in the Valley fault. Regionally, about 5 km southwest of Antamina, the
northeast-trending locus of changes in strike and plunge in the Maran thrust and fold
belt steps left by approximately 8 km (Fig. 5). Thus the step in the Valley fault developed
within, and mimics, a similar, larger-scale, left-stepping jog within the Querococha arch.
The local-scale structural evolution of the Antamina area, including intrusion and
formation of the orebody in the late Miocene, is summarized schematically in Figure 18.
Whereas other models, such as en echelon folding, could explain the apparent dextral
offset of the Antamina anticline, the lateral ramp model is preferred here because it
provides a locus for later, northeast-elongated intrusion and hydrothermal activity. The
northeast-striking fracture set in the host rocks peripheral to the ore at Antamina (Figs. 15
and 18a) was, we contend, formed by deformation associated with thrust translation along
the underlying and similarly oriented transfer fault or lateral ramp (Fig. 14). The overall
form of a thrust sheet does not record its passage over a lateral ramp beyond the limit of
that ramp but, in overlying thrust sheets, longitudinal fractures would form in the lateral
anticline over the ramp. These fractures would have sheared vertically in a north-side
down sense or opened owing to flexure above the lateral ramp or transfer fault and would
persist beyond it (Fig. 15d), providing evidence that a thrust sheet had traversed such a
ramp. At Antamina, the inferred Valley lateral ramp extends only 500 m from one offset
anticlinal axis to the other, but strong, northeast-striking, nearly vertical fracturing,
interpreted herein as evidence of tearing in the overriding thrust sheet, extends farther
northeast (Figs. 15 and 18a) and is interpreted as trace evidence of the Valley lateral
ramp. The upward-fanning axial planar cleavage related to the Antamina anticline also
formed at this time (Fig. 18a). In this model, the left-stepping jog in the trans current
Valley fault localized the Eocene development of the lateral ramp that resulted in
formation of the offset anticlines and also focused the Miocene igneous activity
responsible for the Antamina stock and skarn. Prior to intrusion of the main porphyry
mass, the early east-striking dikes east of the main intrusion formed in association with
east-west shortening localized at the northeast end of the southwestern segment of the
Valley fault (Fig. 18b). Many of the peripheral dikes and mineralized veins beyond
the main porphyry body intrude, or branch from, the same transcurrent structures. The
dike branches were controlled by other preexisting structures, predominantly bedding and
Incaic thrusts, and they may have intruded at this time. In the Jumasha Formation the
42

development of white marble was localized by preexisting axial cleavages, stylolites, and
fractures, providing early fluid pathways. Also, in the Celendn Formation calc-hornfels
formation was localized in many places where the northeast-striking, nearly vertical
fracture set was strongly developed. The primary control on the skarn and some of the
breccia zones was the margin of the main intrusion, which we argue was itself localized by
the postulated step in the Valley fault. The northwest and southeast margins of the main
intrusion appear to be defined by the northeast-trending Valley fault, and the main mass
of the intrusion occupies the left-stepping gap between the two offset segments of this
structure (Figs. 3b and 11). In such a system, any relaxation in east-west compression
could allow sinistral movement on oblique northeast- trending structures such as the
Valley fault and formation of extensional north-south-oriented structures within the leftstepping jog (Fig. 18c), thereby providing a locus for intrusion (Fig. 18d). Development of
northeast-striking breccia zones within skarn on the northwest and southeast margins of
the intrusion and north-striking breccia and endoskarn zones within the intrusion
represent, in this model, structural reactivation of the major transcurrent faults and the
minor north-south extensional faults, respectively.

43

FIG. 18. Schematic diagrams, looking east, illustrating the proposed local-scale structural evolution of the Antamina
area. (a) Structural elements in the Antamina area prior to intrusion and mineralization: frontal thrust ramps offset by
the Valley lateral ramp, which was localized by a jog in the transcurrent Valley fault; fault-bend-fold anticlines
overlying the tops of the offset frontal ramps; strongly developed northeast-striking fracturing parallel to and
overlying the lateral ramp; and upwardfanning axial cleavage in the overlying anticlines. (b) East-west compression
oblique to the Valley fault induced north-south extension and allowed intrusion of dikes in wing cracks localized at the
end of one segment of the Valley fault. (c) Relaxation of east-west compression allowed east-west extension
manifested as north-southstriking normal faulting in the jog between the two segments of the Valley fault. (d)
Monzogranitic magma intruded the parallelogram-shaped extensional regime between the two segments of the
Valley fault, and ultimately generated skarn ore. (e) Renewed east-west compression formed east-weststriking veins
in skarn.

44

45

Although the recrystallization of the host rocks during skarn development obscured
evidence of controls by preexisting fractures or bedding, we deduce that the intersecting
structures that locally control ore grades in skarn may have originated at substantially
different times. Terrones (1958) reported higher exoskarn ore grades where a set of 100
mineralized sheeted veins intersects structures extending from a northwest-striking
anticlinal axis in marble, and which we interpret as upward-fanning axial planar cleavages.
The intersecting sheeted vein set differs in orientation from the northeast-striking fracture
system related to the lateral ramp described above. We propose that these veins, which
cut skarn and were therefore late, formed in tension fractures associated with renewed
east-west shortening (Fig. 18e) after the brief episode of relaxation. Thus an Incaic axial
planar cleavage was intersected by Miocene tension veins and developed a permeable
path for hydrothermal fluids.
Implications for local exploration and ore genesis
The surrounding zones of marble in the Jumasha Formation, and particularly of hornfels
and skarnoid in the Celendn Formation, provide larger exploration targets than the skarn
itself. An isometric block diagram of the simplified geology of the deposit (Fig. 19) shows
the structure and alteration in the adjacent host rocks. Gray marble in the Jumasha
Formation generally extends tens of meters beyond skarn, although locally it extends
beyond 100 m, but the outer limit of the distal, brown, phlogopitic facies of calc-hornfels in
the Celendn Formation typically extends several hundreds of meters from the boundary of
sulfide-bearing skarn. The most extensive halo around the skarn is represented by the
swarm of Agbearing Pb-Zn vein deposits (Fig. 2). These occur within and beyond marble
and calc-hornfels in a 9 km2 area surrounding Antamina and up to a kilometer from the
skarn front.
The development of strong exoskarn mineralization may have been promoted by the
relatively pure calcitic limestone of the Jumasha Formation, which experienced intense
calcite destruction and Ca mobilization. In contrast, the overlying Celendn Formation,
having been metamorphosed to hornfels, may have acted as a cap on this hydrothermal
system, suppressing the upward and outward escape of fluids and thereby fostering
development of extensive endoskarn.
The same lithologic succession of pure carbonate overlain by muddy carbonate to
shale that could contain and concentrate a developing hydrothermal system, also occurs
in the lower transgressive sequence of the Machay Group where the shelly Pariahuanca
Formation limestone is overlain by the marl of the Chulec Formation (Figs. 4 and 16). In
addition, in northern Peru, similar prospective successions may occur at other levels within
the upper transgressive sequence of the Machay Group because at least six shallowingupward marl to limestone sequences (Jaillard, 1987) are recognized in the five formations
correlative with the Jumasha Formation.

Metallogenic and geotectonic implications


Several other ore deposits, in addition to Antamina, coincide with deflections in the
strike of the Maran thrust and fold belt. At the northern extremity of the Cordillera
Blanca (Fig. 1), 175 km to the north-northwest of Antamina, significant middle to late
Miocene intrusion-related mineral deposits, such as the Magistral Cu-Mo skarn prospect
46

and the formerly productive Pasto Bueno W-Cu-Ag vein system (Fig. 1), are associated with
the Casma-Pasto Bueno zone. Farther north, the Yanacocha Au district is localized in the
northeasttrending trans-Andean Chicama-Yanacocha structural corridor (Teal et al., 2002).
Such large-scale structural controls on the location of ore deposits in Peru have been
suggested by Vidal and Noble (1994), Petersen and Vidal (1996), and Rivera (1996), but
not at Antamina. Similarly, in the Appalachians, two of the four lateral ramps studied by
Pohn
(2000) are associated with an unusual abundance of mineral deposits, and many
minor mineral occurrences are associated with another (Coleman, 1988a, b). The Machay
Group has long been recognized as a metallogenically important stratigraphic interval in
central Peru (Petersen, 1965). It hosts many skarn and carbonate replacement-type
deposits such as, from northwest to southeast, Magistral, Antamina, Tuco-Chira,
Pachapaqui, Raura, Uchucchacua, Chungar, Santander, Yauricocha, and Pursima
Concepcin. Equivalent Albian to Turonian carbonate rocks of the Arcurquina and
Ferrobamba Formations in south-central Peru host Oligocene skarn deposits in the
Andahuaylas-Yauri belt (Noble et al., 1984; Soler et al., 1986). The Eocene to Miocene

FIG. 19. Schematic, exploded isometric block diagram of the Antamina deposit. Looking north, showing the major
folds and thrust faults, the Valley fault (VF), the extent of skarn development within and adjacent to the intrusion, the
extensive calc-hornfels and skarnoid development in the Celendn Formation, and the restricted marmorization of the

47

Jumasha Formation. The offset Antamina anticline (AA) is indicated, but the inferred Valley lateral ramp is hidden
from view in this perspective.

Calipuy Supergroup resulted However, along the Querococha arch, an apparent swarm of
intrusive bodies was emplaced during this interval, one of which generated the Antamina
hydrothermal system (Love et al., 2001).
Models for the genesis of giant porphyry Cu systems in the central Andean orogen
(Zentilli et al., 1988; Clark, 1993; Zentilli and Maksaev, 1995; Richards, 2000) postulate
that rapid ascent of magma is an important contributing factor in ore formation because it
would minimize modification through assimilation-fractional crystallization processes,
which are envisaged to decrease the Cu content of melts. Further, an unrestrictive
structure may be necessary to allow small bodies of felsic magma access to the upper
crust, because their low effective buoyancy would normally result in slow ascent, cooling,
and hence deeper crystallization. Thus, during the magmatic lull, the basement structures
controlling the Querococha arch may have provided the conduit necessary for rapid
emplacement of small volumes of fertile melt into the upper crust, allowing the Antamina
porphyry to crystallize in a suitably shallow environment favorable for mineralization.
Therefore, we conclude that the confluence of fertile Miocene magmatism and
reactive carbonate strata, essential for ore genesis at Antamina, was afforded by the
Querococha arch cross-strike structural discontinuity.
from suprasubduction zone magmatism (Noble et al., 1999). The terminal event of this
Supergroup is represented by the middle Miocene Carhuish pluton, dated at 13.7 Ma (U/Pb
zircon date, Mukasa, 1984), and coincided with the formation of many hydrothermal
centers in the Cordillera Negra such as the Pierina high-sulfidation epithermal Au-Ag
deposit (Figs. 1 and 16; Strusievicz et al., 2000). Subsequently, an approximately 5.5 m.y.
hiatus in major igneous activity preceded the late Miocene intrusion of the main phase of
the Cordillera Blanca batholith, the Cohup leucogranodiorite, at 8.2 0.2 Ma (McNulty et
al., 1998). During this period of relative magmatic quiescence, only minor volumes of a
tonalitic to dioritic marginal phase of the Cordillera Blanca batholith were intruded
(Beckinsale et al., 1985), and only a few scattered hydrothermal centers developed in the
Cordillera Negra (Strusievicz et al., 2000).
Conclusions
The world-class late Miocene Antamina skarn deposit is hosted by deformed Upper
Cretaceous carbonate strata of
the Machay Group. The relatively pure marbles and minor intercalated calc-hornfels that
host the ore deposit at surface in the southwest and at depth in its northeast sector
represent the upper part of the Jumasha Formation, whereas the strata that form the
ridges around Antamina and host the uppernortheast part of the deposit are assigned to
the lower part of the overlying, originally muddier, Celendn Formation (Fig. 18), here
uncharacteristically resistant to erosion owing to hornfels formation in proximity to the
Antamina intrusive center. Marble is developed for tens of meters adjacent to the skarn in
the Jumasha Formation, but distinctive calc- hornfelses and skarnoids persist for hundreds
of meters from the skarn front in the overlying Celendn Formation (Fig. 18). Around the
Antamina deposit, both this and the 9 km2 swarm of Pb-Zn-Ag vein deposits provide much
larger exploration targets than the skarn itself. Moreover, the systematic mineralogic
zonation from peripheral phlogopitic through tremolitic to proximal diopsidic facies in the
calc-hornfelses provides a vector toward the intrusion and, by extension, the associated
48

mineralization. The sedimentary succession that hosts the Antamina deposit was folded,
thrust-faulted, and thickened into a complex thrust stack during the late Eocene Incaic
orogeny, which generated the orogen-parallel Maran thrust and fold belt. Within this
stack, a left-stepping jog in the transverse Valley fault apparently controlled the Eocene
formation of the Valley lateral ramp, and together they localized the Miocene intrusion and
skarn. East-west diking east of the main intrusion and the northeast end of one segment of
the Valley fault is interpreted as early and associated with north-south extension related to
east-west shortening. Relaxation of this compression allowed east-west extension, which
would have been focused
in the left-stepping jog in the Valley fault, forming the locus for the main mass of intrusion
and the associated skarn ore. Renewed east-west shortening could have again induced
north-south extension and resulted in the late east-west vein system. The original intrusive
contact unambiguously controlled the location of the skarn, yet was itself controlled by
larger-scale structures. The Antamina hydrothermal activity occurred in a regionalscale,
northeast-trending, cross-strike structural discontinuity, the Querococha arch, which
articulates the Maran thrust and fold belt. About 5 km southwest of Antamina, the locus
of this articulation steps left along strike, a feature mimicked on a smaller scale by the
Valley fault. The arch is inferred to have affected sedimentary and magmatic processes at
least from the Jurassic to the Miocene and to have been controlled by a basement
structure, perhaps a transform segment of the original margin of the West Peruvian
trough. The arch was the northwestern edge of a promontory on which the Cretaceous
Machay Group carbonate rocks were deposited farther west than elsewhere along the belt.
It also allowed Miocene Magmatism to extend toward the foreland and intrude the Machay
Group. We envisage that the carbonate rocks of the Machay Group provided both chemical
and physical traps for ore-forming fluids. Intense exoskarn developed in relatively pure
Jumasha Formation limestone, whereas the Celendn Formation hornfelses capped this
system, promoting recirculation of hydrothermal fluids and extensive endoskarn
development. The Querococha arch provided a suitable structure for the intrusion to reach
the Machay Group at the hypabyssal depths required for fertile fluid release.

References
Allende, T., 1996, Geologa del cuadrngulo de San Pedro de Chonta (18-j), Boletn 68, Serie A:
Carta Geolgica
Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas,
Per, 218 p., 1 map, 1:100,000
scale.
Atherton, M.P., Pitcher, W.S., and Warden, V., 1983, The Mesozoic marginal basin of central Peru:
Nature, v. 305, no. 5932, p. 303306.
Barazangi, M., and Isacks, B., 1976, Spatial distribution of earthquakes and subduction of the Nazca
plate beneath South America: Geology, v. 4, p. 686692.
Beckinsale, R.D., Snchez-Fernndez, A.W., Brook, M., Cobbing, E.J., Taylor, W.P., and Moore, N.D.,
1985, Rb-Sr whole-rock isochron and K-Ar age determinations for the Coastal batholith of Peru, in
Pitcher, W.S.,
Atherton, M.P., Cobbing, E.J., and Beckinsale, R.D., eds., Magmatism at a plate edge: The Peruvian
Andes: Glasgow, Blackie and Son Ltd., p. 177202.
Benavides, V., 1956, Cretaceous System in northern Peru: Bulletin of the American Museum of
Natural History, v. 108, p. 353494.

49

1999, Orogenic evolution of the Peruvian Andes: The Andean cycle, in Skinner, B.J., ed., Geology
and ore deposits of the central Andes: Society of Economic Geologists Special Publication 7, p. 61
107.
Bodenlos, A.J., and Ericksen, G.E., 1955, Lead-zinc deposits of Cordillera Blanca and northern
Cordillera Huayhuash, Peru: United States Geological Survey Bulletin 1017, 102 p.
Broggi, J.A., 1942, Geologa del embalse del Ro Chotano en Lajas: Boletn Sociedad Geolgica del
Per, v. 12, p. 123.
Bussell, M.A., and Pitcher, W.S., 1985, The structural controls of batholith emplacement, in Pitcher,
W.S.,
Atherton, M.P., Cobbing, E.J., and Beckinsale, R.D., eds., Magmatism at a plate edge: The Peruvian
Andes: Glasgow, Blackie and Son Ltd., p. 167176.
Clark, A.H., 1993, Are outsize porphyry copper deposits either anatomically or environmentally
distinctive?, in Whiting, B.H., Hodgson, C.J., and Mason, R., eds., Giant ore deposits: Society of
Economic Geologists Special Publication 2, p. 213283.
Cobbing, E.J., Pitcher, W.S., Wilson, J.J., Baldock, J.W., Taylor, W.P., Mc-Court, W.J., and Snelling, N.J.,
1981, The geology of the western Cordillera of northern Peru: Institute of Geological Sciences
(London) Overseas Memoir 5, 143 p.
Cobbing, E.J., Snchez, A., Martinez, W., and Zrate, H., 1996, Geologa de los cuadrngulos de
Huaraz (20-h), Recuay (20-i), La Unin (20-j), Chiquin (21-i), Yanahuanca (21-j), Boletn 76, Serie A:
Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per,
281 p., 5 maps, 1:100,000 scale.
Coleman, J.L., Jr., 1988a, CSDs of the eastern United States, in Coleman, J.L., Jr., Groshong, R.H., Jr.,
Rheams, K.F., Neathery, T.L., and Rheams, L.J., eds., Structure of the Wills Valley anticline-Lookout
Mountain syncline between the Rising Fawn and Anniston CSDs, northeast Alabama: Alabama
Geological Society Annual Field Trip Guidebook 25, p. 4951.
1988b, Geology of the Anniston CSD, in Coleman, J.L., Jr., Groshong, R.H., Jr., Rheams, K.F.,
Neathery, T.L., and Rheams, L.J., eds., Structure of the Wills Valley anticline-Lookout Mountain
syncline between the Rising Fawn and Anniston CSDs, northeast Alabama: Alabama Geological
Society Annual Field Trip Guidebook 25, p. 4143.
Coney, P.J., 1971, Structural evolution of the Cordillera Huayhuash, Andes of Peru: Geological
Society of America Bulletin, v. 82, p. 18631883.
Cosso, A., 1964, Geologa del cuadrngulos de Santiago de Chuco (17-g) y Santa Rosa (18-g),
Boletn 08, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector
Energa y Minas, Per, 69 p., 2 maps, 1:100,000 scale.
Cosso, A., and Jan, H., 1967, Geologa de los cuadrngulos de Pumape (16-d), Chocope (16-e),
Otuzco (16-f), Trujillo (17-e), Salaverry (17-f) y Santa (18-f), 1967, Boletn 17, Serie A: Carta
Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, 141 p.,
4 maps, 1:100,000 scale.
Egeler, C.G., and De Booy, T., 1956, Geology and petrology of part of the southern Cordillera Blanca,
Peru: Verhandelingen van het Koninklijk Nederlands Geologisch Mijnbouwkundig Genootschap,
Geologische Serie, pt. 17, 86 p.
Einaudi, M.T., 1982a, Description of skarns associated with porphyry copper plutons: Southwestern
North America, in Titley, S.R., ed., Advances in geology of the porphyry copper deposits,
Southwestern North America: Tuscon, Arizona, University of Arizona Press, p. 139184.
1982b, General features and origins of skarns associated with porphyry copper plutons:
Southwestern North America, in Titley, S.R., ed., Advances in geology of the porphyry copper
deposits: Southwestern North America: Tuscon, Arizona, University of Arizona Press, p. 185210.
Einaudi, M.T., 2000, Skarns of the Yerington district, Nevada: A triplog andcommentary, in Dilles,
J.H., Barton, M.D., Johnson, D.A., Proffett, J.M., and Einaudi, M.T., eds., Contrasting styles of
intrusion-associated hydrothermal systems: Society of Economic Geologists Guidebook Series, v.
32, pt. 1, p. 101125.
Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits: ECONOMIC GEOLOGY
SEVENTY-FIFTH ANNIVERSARY VOLUME, p. 317391.
Eskola, P., 1951, Around Pitkranta: Suomalaisen Tiedeakatemian Toimituksia Sarja A (Annales
Academi Scientiarum Fennic Series A), v. III, n. 27, p. 196.

50

Gutscher, M.-A., Olivet, J.-L., Aslanian, D., Eissen, J.-P., and Maury, R., 1999, The lost Inca Plateau:
Cause of flat subduction beneath Peru?: Earth and Planetary Science Letters, v. 171, p. 335341.
Hampel, A., 2002, The migration history of the Nazca Ridge along the Peruvian active margin: A reevaluation: Earth and Planetary Science Letters, v. 203, p. 665679.
Harrison, J.V., 1940, The geology of the central Andes in part of the Province of Junn, Peru:
Quarterly Journal of the Geological Society, London, v. 99, p. 136.
INGEMMET, 1999, Mapa geolgico del Per: Instituto Geolgico, Minero y Metalrgico, Sector
Energa y Minas, Per 1:1,000,000, Digital Version CD-ROM. Jacay, J., 1996, Geologa del
cuadrngulo de Singa (19-j), Boletn 67, Serie A: Carta Geolgica Nacional: Instituto Geolgico,
Minero y Metalrgico, Sector Energa y Minas, Per, 215 p., 1 map, 1:100,000 scale.
Jaillard, E., 1987, Sedimentary evolution of an active margin during middle and upper Cretaceous
times: The north Peruvian margin from late Aptian up to Senonian: Geologische Rundschau, v. 76, p.
677697.
Knopf, A., 1908, Geology of the Seward Peninsula tin deposits, Alaska: United States Geological
Survey Bulletin 358, 71 p.
Kwak, T.A.P., and Askins, P.W., 1981, The nomenclature of carbonate replacement deposits, with
emphasis on Sn-F(-Be-Zn) wrigglite skarns: Journal of the Geological Society of Australia, v. 28, p.
123136.
Love, D.A., Clark, A.H., and Schwarz, F.P., 2000, The Antamina deposit, Ancash, Peru: Anatomy and
petrology of a giant copper skarn [abs.]: Geological Society of America Abstracts with Programs, v.
32, no. 7, p. A137.
Love, D.A., Clark, A.H., Strusievicz, O.R., and Lee, J.K.W., 2001, The regional tectonic setting of the
giant Antamina Cu-Zn skarn deposit, northcentral Peru [abs.]: Geological Society of America
Abstracts with Programs,
v. 33, no. 6, p. A358.
Love, D.A., Clark, A.H., Ullrich, T.D., Archibald, D.A., and Lee, J.K.W., 2003, 40Ar-39Ar evidence for
the age and duration of magmatic-hydrothermal activity in the giant Antamina Cu-Zn skarn deposit,
Ancash, north-central Peru [abs.]: Geological Association of Canada/Mineralogical Association of
Canada/Society of Economic Geologists Abstracts, v. 28, Abstract no. 396 CD-ROM.
Manrique, A.I., 1998, Promocin de la minera del carbn en El Peru: Geologa economica de las
cuencas de Alto Chicama, Santa, Oyn y Jatunhuasi: PROCARBON, Instituto Geolgico, Minero y
Metalrgico, Sector Energa y Minas, Per, 66 p.
McKee, E.H., Noble, D.C., Scherkenbach, D.A., Drexler, J.W., Mendoza, J., and Eyzaguirre, V.R., 1979,
Age of porphyry intrusion, potassic alteration, and related Cu-Zn skarn mineralization, Antamina
district, northern Peru: ECONOMIC GEOLOGY, v. 74, p. 928930.
McLaughlin, D.H., 1924, Geology and physiography of the Peruvian cordillera, Departments of Junn
and Lima: Geological Society of America Bulletin, v. 35, p. 591632.
McNulty, B.A., Farber, D.L., Wallace, G.S., Lopez, R., and Palacios, O., 1998, Role of plate kinematics
and plate-slip-vector partitioning in continental magmatic arcs: Evidence from the Cordillera Blanca,
Peru: Geology, v. 26, p. 827830.
Mgard, F., 1984, The Andean orogenic period and its major structures in central and northern Peru:
Journal of the Geological Society, London, v. 141, p. 893900.
1987, Structure and evolution of the Peruvian Andes, in Schaer, J.-P., and Rodgers, J., eds., The
anatomy of mountain ranges: Princeton, New Jersey, Princeton University Press, p. 179210.
Mourier, T., Bengtson, P., Bonhomme, M., Buge, E., Cappetta, H., Crochet, J.-Y., Feist, M., Hirsch, K.F.,
Jaillard, E., Laubacher, G., Lefranc, J.P., Moullade, M., Noblet, C., Pons, D., Rey, J., Sig, B.,
Tambareau, Y., and Taquet, P., 1988, The Upper Cretaceous-lower Tertiary marine to continental
transition in the Bagua basin, northern Peru: Paleontology, biostratigraphy, radiometry, correlations:
Newsletters on Stratigraphy, v. 19, p. 143177.
Mukasa, S.B., 1984, Comparative Pb isotope systematics and zircon U-Pb geochronology for the
Coastal, San Nicols and Cordillera Blanca batholiths, Peru: Unpublished Ph.D. dissertation,
University of California, Santa Barbara, 362 p.
Myers, J.S., 1974, Cretaceous stratigraphy and structure, western Andes of Peru between latitudes
10 and 10 30': American Association of Petroleum Geologists Bulletin, v. 58, p. 474487.
1975, Vertical crustal movements of the Andes in Peru: Nature, v. 254, p. 672674.

51

1976, Erosion surfaces and ignimbrite eruptions, measures of Andean uplift in northern Peru:
Geological Journal, v. 11, p. 2944.
1980, Geologa de los cuadrngulos de Huarmey (21-g) y Huayllapampa (21-h), Boletn 33,
Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y
Minas, Per, 153 p., 1 map, 1:100,000 scale. Noble, D.C., McKee, E.H., and Mgard, F., 1979, Early
Tertiary Incaic tectonism, uplift and volcanic activity, Andes of central Peru: Geological Society of
America Bulletin, v. 90, p. 903907.
Noble, D.C., McKee, E.H., Eyzaguirre, V.R., and Marocco, R., 1984, Age and regional tectonic and
metallogenetic implications of igneous activity and mineralization in the Andahuaylas-Yauri belt of
southern Peru: ECONOMIC GEOLOGY, v. 79, 172176.
Noble, D.C., Wise, J.M., Vidal, C.E., 1999, Episodes of Cenozoic extension in the Andean orogen of
Peru and their relation to compression, magmatic activity and mineralization, in Machar, J.
Benavides, V., and Rosas, S., eds., 75 Anniversario de la Sociedad Geologica del Peru, Julio, 1999:
Sociedad Geologica del Peru, Volumen Jubilar 5, p. 4566.
Norabuena, E., Dixon, T., Stein, S., and Harrison, C.G.A., 1999, Decelerating Nazca-South America
and Nazca-Pacific motions: Geophysical Research Letters, v. 26, p. 34053408. OConnor, K., 2000,
Yacimiento polimetlico Antamina: historia, exploracin y geologa, in primer volumen de
monografas de yacimientos minerales PeruanosHistoria, exploracin y geologa, volumen Luis
Hochschild Plaut: Lima, Peru, Instituto de Ingenieros de Minas del Peru, p. 231244.
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. 233248.
Petersen, U., 1965, Regional geology and major ore deposits of central Peru: ECONOMIC GEOLOGY,
v. 60, p. 407476.
Petersen, U., and Vidal C., C.E., 1996, Magmatic and tectonic controls on the nature and distribution
of copper deposits in Peru, in Camus, F., Sillitoe, R.S., and Peterson, R., eds., Andean copper
deposits: New discoveries, mineralization, styles and metallogeny: Society of Economic Geologists
Special Publication 5, p. 118. Pilger, R.H., 1981, Plate reconstructions, aseismic ridges, and lowangle subduction beneath the Andes: Geological Society of America Bulletin, v. 92, p. 448456.
Pitcher, W.S., Atherton, M.P., Cobbing, E.J., and Beckinsale, R.D., eds., 1985, Magmatism at a plate
edge: The Peruvian Andes: Glasgow, Blackie and Son Ltd., 328 p. Pohn, H.A., 2000, Lateral ramps in
the folded Appalachians and in overthrust belts worldwideA fundamental element of thrust-belt
architecture: United States Geological Survey Bulletin 2163, 71 p.
Redwood, S.D., 1998, The Antamina copper-zinc skarn deposit, northern Peru [abs.]: Geological
Association of Canada/Mineralogical Association of Canada/Canadian Geophysical Union Program
with Abstracts, v. 23, p.
153.
1999, The geology of the Antmina copper-zinc skarn deposit, Peru: The Gangue, Newsletter of
the Mineral Deposits Division, Geological Association of Canada, issue 60, p. 17.
Reyes, L., 1980, Geologa de los cuadrngulos de Cajamarca (15-f), San Marcos (15-g) y Cajabamba
(16-g), Boletn 31, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico,
Sector Energa y Minas, Per, 70 p., 3 maps, 1:100,000 scale.
Richards, J.P., 2000, Lineaments revisited: Society of Economic Geologists Newsletter, 42, p. 1 and
1420.
Rivera, J.N., 1996, El megafracturamiento pre-Mendaa de Casma-Pasto Bueno y su influencia en la
metalognia Andina: Instituto de Ingenieros de Minas del Per Revista Minera, v. 240, p. 612.
Samam-Boggio, M., 1980, ed., El Per Minero: Instituto Geolgico, Minero y Metalrgico, Sector
Energa y Minas, Lima, Per, 18 volumes.
Snchez, A., 1995, Geologa de los cuadrngulos de Culebras (20-g), Casma (19-g), Chimbote (19-f),
Boletn 59, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector
Energa y Minas, Per, 263 p., 3 maps, 1:100,000 scale.
Snchez, J., lvarez, D., and Lagos, A., 1998, Geologa de los cuadrngulos de Juscusbamba (16-i) y
Plvora (16-j), Boletn 119, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y
Metalrgico, Sector Energa y Minas, Per, 262 p., 2 maps, 1:100,000 scale.

52

Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe, O., Neeraudeau, D.,
Crdenas, J., Rosas, S., and Jimnez, N., 2002, Late Permian-Middle Jurassic lithospheric thinning in
Peru and Bolivia, and its
bearing on Andean-age tectonics: Tectonophysics, v. 345, p. 153181.
Soler, P., Grandin, G., and Fornari, M., 1986, Essai de synthse sur la mtallogenie du Prou:
Godynamique, v. 1, p. 3368.
Somoza, R., 1998, Updated Nazca (Farallon)-South America relative motions during the last 40 m.y.:
Implications for mountain building in the central Andean region: Journal of South American Earth
Sciences, v. 11, p. 211215.
Strusievicz, O.R., Clark, A.H., Lee, J.K.W., Farrar, E., Slauenwhite, M., and Hodgson, C.J., 2000,
Metallogenetic relationships of the Huaraz, Ancash, segment of the precious-base metal
subprovince of northern Peru [abs.]:
Geological Society of America Abstracts with Programs, v. 32, no. 7, p. A504.
Szekely, T.S., 1967, Geology near Huallacocha Lakes, central high Andes, Peru: American
Association of Petroleum Geologists Bulletin, v. 51, p. 13461353.
Teal, L., Harvey, B., Williams, C., and Goldie, M., 2002, Geology of the Yanacocha gold deposits,
northern Peru [extended abs.], in Marsh, E.E., Goldfarb, R.J., and Day, W.C., eds., Global exploration
2002: Integrated methods for discovery: Denver, Colorado, Society of Economic Geologists,
Abstracts, p. 4344.
Terrones, A.J., 1958, Structural control of contact metasomatic deposits in the Peruvian cordillera:
Mining Engineering, Transactions of the American Institute of Mining Engineers, v. 11, p. 365372.
Thomas, W.A., 1977, Evolution of Appalachian-Ouachita salients and recesses from reentrants and
promontories in the continental margin: American Journal of Science, v. 277, p. 12331278.
Tracy, R.J., and Frost, B.R., 1991, Ph ase equilibria and thermobarometry ofcalcareous, ultramafic
and mafic rocks, and iron formations, in Kerrick, D.M., ed., Contact metamorphism: Reviews in
Mineralogy, v. 26, p. 207289.
Vidal, C.E., and Noble, D.C., 1994, Yacimientos hidrotermales controlados por estructura y
magmatismo en la region central del Per: Instituto de Ingenieros de Minas del Per Revista
Minera, v. 230, p. 1619.
Wheeler, R.L., 1978, Cross-strike structural discontinuities, possible exploration tool in detached
forelands [abs.]: Geological Society of America, Southeastern Section Abstracts with Programs, v.
10, no. 4, p. 201.
Wilson, J.J., 1963, Cretaceous stratigraphy of central Andes of Peru: American Association of
Petroleum Geologists Bulletin, v. 47, p. 133.
Wilson, J.J., and Reyes, L., 1964, Geologa del cuadrngulo de Pataz, Boletn 09, Serie A: Carta
Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, 91 p., 1
map, 1:100,000 scale.
Wilson, J.J., Reyes, L., and Garayar, J., 1967, Geologa de los cuadrngulos de Pallasca (17-h),
Tayabamba (17-i), Corongo (18-h), Pomabamba (18-i), Carhuaz (19-h) y Huari (19-i), Boletn 16,
Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y
Minas, Per, 95 p., 6 maps, 1:100,000 scale.
1995, Geologa de los cuadrngu los de Pallasca (17-h), Tayabamba (17-i), Corongo (18-h),
Pomabamba (18-i), Carhuaz (19-h) y Huari (19-i), Boletn 60, Serie A: Carta Geolgica Nacional:
Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, 63 p., 6 maps, 1:100,000
scale.
Zentilli, M., and Maksaev, V., 1995, Metallogenetic model for the late Eocene-early Oligocene
supergiant porphyry event, northern Chile, in Clark, A.H., ed., Giant ore deposits II: Controls on the
scale of orogenic magmatic-hydrothermal mineralization, April 1995, Queens University, Kingston,
Ont., Giant Ore Deposits Workshop, 2nd, Proceedingssecond (corrected) printing, 1996, p. 152
165.
Zentilli, M., Doe, B.R., Hedge, C.E., Alverez, O., Tidy, E., and Daroca, J.A., 1988, Istopos de plomo
en yacimientos de tipo prfido cuprfero comparados con otros depsitos metalferos en los Andes
del norte de Chile y Argentina (English abstract): Actas, V Congreso Geolgico Chileno, Santiago, v.
1, p. B331369.
Zharikov, V.A., 1970, Skarns: International Geology Review, v. 12, p. 541559, 619647, 760775.

53

Zuzunaga, A., 2003, Closing the loopUsing actual concentrator performance to determine the
true value of ore sources [abs.]: Canadian Institute of Mining, Metallurgy and Petroleum, Annual
Meeting, 105th, Montreal, Quebec, May 57, 2003, Program with Abstracts, p. 170.

54

Copper Skarn Deposit at


Coroccohuayco, Tintaya District

FIG. 1. Map of Peru showing the location of Coroccohuayco in the Andahuaylas-Yauri belt (from Maher and Larson, 2007). eyes. (Contour inte

Deposit Setting
Coroccohuayco occurs at an elevation of 4,100 m in the altiplano of Peru. Surface
outcrop is poorly exposed and the majority of the deposit occurs below a depth of 200 m.
As a consequence, the present geologic investigation is based largely on drill core. The
presence of a large hydrothermal system is only weakly manifested at the surface through
outcrops of altered intrusive porphyries and small outcrops (<500 m 2) of oxidized garnet
skarn and weathered massive
magnetite. The Pre-Quaternary outcrop at Coroccohuayco consists of mostly unaltered
diorite which overlies the majority of the deposit.
Stratigraphy

55

Regional stratigraphy in the Tintaya district includes, from oldest to youngest,


Cretaceous clastic to carbonate sedimentary units, Tertiary diorite sills, and EoceneOligocene intrusive rocks of intermediate to felsic composition of the Andahuaylas-Yauri
batholith (De La Cruz B., 1995). These rocks are unconformably overlain by Cenozoic
volcaniclastic, lacustrine, alluvial, and glacial sedimentary rocks (Fig. 2). At Coroccohuayco
the Cretaceous sedimentary rocks consist of fine-grained clastic and relatively pure
carbonate rocks of the Mara and Ferrobamba Formations, respectively. The Mara Formation
(regionally referred to as the Murco Formation) comprises shale, siltstone, minor
sandstone, and minor rock gypsum and/or anhydrite in the Tintaya district. Drill
intersections at Coroccohuayco show the Mara Formation to consist of mudstone with
laterally discontinuous, lensoidal, sandy to silty beds of variable thickness, and up to 5 m
of laterally discontinuous massive anhydrite and/or gypsum near the Mara-Ferrobamba
contact. The formation is metamorphosed to biotite hornfels or quartzite, depending on
the local composition. Sedimentary bedding is preserved in some areas, but the biotite
hornfels is nonfissile and hard, with a dark brown to greenish color. The Mara Formation is
nonreactive and does not host significant mineralization at Coroccohuayco, although
elsewhere in the Tintaya district it is an important host to vein-style mineralization (e.g.,
Quechuas; Perell et al., 2003). The most important ore host at Coroccohuayco is the
Ferrobamba Formation (regionally known as the Arqurquina Formation), which is also the
principal host rock to mineralization at Tintaya and Las Bambas (Zweng et al., 1997;
Perell et al., 2003). Hornfels is less strongly developed in the thin discontinuous
mudstone zones in the Ferrobamba Formation than in the Mara Formation.

56

FIG. 2. Stratigraphic column of the Tintaya district, Peru. Skarn forms in the Ferrobamba Formation. Adapted from De La Cruz (1995) and Zwe

Structural geology
The altiplano of southern Peru has been affected by tectonic stresses transmitted from the
convergent boundary since the deposition of the Cretaceous sedimentary rocks up to the
present. The largest scale structures in the Tintaya district are generally gentle folds of
large wavelength of probable Inca I deformation phase (Benavides-Caceres, 1999; Espirilla
R., 2004). The fold orientations mostly trend from northwest to southeast in the Tintaya
district (Fig. 3). Much of the mineralization and alteration at Coroccohuayco occupies the
hinge zone and limbs of this syncline. Structural analysis by Espirilla R. (2004) at Tintaya
indicates that four main systems of faults are present in the district. These fault systems
strike northwest, northeast, east, and north. This faulting is interpreted to have been
produced during and since the later Incaic deformation phases, and the regional
intersections of the northwest- and northeast-striking fault systems controlled
emplacement of separate monzonite cupolas above an inferred large batholith at depth.
The east-striking fault system was important in controlling postskarn mineralization at
Tintaya (Espirilla R., 2004) and also possibly at Coroccohuayco. The earlier fault systems
were likely reactivated by subsequent tectonic activity. Small displacement faults at low
angles-to-core axis are common in all rock types at Coroccohuayco. Slickensides
commonly record dip-slip movement, although oblique-slip movement is also recognized.

57

Many of the small displacement faults and argillized fault zones are postskarn and
postmineralization.
Similar features at Tintaya are generally normal in their sense of displacement (Zweng et
al., 1997). Late basaltic
andesite dikes were emplaced along these structures and have a north -northeast to south
-southwest strike orientation comparable to similar late mafic dikes at Tintaya (Espirilla R.,
2004). Modeling based on drill core interceptions suggests the presence of a few large
(>100-m vertical) offsetting structures, based on the offset of the Mara-Ferrobamba
contact (Fig. 4) or by correlation of igneous dikes and sills (Fig. 5). Most of these structures
are interpreted to have a normal sense of displacement.
Igneous rocks
Coroccohuayco records a polyphase intrusive history similar to that observed at
Tintaya (Zweng et al., 1997) and other orphyry-related intrusive centers (Gustafson and
Hunt, 1975; Lanier et al., 1978; Lang and Titley, 1998; Richards et al., 2001). Igneous
rocks at Coroccohuayco are texturally and mineralogically similar to intrusive rocks
observed at Tintaya (Fernandez B., 2002). An early, composite dioritic phase present
throughout the Tintaya district intruded the sedimentary package mainly as sills that were
later intruded by monzonitico dikes. Diorite is volumetrically the most abundant igneous
rock at Coroccohuayco and occurs as intrusive bodies up to 300 m thick and overlies
mineralized skarn zones in the Ferrobamba Formation (Figs. 4, 5).

58

FIG. 3. Generalized geologic map of the Tintaya district, Peru. Adapted from De La Cruz (1995) and Espirilla R. (2004).
Shown is the outline of the deposit at Coroccohuayco and the location of Figure 4 (mine section 1400) and Figure 5.

59

FIG. 4. Geologic cross section along mine section 1400 at Coroccohuayco, showing the location of drill hole traces (numbered
above the ground surface). Note the general layered nature of the skarn alteration in the Ferrobamba Formation which lies
between the overlying massive diorite sill and the underlying biotite hornfels of the Mara Formation.

60

61

Major modal constituents are hornblende, oscillatory zoned plagioclase, magnetite, and
locally minor quartz and pyroxene (Table 1). The diorite in the deposit is texturally
heterogeneous, including fine-grained (microdiorite) to coarse, mafic-poor leucodiorite,
and is generally more coarsely crystalline than diorite at Tintaya. In some interceptions
diorite abruptly transitions to plagioclase cummulate rock containing intercummulate
hornblende. Internal contacts and apparent layering suggest that the diorite may be a
series of stacked sills or differentiated magma bodies resulting from crystal fractionation.
Lateral continuity of thin diorite sills is particularly limited in the southern part of the
deposit so hole to hole geologic correlation of this rock is unclear, and this area potentially
represents a zone of irregular interfingering of sills.
Similar to observations in the Tintaya deposit (Zweng et al., 1997; Fernandez B.,
2002), various porphyritic rocks intrude earlier rocks at Coroccohuayco with at least seven
mappable monzonitic subtypes, some of which are genetically related to skarn formation
and/or mineralization. These rocks are as yet undated but are inferred to be similar in age
to monzonitic rocks at Tintaya that yielded K-Ar ages of 33 to 34 Ma (Noble et al., 1984).
Porphyritic intrusions at Coroccohuayco are texturally and mineralogically similar to those
at Tintaya and are mainly distinguished by lower quartz and higher K-feldspar modes and a
finer grained groundmass (Fig. 6). These intrusions have been subdivided by timing
relative to skarn alteration into two main families referred to
as early synskarn monzonites A-B and late synskarn monzonites F-G-H. All early porphyries
are termed monzonite porphyries and are compositionally monzonite to granodiorite.
Postskarn monzonitic rocks consist of the locally mineralizing dike rock called monzonite
porphyry M and a similar barren monzonitic rock referred to as latite for exploration
purposes. Fernandez B. (2002) classified most skarn- and

62

FIG. 5. Geologic section showing generalized geology and skarn alteration through Coroccohuayco parallel to the
major fold axes in the district (N23W) at easting 16.3, from northing 1600 to 500 (approx 1 km). Not shown is
approximately 100 m of the overlying diorite sill covering this part of the deposit up to the surface. Drill traces
marked with * are projected onto the section from 70 m; all others are within 50 m of the section.

63

64

FIG. 6. Photographs showing textures and alteration in porphyries at Coroccohuayco. A. Monzonite F (synskarn) with weak K-feldspar veinlets a

mineralization-related porphyries at Tintaya as monzodiorite to monzonite. The latite


intrusion is similar to monzonite porphyry M in texture but is paragenetically later and
lacks evidence for copper mineralization. The timing relationships and modal mineralogy
for diorite, syn and postskarn monzonite
porphyries, and later dike rocks at Coroccohuayco are
presented in Table 1, with examples shown in Figure 6. Dikes of monzonite porphyry A-B
are the earliest synalteration igneous rocks and are primarily found in the southern part of
the deposit. These dikes tend to be irregular or have a north-northeast orientation. Dikes
of monzonite porphyry family F-G-H are late synskarn dikes found throughout the deposit
but are more common in the northern half. Monzonite G is a large, nearly equigranular
intrusion intercepted by drilling in the northeastern part of Coroccohuayco and has no
recognized genetic relationship to skarn or mineralization.
Dikes of monzonite porphyries F-H have an irregular geometry and trend generally
north-northwest to northwest. Monzonite
porphyry M is postskarn, volumetrically more important in the northern part of
Coroccohuayco, and tends to have a west-northwest orientation. Its modal composition
and texture is similar to the A-B monzonite rocks (Table 1), and it
tends to have stronger sericitic alteration than the early synskarn
rocks (see below). Although monzonite porphyry M

65

postdates the formation of skarn, it does locally contribute copper to the deposit. Dikes of
latitic composition as well as basaltic andesite also occur in Coroccohuayco. Latite is a
light-colored fine-grained porphyry with phenocryst plagioclase having a distinctive seriate
texture. Hornblende dominates the mafic mineral assemblage
and magnetite may be visually conspicuous because of the light color of the rock. The
basaltic andesite is a weakly porphyritic rock with phenocrysts of clinopyroxene. It is black
where fresh but is typically argillized as it occupies late fault zones. These rocks are
texturally and mineralogically similar to rocks observed cutting skarn and mineralization at
Tintaya (Fernandez B., 2002).
Deposit Alteration
Igneous rocks
The dominant alteration of diorite is weak propylitization (chlorite-amphibole-epidote) and
small zones of vein envelope- related albite-amphibole magnetite alteration in which
secondary amphibole locally replaces primary hornblende. Endoskarn is locally observed in
diorite near the contact with the underlying limestone and consists of an assemblage of
epidote plagioclase pyroxene zeolite titanite. Up to 20 m of massive texturally
destructive epidote (>80% mode) occurs locally, usually at the base of the diorite sill. This
calc-silicate alteration is not always indicative of massive exoskarn below. Endoskarn in
diorite may be weakly copper mineralized in veinlets or stringers.
Alteration in monzonitic porphyries is variable with potassic, albitic, and sericitic
alteration as the most common types apart from calc-silicate alteration. Locally, weakly
developed secondary biotite alteration of primary biotite and hornblende is present as well
as secondary K-feldspar in and around vein envelopes, however, K-silicate alteration is not
as important at Coroccohuayco as it is in other porphyry skarn systems (Meinert et al.,
2005) or in the principal stock at Tintaya (Espirilla R., 2004). Pyrite and chalcopyrite locally
replace mafic minerals in trace amounts in the more intense potassic alteration.
Sericitization is primarily manifested as chloritization of hornblende and biotite,
weakly to strongly developed sericitization of plagioclase, and titanite locally altered to
rutile, sericite, and calcite. K-feldspar phenocrysts are mostly unaltered where present.
Calcite is present in more weakly sericitized zones. Sericitization overprints the earlier
albitic and potassic alteration.
Endoskarn in porphyries consists of pale garnet, which may be strongly color zoned
with late, red-brown rims, and is accompanied by minor pyroxene, epidote, and
plagioclase (Fig. 6). Endoskarn is generally zoned from garnet dominant at the skarnintrusive contact to pyroxene and plagioclase into the less altered intrusion. Endoskarn is
mostly observed in monzonite porphyries F and H in the northern part of Coroccohuayco
and in porphyries A and B in the southern part. No endoskarn development has been
recorded in later intrusive phases.
Hornfels
Metamorphic alteration has strongly affected the mudstone of the Mara Formation.
In each deposit in the Tintaya district mudstone has been metamorphosed to a dense,
biotitic hornfels (Zweng et al., 1997). At Coroccohuayco, the metamorphism of the
66

mudstone has made it impermeable so that hydrothermal alteration is primarily confined


to vein envelopes. Veins present are mostly pyrite or anhydrite and/or gypsum, with
secondary biotite or sericitic alteration envelopes. Some late veins contain calcite. No
veins have been encountered in the Mara Formation with significant copper.

FIG. 7. Photographs of the weaker alteration styles in the limestone of the Ferrobamba Formation. A. Recrystallization
in limestone parallel to bedding laminations. ml are remnant micritic carbonate layers interlayered with
recrystallized limestone layers. B. Sample showing bleaching effects in limestone parallel to bedding laminations and
also a vein-related patch of bleaching and recrystallization. C. Weakly developed acicular pyroxene in gray marble
with bleached bands. Inset shows magnification of the white pyroxene in the marble. D. Light green pyroxene
developed along bedding laminations in calcsilicate marble.

67

Marble and skarn

The alteration within the Ferrobamba Formation at Coroccohuayco can be


subdivided into marble, calc-silicate marble, and massive skarn, typically with
gradational variations between each. The calc-silicate alteration at Coroccohuayco
is spatially associated with monzonite porphyries which are typically
hydrothermally altered and locally show strong calc-silicate alteration adjacent to
their contacts. Calc-silicate marble is a gray to tan-white rock, similar to recrystallized
limestone but containing fine-grained phyllosilicates, calc-silicates, silica, albite, hematite
and/or sulfides. The most common silicate mineral observed is fine-grained disseminated
acicular diopsidic pyroxene. This pyroxene occurs in both bleached and unbleached marble
(Fig. 7). Rarely, garnet is observed disseminated in bleached marble. Calc-silicate marble
represents a transitional alteration between marble and skarn and may grade into wellmineralized garnet over only tens of centimeters in drill interceptions. Skarn is hosted by
the Ferrobamba Formation at Coroccohuayco. Mineralogically, it consists dominantly of
garnet, pyroxene, magnetite, hematite, amphibole, quartz, and carbonate minerals. Early,
high-temperature prograde anhydrous minerals are dominated by garnet and pyroxene,
with later,
lower temperature magnetite locally replacing both of these minerals. A general
paragenesis of the alteration and mineralization in the Ferrobamba Formation is presented
in Figure 8. Amphibole, chlorite, quartz, carbonate, Fe oxide, and serpentine minerals are
generally found in pyroxene skarn that has been subjected to retrograde alteration and is
not particularly widespread. Locally, fine-grained clay-hematite-carbonate alteration of
skarn is observed around fault zones.

68

The garnet in skarn is color zoned relative to proximity to fluid source or principal
fluid pathways as in other copper skarn deposits, with proximal garnet being dark redbrown and massive and becoming more green and granular in habit with distance The

FIG. 8. Schematic diagram of skarn paragenesis and later alteration at Coroccohuayco. Data temperatures are general ranges and have been

granularity is defined here as aggregates of crystals less than 5 mm in diameter (Fig. 9).
Massive garnet tends to lack this texture. Green garnet tends to be granular, locally
replaces early pale pyroxene, and is commonly rimmed and replaced by later darker redbrown or tan garnet. Garnet skarn contains disseminated mineralization and is the major
ore at Coroccohuayco. In general, garnet more distal to fluid sources tends to have a more
granular and less massive appearance.
The proportion of noncalc-silicate material consisting primarily of albite, quartz,
chalcedony, magnetite, and Cu-Fe sulfides is lower in proximal garnet compared to more
distal garnet skarn. Zones of massive honey-brown garnet
occur in deep skarn to the east of easting 22 in the northern part of Coroccohuayco (Fig.
4). Mineralization in this particular
garnet alteration is minor with pyrite, specular hematite, and chalcopyrite. This is quite
different from skarn mineralization elsewhere. Light-colored pyroxene as patches with
garnet is common in more distal alteration at Coroccohuayco. However, pyroxenedominant skarn is volumetrically less important than garnet- dominant skarn. Pyroxene as
massive skarn may contain minor carbonate, quartz, Cu-Fe sulfides, and garnet. A horizon
of generally fine-grained pyroxene skarn is found immediately above the Ferrobamba
-Mara contact in Coroccohuayco (Fig. 4). This alteration may have been developed in a
higher clastic or possibly dolomitic zone within limestone immediately above the
Ferrobamba-Mara contact.
69

At Coroccohuayco there are two centers of intense skarn alteration observed in


the southern and northern parts of the deposit. The skarn sequence in the southern
part is complicated by a large mass of endoskarn in monzonite porphyry, as well as
abundant smaller diorite sills (Fig. 5). The northern part of the deposit shows a thick
package of exoskarn, which The skarn zones are characterized by generally massive
calc-silicate alteration from the base of the Ferrobamba Formation to the base of the
diorite and contain garnet much greater volumetrically than pyroxene. Laterally to
these massive skarn zones the alteration grades from garnet dominant into garnetpyroxene skarn, and pyroxene dominant, with increasingly higher volumes of calcsilicate marble present in layers between skarn. Locally, garnet-dominant alteration
abruptly changes to marble or calc-silicate marble without an intervening pyroxenedominant zone. Interestingly, no large intrusive masses potentially responsible for the
massive skarn alteration in or around easting 16 of this section have been identified
and only postskarn monzonite porphyry M dikes are present. Trends in garnet color and
texture suggest that on a local scale within a single continuous package of calc-silicate
alteration garnet zonation occurs vertically over tens of meters from proximal to more
distal garnet. An example of such zonation is observed in drill hole 1400 17.7 from a
depth of 315 to 345 m. This interval occurs in a garnet-dominant package of
alteration, but it contains five distinct zones of redbrown rimed green garnet
interlayered with either red-brown garnet or green garnet containing pyroxene
patches. Commonly the mineralization in the skarn host increases from massive garnet
to granular garnet (Fig. 9) as the replaceable material increases.
Retrograde and low-temperature alteration
Retrograde alteration is manifested by locally intense and nearly complete
replacement of garnet and pyroxene by massive magnetite which may locally exceed
10 m vertically. Drill core and petrographic observations show that magnetite
mineralization occurs with other phases such as quartz, carbonate, amphibole, or
serpentine.
Because magnetite replaces calc-silicate minerals, particularly garnet, and is
associated with silica (mostly as quartz) and calcite, it is considered a retrograde
alteration product of garnet although this reaction assemblage is not typical of the
hydrous silicate (e.g., epidote) retrograde alteration observed in many skarns (Meinert
et al., 2005).
In some cases the magnetite alteration contains little of the other two
retrograde alteration products. Much of the magnetite is intergrown with copper
sulfides and in some zones of the deposit the magnetite has pseudomorphically
replaced bladed hematite which itself formed paragenetically after honey-brown
garnet or white pyroxene. In a few drill intersections magnetite is observed as small
(<0.5 m) zones near skarn-marble contacts which replaces marble. The lowest
temperature retrograde alteration in garnet is characterized by calcite-hematite
alteration along microfractures and in pyroxene by patchy calcite-silica amphibole
chlorite serpentine, but this alteration is not as extensive in Coroccohuayco as in
other shallow skarn systems (Meinert et al., 2005).
The most intense low temperature clay-carbonate alteration of calc-silicates is
related to large fault zones in the deposit. There is little introduced mineralization
associated with low-temperature alteration at Coroccohuayco, although supergene70

related chalcocite exists in and near some fault zones. Meteoric oxidation or
weathering of the principal mineralized zones is effectively absent.

Mineralization

Igneous rocks
Most of the intrusive rocks at Coroccohuayco are poorly mineralized even
though quartz veins may be present. Mineralization in the monzonite porphyry rocks A,
B, F, G, and H is generally less than 0.5 percent Cu, although monzonite B may locally
contain up to 1.5 percent Cu in short drill intercepts as chalcopyrite in veins and
disseminated in mafic sites. Monzonite porphyry M is generally unmineralized except
in the northern part of the deposit where it is locally highly veined and achieves high
copper grades (>5 % Cu) where it cuts existing skarn. This mineralization originates in
the interior of the porphyry as finely sheeted quartz-chalcopyrite molybdenite
veinlets (1-mm width, with a density up to 4/cm) and becomes more massive toward
the periphery of the dike with decreasing molybdenite. Quartz-bornite chalcopyrite
veins (up to several cm in width) increase in width and copper content in the same
direction. Outside the dike mineralization occurs with silicification and overprints
earlier calcsilicate alteration, locally with magnetite mineralization (including
retrograde alteration of skarn). Although quartz K-feldspar veinlets occur in most
monzonite porphyry M dikes in section 1400, only in parts of drill holes 17.7 and 18.9
has high-grade copper mineralization been encountered. In drill hole 1400 18.9
greater than 3 percent copper postskarn vein-related mineralization occurs in a
vertical intercept of about 80 m, although its true width is less than this. Molybdenite
mineralization in quartz veins is common but weak in monzonite porphyry M and not
present in other igneous rocks at Coroccohuayco.
Skarn
The bulk of the mineralization at Coroccohuayco is disseminated in skarn and
closely correlates with the distribution and style of calc-silicate alteration. As such it
has a complex distribution in the deposit in terms of mineralogy and concentration.
Green-tan granular garnet is the most significant ore host because it tends to have
more bornite than chalcopyrite relative to proximal garnet and the sulfides are
volumetrically greater. This style of alteration is located between the massive darkcolored proximal garnet and skarn-marble front or the pyroxene alteration zone if
present. Cu grades greater than 2 percent are common due to the presence of more
abundant replaceable interstitial carbonate in the green granular garnet skarn. In
some places a porous or vuggy texture occurs where only part of the calcite was
replaced by Cu sulfides and the rest was removed. Mineralization in pyroxene is
generally bornite-chalcopyrite dominant but pyroxene is generally a host of poor grade
and irregular distribution at Coroccohuayco. Mineralized veins which cut skarn are
generally confined to zones nearest late mineralizing intrusions and are also observed
71

in biotite hornfels. Vein-type mineralization is typically of lesser importance than the


disseminated mineralization in skarn, although exceptions exist near some monzonite
porphyry M dikes as described above. Manto-style or massive mineralization outside of
skarn is rare and only locally manifested as 10-to 20- cm massive sulfide specular
hematite zones at the skarnmarble contact with pyrite and minor chalcopyrite. Locally,
poorly mineralized massive magnetite is also observed at the skarn-marble contact.
Ore petrography

In thin section the most important ore minerals identified include chalcopyrite, bornite,
chalcocite, and digenite (Fig. 10). It is clear from the ore textures and field relationships
that bornite, and the majority of chalcocite and/or digenite, are hypogene in origin. Bornite
commonly occurs as inclusions, blebs, stringers, and patches in chalcopyrite grains in
chalcopyrite-dominant mineralized zones, and chalcopyrite occurs as similar textures in
bornite in the bornite-dominant mineralized zones. In bornite-chalcocite digenite zones
there are two main textural styles with chalcocite as inclusions, blebs, patches in
myrmekitic textures, and oriented intergrowths in bornite and locally as late, possibly
supergene enriched rims on bornite in zones near fault zones.

72

FIG. 10. Photomicrographs of sulfide textures locally observed at Coroccohuayco. A. Possible subsolidus exsolution
texture of bornite from chalcopyrite (cpy). B. Bornite (born) grain with zones of crystallographically exsolved
intermediate Cu- Fe sulfides (circled) with chalcopyrite. C. Complex sulfide grain with a chalcopyrite core and bornitechalcocite (cc) mixed outer zone. D. Bornite grain with minor chalcopyrite domains scattered along the grain
boundary. E. Bornite grain with minor chalcocite and scattered Ag-Au telluride phases (Ag). F. Bornite-chalcocite grain
with electrum (Au) on the rim with chalcocite. Gangue is garnet-carbonate.

Pyrite is also found in certain zones of honey-yellow garnet skarn with specular
hematite. Gold and silver minerals are observed as small 1- to 5-m inclusions in other
sulfide grains. They tend to be most abundant in the bornite-rich mineralization and occur
close to the sulfide grain rim (Fig. 10). Because of the common association with bornitechalcocite, gold and silver tend to be highest in the green garnet alteration zones where
the bornite is highest.
Magnetite

73

Magnetite occurs in several contexts at Coroccohuayco. It may occur as very fine grained
laminations and bands with pyroxene with no evidence of it being a replacement phase. It
more commonly replaces garnet and pyroxene as late alteration. In this latter context
magnetite alteration may occur massively, as patches, or banded in contorted forms which
are unrelated to sedimentary bedding. Magnetite locally replaces earlier formed bladed
hematite, although this is much rarer than replacement of calc-silicate minerals. Massive
magnetite in relatively small bodies (<2-m drill intercept) is observed near the marbleskarn contact, where pH is likely to increase due to fluid interaction with carbonate.
Locally, massive magnetite is brecciated and may contain a calc-silicate matrix. The
matrix of this breccia is locally a two-stage event, with early calcite matrix being replaced
by later calcsilicate
minerals. As discussed earlier, magnetite is commonly associated with sulfide
mineralization. In some cases the sulfide minerals petrographically appear interstitial to
magnetite grains, indicating a slightly later paragenesis. In other cases, the magnetite
contains
chalcopyrite
and
bornite
inclusions,
which
suggest
it
formed
contemporaneously with sulfide mineralization.
Igneous Rock Geochemistry
Diorite and monzonite porphyry rocks at Coroccohuayco are calc-alkaline in
geochemical affinity using the Al 2O3 (wt %) versus normative plagioclase division of Irvine
and Baragar (1971), although the most primitive diorite at Coroccohuayco falls in the
tholeiitic field on their AMF diagram. There is little difference in major or rare earth
elements between the fresher monzonite rocks (Table 2). The most outstanding feature of
these rocks relative to rocks of similar composition in other porphyry systems is their
elevated Na2O . Porphyry rocks in the Tintaya district (Fernandez B., 2002) show similar
characteristics in alkalis indicating that these chemical features are common to the
district-wide igneous petrogenesis and not simply the result of local alteration effects.
In general, these rocks show limited enrichment in REE indicating a relatively
primitive source magma. The monzonitic rocks show slight enrichment in light rare earth
elements and slight depletions in the middle and heavy rare earth elements relative to two
diorite samples.

74

Mineral Chemistry

FIG. 11. Variation diagram in Na2O vs. SiO2 for igneous rocks from Coroccohuayco and other porphyry systems. Rocks from Coroccohuayco an

Calc-silicate minerals in copper skarn systems are compositionally zoned in major


elements such as Fe, Al, Mg, and Mn (Meinert et al., 2005). Garnets from different parts of
Coroccohuayco were analyzed by electron microprobe to quantify spatial and temporal
compositional variations and to confirm the presence of a metamorphic garnet
component. Pyroxene at Coroccohuayco was also analyzed to identify any elemental
trends present, primarily due to the presence of late green pyroxene locally observed in
more proximal skarn zones. Microprobe analyses are shown on the calcic garnet and
pyroxene end-member ternary diagrams in Figure 14.
Garnet
Green, honey yellow, tan, brown, and red-brown garnets all have highly andraditic
compositions (An >90; Table 3) except in the darker optically isotropic cores which tend to
have a higher grossular content. Distinct color zones within the same

75

garnet crystal commonly have similar major element chemistry and little systematic
compositional variation is recognized in the samples analyzed, apart from the rim-core
relationship. The highly andraditic compositions of exoskarn garnets indicate that the
hydrothermal fluid was oxidized and Fe rich during the replacement of carbonate with
garnet. There is a population of brown garnets with higher grossular compositions found in
the northernmost part of the deposit and it is possible that these aluminous garnets were
formed in a protolith of slightly higher clastic component than the others.
Endoskarn garnet is a pale tan color although commonly with darker rims. The
composition of endoskarn garnets is much more aluminous than the great majority of
exoskarn garnets (Table 3). Chemical zonation is more strongly pronounced and
76

corresponds to iron content with the pale tan cores more grossularitic and the darker
garnet rims more andraditic. The pyralspite component of all garnets is generally less than
10 percent.
Pyroxene
The early massive pyroxene is white and diopsidic in composition with no clear spatial
trends. It was confirmed by microprobe analysis that the acicular mineral in calc-silicate
marble is diopside (Fig. 14). Compositionally the late green pyroxene is more hedenbergitic
with elevated sodium concentration indicating that these pyroxenes were a product of less
oxidized, Fe- and Na-rich fluids (Fig. 15) that infiltrated locally overprinted earlier skarn. In
all types of pyroxene the johanssenitic component is very low, typical of pyroxene from
oxidized Cu skarns (Meinert et al., 2005).

FIG. 12. Major element data graphs of Coroccohuayco intrusive rocks relative to the labeled average values of igneous rocks from different ska

77

78

FIG. 13. Rare earth element plot of Coroccohuayco intrusions normalized to chondrite values taken from McDonough and Sun (1995). All the

Sulfide and precious metal mineralization chemistry


Textural and mineralogical characteristics of disseminated and vein mineralization were
evaluated by reflected light microscopy. Colors of bornite and chalcocite and/or digenite
appear to vary between samples in freshly polished
sections. Because of these
differences, chemical analyses by electron microprobe of ore minerals were performed at
the Geoanalytical Laboratory at Washington State University. Electron microprobe analyses
were made on polished areas away from phase intermingling. Table 4 lists representative
analyses of bornite, blue bornite, chalcocite, digenite, and chalcopyrite.

79

Two populations of chalcopyrite were dentified: those close to ideal chalcopyrite


composition and those which were enriched in Cu relative to S (Fig. 16), although they are
visually similar. The blue bornite contains a slightly higher atomic ratio of copper/sulfur
than typical bornite. The bornite compositions as a whole lie closer to ideal bornite than
either the analyzed chalcopyrite or chalcocite-digenite compositions. When the

FIG. 14. Ternary diagrams of garnet and pyroxene end members for skarn minerals from Coroccohuayco. Late pyroxenes are distinctly enriche

chalcopyrite-bornite pairs from the same sample are compared, they show that the
samples with Cu-enriched chalcopyrite have a slightly higher Cu:S in bornite, and those
with slightly Cu depleted or close to ideal chalcopyrite also showed ideal bornite
compositions.
Chalcocite and/or digenite generally contains some iron relative to ideal chalcocite,
with the lowest Fe bearing chalcocite from minor late, low-temperature supergene
chalcocite veins cutting bornite in retrograde altered skarn. Gold and silver mineralization
in Coroccohuayco is associated with bornite-chalcocite digenitedominant hypogene
mineralization. The precious metals occur as electrum and silver-gold tellurides (Table 4),
both of which are very fine grained, typically <10 m in size (Fig. 10). As discussed before,
these grains are mostly observed near the rims of sulfide grains.

80

FIG. 15. Variation diagram of Na vs. Fe for pyroxenes from Coroccohuayco. Green (more hedenbergitic) pyroxene also contains more sodium. T

81

Fluid Inclusions
The inclusions in this study are classified as primary, pseudosecondary, and
secondary principally based on their spatial relationship to other fluid inclusions and
growth zones in crystals as well as their composition. This includes the presence of
daughter minerals and the approximate ratio of the vapor bubble to inclusion volume.
These tend to have a relatively uniform vapor/liquid ratio of about 0.25 to 0.4 and are
three-phase consisting of vapor, liquid, and one or more daughter minerals. These
daughters are dominated by halite, but sylvite, hematite, opaque phases, and unidentified
anisotropic daughters may also be present. Primary inclusions in pyroxene and garnet tend
to have more than one daughter mineral. Threephase fluid inclusions are also observed as
planar arrays cutting calc-silicate crystals. These are interpreted as pseudosecondary
inclusions based on their context (Fig. 17) and are considered to have formed
contemporaneously with the alteration. Pseudosecondary inclusions have similar
vapor/liquid ratios to primary inclusions observed in the same and
nearby crystals. These inclusions are grouped with the primary fluid inclusions in data
analysis because of their similarity in salinity and homogenization temperatures (T h).

Vapor/liquid ratios in primary inclusions in quartz tend to be less than 0.5 and halite
daughter minerals are common. In some samples of vein quartz, however, irregularly
dispersed (primary) inclusions have highly variable vapor/liquid ratios and many lack
halite daughters. The presence of these inclusions indicates that an H 2O-boiling fluid was
locally trapped in quartz. These types of inclusions are generally rare in calc-silicates.
Secondary fluid inclusions are identified as those present in trails that cut through garnet,
quartz, and pyroxenes, are vapor rich, and lack halite daughters. As such the secondary
82

inclusions are visually distinct from the irregularly dispersed primary inclusions as well as
the pseudosecondary inclusions described above.

FIG. 16. Diagram from the Cu-Fe-S ternary with microprobe analyses of
Cu mineralization at Coroccohuayco. The diagram shows the region between 30 and 60 percent S. End-member compositions are shown by th

Fluid inclusion analyses of primary inclusions in prograde skarn and quartz were
performed on a USGS-style heatingfreezing stage at Washington State University, which
was calibrated using synthetic inclusion standards (Sterner and Bodnar, 1984). The results
represent homogenization temperatures and no pressure correction has been applied to
the presented data. A pressure estimate of about 25 MPa, corresponding to a 1- to 3-km
depth, is assumed based on the generally fine-grained nature of the garnet and pyroxene
alteration and common hydrothermal brecciation, both of which indicate relatively shallow
depths of formation. The high salinity of these fluid inclusions and the above pressure
correction would add 25C (Potter, 1977) to the T h to give an approximate trapping
temperature. Fluid salinity was determined from melting temperatures of halite daughter
minerals, using the equation of Sterner et al. (1988). Where sylvite melting temperatures
could be determined the ternary diagram of Roedder (1984) was employed to calculate
fluid salinity. earlier light green and white pyroxene partially overlap in T h range (Fig. 18),
with the green pyroxene extending to lower temperatures. Average T h for quartz in veins is
lower than for white pyroxene and garnet with the high-temperature initiation of
precipitation of quartz in these veins occurring at
about 420C (Fig. 18). Some of the quartz veins analyzed contained magnetite and
chalcopyrite, indicating the general temperature range of precipitation of these two
minerals. Quartz interstitial to garnet occupies the same paragenetic position as
magnetite and sulfides in skarn and has a T h range of 220 to 400C, similar to that of vein
quartz.
Discussion
Role of igneous rocks

83

The similarities between the REE patterns of diorite and monzonite porphyries
suggest that the diorite and monzonitic porphyry families are petrogenetically related,
with the monzonites and diorite potentially derived from the same differentiating parent
magma. Hornblende fractionation may have preferentially removed the middle to heavy
rare earth elements from the residual magma to produce the monzonite porphyry
magmas. The change in Eu anomalies in the diorite samples, from a large positive
anomaly in the most primitive diorite to a slight negative anomaly similar to that seen in
the monzonite porphyries, suggests the involvement of plagioclase
fractionation in the diorite sills, potentially at the site of intrusion. As the earliest intrusive
product, the diorite may have intruded into the upper crust under a convergent stress
regime, with the later monzonitic rocks intruding in an extensional regime caused by a
reorientation of principal stresses. This likely corresponds to a period of upper crustal
extension. The diorite may have acted as a heat source for localized limestone
recrystallization, but skarn is not generally zoned away from diorite contacts and the

diorite itself shows little anhydrous calc-silicate development that would be consistent with
a genetic relationship to skarn.

84

FIG. 17. Photomicrographs of typical fluid inclusions at Coroccohuayco. A. Irregularly shaped three-phase inclusions in
garnet interpreted as primary in origin; note relationship to crystal grow zones. B. Larger isolated primary inclusion in
garnet, with a trail of secondary inclusions. C. Trail of lower salinity inclusions in garnet interpreted as secondary
(parallel to white line). D. A trail of three-phase fluid inclusions in pyroxene, interpreted as pseudosecondary; note
relationship to grain boundary. E. Multi-phase fluid inclusions in pyroxene with two halides, hematite, and an
anisotropic daughter (not visible). F. Dispersed three-phase fluid inclusions in quartz with a general consistency in
vapor/fluid ratios.

It may, however, have influenced the formation of magnetite in parts of the deposit (see
below) as well as buffered the effect of any descending meteoric fluids on the deposit. The
petrographic and chemical similarities of rocks from Coroccohuayco and Tintaya suggest
that the intrusive rocks at both deposits were related to compositionally similar magmas
and the deposits represent upper level loci of intrusive activity associated with a larger
differentiating batholithic body under the district. One of the important metallogenic
features from Coroccohuayco is that monzonitic dikes F-G-H locally cut skarn but in other
areas form skarn. This feature suggests that skarn formation was a continuing but pulsed
process during the polyphase intrusive history of the deposit.
Skarn zonation
The geology of the skarn alteration and mine section modeling (section 1400; Fig. 4)
indicates that skarn-forming fluids infiltrated the carbonate host rock of the Ferrobamba
Formation by lateral movement along more permeable and/or reactive layers. The weak
calc-silicate marble alteration at Coroccohuayco may extend hundreds of meters laterally
85

away from the mineralized massive skarn (Fig. 4), although the vertical gradation is much
shorter. The expression of higher permeability is shown by the layered nature of the
recrystallization, bleaching, and weak calc-silicate development observed parallel to
bedding laminations (Maher, 1999; Fig. 7).

86

This layered alteration style is a smaller scale equivalent of the replacement of specific
carbonate beds by calc-silicate alteration in other skarn deposits (cf. Koski and Cook,
1982). At Coroccohuayco this layering is related to the bedding structure in the protolith,
although the large lateral extent of alteration may also be partly related to the reactive
infiltration instability process which tends to focus reactive fluids into specific channels
(Ortoleva et al., 1987; Meinert et al., 2005). In addition to the permeability variations
related to original sedimentary characteristics, increased permeability may have been
induced by subtle deposit-scale thermal recrystallization along certain layers and
laminations. The mass of skarn in the northern part of Coroccohuayco (as described
above) is volumetrically significant. However, it is not entirely clear what the actual source
of the fluid was which produced this as few synskarn porphyries are present in this area,
particularly around easting 16. It is hypothesized that at least some of the skarn in this
zone may have been produced by fluids derived from a deeper intrusive body. Fluid
originating from this intrusion could have ascended hypothesized faults until it reached the
reactive Ferrobamba Formation and spread laterally along the permeable bedding
conduits. Although there is no direct drilling evidence for the existence of this intrusion,
the common presence of pyrite veins with biotitic envelopes encountered in the Mara

87

FIG. 18. Histograms of fluid inclusion homogenization temperatures (Th) from Coroccohuayco. A. The combined fluid inclusion data of garnet, p

Formation in the deepest drilling in this area suggest the potential for a deeper fluid
source.
Skarn-forming conditions
Fluid inclusion data indicate that both garnet and pyroxene commenced precipitation from
highly saline fluids at temperatures exceeding 600C and ended around 400C, with
individual skarn samples showing the entire range of temperatures. Several observations
indicate that the early prograde skarn alteration formed under oxidized conditions,
including the highly andraditic composition of garnet and diopsidic composition of early
pyroxene, lack of a strongly colored distal pyroxene zone (Meinert et al., 2005), garnet
volumetrically greater than pyroxene, and the presence of hematite daughter minerals in
primary fluid inclusions in both garnet and pyroxene. These features indicate that the
primary control over the composition of alteration was the hydrothermal fluid since the
protolith is a relatively clean carbonate.
Wollastonite has not been identified at Coroccohuayco and its absence may be due
to several factors. The most important being that the Ferrobamba Formation in the deposit
is a relatively pure carbonate unit with little disseminated sedimentary silicate material or
88

chert present and, to a lesser extent, that metamorphic heating was not a particularly
important alteration process in the Ferrobamba Formation. The geochemistry and texture
of skarn also indicates that metasomatic reactions dominated over isochemical reactions
during genesis of the alteration in the deposit.
Mineralization
Copper mineralization commenced as temperatures decreased to the saturation of
chalcopyrite bornite in the fluid and eventually to the stability of Cu-rich phases, with
precipitation probably commencing around 400C. This is inferred from the fluid inclusion
data derived from vein and disseminated quartz which either occurs with, or occupies, a
paragenetic position similar to sulfides.
The mineralization at Coroccohuayco is dominated by disseminated chalcopyritebornite hosted by garnet and, to a lesser extent, pyroxene. The close spatial relationship
between the alteration and mineralization indicates that the ore forming process was
genetically related to the alteration process rather than being unrelated superimposed
events. In addition, skarn distal to known dikes and garnet veins in marble contains
disseminated sulfide mineralization, pointing toward a genetic link between the calcsilicates and copper mineralization. The actual initiation of sulfide precipitation in skarn
from an evolved prograde hydrothermal fluid may have occurred in the more distal green
granular garnet zone, since lower temperatures or chemical instability of the transporting
complexes would have been encountered here prior to the more proximal zones during a
thermal collapse of the prograde hydrothermal system. The abundance of remnant
interstitial calcite in distal garnet could facilitate metal precipitation from chloride
complexes by consuming acid in a decarbonation reaction like,

driven by acid-producing sulfide precipitation reactions such as,

thereby resulting in a feedback to the decarbonation reaction. With abundant calcite in the
green garnet skarn, this process will proceed until the calcite is exhausted, porosity is
filled, or the metallic elements in the fluid are spent.
Some sulfide ore textures probably represent exsolution from higher temperature solid
solutions and others from
modification of deposited sulfides by evolved hydrothermal fluids. The myrmekitic and/or
lamellar textures of bornitechalcopyrite and bornite-chalcocite may be due to the
exsolution of an originally homogeneous Cu-Fe sulfide phase during cooling. Other
complex sulfide textures are most likely related to fluctuations in fluid chemistry during
deposition. Gold and silver mineralization is mostly observed in bornite chalcocite
grains. This style of mineralization is most common in the granular green garnet host and
thus the higher grades of Cu mineralization in this host also correspond to higher precious
metal content. The common occurrence of electrum and silver-gold tellurides near the
rims of sulfide grains indicate that gold and silver were precipitated during
the later stages of copper mineralization at any specific location. There is no textural
evidence to suggest that precious metal mineralization is genetically distinct from the
copper mineralization. Ultimately, the gold and silver deposition was likely triggered by
89

instability in gold-bisulfide complexes due to changes in temperature, fO2, pH, or


consumption of S from bisulfide complexes to form S-rich copper phases during sulfide
mineralization.
Magnetite
A minor volume of magnetite is observed near skarn-marble contacts, suggesting that the
precipitation of magnetite was initiated by changes in fluid physicochemical conditions
resulting from reactions between carbonate protolith and the hydrothermal fluid. However,
the majority of magnetite replaces calc-silicate minerals. Retrograde magnetite alteration
of skarn likely occurred for several reasons although the processes may not have been
mutually exclusive. In many different zones in the deposit magnetite occurs in green
garnet and pyroxene skarn distal to fluid sources. This suggests a change leading to
magnetite deposition that consistently occurred
in certain zones with similar geochemical conditions throughout the deposit at both high
and low elevations. For
magnetite that replaces more distal skarn minerals, this change is likely related to
changes in temperature and oxidation state of the fluid during prograde alteration
(Johnson and Norton, 1985). For example, the oxidation state of the fluid will change as
ferric ion is consumed during early precipitation of andraditic garnet, during cooling, and
by fluid interaction with a potential reductant like elemental carbon in limestone. Increases
in PCO2 resulting from carbonate dissolution could lead to destabilization of garnet following
a reaction like:

Reaction (3) would require the reduction of some ferric ion but the associated calcite and
quartz (if precipitated) could effectively decrease rock permeability and seal the system to
further reaction (Seward and Barnes, 1997). A reaction involving the addition of ferrous ion
by the fluid and removal of silica and carbonate as dissolved species could likewise
produce massive magnetite in a reaction like:

A reaction like this is especially favored if the carbonate and silica reaction products are
removed, thereby resulting in nearly pure magnetite. Reaction (4) would initiate a
feedback process where further acid production could facilitate decarbonation reactions
and then continued magnetite precipitation if the evolved CO 2 could not quickly escape.
Thus magnetite would be expected as a retrograde alteration product as evolved (more
reduced) skarn-forming fluids decreased in temperature and P CO2 increased locally during
thermal collapse, with the volume of magnetite produced at any location controlled by the
availability of reduced iron in the fluid. The precipitation of Cu-Fe sulfides from Cu
complexes nearly contemporaneously with magnetite in the same zone indicates
thermodynamic instability of both the Fe- and Cu-transporting complexes and high
chlorine to sulfur ratios in the fluid (Barton and Johnson, 2000). Based on fluid inclusion
evidence in silicates, this likely occurred at 400(25)C. At Coroccohuayco, magnetite is
also related to a relatively reduced postprograde skarn hydrothermal fluid that overprinted
existing skarn adjacent to the postskarn monzonite porphyry M dikes. This unusual
90

alteration is locally associated with dark green late pyroxene alteration that overprints
earlier white pyroxene. Microprobe analyses of this green pyroxene correlate it to a fluid
more reduced and enriched in Na (Table 3; Fig. 15), consistent with a derivation from a
relatively reduced hydrothermal fluid with the potential to produce albitic alteration and
magnetite. The area of influence by these late fluids does not appear to exceed 50 m from
the dikes and magnetite generally tends to occur as disseminations and patches so this
process produced only a small volume of the magnetite formed at Coroccohuayco.
A significant volume of massive magnetite was produced near the skarn-diorite
contact in the upper parts of the deposit (Fig. 4). Although diorite is not genetically related
to skarn alteration and mineralization, it may have acted as a barrier to fluid escape,
buffering the composition and oxidation state of the skarn-forming fluids, and/or acting as
a seal to the escape of CO2 thereby permitting increases in PCO2 and replacement of
existing garnet following a reaction such as (3) above. The release of ferrous ion from
altered primary silicates in diorite could have also locally contributed to the formation of
magnetite. Likewise the sill could have acted as a barrier or buffer for downward
circulating oxygenated meteoric fluids that potentially could have produced hydrous
retrograde alteration common in other copper skarns systems.
Conclusions
Coroccohuayco differs from other Cu skarn systems in that it has only weak potassic
alteration in the intermediate igneous rocks and common albitic and generally weak late
sericitic alteration in porphyries. Part of this is likely related to the original geochemistry of
the monzonitic phases and the stage and/or temperature at which complete crystallization
occurred, but the abundant albitic alteration is probably a consequence of the
petrogenetic character of the parent magma. Polyphase intrusive activity was important in
forming the large extent of mineralization, with several early and late synskarn porphyries
concentrated in a small area. Mineralization in the causative intrusions is inconsistent and
generally low. Skarn-forming fluids migrated distances greater than 100 m away from fluid
sources at several stratigraphic levels because of the incipient protolith permeability and
favorable geometry of the host structure. Andraditic garnet alteration predominates in
skarn because of the composition of the relatively pure carbonate protolith and the
oxidized iron-rich character of the prograde skarn-forming fluid. Retrograde alteration in
skarn is dominated by magnetite carbonate silica with hydrous retrograde alteration
of garnet poorly developed. This is probably related to several factors including the
possibility of the absence of late-stage infiltrating meteoric fluids, the geochemical effect
on the evolution of hydrothermal and meteoric fluids by the large diorite sill overlying the
deposit, and the control by oxidized iron-rich metasomatism over resulting prograde skarn
composition. Copper mineralization is dominantly disseminated interstitial to calc-silicates
and highest where remnant carbonate or porosity was also higher. Magnetite alteration is
an important feature at Coroccohuayco, as it is in other deposits in the Andahuaylas-Yauri
belt, such as Tintaya and Las Bambas (Perell et al., 2003) because of its relationship to
calc-silicate alteration and mineralization.As such, magnetite is a valuable geophysical
exploration tool because it appears to be a common product of the retrograde alteration of
calc-silicates in oxidized iron-rich skarn environments as well as being introduced to skarn
by late Na-rich fluids derived from sodium-rich intrusive phases such as those observed at
Coroccohuayco.

91

References
Atkinson, W.W., Jr., and Einaudi, M.T., 1978, Skarn formation and mineralization in the contact
aureole at Carr Fork, Bingham, Utah: ECONOMIC GEOLOGY, v. 73, p. 13261365.
Barton, M.D., and Johnson, D.A., 2000, Alternative brine sources for Feoxide(- Cu-Au) systems:
Implications for hydrothermal alteration and metals, in Porter, T.M., ed., Hydrothermal iron oxide
copper-gold and related deposits: A global perspective: Australian Mineral Foundation, v. 1, p.
4360.
Benavides-Caceres, V., 1999, Orogenic evolution of the Peruvian Andes: the Andean cycle: Society
of Geologists Special Publication 7, p. 61107.
Bernstein, M., and Ly, P., 1993, Perus precious and base metals status and future prospects: Mining
Engineering, July, p 705709.
BHP Billiton Ltd., 2000, Tintaya oxide project engineering work progresses: Melbourne, Australia:
News Report, 30 Mar 2000, Ref. 27/00.

92

Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., Arenas F., M.J., France, L.J., McBride, S.L.,
Woodman, P.L., Wasteneys, H.A., Sandeman, H.A., and Archibald, D.A., 1990, Geologic and
geochronologic contraints on the metallogenic evolution of Andes of southeastern Peru: E CONOMIC
GEOLOGY, v. 85, p. 15201583.
De La Cruz B., N., 1995, Geologia de los cuadrangulos de Velille, Yauri, Ayaviri y Azangaro: Lima
(Peru), Carta Geologica Nacional, Instituto Geolgico Minero y Metalurgico Boletin 58, Serie A, 144
p. Espirilla R., C.R., 2004, Controles estructurales sobre emplazamiento de intrusivos, y
mineralizacin en Tintaya, Per: Unpublished Titulo Profesional de Ingeniera Geolgica, Arequipa,
Per, Universidad Nacinal San Agustin, 68 p.
Fernandez B., J.C., 2002, Geoqumica y petrografia de las rocas igneas del yacimiento de Tintaya:
Unpublished Titulo Professional de Ingeniera Geolgica, Arequipa, Per, Universidad Nacional San
Agustin, 105 p.
Gustafson, L.B. and Hunt, J.P., 1975, The porphyry copper deposit at El Salvador, Chile: E CONOMIC
GEOLOGY, v. 70, p. 857912.
Irvine, T.N., and Baragar, W.R.A., 1971, A guide to the chemical classification of the common
volcanic rocks: Canadian Journal of Earth Sciences, v. 8, p. 523548.
Jaillard, E., Herail, J., Monfret, T., Diaz-Martinez, E., Baby, P., Lavenu, A., and Dumont, J.F., 2000,
Tectonic evolution of the Andes of Ecuador, Peru, Bolivia, and northernmost Chile: International
Geological Congress, 31st, Rio de Janeiro, Brazil, August 1617, Proceedings, p. 481559.
Johnson, J.W., and Norton, D., 1985, Theoretical prediction of hydrothermal conditions and chemical
equilibria during skarn formation in porphyry copper systems: E CONOMIC GEOLOGY, v. 80, p.
17971823.
Koski, R.A., and Cook, D.S., 1982, Geology of the Christmas porphyry copper deposit, Gila County,
Arizona, in Titley, S.R., ed., Advances in geology of the porphyry copper deposits, southwestern
North America: Tucson, University of Arizona, p. 353374.
Lang, J.R., and Titley, S.R., 1998, Isotopic and geochemical characteristics of Laramide magmatic
systems in Arizona and implications for the genesis of porphyry copper deposits: E CONOMIC GEOLOGY,
v. 93, p. 138170.
Lanier, G., Raab, W.J., Folsom, R.B., and Cone, S., 1978, Alteration of equigranular monzonite,
Bingham mining district, Utah: ECONOMIC GEOLOGY, v.73, p. 12701286.
Maher, K.C., 1999, Geology of the Cu-skarn at Coroccohuayco, Peru: Unpublished M.Sc. thesis,
Pullman, Washington State University, 133 p.
Maher K.C., and Larson, P.B., 2007, Variation in copper isotope ratios and controls on fractionation in
hypogene skarn mineralization at Coroccohuayco and Tintaya, Per: E CONOMIC GEOLOGY, v. 102, p.
225237.
McDonough, W.F., and Sun, S.S., 1995, The composition of the earth: Chemical Geology, v. 120, p.
223253.
Meinert, L.D., 1995, Compositional variation of igneous rocks associated with skarn deposits
chemical evidence for a genetic connection between petrogenesis and mineralization: Mineralogical
Association of Canada Short Course Series, v. 23, p. 401418.
Meinert, L.D., Dipple, G.M., and Nicolescu, S., 2005, World skarn deposits: E CONOMIC GEOLOGY 100TH
ANNIVERSARY VOLUME, p. 299336.
Noble, D.C., McKee, E.H., Eyzaguirre, V. R., and Marocco, R., 1984, Age and regional tectonic and
metallogenetic implications of igneous activity and mineralization in the Andahuaylas-Yauri belt of
southern Peru: ECONOMIC GEOLOGY, v. 79, p. 172176.
Ortoleva, P., Chadam, J., Merino, E., and Sen, A., 1987, Geochemical self-organization. II: The
reactive-infiltration instability: American Journal of Science, v. 287, p. 10081040.
Perell, J., Carlotto, V., Zrate, 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, Per: ECONOMIC GEOLOGY, v. 98, p. 15751605.
Potter, R.W., 1977, Pressure corrections for fluid-inclusion homogenization temperatures based on
the volumetric properties of the system NaCl-H 2O: U. S. Geological Survey Journal of Research, v. 5,
p. 603607. Richards, J.P., Boyce, A.J., and Pringle, M.S., 2001, Geologic evolution of the Escondida
area, northern Chile: A model for spatial and temporal localization
of porphyry Cu mineralization: ECONOMIC GEOLOGY, v. 96, p. 271305.

93

Roedder, E., 1984, Fluid inclusions: Reviews in Mineralogy, v. 12, 644 p. Seward, T.M., and Barnes,
H.L., 1997, Metal transport by hydrothermal fluids, in Barnes, H.L., ed., Geochemistry of
hydrothermal ore deposits, 3rd ed.: New York, John Wiley and Sons, p. 435486.
Singer, D.A., 1986, Grade and tonnage model of porphyry Cu skarn-related deposits: U.S. Geological
Survey Bulletin 1693, p. 7781.
Sterner, S.M., and Bodnar, R.J., 1984, Synthetic fluid inclusions in natural quartz: 1. Compositional
types synthesized and applications to experimental geochemistry: Geochimica et Cosmochimica
Acta, v. 48, p. 26592668.
Sterner, S.M., Hall, D.L., and Bodnar, R.J., 1988, Synthetic fluid inclusions: V. Solubility relations in
the system NaCl-KCl-H2O under vapor-saturated conditions: Geochimica et Cosmochimica Acta, v.
52, p. 9891005.
Zweng, P.L., Yagua P., J., Fierro R., J., Gamarra R., H., Jordan G., L., Brooks, J., Yurko, E., and
Mulhollen, R., 1997, The Cu-(Au,Ag) skarn deposits at Tintaya, Peru: Socieded Geologica del Peru
Special Volume 1, p. 237242.

94

Projects and Mining


Prospects

95

You might also like