You are on page 1of 30

2004 by Economic Geology

Vol. 99, pp. 887916

The Lithologic, Stratigraphic, and Structural Setting of the


Giant Antamina Copper-Zinc Skarn Deposit, Ancash, Peru
DAVID A. LOVE, ALAN H. CLARK,
Department of Geological Sciences and Geological Engineering, Queens University, Kingston, Ontario, Canada K7L 3N6

AND J. KEITH GLOVER*


Glover Consulting Ltd., 146 Simcoe St., Victoria, B.C., Canada V8V 1K4

Abstract
Antamina, located at latitude 9 32' S and longitude 77 03' W in the Ancash Department of north-central
Peru, is the largest known Cu-Zn skarn ore deposit. It incorporates a mineral reserve of 561 Mt, which has an
average grade of 1.24 percent Cu, 1.03 percent Zn, 13.71 g/t Ag and 0.029 percent Mo, calculated at a 0.7 per-
cent Cu equiv cutoff grade. The grandite-dominated calcic skarn formed in and around an upper Miocene por-
phyritic monzogranite stock emplaced into Upper Cretaceous carbonate strata that had experienced thin-
skinned, northeast-verging thrusting and folding in the late Eocene Incaic orogeny. The exoskarn Cu-Zn ore is
discordant to the strata of the Jumasha and overlying Celendn Formations, which comprise, respectively, mas-
sive to thick-bedded, relatively pure limestones and thin-bedded, predominantly marly limestones. The Ju-
masha Formation, the upper contact of which is locally defined as the top of the uppermost thick-bedded lime-
stone or marble unit, hosts approximately three-quarters of the known exoskarn. Approximately the same
fraction of the contiguous endoskarn Cu ore occurs adjacent to this formation. The overlying Celendn For-
mation is less extensively mineralized but, because it is widely metamorphosed to hornfels and locally con-
verted to diopsidic skarnoid, may have inhibited the upward and outward migration of hydrothermal fluids,
thereby promoting the development of the unusually large endoskarn ore zone. Ore also occurs in late hy-
drothermal breccias emplaced during the formation of mineralized endoskarn.
The preskarn thermal metamorphic aureole around the ore deposit is expressed differently in the two host
formations. Jumasha Formation limestone is coarsened and bleached to banded gray marble and locally to
white marble peripheral to the intrusion and skarn. Minor scapolite occurs in dark gray bands in marble, con-
centrated in a discontinuous halo tens of meters wide and commonly separated from the skarn by tens of me-
ters. Three facies of calc-hornfels are recognized in the marl beds of the Celendn Formation adjacent to the
intrusion extending hundreds of meters beyond sulfide-bearing skarn: a peripheral, very fine grained, light
brown phlogopitic facies; an intermediate, fine-grained, gray tremolitic facies; and a proximal, medium-grained,
light green diopsidic facies. At an XCO2 of 0.1 to 0.9 and P = 100 MPa, these zones reflect temperatures in-
creasing to circa 495C adjacent to the intrusion. In addition, in nodular beds of the Celendn Formation that
have been metamorphosed to hornfels, diagenetic calcite nodules are selectively replaced by diopside for dis-
tances of tens of meters beyond the skarn front. Such calc-silicate formation through both metamorphism and
metasomatism, together with a 9 km2 cluster of Pb-Zn-Ag vein deposits, provides district-scale vectors to ore.
The Antamina deposit lies on a newly recognized cross-strike structural discontinuity in the segmented In-
caic Maran thrust and fold belt, the northeast-trending Querococha arch. Southeast of the arch, Incaic folds
and thrust faults strike north-northwest, but northwest of the arch they strike northerly. The plunge of fold axes
concomitantly changes from south-southeast to north. Stratigraphic relationships indicate that the arch was a
paleohigh, at least in the Jurassic and possibly throughout the late Paleozoic-early Mesozoic interval. The mid-
dle Miocene Carhuish pluton is exposed on the arch 30 km southwest of Antamina, whereas coeval Calipuy Su-
pergroup volcanic units lie at similar altitudes to the north and south. Only scattered hydrothermal centers of
late Miocene age are known in the Cordillera Negra, but an apparent swarm of intrusions, including the Anta-
mina stock, occurs along the Querococha arch.
Antamina is situated where the locus of changes in the strike of folds and faults and the plunge of folds steps
left along the arch. At Antamina, a pair of fault-bend folds above frontal thrust ramps show approximately 500
m of dextral apparent offset across the deposit and are inferred to have been separated by a northeast-strik-
ing transfer fault or lateral ramp, itself localized by a left-stepping jog in the Valley fault, an underlying, sim-
ilarly oriented transverse structure. The jog in the Valley fault is inferred to have also controlled intrusion and
skarn development. This local-scale jog in the Valley fault mimics the regional step along the arch. The arch
may reflect a transform segment of the originally jagged, rifted continental margin, which persisted as a trans-
verse basement weakness. Northeast-striking, originally sinistral, basement structures affected regional-scale
sedimentation and structural patterns, including articulation of the thrust and fold belt. At a local scale, they
Corresponding author: e-mail, love@geol.queensu.ca
*In Memoriam. On July 17, 2001, Keith Glover died unexpectedly but peacefully in his sleep, immediately after returning from fieldwork. He had an in-
fectious passion for rocks and great patience in teaching about them. He was a skilled structural geologist and a professional ore deposit specialist with acute
perception and a love for walking the rocks. Keith had been consulting internationally for 14 years and was respected for his intellect and breadth of geolog-
ical understanding. He also managed admirably to balance his love for geology with that for his family. Keith was well liked for his generous spirit and hu-
manity, and he is greatly missed.

0361-0128/01/3439/887-30 $6.00 887


888 LOVE ET AL.

influenced lateral ramp formation and related fracture development in the overlying thrust sheets. In the pro-
posed model, they also localized later uplift and the rapid transit of small volumes of productive melt into a
shallow crustal setting, conditions favorable for formation of a giant magmatic-hydrothermal ore deposit.

Introduction from latitudes 2 S to 15 S they are now underlain by a flat


DESPITE their potential ore genetic and metallogenic impor- subduction zone widely ascribed to underthrusting of the
tance, the lithologic, stratigraphic, and structural settings of Nazca Ridge (Barazangi and Isacks, 1976; Pilger, 1981; Ham-
skarn mineralization have rarely been comprehensively docu- pel, 2002) and, in the northern part, the postulated Inca
mented. It is therefore difficult to assess their influence on Plateau (Gutscher et al., 1999). This major flat-slab domain
the localization of skarn-generating hydrothermal systems separates the Northern and Central Volcanic zones of the
and, in particular, to envisage the specific environments in Andes and has been apparently amagmatic since emplace-
which exceptional deposits have developed. In this paper, we ment of the last phase of the Cordillera Blanca batholith at
describe the host rocks and structural relationships of the An- 6.3 to 8.2 Ma (Mukasa, 1984; McNulty et al., 1998). Intrusion
tamina Cu-Zn(-Ag-Mo) deposit, north-central Peru, and pro- and mineralization at Antamina took place shortly before this
pose a model for the stratigraphic and tectonic environment terminal magmatism. Although the carbonate rocks that host
in which this largest known Cu-Zn skarn orebody formed. the deposit have long been recognized as Upper Cretaceous,
These aspects are controversial, in part because of the poorly they have been assigned to various formations. The strati-
defined local stratigraphic succession and because of the de- graphic relationships in the mine area are herein clarified
formation, metamorphism, and metasomatism imposed on through examination of the carbonate rocks around the skarn
the ore-hosting strata. Following a brief summary of the geol- and their comparison with well-described measured sections
ogy of the deposit, we document the regional-scale (ca. 5,000 elsewhere in north-central Peru. This analysis permits both
km2) geologic and geodynamic setting of the mineralization elucidation of the structure of the area and characterization of
before focusing on the district scale (ca. 120 km2). the types of rocks replaced by the skarn.
Copper mineralization was known at Antamina (anta: cop- The Antamina deposit shares numerous common features
per in Quechua) in pre-Colonial times, but only modest with other large porphyry-related Cu skarns (Einaudi, 1982a,
amounts of Pb and Ag are known to have been produced in b), but it differs from most in the exceptional development of
the district prior to 2001 (Redwood, 1999). The skarn con- mineralized endoskarn and the association of ore-grade Cu,
tains proven and probable reserves of 561 Mt with an average Zn, and Mo in contiguous zones. An additional unusual fea-
grade of 1.24 percent Cu, 1.03 percent Zn, 13.71 g/t Ag, and ture is the widespread development of chalcopyrite-rich hy-
0.029 percent Mo (calculated at a 0.7% Cu equiv cutoff drothermal breccias, the extent and above-average Cu con-
grade). Compaa Minera Antamina S.A., which operates the tent of which were unrecognized prior to our research. Many
Antamina open-pit mine, is owned by BHP Billiton (33.75%), of the observations on the deposit- and district-scale geology
Noranda (33.75%), Teck Cominco (22.5%), and Mitsubishi recorded herein, and their interpretation, were introduced in
(10%). Production of copper-silver and zinc concentrates, as unpublished reports prepared by the authors (D.A. Love and
well as lead, molybdenum, and bismuth byproducts, began in A.H. Clark, 1998a, b, 2000, unpublished reports to Compana
July 2001 (Zuzunaga, 2003). Minera Antamina S.A., Lima; J.K. Glover, 1998a, 1998b,
The Antamina mine is located at approximately 9 32' S and 1998c, unpublished reports to Compana Minera Antamina
77 03' W, 270 km north of Lima and 130 km from the Pacific S.A., Lima).
coast, in Ancash Department in north-central Peru (Fig. 1). It
lies in the eastern part of the Cordillera Occidental, east of The Antamina deposit
the Cordillera Blanca and west of the Ro Maran valley. The Antamina deposit comprises endoskarn and exoskarn,
The skarn is exposed between approximately 4,200 and 4,800 with subordinate breccia bodies that cut both skarn and in-
m a.s.l., at the head of a southwest-draining glacial valley, but trusion within the perimeter of skarn. The mineralized skarn
prior to mining much of the orebody was covered by Lago is dominated by grandite garnet, which grades from brown to
Antamina, a glacial tarn (Fig. 2). The history of exploration at green nearer the host limestone (Petersen, 1965). Thus the
Antamina and the general geology of the deposit are summa- deposit conforms to the oxidized calcic clan of Einaudi et al.
rized by Redwood (1998, 1999). OConnor (2000) reviewed (1981). The geology of the deposit and its immediate sur-
the geologic and geophysical approaches that delineated the roundings is summarized in Figure 3a on three northwest-
ore and outlined the development of the mine, metallurgical southeast drill-hole cross sections, spaced 50 m apart,
testing, and resource calculations. Both authors generally through the middle of the orebody. The simplified surface ge-
supported previous geologic descriptions and interpretations ology map in Figure 3b was constructed by projecting the
(e.g., Petersen, 1965). lithologic boundaries defined on these and thirty other cross
The Antamina deposit formed at 9.86 to 10.18 Ma sections. The area delimited by the mineralized skarn front is
(40Ar/39Ar step-heating data of Love et al., 2003) around a approximately 1.18 km2 overall. The skarn zone straddles the
small monzogranitic porphyry intrusion. It is hosted by Upper original intrusive contact and surrounds a circa 0.24 km2 core
Cretaceous carbonate strata within the Maran thrust and of porphyry that forms a crudely parallelogram-shaped prism
fold belt, formed by the late Eocene Incaic orogeny (Fig. 1; with a vertical axis (Fig. 3b). The outer boundary of the min-
Noble et al., 1979; Mgard, 1984). The western Andes of eralized skarn is more elongated northeast-southwest than
Peru were the site of episodic arc magmatism from the Late this core because it expands around faults and dikes extending
Triassic to the late Miocene (Cobbing et al., 1981). However, to the east, northeast, and southwest. It is therefore roughly

0361-0128/98/000/000-00 $6.00 888


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 889

81W
MTFB 0
operating mine

past producer

prospect
Peru
8S Pasto Bueno
Magistral La
Libertad
Trujillo
Ancash
PACIFIC
LIMA
OCEAN

0 25 50 100 km
18S

CO
69W

RD
ILL
ER
79 W
Fig. 6

A
Pliocene - Quaternary clastic sediments 9S
& glaciers

BL
Chimbote CO Llamelln
Upper Miocene - Pliocene volcanic rocks

AN
(Yungay & Fortaleza Fms.)
RD

CA
ILL
Upper Miocene granitoid intrusions
Pierina
ER

Lower - Middle Miocene granitoid intrusions Fig. 5 Antamina


Antamina
A

Middle Eocene - Middle Miocenevolcanic rocks


NE

(Calipuy Supergroup)
Casma
GR

Upper Cretaceous - Paleocene red beds (Casapalca Gp.)


A

Huaraz
Cretaceous - Paleogene intrusions
Ta

Ro
Ma
pa

Albian - Upper Cretaceous carbonates (Machay Group and


co

ra
equivalents)
ch

n
a A

Cretaceous volcanic rocks


R
xis

oS
Upper Jurassic - Lower Cretaceous (Berriasian - Aptian) an
siliciclastic rocks (Chicama Gp. & Goyllarisquisga Gp.) 10S ta

Mississippian - Lower Jurassic (including Ambo, Mitu & Huarmey


Pucar Groups)

pre-Ordovician (Maran metamorphic complex)


78 W 77 W MTFB
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) men-
tioned in the text are shown, as are the areas illustrated in Figs. 5 and 6. MTFB = Maran thrust and fold belt.

elliptical in plan and has a northwest-southeast width of up to exoskarn comprising the copper-zinc ore. Molybdenite is dis-
1,000 m, and a northeast-southwest length of more than 2,500 seminated in irregular zones within and at the margin of the
m. The long axis parallels the Antamina valley and is perpen- intrusion. Chalcopyrite is the dominant copper mineral ex-
dicular to the regional structural grain of the deformed car- cept at shallow depths in the southwestern part of the deposit,
bonate host rocks. As recognized by Petersen (1965) the where bornite predominates in wollastonitic exoskarn, which
outer limit of skarn is generally subvertical. The skarn nar- forms an enclave in green garnet Cu-Zn exoskarn.
rows with depth as the core of porphyry widens (Fig. 3a), but Hydrothermal breccia is common at or near the endoskarn-
there is no significant change in the Cu and Zn grades to exoskarn contacts along the northwest and southeast sides of
depths of at least 400 m below the original valley floor. Except the deposit. Breccia also cuts the porphyry core as anasto-
at high elevations in the eastern part of the deposit, the ex- mosing sheets and pipes that are commonly enveloped by
oskarn is almost everywhere mineralized. fine-grained maroon garnet endoskarn (Fig. 3). The breccia
In detail, the individual skarn facies and breccia bodies are zones contain minor chlorite and were originally described by
complexly shaped and discontinuous, but the deposit can be Petersen (1965) as chlorite skarn. The breccias are generally
simplified as comprising an inner shell of endoskarn, stock- poorly sorted and comprise angular to rounded fragments of
work, breccia, and brown garnet exoskarn that contains the the skarn and metallic minerals supported in a sand-sized
copper molybdenum ore and an outer shell of green garnet matrix of similar composition. As in skarn, sphalerite and

0361-0128/98/000/000-00 $6.00 889


890 LOVE ET AL.

0 western South America (Sempere et al., 2002). This Late

77 05W
460
Jurassic to Paleogene subsidence formed the West Peruvian

00
trough, which separated the magmatic arcs to the west from

42

00
40
the eastern geanticline now represented by the Maran
9 30 S
metamorphic complex (Benavides, 1956; Wilson, 1963;
Atherton et al., 1983; Mgard, 1987). The Cretaceous sedi-
mentary rocks that crop out in the Antamina area accumu-
lated in the shallow-water portion of this trough, the Yauli
4400

shelf (Szekely, 1967). The Tapacocha axis, now a north-north-


42

46

4400
00

00

westtrending high strain zone, separates the western,


deeper-water portion of the trough from the shelf sedimen-

4000
San Marcos 42
00
<== 5 km
tary rocks (Fig. 1; Myers, 1974, 1975).

46
00 Stratigraphic relationships
44
00 The host Machay Group (Figs. 1 and 4) includes all Albian
to mid-Campanian carbonate rocks south of 9 S and east of
00 the Tapacocha axis in north-central Peru. Szekely (1967) in-
4200

44
K EY troduced the Machay Group in central Peru and Samam-
4400

Lake Boggio (1980) applied it throughout Peru. These rocks also


0 1 2 3 4 P hoto viewpoint have been referred to informally as the middle Cretaceous
42 46
abandoned mine
working
limestone series (Harrison, 1940), the upper Cretaceous
00 00 Y anacancha
kilometers and Albian carbonate series (Mgard, 1984), and the upper
FIG. 2. Premine physiography of the Antamina area, with place names re- carbonate sequence (Manrique, 1998). The group is under-
ferred to in the text, illustrating the clustering of Ag-bearing Pb-Zn vein de- lain by a predominantly siliciclastic sequence comprising
posits around Lago Antamina documented by Bodenlos and Ericksen (1955). Upper Jurassic marine black shales of the Chicama Group
Also indicated are the Contonga Pb-Zn-Cu-Ag mine 5 km to the north-north-
west of Lago Antamina and veins about 500 m northeast of Contonga. B =
and Lower Cretaceous (Berriasian to Aptian) continental to
Barrn, C = Casualidad, Cc = Condorcoccha, F = Fortuna, JE = Julia Eloisa, shelf sandstones, shales, and minor limestones of the Goyllar-
P = Poderosa, Pp = Putapuquio, R = Recompensa, RdO = Rosita de Oro, SF isquisga Group (Fig. 4). It is widely overlain, conformably or
= San Francisco, SR = Santa Rosa, UP = Usu Pallares. The viewpoints for slightly unconformably, by red beds (Wilson, 1963), which are
photographs in Figures 7, 8, 12, and 15 are shown as eyes. (Contour inter- mainly Campanian to Paleocene but as old as Santonian in
val is 200 m.)
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
molybdenite do not commonly occur together in the breccias; km southeast of Antamina at La Unin (Wilson, 1963, after
sphalerite is found in breccias in or near exoskarn, but molyb- McLaughlin, 1924), the Chota Formation, 20 km north of An-
denite occurs in breccias in endoskarn. The common metallic tamina (Benavides, 1956: after Broggi, 1942), and the Cas-
minerals in the breccias are pyrite, chalcopyrite, and mag- apalca Group, 25 km southwest of Antamina in the Cordillera
netite, which are mostly comminuted but also occur as veins, Huayhuash (Coney, 1971, after McLaughlin, 1924).
massive bodies, and large fragments. This mineral assemblage The Machay Group contains two transgressive sequences
is rare in exoskarn but forms widespread stockworks and separated by a disconformity ascribed to late-middle Albian
sheeted vein swarms in fine-grained maroon garnet en- uplift and erosion related to the Mochica orogeny (Mgard,
doskarn, in many places grading into crackle, mosaic, and ma- 1984). In the lower part of the group, the successive Pari-
trix-dominated breccia. We estimate that approximately one- ahuanca, Chulec, and Pariatambo Formations (Fig. 4) record
third of the ore at Antamina may have formed during this late a transition from near-shore, calcareous sandstone and mas-
brecciation, veining, and endoskarn-forming stage. sive, shelly limestone, through thin-bedded limestone and
Widely spaced, late calcite-tetrahedrite sphalerite marl, to deep-water, thin-bedded, bituminous, dark gray marl
galena veinlets are common and cut all skarn and breccia and limestone (Benavides, 1956, 1999; Wilson, 1963; Jaillard,
types, although they are also locally dismembered in breccia, 1987). Following the late-middle Albian hiatus, carbonate
probably because of settling. Scarce realgar veinlets occur in sedimentation on the platform resumed with the deposition
calc-hornfels above and peripheral to the skarn. of the shallow-water, upper Albian to upper Turonian Ju-
masha Formation (Jaillard, 1987), originally defined by
Regional Geologic Setting McLaughlin (1924) in central Peru. This formation is overlain
The Upper Cretaceous strata enclosing the Antamina de- by the muddier, deeper-water Celendn Formation (Bena-
posit are part of a metallogenically important Albian and vides, 1956), largely Coniacian to Santonian in age (Jaillard,
Upper Cretaceous package of carbonate rocks, the Machay 1987) but attaining the mid-Campanian in northern Peru
Group, that hosts many ore deposits in the polymetallic skarn (Mourier et al., 1988). The upper part of the group, compris-
and carbonate-replacement belt of central Peru (Soler et al., ing the Jumasha and Celendn Formations (Fig. 4), thus rep-
1986). These rocks formed during the later of two Permian to resents a second major transgressive sequence. The lower to
Paleocene episodes of basin development that deposited a middle Albian carbonate strata (the first transgressive sequence)
succession of alternating siliciclastic and carbonate facies in are similar in lithology and thickness in both northern and

0361-0128/98/000/000-00 $6.00 890


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 891

Se
a cti
on
20
200 m ,55
0N
Scale

100
0
Se
10 cti
0 on
0 20
20 ,50
0m 0N

Se
cti
on
20
,45
0N

LEGEND

Limestone & Marble, Ce: Celendn Fm., J: Jumasha Fm.


Wollastonite Exoskarn
Green, Green & Brown, and Brown Garnet Exoskarn
Pink Garnet Endoskarn and Brown Garnet Endoskarn
Hydrothermal Breccia
Porphyritic Monzogranite

00
b N
44 43
50

a VF Lago Antamina, before mining


in
am
nt
oA
Anticline Syncline
Lag
20

Cross-sections shown in Fig. 3a


55
0
N
44
00

455
445

4600
0

4500

AA
43
50

4450
465

50
46
0

4400

00
44

4350 46
0 5

Ce J 00
47
44

50
00

45

50
42 R
VL
46

50
47
00

50
46

00
46
20
55
20

0
50

N
20
43

VF
00

0
45

N
0
N

50
45 Scale
00
47
0 100 200 m
44

J
50

Ce
455

46

50
50
0

AA
46

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 map-
ping by D.A.L. and J.K.G., combined with that by L. Hathaway (Inmet) and M. Wunder (Noranda). Prominent northwest-
striking Incaic folds and the left-stepping, transverse Valley fault (VF) are indicated. The location of the proposed Valley lat-
eral ramp (VLR), inferred to be responsible for the apparent dextral offset of the Antamina anticline (AA), is also shown.

0361-0128/98/000/000-00 $6.00 891


892 LOVE ET AL.

Antamina Mine Area Uchupata Type-Sections


Campanian
Upper Sequence Celendn Formation (163 m, from Section 19) - ? - - - ? - - - ? -
Thin-bedded shaly limestone: dark gray to black, weathers Marl: light gray, nodular, soft, white weathering Santonian
medium to light gray, thin- to thick-bedded, generally very fine Shale: calcareous, slightly silty, yellowish, with
grained, dominated by biomicrite to microsparite, ranges in a few interbeds of dark-brown limestone
carbonate content from muddy calcisiltite with 50 - 75 % Marl: light gray, nodular, white weathering, with
calcite to calcareous siltstone with < 50 % calcite, variably sparse interbeds of limestone
nodular, 10-90% nodules 1-15 cm diameter, nodules are Marl: light gray, nodular, soft, white weathering -?---?---?-
rounded to multilobate and have sharp to indistinct contacts Marl: light gray to tan, nodular, with few interbeds Coniacian
with the matrix of massive light-gray limestone
Lower Sequence Jumasha Formation (820 m, from Section 20)
Impure limestone interbedded near the top:
rare, medium to thick beds (20-50 cm), very dark gray to Turonian
black, silty limestone, weathers medium gray Limestone: medium gray, thick bedded,
Predominantly thick-bedded, relatively pure limestone: S weathering dark dove gray,
dark gray, medium- to very thick-bedded, ranges in grain size K Foraminifera-bearing
from mudstone to wackestone, weathers light to pale gray A 0
Locally fossiliferous bioclastic wackestone:
with broken pelecypods and gastropods, but no apparent
R
diagnostic fauna N
100
Outline of Regional Stratigraphy -?---?---?-

M Eocene - M Miocene Calipuy Supergroup


200 m
Campanian - Paleocene Casapalca Group
Limestone: argillaceous Cenomanian
Albian - U Cretaceous Machay Group Limestone: medium gray, massive, thick
Coniacian - Santonian Celendn Formation bedded, weathering dark-brownish gray
U Albian - Turonian Jumasha Formation Dolostone: thin bedded, brown
M Albian Pariatambo Formation
L - M Albian Chulec Formation
L Albian Pariahuanca Formation Dolostone: light gray to orange-brown,
massive, thick-bedded, karstic, weathering
L Cretaceous Goyllarisquisga Group dark orange-brown
(Berriasian - Aptian)
U Jurassic Chicama Group -?---?---?-

U Triassic - L Jurassic Pucar Group


Upper Albian
L Permian - L Triassic Mitu Group
Dolostone: silty, medium gray, somewhat nodular
Mississippian Ambo Group

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 km north
of Antamina. The contact between the Jumasha and Celendn Formations coincides with the boundary between the Turon-
ian 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.

central Peru (Benavides, 1956; Jaillard, 1987). However, Love et al. (2001) termed the structure the Querococha arch
north of approximately 9 S, the overlying upper Albian to because its southwestern limit at the margin of the Callejon
mid-Campanian carbonate rocks are much thicker, more fos- de Huaylas lies close to Laguna Querococha (Figs. 5 and 6).
siliferous, and lithologically more variable than in central Its influence on the abundance of Neogene intrusions, the re-
Peru, and the Jumasha Formation interval is divided into five gional strikes in the Maran thrust and fold belt, and the
formations (Benavides, 1956). Jaillard (1987) provides corre- stratigraphic relationships of underlying Mississippian to
lations between these stratigraphic sections in northern and Lower Jurassic strata (Fig. 4) is described below.
central Peru. The overall strike of the Maran thrust and fold belt
changes, and the common plunge directions reverse, across
Tectonic relationships the proposed northeast-trending cross-strike structural dis-
Published descriptions of the structural setting of the Anta- continuity (Fig. 5). These thin-skinned Eocene structures, at-
mina deposit (Bodenlos and Ericksen, 1955; Terrones, 1958; tributed to the Incaic orogeny, are the dominant tectonic ele-
Petersen, 1965; Redwood, 1999) have focused on the ments in the region, although two regional unconformities in
Maran thrust and fold belt, which developed circa 30 m.y. the Cretaceous succession in central Peru, one at the base of
before mineralization, with scant consideration of the re- the Jumasha Formation and the other below the Casapalca
gional tectonic environment that existed in the mid-Miocene. Group and its equivalents, represent earlier emergence dur-
We argue, however, that Antamina lies athwart a large-scale, ing, respectively, the Mochica and Peruvian tectonic phases
cross-strike (northeast-southwest) structural discontinuity (Noble et al., 1979; Mgard, 1984). The Maran thrust and
(Wheeler, 1978) that was tectonically active at the time of fold belt extends from 5 S to 12 30' S, generally striking
skarn formation and hence has metallogenic significance. north-northwestsouth-southeast, parallel to the present

0361-0128/98/000/000-00 $6.00 892


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 893

a
chc
Pu
o
R

Nevado ANTAMINA
930'S

osna
Huantsn

Ro M
Cordillera

Blanca

Pampa
940'S
Junin
Carhuish

Laguna
Querococha Pluton
Ca
lle
jn
de
Hu

Cordillera
ay

Huayhuash
las

7720'W 7710'W 7700'W


0 4 8 12 16 20 km
Pliocene - Quaternary clastic sediments
Middle Eocene - Upper Miocene Calipuy-equivalent and younger granitoid intrusions
Middle Eocene - Middle MioceneCalipuy Supergroup volcanic rocks
Upper Cretaceous - Paleocene Casapalca Group & equivalent red beds
Albian - Upper Cretaceous Machay Group carbonate rocks
Upper Jurassic - Lower Cretaceous (Berriasian - Aptian) Chicama & Goyllarisquisga Groups siliciclastic rocks
pre-Ordovician Maran Metamorphic Complex
anticline
syncline
9S
plunge direction
19i 19j
thrust fault
normal fault
strike-slip fault

20i 20j approximate locus of change of plunge


10S
7730'W 7630'W of folds and strike of folds and faults

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-south-
westtrending loci of these changes in structural attitude is indicated by the heavy dashed line, which is offset in a left-step-
ping sense in the vicinity of Antamina. This area is the location of the proposed cross-strike structural discontinuity discussed
in the text. Inset shows the location of the map area relative to the Huari (19-i), 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 as-
signed to the Calipuy Supergroup (see text for discussion).

0361-0128/98/000/000-00 $6.00 893


894 LOVE ET AL.

77 W

N
900' S

Llamelln

CS
D
Nevado
Huantsan
ANTAMINA
930' S

Pampa
Junin

Laguna
Querococha

Carhuish
pluton
Cordillera
Huayhuash

Structures
faults
fold axes
Rock types
0 20 km
Middle Eocene - Middle Miocene granitoid rocks
Middle Eocene - Middle Miocene Calipuy Supergroup subaerial volcanic &
associated rocks
Upper Jurassic - Cretaceous - Paleocene Chicama & Goyllarisquisga Gps. silici-
clastic rocks, Machay Gp. carbonate rocks, & Casapalca Gp. red beds
Mississippian - Lower Jurassic, including Ambo, Mitu & Pucar Groups

pre-Ordovician Maran Metamorphic Complex

FIG. 6. Simplified geology of approximately 8,000 km2 of the Maran thrust and fold belt in the region around Antam-
ina; 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 disconti-
nuity 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 disconti-
nuity but are present north and southeast of the cross-strike structural discontinuity, except where cut out by faulting. An un-
usual 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).

0361-0128/98/000/000-00 $6.00 894


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 895

plate boundary, and comprises structures that predomi- Northeast of Antamina, Mississippian to Lower Jurassic
nantly verge northeast (Mgard, 1984; Fig. 5). However, strata (Fig. 4) are absent along the proposed cross-strike
southeast of the cross-strike structural discontinuity, folds structural discontinuity, and the clastic rocks of the Lower
and thrust faults strike north-northwest, whereas to the Cretaceous Goyllarisquisga Group lie unconformably on the
northwest of it, they strike northerly (Figs. 1, 5, and 6). pre-Ordovician Maran metamorphic complex (Figs. 5 and
Moreover, fold plunges are reversed across this zone: to the 6). In contrast, north and southeast of the cross-strike struc-
southeast, most major anticlines and synclines plunge to the tural discontinuity, a relatively thick sequence of Mississip-
south-southeast, whereas to the northwest, they plunge pian to Lower Jurassic strata separates these units (Fig. 6).
north (Fig. 5). The locus of changes in strike and plunge ex- Sandstones and shales of the Mississippian Ambo Group lo-
tends northeast from Laguna Querococha, but about 5 km cally unconformably overlie the Maran complex near its
southwest of Antamina it steps 8 km to the north before con- western limit. Similarly, continental sedimentary rocks and al-
tinuing northeastward (Fig. 5). Faults with the same overall kaline to subalkaline volcanic rocks of the Lower Permian
northeast strike as the cross-strike structural discontinuity Mitu Group commonly overlie the Ambo Group and also lo-
control some present-day drainages, such as the northeast- cally lie unconformably on the Maran complex along its
trending Ro Puchca valley, 20 km north of Antamina, which western edge but are absent northeast of Antamina. Both
is discordant to the overall north to north-northwest grain of north and south-southeast of the cross-strike structural dis-
the terrain (Fig. 5). continuity, these siliciclastic sedimentary rocks are overlain by
The deflection in the Antamina area is one of several that Upper Triassic to Lower Jurassic carbonate rocks and shales
articulate the Maran thrust and fold belt. Northerly strikes (Pucar Group). The absence of these three groups directly
continue to Llamalln, 50 km to the north of Antamina in the northeast of Antamina records either a sub-Cretaceous ero-
eastern part of the belt (Fig. 1). Still farther north, the over- sional unconformity or nondeposition from Mississippian to
all north-northwest regional strike of the Maran thrust and Late Jurassic times. The cross-strike structural discontinuity
fold belt resumes. A comparable sharp deflection to north- therefore also coincided with a topographic high or arch, at
south strikes occurs at the northern end of the Cordillera least in the Middle to Late Jurassic, but possibly persisting
Blanca (Fig. 1), 175 km to the north-northwest of Antamina, throughout the Mississippian to Jurassic interval.
where the Casma-Pasto Bueno fault zone intersects the re- The cross-strike structural discontinuities probably also in-
gional north-northwest strikes (Rivera, 1996). Benavides fluenced the distribution of the uppermost Cretaceous to Pa-
(1999) identified many segments in the fabric of the Maran leocene red beds that unconformably overlie the Cretaceous
thrust and fold belt, including the two described above, al- strata (Fig. 4). These are absent near the proposed cross-
though he proffered a different mechanism for their forma- strike structural discontinuity through the Antamina area, and
tion, as discussed below. they also thin significantly near the more northerly Casma-
Stratigraphic variations across the cross-strike structural Pasto Bueno deflection, but they attain a considerable thick-
discontinuity: The contact relationships of the upper Paleo- ness between these two transverse zones as well as to the
zoic and Mesozoic strata (Fig. 4) to the lower Paleozoic south of the arch (Fig. 1).
Maran metamorphic complex vary in accordance with the Igneous activity along the cross-strike structural disconti-
segmentation of the thrust and fold belt. Figure 6 shows the nuity: Between the Cordillera Blanca and the Maran meta-
relationships along the western margin of the Maran com- morphic complex, the 1:100,000 quadrangle maps of Cobbing
plex throughout an area more extensive than that shown in et al. (1996) and Wilson et al. (1967, 1995) record a greater
Figure 5. Mississippian to Lower Jurassic strata that normally abundance of Tertiary hypabyssal and extrusive rocks along
separate the pre-Ordovician metamorphic rocks from the the proposed cross-strike structural discontinuity than to the
Cretaceous sedimentary rocks are absent near the proposed northwest or southeast (Fig. 6). In addition, west of the thrust
cross-strike structural discontinuity. In the north-striking seg- and fold belt, volcanic rocks of the Calipuy Supergroup
ment of the fold belt immediately north of the Antamina area, (Strusievicz et al., 2000) are abundant northwest and south-
the eastern limit of the Mississippian to Cretaceous succes- east of this cross-strike structural discontinuity (e.g., in the
sion is a subhorizontal unconformity that strikes north overall Nevado Huantsn area and the Cordillera Huayhuash; Fig.
but has an irregular surface trace owing to the incised topog- 6). These rocks are interpreted to be lower-middle Miocene
raphy (Figs. 5 and 6). In the north-northweststriking seg- because hypabyssal intrusions associated with similar volcanic
ments of the belt, however, the eastern extent of the Meso- rocks elsewhere in the region (Huaraz Group of the Calipuy
zoic rocks is generally delimited by north-northweststriking Supergroup; Fig. 5) have been shown by 40Ar/39Ar step-heat-
faults (e.g., southeast of Antamina, and northwest of Llamal- ing geochronology to have persisted to 14.2 Ma (Strusievicz et
ln; Fig. 6). Southeast of Antamina, the southwest margin of al., 2000; Love et al., 2001). However, the 115 km2 middle
the Maran metamorphic complex is largely defined by a se- Miocene granodioritic Carhuish pluton crops out on the axis
ries of major northeast-verging reverse faults that involved of the proposed cross-strike structural discontinuity (Fig. 5),
basement, but the Mississippian to Lower Jurassic strata are representing deeper-seated rocks of broadly equivalent age
preserved between the pre-Ordovician metamorphic rocks (13.7 Ma U/Pb zircon date on the main phase, Mukasa, 1984;
and the Cretaceous sedimentary rocks (Fig. 6). However, in 16.5 Ma K/Ar date on the marginal phase, Cobbing et al.,
the north-northweststriking segment northwest of Llamal- 1981). The volcanic rocks of Pampa Junn (Egeler and de
ln, the Maran complex is backthrust over the Cretaceous Booy, 1956) northeast of the Carhuish pluton (Fig. 5) have
rocks and the thickness of the Mississippian to Lower Juras- not been dated, so their significance with regard to the trans-
sic strata is only locally apparent (Fig. 6). verse structure is uncertain. However, we propose that the

0361-0128/98/000/000-00 $6.00 895


896 LOVE ET AL.

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

a NE SW

Ce

<- Lago Antamina 100m


skarn

b NW SE

Ce

skarn

Lago Antamina

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 pho-
tographs were taken are indicated in Fig. 2. Photographs taken in 1997 through 1999; this ridge has now been largely re-
moved 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 lime-
stone (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.

0361-0128/98/000/000-00 $6.00 896


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 897

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

a b

c d

e f

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) pele-
cypod 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 Ce-
lendn Formation, consisting of light gray, coalescing, carbonate-rich nodules separated by wisps of dark gray calcareous silt-
stone (Fortuna Mine area, see Fig. 2; hammer for scale). (f) Upper: Celendn Formation nodular limestone (DDH CMA-
C3, 221 m). Lower: lenticular bedding interpreted as sheared nodular limestone of the Celendn Formation (DDH
CMA-C5, 87.5 m).

0361-0128/98/000/000-00 $6.00 897


898 LOVE ET AL.

(Figs. 4 and 7a). This sequence hosts ore at the surface and at characteristic of its upper part. The relatively pure calcitic
depth in the southwest part of the deposit, but only in the marble of the Jumasha Formation that hosts ore at surface in
subsurface in the northeast. Few whole fossils are preserved the southwest part of the deposit and at depth in the north-
(Fig. 8c) and no ammonites were observed in the mine area, east sector does not contain the thick dolostones characteris-
precluding biostratigraphic correlations. However, we assign tic of the lower part of that formation, and no magnesian
this sequence to the Jumasha Formation on lithologic skarn has been found. On this basis, we conclude that the
grounds. Sheared fossiliferous carbonate wackestone occurs skarn developed in the upper part of the Jumasha Formation
locally in this sequence (Fig. 8c). In north-central Peru, Be- and the lower part of the Celendn Formation (Fig. 4). How-
navides (1956) described the Jumasha Formation as domi- ever, we estimate that approximately three-quarters of the
nated by massive, thick-bedded, light orange-brown to yel- mineralized exoskarn, and all of the wollastonite exoskarn, de-
lowish-brown and gray, fossil-poor dolostones and limestones veloped in the Jumasha Formation, and that a similar propor-
that weather dark yellowish-brown to brownish-gray. The for- tion of the contiguous and genetically related endoskarn was
mation has been further described as comprising topographi- formed in the stock adjacent to that formation. It is not
cally prominent, cliff-forming, light-gray limestones and yel- known whether magnesian skarn developed in dolostone
lowish dolostones that are characteristically bioclastic below the limit of exploration and development drilling.
(Wilson, 1963). In the Uchupata section (Fig. 4), the upper
437 m is limestone, and the lower 353 m is dolostone. Unlike Alteration of sedimentary rocks adjacent to the skarn
the overlying Celendn Formation, this formation only rarely The outer contact of coarse-grained mineralized exoskarn is
contains calcareous siltstones, although it incorporates marly abrupt, convoluted, and uninfluenced by bedding in the mar-
limestone beds near its top (Jaillard, 1987). The upper limit of bles formed from the limestones of the Jumasha Formation
the Jumasha Formation was defined lithostratigraphically (Fig. 7a), but it is more gradational and stratigraphically con-
where medium- or thick-bedded limestones pass upward into trolled in the fine-grained calc-silicate rocks developed in the
thin-bedded marls and limestones (Wilson, 1963). interbedded limestones and marls of the overlying Celendn
The upper sequence in the mine area (Figs. 4, 7, and 8d) Formation (Fig. 7b). In the Jumasha Formation, the mineral-
comprises thin- to thick-bedded impure limestone-marl that ized skarn is juxtaposed with marbles containing local hori-
varies in carbonate/silicate ratio from relatively calcite-rich zons rich in calc-silicate minerals that may be ascribed to
muddy limestone (generally 5075% carbonate) to calcareous thermal metamorphism. In contrast, the calc-silicate-bearing
siltstone (less than 50% carbonate). Many units contain light- rocks developed in the marly Celendn Formation include
gray, calcitic nodules, composing 10 to 90 percent of the rock, rock types that are interpreted as the products of either meta-
enclosed by dark, silty calcareous mudstone (Fig. 8e, f). No morphism or metasomatism. The latter do not host sulfide
limestone beds with siliceous nodules have been observed. minerals and differ radically in texture and mineralogy from
The nodular texture is interpreted to be diagenetic. Sheared the exoskarn ore. Their wide distribution around the upper
limestone with lenticular bedding exposed at the northeast part of the orebody constitutes a significant exploration tar-
head of Quebrada Antamina represents deformed nodular get.
limestone in which the nodules have been flattened into Jumasha Formation: On approaching skarn, dark gray Ju-
lenses during folding and faulting (Fig. 8f). The upper se- masha limestone is converted to coarsely crystalline, gray to
quence lacks identifiable fossils but is assigned to the Ce- white calcitic marble (Table 1, Fig. 9), forming an aureole
lendn Formation on lithologic grounds. In the Uchupata sec- ranging from tens to hundreds of meters wide. Within the au-
tion, 20 km north of the mine, the Celendn Formation reole, local development of sparse diopside, wollastonite, or
comprises very soft, friable, fossil-poor, light greenish-gray, scapolite porphyroblasts (Fig. 9a) or slight differential erosion
nodular, moderately silty marls and calcareous shales (Fig. 4; and color variation (Fig. 9b) in calcite marble record minor
Benavides, 1956). This formation is generally medium-bed- compositional variations in the limestone. Rare, fine-grained,
ded (0.30.8 m) and variably dolomitic, and it ranges from light green, diopsidic layers in calcite marble (Fig. 9c), the
fine-grained to pelletal (Wilson, 1963). porcellanite of Terrones (1958), are interpreted as calc-horn-
The contact between the Jumasha and Celendn Forma- fels, representing thermally metamorphosed, medium to
tions in the mine area is conformable and generally grada- thick beds of dolomitic, muddy, fine-grained limestone near
tional throughout several meters and is marked by upward-in- the top of the formation.
creasing siltiness and decreasing bed thickness. This contact Within the marble aureole, medium-grained, mottled, or
is interpreted to be the top of the uppermost thick-bedded banded gray marble predominates, locally grading inward to
limestone or marble unit on the basis of lithology and bedding medium- to coarse-grained, pure white marble, reflecting ei-
characteristics. This stratigraphic position is illustrated in Fig- ther degraphitization or metamorphism of organic matter to
ure 7a for outcrops on the southeast wall of the Antamina val- clear vitrinite. Rare, medium- to coarse-grained, thin, buff
ley. The contact is folded by the anticlines exposed on the val- garnet layers in both white and mottled facies of marble
ley walls adjacent to the southwestern part of the deposit, and mimic the forms of folded silicate-rich laminae (Fig. 9d) and
it dips northeast in the subsurface in the northeast part of the appear to have replaced them. This type of garnet develop-
deposit, as do the overlying strata exposed in that area. The ment is interpreted as a bimetasomatic reaction skarn. In gray
thin-bedded calc-hornfelses that predominate on the ridge marble, scapolite is locally developed in some darker-gray
crests flanking Quebrada Antamina grade laterally away from bands (Fig. 9e) but does not persist into white marble or
skarn into the marly, variably nodular sequence assigned to skarn (Table 1). Scapolite therefore forms a discontinuous
the Celendn Formation, but they do not contain the shales halo up to tens of meters wide and commonly separated from

0361-0128/98/000/000-00 $6.00 898


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 899

TABLE 1. Mineral Assemblages in Calc-Hornfels in the Celendn Formation, and in Marbles of the Jumasha Formation,
Adjacent to Ore-Bearing Garnet Skarn, Antamina, Peru1

Formation, rock type, Facies


and mineralogy Distal Intermediate Proximal

Celendin Formation2
Calc-hornfels Brown, very fine grained Gray, fine grained Green, fine to medium grained

Common minerals Calcite, anorthite, calcite, Calcite, quartz, tremolite Diopside, calcite, quartz, microcline
quartz, phlogopite

Minor minerals Diopside Diopside, Vesuvianite,


wollastonite, grossular,
phlogopite, hedenbergite
orthoclase, ferrosilite

Jumasha Formation
Marble Gray; fine to medium grained, Gray; medium grained, banded Gray; medium to coarse grained, massive
banded

Common minerals Calcite Calcite Calcite

Minor minerals Vitrinite or other organic matter Scapolite, vitrinite, graphite, Colorless vitrinite,
or other organic matter wollastonite,
diopside

1Determined by X-ray powder diffraction using a Philips PANalytical XPert PRO diffraction system with XPert Plus and XPert HighScore software for

peak matching and phase identification


2
Distal, intermediate, and proximal facies zones are broader in the Celendn Formation than in the Jumasha Formation

the skarn front by tens of meters. White marble develops in fine grained, light-brown phlogopitic facies; an intermediate
patches (Fig. 9d and f) and is concentrated along stylolites fine-grained, light-gray, tremolitic facies; and a proximal
and fractures (Fig. 9f). The latter probably represent fluid medium-grained, light-green diopsidic facies (Table 1, Fig.
pathways, suggesting that development of white marble in- 10a, b). The boundaries between these facies are commonly
volved the local channeling of fluids. Such pathways were not sharp and smooth but are locally irregular where controlled
consistently used by later skarn-related fluids and are locally by stockwork fractures (Fig. 10a, b). The outer limit of the
cut by fractures with wollastonite-rich selvages (Fig. 9f). distal brown facies is typically hundreds of meters from the
Immediately southwest of the skarn, the Jumasha Forma- boundary of sulfide-bearing skarn, and the progression from
tion has a distinct planar banded fabric, locally a spaced cleav- brown through gray to green facies occurs over distances
age, that dips northeast at a high angle to bedding (Fig. 9a). ranging from tens of meters adjacent to the main intrusion to
To the northeast, this fabric steepens and becomes vertical tens of centimeters (Fig 10b) adjacent to dikes extending be-
adjacent to the skarn and along the strike of a major fold axis. yond the main intrusion.
The consistent strike, systematically changing dip, and rela- Although local fluid channeling and probably metasoma-
tionship to folds visible in the adjacent limestones indicate tism occurred at the boundaries of the facies, the systematic
that this is an upward-fanning cleavage associated with local mineralogic zonation (Table 1) most likely reflects increasing
folding. In the adjacent skarn, alternating andradite-rich and temperature. The first appearance of light brown calc-horn-
sphalerite-rich bands, generally 5 to 50 cm wide, are subpar- fels corresponds with the phlogopite-in reaction, Dol + Kfs =
allel to this fabric, suggesting that at least some preexisting Phl + Cal, which occurs in the range 350 to 465C under ge-
structures influenced mineralization. This banding may cor- ologically reasonable conditions (i.e., P = 1,000 bars and XCO2
respond with the coarse-grained garnet-defined fabric ob- = 0.1 to 0.9: Tracy and Frost, 1991). The gray calc-hornfels
served by Terrones (1958). contains tremolite with or without phlogopite, whereas brown
Celendn Formation: In the Celendn Formation, on the calc-hornfels is tremolite free (Table 1). The transition from
ridges around Lago Antamina, the skarn has a halo of miner- brown to gray calc-hornfels therefore is interpreted to repre-
alogically diverse fine-grained calc-silicate rocks with subcon- sent a phlogopite-out reaction that coincides with the forma-
choidal to conchoidal fractures. With increasing distance tion of tremolite, probably through the reaction Phl + Cal +
from the skarn, these rocks grade into interbedded limestone Qtz = Tr + Kfs (Tracy and Frost, 1991; Table 1). The light-
and marl that reacted differentially to thermal metamorphism green calc-hornfels records the first appearance of massive
around the porphyry intrusion and orebody, thus highlighting diopside (Table 1), although veinlets and small blebs of this
the bedding (Fig. 7b). mineral locally occur farther from the intrusive contacts. This
Dark gray massive units change in color, mineralogy, and zone probably formed through the reaction Tr + Qtz + Cal =
grain size proximal to monzogranitic dikes. Although we have Di, which under similar pressure and XCO2 occurs at temper-
not mapped the various facies of these rocks in detail, we rec- atures of between 405 and 495C (Tracy and Frost, 1991).
ognize the development of three distinct zones: a distal very The local narrowness of the gray calc-hornfels facies and the

0361-0128/98/000/000-00 $6.00 899


900 LOVE ET AL.

a b

c d

e f

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-con-
trolled 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.

0361-0128/98/000/000-00 $6.00 900


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 901

a b

c d

e f

FIG. 10. Hornfels, skarnoid, and wrigglite developed in the Celendn Formation. (a) Irregular stockwork of sealed frac-
tures 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 phl-
ogopitic 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, ap-
proximately 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 calcite-bearing calc-hornfels matrix (base of cliffs, east of Lago
Antamina; 15 cm pencil for scale). (e) Irregular concentric 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).

0361-0128/98/000/000-00 $6.00 901


902 LOVE ET AL.

absence of tremolite in the phlogopite facies suggest that the steep slopes around the Antamina valley provided clear,
metamorphism took place at high XCO2. Under such condi- three-dimensional images of the local structural relationships
tions the diopside-in and phlogopite-in reactions are sepa- (Fig. 12), summarized in an isometric block diagram in Fig-
rated by only circa 10C. Also, the development of tremolite ure 13. The vergence of the thrust faults are interpreted from
through the elimination of phlogopite is promoted under cutoff angles, and it is assumed that the strata are all upright,
these conditions, but its formation within the field of phlogo- unless obviously overturned. The local thrusting and related
pite stability through the reaction Dol + Qtz = Tr + Cal is in- folding are described below in sequence from southwest to
hibited. northeast.
In contrast to these calc-hornfelses that formed under es- Southwest of Antamina, a northeast-verging thrust, herein
sentially thermal metamorphic conditions, calc-silicate devel- named the Yaquirsh-Buque Punta thrust (YBPT in Figs. 11,
opment through metasomatism is revealed by development of 12, and 13), has caused structural repetition of the Jumasha
skarnoid in some beds of the Celendn Formation, extending Formation. A few hundred meters to the east of this fault, the
tens of meters beyond the mineralized skarn front. Skarnoid Jumasha Formation is thrust over the younger Celendn For-
is a descriptive term for calc-silicate rocks that are relatively mation on the northeast-verging Antamina thrust fault (mine
fine grained, calcium rich, and iron poor, and that reflect, at terminology; AT in Figs. 11, 12, and 13). The imbricate thrust
least in part, an aluminosilicate component in the protolith stack is developed in massive limestones of the Jumasha For-
(Zharikov, 1970). It is genetically intermediate between a mation (Fig. 8a) in the next transcurrent valley to the north-
purely metamorphic hornfels and a purely metasomatic, west of Antamina (Quebrada Callapo in Fig. 2). This is part of
fluid-controlled skarn, which is typically coarser grained and a duplex in which the Antamina thrust is the sole and one of
does not as closely reflect the composition or texture of the the unnamed thrusts west of the Yaquirsh-Buque Punta
immediately surrounding rocks (Einaudi, 2000). In some thrust is the roof, although the latter has been extensively
beds in the Celendn Formation, diopsidic nodules stand out eroded. These imbricate slices do not all continue south from
in relief against the calcite-bearing calc-silicate matrix (Fig. Quebrada Callapo into Quebrada Antamina because some
10c, d). The nodules, locally concentrically banded, range fault surfaces merge.
from >10 cm to <1 cm in diameter and from sparse to abun- On the northwest slope of the Antamina valley, only the Ju-
dant, independent of their size (Fig. 10c, d). They are inter- masha Formation occurs between the Antamina thrust and
preted as products of the preferential metasomatism of cal- the Yaquirsh-Buque Punta thrust, whereas the Jumasha and
citic nodules (cf. Fig. 8e, f). Celendn Formations in conformable stratigraphic sequence
Both close to the skarn front and tens to hundreds of me- separate these two thrusts on the opposing southeast slope
ters from it, sulfide-free concentrically layered structures (Figs. 11, 12, and 13). The immediately underlying thrust
occur in some massive (nonnodular) beds of altered Celendn slice, carried on the northeast-verging Fortuna thrust (mine
limestone (Fig. 10e). These were interpreted as algal struc- terminology: FT), occurs only on the northwest slope of the
tures by Bodenlos and Erickson (1955). However, they do not valley, east of and below the Antamina thrust. In this area, the
have the requisite three-dimensional geometry, and no un- Fortuna thrust repeats the Celendn Formation and has a
ambiguous algal structures have been seen in unaltered lime- small recumbent anticline in its hanging wall (Figs. 11, 12a,
stones in the area. Also, at the transition from brown to gray and 13). However, neither this thrust nor its hanging-wall an-
facies in calc-hornfels, rhythmic banding is developed adja- ticline can be traced across the valley to its southeast side
cent to through-going, northeast-striking fractures (Fig. 10f). (Fig. 12b, c), either because the thrust surfaces merged or be-
Similar laminated features occur in ore-bearing wollastonite cause the branch line of the two thrusts plunges north and
skarn. Because of their similarity to banded features de- therefore climbs above the erosion level to the south. The An-
scribed in other skarns (e.g., Knopf, 1908; Eskola, 1951), we tamina thrust cuts the shallowly dipping Fortuna thrust (Figs.
interpret the layering as metasomatic wrigglite (Kwak and 11, 12a, 13) as well as other minor, bedding-parallel thrusts in
Askins, 1981). From these relationships, we conclude that the Celendn Formation.
minor metasomatism occurred beyond the extent of metallic Most of the folds and thrust faults at Antamina are north-
mineralization in the Celendn Formation. east verging, as is characteristic of the Maran thrust and
fold belt (Mgard, 1984), but two southwest-verging thrusts
Structural geology of the mine area occur southeast of Lago Antamina (Figs. 11, 12c, and 13).
The Upper Cretaceous strata of the Machay Group that These thrusts are interpreted as backthrusts related to the un-
host the skarn were intensely deformed in the Maran derlying northeast-verging Oscarina thrust (mine terminol-
thrust and fold belt. Southwest of the mine, the Jumasha For- ogy: OT), which is the lowest in the immediate mine area.
mation is thrust over the younger Celendn Formation, and This thrust can be traced along the crest of the southwest
the Lower Cretaceous formations of the Goyllarisquisga slope of the Tucush valley (Fig. 2), east of Antamina. It also
Group are thrust both over the Celendn Formation and over repeats the Celendn Formation and deforms it into a hang-
each other in reverse stratigraphic order (Fig. 11). ing-wall anticline (Figs. 11; 12a, b, d; and 13). Many other
Local folds and faults: The Antamina deposit is hosted by a minor thrust faults are recorded by ramps with bedding cut-
relatively flat-lying section of the Jumasha and Celendn For- offs for distances of tens of meters, especially in the Celendn
mations, is deformed by open folds, and is cut by numerous Formation, but are not illustrated here.
small thrust-fault ramps and bedding-parallel thrust faults Prominent northwest-striking, open, upright anticlines are
(Fig. 11), all interpreted to have developed during the late exposed above the deposit on the northwest and southeast
Eocene Incaic orogeny. Prior to open-pit development, the walls of the valley (Figs. 7a; 12a, c; and 13). They become

0361-0128/98/000/000-00 $6.00 902


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 903

Contonga Cu-Pb-Zn-Ag Mine

Contonga
Stock
Taully
Stock

Os
ca
rin
at
hr
us
t
8948000N
Fortuna Mine

Q
AT

ue
bra
da
Tu

sh
cu

a
YBP

Ay
sh

da
T

ra
eb
FT

Qu
VF
Rosita de Oro Mine

VF
ina
am
Ant
a 8944000N
ad
ebr
An ticli

Qu
an
tam ne
Anta

ina
mina
YB
PT

thr

Yanacancha
us
t

272000E 276000E

Miocene Antamina planned pit outline


skarn
thrust faults
granitoid intrusions of known or inferred Miocene age
folds: anticlines, synclines
Machay Group
Upper Cretaceous bedding traces
Celendn Fm. (thin-bedded marl, limestone & calcareous shale)
past-producing mine or prospect
Jumasha Fm. (thick-bedded limestone)

Lower Cretaceous
Pariatambo Fm. (thin-bedded, dark-gray bituminous limestone & shale)
Chulec Fm. (thin-bedded, light-gray, carbonates, calcareous shale & calcareous sandstone)
Pariahuanca Fm. (medium- to thick-bedded limestone & calcareous sandstone)

Goyllarisquizga Group
Carhuaz Fm. (thin- to medium-bedded sandstone, siltstone & shale)
Santa Fm. (thin- to medium-bedded limestone, shale & sandstone) 0 1 2 3

Chimu Fm. (quartz sandstone) kilometers

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.

0361-0128/98/000/000-00 $6.00 903


904 LOVE ET AL.

a
SW Cerro Yaquirsh YBPT NE

J J
AT FT OT
Ce
Ce

Ce
+

YBPT J VF
Lago Antamina

b
SW YBPT AT NE
Cerro J Ce
Buque Ce Ce
Punta +
J + OT
+
+ Ce

c
NE AT YBPT SW
Ce Cerro Buque Punta
J J
Ce Ce
Ce Ce
Ce Ce
Ce
+
+
+
+ J

+ +
+
Lago Antamina
d
NE Cerro Yaquirsh SW
Cerro Jatunpunta
YBPT
J
FT J
OT Ce AT J AT
Ce

Ce Ce

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 con-
tact 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 Ju-
masha (J) and Celendn (Ce) Formations. (d) The northwest flank of the ridge extending northeast of 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.

0361-0128/98/000/000-00 $6.00 904


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 905

e e

OT
C C
e
C

OT
e
C
e
C

AA
Ce

FT
e Ce
Ce
C
e
C

AT
C
e NE
YB J
FT

Ce
PT

AT Ce C
e
F
V
J

OT
J
Ce
e
C
e
C
PT J e

AA
C
YB OT
J e

OT
C
J
e
LR
C Ce C
e
Ce
V
J C
e J
J J
e
C
J
V
F
OT

AA
SW J NE
e
AT

J
e
Ce
C

J
J J
J
e AT Ce
C
J

m
Ce 00
PT
J
10
YB

0
50
SW 0

FIG. 13. Schematic, isometric block diagram of the pre-Miocene geology of the Antamina area, looking north, summa-
rizing the major Eocene Incaic folding and thrust faulting in the host Cretaceous strata. In the late Miocene, intrusion and
formation of the Antamina deposit occurred in the central block on this diagram, localized by the Valley lateral ramp (VLR)
and the step in the Valley fault (VF). The locations of the three blocks are indicated by rectangles in Fig. 11. The Jumasha
(J) and Celendn (Ce) Formations are indicated. Faults and fold axes are shown by solid black lines, bedding by black dashes.
Other features abbreviated as in Figures 11 and 12.

broader and more open upward but the intervening synclines faulting and folding and that of intrusion and mineralization.
are tight. Such concentric folds are common elsewhere in the The orientation of these dikes changes considerably along
Maran thrust and fold belt (e.g., Coney, 1971). The largest strike; near the main stock they are nearly vertical, strike east,
of the parallel folds visible on the flanks of the Antamina val- and cut strata (Fig 12c), whereas farther east beyond the
ley has been widely referred to as the Antamina anticline, the ridge crest they become sills and follow either strata or bed-
axial plane of which has an apparent dextral offset of approx- ding-parallel thrust flats, strike south, and dip moderately to
imately 500 m in a northeast-southwest direction across the the west (Fig. 12b). The dikes become progressively more al-
valley. This offset has been variously ascribed to local bends tered to endoskarn toward the main intrusion, and although
in the strike of the folds (Bodenlos and Ericksen, 1955), dex- they transect zoned exoskarn they do not appear to have in-
tral offset on a postulated northeast-striking Valley fault (Ter- truded it. A similarly variable orientation is shown by a dike
rones, 1958), transverse normal faulting (Petersen, 1965,) and that extends northeast from the northern shore of Lago Anta-
cross-folding (McKee et al., 1979). However, in this study we mina; it locally crosses strata at a steep angle but also follows
show that the offset is best explained by a lateral ramp model. flat and moderately dipping strata (Fig. 12a). The parallel fea-
In this model (Fig. 14), the anticlines represent fault-bend tures at opposite ends of the main intrusion and orebody sug-
folds related to hanging-wall cutoffs folded over the top of gest an offset, left-stepping, northeast-southwest fracture
northwest-striking frontal ramps that are offset because a zone that is longer than the 500 m lateral ramp, the postu-
northeast-striking, 500-m-long transfer fault or lateral ramp lated Valley fault (Figs. 3b and 11).
separates the frontal ramps in an apparent dextral sense. This A well developed northeast-striking, widely spaced, nearly
postulated structure, the Valley lateral ramp, would account vertical fracture set occurs in the sedimentary host rocks
for the absence of corresponding offsets of other structures to throughout the area (e.g., on the southwest flank of Cerro
the southwest and northeast (Fig. 11). Racpe; Fig. 15a), and we interpret it as regional a-c jointing
Minor diking and skarn alteration extend beyond the main related to folding. A similarly oriented fracture set is in-
intrusion and orebody to the southwest down Quebrada An- tensely developed near the deposit, especially along the
tamina and to the northeast and east at the head of the valley. southeast side of the Antamina valley (Fig. 15b, c). In many
The vertical dikes that extend east from the main porphyry places, calc-hornfels is more intensely developed and wide-
mass to the Rosita de Oro area (Figs. 11 and 12c) offset two spread where these fractures are closely spaced. All of the
minor forethrusts and backthrusts, indicating that there was northeast-striking fractures may have been related to Eocene
minor, apparently vertical, faulting between the time of thrust folding and thrust faulting, but they were better developed in

0361-0128/98/000/000-00 $6.00 905


906 LOVE ET AL.

Future
Hanging-wall
Frontal Ramp
Block Foot-wall Block

Future Flat mp
l Ra
Foot-wall era
Block Lat
Trace of Future Fault
transport
direction

a b
Fault bend fold
Anticline
Anticline Hangingwall Cutoff
Anticline

line
ntic
al A
Later

c d
FIG. 14. Schematic diagram illustrating the lateral ramp model for the structural setting of the Antamina deposit. Look-
ing 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) Hang-
ing-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.

the mine area because of flexure and tearing of strata above Valley fault. The hydrothermal breccia sheets that are com-
a transfer fault or lateral ramp (discussed below). Following mon at or near the endoskarn-exoskarn contact also have this
intrusion, they provided access for fluids to the hornfels- orientation. The irregular zones of breccia and endoskarn
bearing strata. within the intrusion are interpreted to strike predominantly
In the Antamina district, therefore, the Jumasha and Ce- north-south (Fig. 3b) and may have been controlled by cross-
lendn Formations have been thrust-faulted, folded, and jux- cutting structures. The western end of the main body of the
taposed into a thick, complex thrust stack. The total thickness intrusion also generally strikes north-south, as does the skarn
of rock that overlay the site of mineralization in the late front in that area (Fig. 3b). The parallelogram shape, in plan,
Miocene cannot be estimated because the local thickness of of the main body of the intrusion and the surrounding skarn
overlying uppermost Cretaceous and Paleocene red beds is is complicated by the network of anastomosing dikes with en-
unknown and because it is unclear if the subaerial volcanic velopes of fine-grained garnet skarn extending to higher ele-
rocks and associated sedimentary rocks of the Eocene to mid- vations to the east of Lago Antamina (Fig. 12c). At least one
dle Miocene Calipuy Supergroup (Strusievicz et al., 2000) ex- of these dikes intrudes a normal fault on which movement oc-
tended into this area. curred between the time of thrust faulting and folding and
Structural control of mineralization: As in most intrusion- that of intrusion and mineralization (Fig. 12c). Several other
related skarn deposits, the most obvious feature controlling dikes with skarn envelopes also extend beyond the main mass
mineralization at Antamina is the contact between the main of porphyry and most are controlled by Incaic structures.
stock and the host limestones (Terrones, 1958; Petersen, Along the southern edge of the porphyry and skarn, minor pe-
1965; Redwood, 1999). However, the southeast and northwest ripheral rhyodacitic dikes and sills and their skarn envelopes
intrusive contacts are themselves parallel to, and along strike locally follow bedding and thrust faults. Several minor Pb-Zn-
from, the two segments of the proposed northeast-striking Ag veins (Fig. 2), some of which have been mined on a small

0361-0128/98/000/000-00 $6.00 906


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 907

scale, occur at or near the contacts between these dikes and faults. Lateral ramps are thought to be largely the result of
the host limestone, describing a circa 3 3 km area of dis- the interaction of thrust sheets and old basement fracture sys-
persed hydrothermal activity centered on the Antamina skarn tems (Pohn, 2000). Both Mgard (1987) and Benavides
(Bodenlos and Ericksen, 1955). (1999) proposed that the segmentation and articulation of the
Maran thrust and fold belt probably reflect control by base-
Discussion ment structures, although they differ in their interpretations
of the mechanism. Benavides (1999) attributed the segmen-
Regional stratigraphic and tectonic relationships tation of the Maran thrust and fold belt to major northeast-
The postulated Valley lateral ramp at Antamina may have striking, dextral basement faults, without specifying when the
been controlled by underlying, northeast-striking basement faults were active. Further, Bussell and Pitcher (1985) sug-
gested that a well developed set of northeast-striking, en ech-
elon, dextral faults in the Cretaceous-Paleogene Coastal
a batholith may have controlled some of the contacts of early
Paleocene intrusive ring complexes (Fig. 16). However, no
J significant tear faulting parallel to the northeast-trending de-
Ce
flections 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.
b 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
Ce the margin. It is evident that the Cretaceous strata exhibit no
clear relationship between thickness and proximity to a cross-
strike structural discontinuity.
A model for the geologic history of the Antamina region
The geologic history of the Antamina region from the for-
mation of the West Peruvian trough in the Jurassic to the
c Fractures:

steep,
FIG. 15. Northeast-striking, nearly vertical fracturing in the eastern and
vertical
southeastern part of the Antamina mine area. Locations from which the pho-
tographs 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 east-
northeast at the southwest flank of Cerro Racpe where all bedding dips
southwest, toward the point of view. Widely spaced northeast-striking frac-
tures are expressed as lineaments that converge in the distance and are per-
pendicular to the strike of the Celendn (Ce) and upper Jumasha (J) Forma-
tions on the lower slope, but are more obvious as vertical fractures in the
Ce lower Jumasha Formation on the steep upper slope. Long black dashes indi-
cate the stratigraphic contact between the Celendn and Jumasha Forma-
tions. (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 in-
tensely 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 northeast-
striking 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).

0361-0128/98/000/000-00 $6.00 907


908 LOVE ET AL.

ANTAMINA

8S

S
9S
a

W
W
Cerro de Pasco

W
Huancayo

10

11

12
75
77

76
W

Rio Maran PASTO La Oroya


Hunuco
78

BUENO

B-16
B-19
YANACOCHA B-14,15
6
9 Huancavelica
11 14 20 22 D
Cajamarca
Cordillera Bla 17
nca KEY
W

Casma - Pasto Huaraz


7 Section
79

Bueno zone PIERINA Querococha


Arch Intrusion
Anticline
Fault
Coastal Batholith Segment boundaries
discussed in text
Trujillo Other segment boundaries
N

Chimbote Casma of Benavides (1999)


Huarmey 0 100 200 kms
Lima
Pacific Ocean
0 100 200 miles

b B-16
14
>1000m
>920m

ANTAMINA 11
>500m 17 20 22
B-19 6 D
KsP-Ch 9
Campanian and >250m
>180m >200m >200m >200m
younger (< 83 4 Ma) KsP-Ch KsP-C >150m KsP-C KsP-P ~250m
25m 25m
163m 131m 115m 135m
174m Ks-J ~340m
Ks-Ce 443m 185m 59m 337m
70m 617m
558m 52m
calcareous 433m 122m 230m 44m Ki-Pt ~50m
sequence 750m 39m 232m
167m 280m Ki-Cl ~280m
nian 2) Ki-Pt 25m
Ks-J 107m
Ks-Ca oma 138m 600m 44m
Cen n (97.0 Ki-Cl 138m
800m 266m
Albia >280m
(equivalent to
top 1/5 of Ks-J ) Ki-G
Ki-Cr 190m 851m
>250m ~850m
B-14, 15 >485m

Ks-R (equivalent to NORTH NORTH-CENTRAL CENTRAL


lowest 1/5 of Ks-J) >413m
(north of approx. 930' S) (south of approx. 930' S) Chunumayo Fm.
Paleogene Carcapuquio Fm.
~~~~~~~~~~ Chota Fm. (KsP-Ch), Casalpalca Gp. (KsP-C) Pocobamba Fm. (KsP-P)
135m Pucar Gp.
Celendn Fm.(Ks-Ce) Celendn Fm.(Ks-Ce) ~700m

Ki-Cr ) Cretaceous Otuzco Gp.


1
3. Upper Cajamarca Fm. (Ks-Ca)
613m
clastic us 4.8 Quilquian Gp. Coor Fm.
eo 14 Romirn Fm. Jumasha Fm. (Ks-J) Machay Group Jatunhuasi
sequence ac (
et sic Mujarrn Fm. (north-central and central) Basin,
Cr ras Pulluicana Gp.
J u ~~~~~~~~~~ Rosa Fm. (Ks-R) S of La Oroya
Lower & SW of
Ki-G Pariatambo Fm. (Ki-Pt) Huancayo
>516m Crisnejas Fm. (Ki-Cr)
Chulec Fm. (Ki-Cl)

Inca Fm. Pariahuanca Fm.

Goyllarisquisga Gp. (Ki-G) Goyllarisquisga Gp. (Ki-G)

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. Sec-
tions 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 sec-
tion D from Manrique (1998). Antamina is located between sections 6 and B-19.

Miocene intrusion and formation of the orebody is summa- The Querococha arch coincided with an intermittent topo-
rized schematically in Figure 17. In this model, the margin- graphic high that developed at least in the Middle Jurassic, or
parallel West Peruvian trough formed during the Middle or even throughout the Mississippian to Middle Jurassic inter-
Late Jurassic by extension on en echelon, northwest-striking val, and which also influenced the distribution of Cretaceous
normal faults separated by northeast-striking transform faults carbonate rocks, Paleocene red beds, and Miocene igneous
(Fig. 17a, b; Mgard, 1987). The normal faults on the western rocks. It is apparent that the development of the northeast-
margin of the basement are thus right-stepping, but the seg- striking basement structures predated Jurassic-Cretaceous
ments experienced no relative displacement and each north- sedimentation. The Querococha arch apparently influenced
east-striking transform fault experienced only sinistral move- the distribution of Mississippian to Lower Jurassic rocks
ment. This distribution of growth faults in the Jurassic would northeast of Antamina, resulting in either local nondeposition
result in promontories and reentrants in the margin of the in the Mississippian to Early Jurassic or a Middle to Late
West Peruvian trough, similar to those of the larger-scale Jurassic erosional unconformity (Fig. 17b). Additional minor
early Paleozoic eastern margin of North America (Thomas, extension at any angle discordant to the northeast-trending
1977). faults could have resulted in reactivation of the originally

0361-0128/98/000/000-00 $6.00 908


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 909

t
a men b ra ssic
base le Ju
Midd
N Q
CPB

eous
c retac d
- L ate C ocen
e
midd
le Pale

ne ene
e Eoce f Mioc

A
Q
CPB

Miocene Faults
intrusions transform
Paleocene
Casapalca Gp. normal
Albian - Upper Cretaceous
Machay Gp. thrust
Upper Jurassic - Lower Cretaceous
Chicama & Goyllarisquizga Gps. Fold axis, with regional plunge
Upper Triassic - Lower Jurassic
Pucar Gp. Regional tectonic forces
Mississippian - Lower Triassic
Ambo & Mitu Gps.
pre-Ordovician
Maran metamorphic complex

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 ech-
elon 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 M-
gard, 1987). (b) Also in the Middle or Late Jurassic, minor extension at an angle to the transform faults segmented the con-
tinental 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 al-
though carbonate facies may have extended farther seaward 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 al-
lowed 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 In the Cretaceous, the Querococha arch and other trans-
erosion of preexisting sedimentary rocks from the transverse verse structures that segment the Maran thrust and fold
highs (Fig. 17b). The effect of the Querococha arch in the belt had little apparent effect on the thickness of clastic
Late Jurassic cannot be deduced because the Upper Jurassic (Goyllarisquisga Group) and carbonate (Machay Group) sed-
Chicama Group is not exposed in the eastern part of the imentary rocks on the Yauli shelf (Figs. 16b and 17c), but it
Maran thrust and fold belt. appears to have affected the distribution of the carbonate

0361-0128/98/000/000-00 $6.00 909


910 LOVE ET AL.

rocks. The Machay Group does not extend as far west to the Regional effects on local structure and mineralization
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 At Antamina, the northeast-trending fracture set, the Valley
Machay Group therefore crosses the belt in a northeast- fault and the Valley lateral ramp are parallel to the regional
southwest direction (Fig. 1). Thomas (1977) noted a similar cross-strike structural discontinuity, the Querococha arch,
relationship between the extent of carbonate facies and the and may have been controlled by similarly oriented underly-
locations of structural salients and recesses in the Appalachi- ing basement structures. At the local scale, the Valley lateral
ans. The distribution of latest Cretaceous and Paleocene red ramp and the Antamina intrusion have been localized by the
beds has also been influenced by the transverse structures left-stepping jog in the Valley fault. Regionally, about 5 km
(Fig. 17d). The red beds pinch out toward both the Quero- southwest of Antamina, the northeast-trending locus of
cocha arch and the Casma-Pasto Bueno zone, and they at- changes in strike and plunge in the Maran thrust and fold
tain their greatest thickness between these arches in a struc- belt steps left by approximately 8 km (Fig. 5). Thus the step
tural salient, which would have been a reentrant and in the Valley fault developed within, and mimics, a similar,
depocenter in the margin of the West Peruvian trough at the larger-scale, left-stepping jog within the Querococha arch.
time of sedimentation. The local-scale structural evolution of the Antamina area,
Shortening and variations in the regional attitudes of folds including intrusion and formation of the orebody in the late
and thrust faults generated in the Eocene Incaic orogeny are Miocene, is summarized schematically in Figure 18. Whereas
represented in Figure 17e. The folds and thrusts in the An- other models, such as en echelon folding, could explain the
tamina region constitute an articulated structural recess in apparent dextral offset of the Antamina anticline, the lateral
the margin of the craton, which probably formed through de- ramp model is preferred here because it provides a locus for
formation around a basement promontory, the northwestern later, northeast-elongated intrusion and hydrothermal activ-
edge of which is now delineated by the Querococha arch. ity. The northeast-striking fracture set in the host rocks pe-
Because the orientation of folds and thrust faults in thin- ripheral to the ore at Antamina (Figs. 15 and 18a) was, we
skinned tectonic belts generally reflects the underlying contend, formed by deformation associated with thrust trans-
ramps rather than the translation direction of deformation lation along the underlying and similarly oriented transfer
(Pohn, 2000), the strike and articulation of the Maran fault or lateral ramp (Fig. 14). The overall form of a thrust
thrust and fold belt mimic the geometry of the basement, de- sheet does not record its passage over a lateral ramp beyond
spite variations in the Cenozoic plate convergence direction the limit of that ramp but, in overlying thrust sheets, longitu-
and rate (Pardo-Casas and Molnar, 1987; Somoza, 1998; dinal fractures would form in the lateral anticline over the
Norabuena et al., 1999). Old transform faults in the margin ramp. These fractures would have sheared vertically in a
of the West Peruvian trough did not experience extensive north-sidedown sense or opened owing to flexure above the
later strike-slip movement because the maximum shortening lateral ramp or transfer fault and would persist beyond it (Fig.
direction was not parallel to the orientation of movement 15d), providing evidence that a thrust sheet had traversed
during rifting. We propose that the sinuous configuration of such a ramp. At Antamina, the inferred Valley lateral ramp ex-
the mountain belt generally reproduces the original zig-zag tends only 500 m from one offset anticlinal axis to the other,
margin, although the articulation is pronounced on the east- but strong, northeast-striking, nearly vertical fracturing, in-
ern side of the belt, and shortening in the interior of the belt terpreted herein as evidence of tearing in the overriding
had a smoothing effect on the segmentation. Thus, the re- thrust sheet, extends farther northeast (Figs. 15 and 18a) and
gional strike at the western extent of the Cretaceous carbon- is interpreted as trace evidence of the Valley lateral ramp.
ate strata gradually changes throughout a distance of approx- The upward-fanning axial planar cleavage related to the An-
imately 175 km along strike, from northerly near Antamina tamina anticline also formed at this time (Fig. 18a).
to north-northwesterly at the northwest end of the Cordillera In this model, the left-stepping jog in the transcurrent Val-
Blanca (Fig. 1). As Thomas (1977) concluded in the context ley fault localized the Eocene development of the lateral
of eastern North America, the Maran thrust and fold belt ramp that resulted in formation of the offset anticlines and
has formed a best-fit curve around old promontories and also focused the Miocene igneous activity responsible for the
reentrants. Antamina stock and skarn. Prior to intrusion of the main por-
The persistence of Miocene igneous rocks farther from the phyry mass, the early east-striking dikes east of the main in-
main axis of the magmatic arc into the foreland thrust and trusion formed in association with east-west shortening local-
fold belt along a proposed basement transverse structure ized at the northeast end of the southwestern segment of the
(Fig. 17f) is consistent with observations in other belts. Ig- Valley fault (Fig. 18b). Many of the peripheral dikes and min-
neous intrusions have been documented directly over, and eralized veins beyond the main porphyry body intrude, or
elongated parallel to, three of the four lateral ramps in the branch from, the same transcurrent structures. The dike
Appalachians analyzed in detail by Pohn (2000). Further- branches were controlled by other preexisting structures, pre-
more, the difference in exposure level of the Miocene ig- dominantly bedding and Incaic thrusts, and they may have in-
neous rocks on and adjacent to the southwest end of the arch truded at this time. In the Jumasha Formation the develop-
implies that post-Eocene uplift enhanced the plunge rever- ment of white marble was localized by preexisting axial
sals of the Eocene folds across the arch (Love et al., 2001). cleavages, stylolites, and fractures, providing early fluid path-
Thus, faults parallel to the arch may have accommodated its ways. Also, in the Celendn Formation calc-hornfels forma-
uplift and furnished the structural anisotropies that provided tion was localized in many places where the northeast-strik-
conduits for magma ascent in the late Miocene. ing, nearly vertical fracture set was strongly developed.

0361-0128/98/000/000-00 $6.00 910


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 911

fra lt an
fa u
a

Va

ctu d
b

lley
rds-

res late
upwa g

Fa

pa ral r
fa n n in

u
age

ral am
cleav

lt

lel
fold

to p
line bend
ntic fault
ina A cline
Antam end fold ina Anti
fault
b Antam

p
p l ram
l ram fronta
fronta

ral Ra
mp Va
lley
y Late
Valle Fa
u
N
lt

c d

Strike and dip of cleavage


e Anticline
Thrust fault
Normal fault
Fractures
Dike
Intrusion

Regional east-west shortening

Regional east-west extension

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 upward-
fanning 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 be-
tween the two segments of the Valley fault, and ultimately generated skarn ore. (e) Renewed east-west compression formed
east-weststriking veins in skarn.

The primary control on the skarn and some of the breccia gap between the two offset segments of this structure (Figs.
zones was the margin of the main intrusion, which we argue 3b and 11). In such a system, any relaxation in east-west com-
was itself localized by the postulated step in the Valley fault. pression could allow sinistral movement on oblique north-
The northwest and southeast margins of the main intrusion east-trending structures such as the Valley fault and forma-
appear to be defined by the northeast-trending Valley fault, tion of extensional north-south-oriented structures within the
and the main mass of the intrusion occupies the left-stepping left-stepping jog (Fig. 18c), thereby providing a locus for

0361-0128/98/000/000-00 $6.00 911


912 LOVE ET AL.

intrusion (Fig. 18d). Development of northeast-striking brec- Implications for local exploration and ore genesis
cia zones within skarn on the northwest and southeast mar-
gins of the intrusion and north-striking breccia and endoskarn The surrounding zones of marble in the Jumasha Forma-
zones within the intrusion represent, in this model, structural tion, and particularly of hornfels and skarnoid in the Celendn
reactivation of the major transcurrent faults and the minor Formation, provide larger exploration targets than the skarn
north-south extensional faults, respectively. itself. An isometric block diagram of the simplified geology of
Although the recrystallization of the host rocks during skarn the deposit (Fig. 19) shows the structure and alteration in the
development obscured evidence of controls by preexisting adjacent host rocks. Gray marble in the Jumasha Formation
fractures or bedding, we deduce that the intersecting struc- generally extends tens of meters beyond skarn, although lo-
tures that locally control ore grades in skarn may have origi- cally it extends beyond 100 m, but the outer limit of the dis-
nated at substantially different times. Terrones (1958) re- tal, brown, phlogopitic facies of calc-hornfels in the Celendn
ported higher exoskarn ore grades where a set of 100 Formation typically extends several hundreds of meters from
mineralized sheeted veins intersects structures extending the boundary of sulfide-bearing skarn. The most extensive
from a northwest-striking anticlinal axis in marble, and which halo around the skarn is represented by the swarm of Ag-
we interpret as upward-fanning axial planar cleavages. The in- bearing Pb-Zn vein deposits (Fig. 2). These occur within and
tersecting sheeted vein set differs in orientation from the beyond marble and calc-hornfels in a 9 km2 area surrounding
northeast-striking fracture system related to the lateral ramp Antamina and up to a kilometer from the skarn front.
described above. We propose that these veins, which cut The development of strong exoskarn mineralization may
skarn and were therefore late, formed in tension fractures as- have been promoted by the relatively pure calcitic limestone
sociated with renewed east-west shortening (Fig. 18e) after of the Jumasha Formation, which experienced intense calcite
the brief episode of relaxation. Thus an Incaic axial planar destruction and Ca mobilization. In contrast, the overlying
cleavage was intersected by Miocene tension veins and devel- Celendn Formation, having been metamorphosed to horn-
oped a permeable path for hydrothermal fluids. fels, may have acted as a cap on this hydrothermal system,

VF

Ce

Ce

Ce
0 Ce
Ce
20
0 AA Ce
m VF
J

J J
J
Ce
Ce

J J
Ce J Former Lago Antamina
VF
J VF
Plunging anticline
? Thrust fault
Trace of Valley Fault
J ? AA
Intrusion
J
Ce Skarn with interpreted position of
J original intrusive contact
VF Ce
J Celendn Approximate extent of macroscopically visible
J Formation
J Ce hornfels and skarnoid development in
Celendn Fm.
J J Jumasha
Formation Approximate limit of macroscopically visible
marmorization in Jumasha Fm.

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 exten-
sive calc-hornfels and skarnoid development in the Celendn Formation, and the restricted marmorization of the Jumasha
Formation. The offset Antamina anticline (AA) is indicated, but the inferred Valley lateral ramp is hidden from view in
this perspective.

0361-0128/98/000/000-00 $6.00 912


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 913

suppressing the upward and outward escape of fluids and the Querococha arch, an apparent swarm of intrusive bodies
thereby fostering development of extensive endoskarn. was emplaced during this interval, one of which generated
The same lithologic succession of pure carbonate overlain the Antamina hydrothermal system (Love et al., 2001).
by muddy carbonate to shale that could contain and concen- Models for the genesis of giant porphyry Cu systems in the
trate a developing hydrothermal system, also occurs in the central Andean orogen (Zentilli et al., 1988; Clark, 1993; Zen-
lower transgressive sequence of the Machay Group where the tilli and Maksaev, 1995; Richards, 2000) postulate that rapid
shelly Pariahuanca Formation limestone is overlain by the ascent of magma is an important contributing factor in ore
marl of the Chulec Formation (Figs. 4 and 16). In addition, in formation because it would minimize modification through
northern Peru, similar prospective successions may occur at assimilation-fractional crystallization processes, which are en-
other levels within the upper transgressive sequence of the visaged to decrease the Cu content of melts. Further, an un-
Machay Group because at least six shallowing-upward marl to restrictive structure may be necessary to allow small bodies of
limestone sequences (Jaillard, 1987) are recognized in the felsic magma access to the upper crust, because their low ef-
five formations correlative with the Jumasha Formation. fective buoyancy would normally result in slow ascent, cool-
ing, and hence deeper crystallization. Thus, during the
Metallogenic and geotectonic implications magmatic lull, the basement structures controlling the Que-
Several other ore deposits, in addition to Antamina, coin- rococha arch may have provided the conduit necessary for
cide with deflections in the strike of the Maran thrust and rapid emplacement of small volumes of fertile melt into the
fold belt. At the northern extremity of the Cordillera Blanca upper crust, allowing the Antamina porphyry to crystallize in
(Fig. 1), 175 km to the north-northwest of Antamina, signifi- a suitably shallow environment favorable for mineralization.
cant middle to late Miocene intrusion-related mineral de- Therefore, we conclude that the confluence of fertile
posits, such as the Magistral Cu-Mo skarn prospect and the Miocene magmatism and reactive carbonate strata, essential
formerly productive Pasto Bueno W-Cu-Ag vein system (Fig. for ore genesis at Antamina, was afforded by the Querococha
1), are associated with the Casma-Pasto Bueno zone. Farther arch cross-strike structural discontinuity.
north, the Yanacocha Au district is localized in the northeast-
trending trans-Andean Chicama-Yanacocha structural corri- Conclusions
dor (Teal et al., 2002). Such large-scale structural controls on The world-class late Miocene Antamina skarn deposit is
the location of ore deposits in Peru have been suggested by hosted by deformed Upper Cretaceous carbonate strata of
Vidal and Noble (1994), Petersen and Vidal (1996), and the Machay Group. The relatively pure marbles and minor in-
Rivera (1996), but not at Antamina. Similarly, in the Ap- tercalated calc-hornfels that host the ore deposit at surface in
palachians, two of the four lateral ramps studied by Pohn the southwest and at depth in its northeast sector represent
(2000) are associated with an unusual abundance of mineral the upper part of the Jumasha Formation, whereas the strata
deposits, and many minor mineral occurrences are associated that form the ridges around Antamina and host the upper-
with another (Coleman, 1988a, b). northeast part of the deposit are assigned to the lower part of
The Machay Group has long been recognized as a metallo- the overlying, originally muddier, Celendn Formation (Fig.
genically important stratigraphic interval in central Peru (Pe- 18), here uncharacteristically resistant to erosion owing to
tersen, 1965). It hosts many skarn and carbonate replace- hornfels formation in proximity to the Antamina intrusive
ment-type deposits such as, from northwest to southeast, center. Marble is developed for tens of meters adjacent to the
Magistral, Antamina, Tuco-Chira, Pachapaqui, Raura, Uchuc- skarn in the Jumasha Formation, but distinctive calc-horn-
chacua, Chungar, Santander, Yauricocha, and Pursima Con- felses and skarnoids persist for hundreds of meters from the
cepcin. Equivalent Albian to Turonian carbonate rocks of skarn front in the overlying Celendn Formation (Fig. 18).
the Arcurquina and Ferrobamba Formations in south-central Around the Antamina deposit, both this and the 9 km2 swarm
Peru host Oligocene skarn deposits in the Andahuaylas-Yauri of Pb-Zn-Ag vein deposits provide much larger exploration
belt (Noble et al., 1984; Soler et al., 1986). targets than the skarn itself. Moreover, the systematic miner-
The Eocene to Miocene Calipuy Supergroup resulted alogic zonation from peripheral phlogopitic through
from suprasubduction zone magmatism (Noble et al., 1999). tremolitic to proximal diopsidic facies in the calc-hornfelses
The terminal event of this Supergroup is represented by the provides a vector toward the intrusion and, by extension, the
middle Miocene Carhuish pluton, dated at 13.7 Ma (U/Pb associated mineralization.
zircon date, Mukasa, 1984), and coincided with the forma- The sedimentary succession that hosts the Antamina de-
tion of many hydrothermal centers in the Cordillera Negra posit was folded, thrust-faulted, and thickened into a complex
such as the Pierina high-sulfidation epithermal Au-Ag de- thrust stack during the late Eocene Incaic orogeny, which
posit (Figs. 1 and 16; Strusievicz et al., 2000). Subsequently, generated the orogen-parallel Maran thrust and fold belt.
an approximately 5.5 m.y. hiatus in major igneous activity Within this stack, a left-stepping jog in the transverse Valley
preceded the late Miocene intrusion of the main phase of fault apparently controlled the Eocene formation of the Val-
the Cordillera Blanca batholith, the Cohup leucogranodior- ley lateral ramp, and together they localized the Miocene in-
ite, at 8.2 0.2 Ma (McNulty et al., 1998). During this pe- trusion and skarn. East-west diking east of the main intrusion
riod of relative magmatic quiescence, only minor volumes of and the northeast end of one segment of the Valley fault is in-
a tonalitic to dioritic marginal phase of the Cordillera Blanca terpreted as early and associated with north-south extension
batholith were intruded (Beckinsale et al., 1985), and only a related to east-west shortening. Relaxation of this compres-
few scattered hydrothermal centers developed in the sion allowed east-west extension, which would have been fo-
Cordillera Negra (Strusievicz et al., 2000). However, along cused in the left-stepping jog in the Valley fault, forming the

0361-0128/98/000/000-00 $6.00 913


914 LOVE ET AL.

locus for the main mass of intrusion and the associated skarn REFERENCES
ore. Renewed east-west shortening could have again induced Allende, T., 1996, Geologa del cuadrngulo de San Pedro de Chonta (18-j),
north-south extension and resulted in the late east-west vein Boletn 68, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero
system. The original intrusive contact unambiguously con- y Metalrgico, Sector Energa y Minas, Per, 218 p., 1 map, 1:100,000
scale.
trolled the location of the skarn, yet was itself controlled by Atherton, M.P., Pitcher, W.S., and Warden, V., 1983, The Mesozoic marginal
larger-scale structures. basin of central Peru: Nature, v. 305, no. 5932, p. 303306.
The Antamina hydrothermal activity occurred in a regional- Barazangi, M., and Isacks, B., 1976, Spatial distribution of earthquakes and
scale, northeast-trending, cross-strike structural discontinuity, subduction of the Nazca plate beneath South America: Geology, v. 4, p.
686692.
the Querococha arch, which articulates the Maran thrust Beckinsale, R.D., Snchez-Fernndez, A.W., Brook, M., Cobbing, E.J., Tay-
and fold belt. About 5 km southwest of Antamina, the locus of lor, W.P., and Moore, N.D., 1985, Rb-Sr whole-rock isochron and K-Ar age
this articulation steps left along strike, a feature mimicked on determinations for the Coastal batholith of Peru, in Pitcher, W.S., Atherton,
a smaller scale by the Valley fault. The arch is inferred to have 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.
affected sedimentary and magmatic processes at least from Benavides, V., 1956, Cretaceous System in northern Peru: Bulletin of the
the Jurassic to the Miocene and to have been controlled by a American Museum of Natural History, v. 108, p. 353494.
basement structure, perhaps a transform segment of the orig- 1999, Orogenic evolution of the Peruvian Andes: The Andean cycle, in
inal margin of the West Peruvian trough. The arch was the Skinner, B.J., ed., Geology and ore deposits of the central Andes: Society of
northwestern edge of a promontory on which the Cretaceous Economic Geologists Special Publication 7, p. 61107.
Bodenlos, A.J., and Ericksen, G.E., 1955, Lead-zinc deposits of Cordillera
Machay Group carbonate rocks were deposited farther west Blanca and northern Cordillera Huayhuash, Peru: United States Geologi-
than elsewhere along the belt. It also allowed Miocene mag- cal Survey Bulletin 1017, 102 p.
matism to extend toward the foreland and intrude the Broggi, J.A., 1942, Geologa del embalse del Ro Chotano en Lajas: Boletn
Machay Group. Sociedad Geolgica del Per, v. 12, p. 123.
Bussell, M.A., and Pitcher, W.S., 1985, The structural controls of batholith
We envisage that the carbonate rocks of the Machay Group emplacement, in Pitcher, W.S., Atherton, M.P., Cobbing, E.J., and Beckin-
provided both chemical and physical traps for ore-forming sale, R.D., eds., Magmatism at a plate edge: The Peruvian Andes: Glasgow,
fluids. Intense exoskarn developed in relatively pure Jumasha Blackie and Son Ltd., p. 167176.
Formation limestone, whereas the Celendn Formation horn- Clark, A.H., 1993, Are outsize porphyry copper deposits either anatomically
felses capped this system, promoting recirculation of hy- or environmentally distinctive?, in Whiting, B.H., Hodgson, C.J., and
Mason, R., eds., Giant ore deposits: Society of Economic Geologists Spe-
drothermal fluids and extensive endoskarn development. The cial Publication 2, p. 213283.
Querococha arch provided a suitable structure for the intru- Cobbing, E.J., Pitcher, W.S., Wilson, J.J., Baldock, J.W., Taylor, W.P., Mc-
sion to reach the Machay Group at the hypabyssal depths re- Court, W.J., and Snelling, N.J., 1981, The geology of the western Cordillera
quired for fertile fluid release. 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
Acknowledgments los cuadrngulos de Huaraz (20-h), Recuay (20-i), La Unin (20-j),
We thank Inmet Mining Corporation and Rio Algom Ltd., Chiquin (21-i), Yanahuanca (21-j), Boletn 76, Serie A: Carta Geolgica
Nacional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y
and especially Frank Balint and Kelly OConnor, for initial Minas, Per, 281 p., 5 maps, 1:100,000 scale.
support of this project, a contribution to the Queens Uni- Coleman, J.L., Jr., 1988a, CSDs of the eastern United States, in Coleman,
versity Central Andean Metallogenetic Project (QCAMP), J.L., Jr., Groshong, R.H., Jr., Rheams, K.F., Neathery, T.L., and Rheams,
and for repeatedly employing J.K.G. to examine various L.J., eds., Structure of the Wills Valley anticline-Lookout Mountain syn-
cline between the Rising Fawn and Anniston CSDs, northeast Alabama:
structural aspects of Antamina during 1997 and 1998. A post- Alabama Geological Society Annual Field Trip Guidebook 25, p. 4951.
doctoral fellowship at Queens University for the senior au- 1988b, Geology of the Anniston CSD, in Coleman, J.L., Jr., Groshong,
thor was funded in 1997 and 1998 by Inmet Mining Corpo- R.H., Jr., Rheams, K.F., Neathery, T.L., and Rheams, L.J., eds., Structure
ration and Rio Algom Ltd., and in 1998 through 2000 by Rio of the Wills Valley anticline-Lookout Mountain syncline between the Ris-
Algom. James Macdonald of BHP Billiton, Bill Mercer of ing Fawn and Anniston CSDs, northeast Alabama: Alabama Geological So-
ciety Annual Field Trip Guidebook 25, p. 4143.
Noranda, and John Thompson of Teck Cominco subse- Coney, P.J., 1971, Structural evolution of the Cordillera Huayhuash, Andes of
quently provided support and encouragement. This research Peru: Geological Society of America Bulletin, v. 82, p. 18631883.
was also funded by Natural Science and Engineering Re- Cosso, A., 1964, Geologa del cuadrngulos de Santiago de Chuco (17-g) y
search Council of Canada Discovery Grants to A.H.C. We Santa Rosa (18-g), Boletn 08, Serie A: Carta Geolgica Nacional: Instituto
Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, 69 p., 2
also thank Ca. Minera Antamina S.A. for unstinting logistic maps, 1:100,000 scale.
support and are grateful to the many geologists involved in Cosso, A., and Jan, H., 1967, Geologa de los cuadrngulos de Pumape
the mine development, including Leo Hathaway, Stewart (16-d), Chocope (16-e), Otuzco (16-f), Trujillo (17-e), Salaverry (17-f) y
Redwood, Manuel Pacheco, Jose Sales, Richard Ct, Diane Santa (18-f), 1967, Boletn 17, Serie A: Carta Geolgica Nacional: Instituto
Nicolson, Matt Wunder, Rick Schwarz, Jeff Hussey, Scott Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, 141 p., 4
maps, 1:100,000 scale.
Smith and, especially, Eric Lipten, for stimulating discus- Egeler, C.G., and De Booy, T., 1956, Geology and petrology of part of the
sions of mine geology and ore genesis. We would also like to southern Cordillera Blanca, Peru: Verhandelingen van het Koninklijk Ned-
thank the Economic Geology reviewers Gerry Ray, Greg erlands Geologisch Mijnbouwkundig Genootschap, Geologische Serie, pt.
Dipple, Larry Meinert, and Andreas Mueller, associate edi- 17, 86 p.
Einaudi, M.T., 1982a, Description of skarns associated with porphyry copper
tors Dave Cooke and Steve Garwin, and especially editor plutons: Southwestern North America, in Titley, S.R., ed., Advances in ge-
Mark Hannington for their large contributions to the prepa- ology of the porphyry copper deposits, Southwestern North America: Tus-
ration of the final manuscript. con, Arizona, University of Arizona Press, p. 139184.
1982b, General features and origins of skarns associated with porphyry
September 10, 2002; March 22, 2004 copper plutons: Southwestern North America, in Titley, S.R., ed., Advances

0361-0128/98/000/000-00 $6.00 914


LITHOSTRATIGRAPHY AND STRUCTURE, ANTAMINA COPPER-ZINC DEPOSIT, PERU 915

in geology of the porphyry copper deposits: Southwestern North America: transition in the Bagua basin, northern Peru: Paleontology, biostratigraphy,
Tuscon, Arizona, University of Arizona Press, p. 185210. radiometry, correlations: Newsletters on Stratigraphy, v. 19, p. 143177.
Einaudi, M.T., 2000, Skarns of the Yerington district, Nevada: A triplog and Mukasa, S.B., 1984, Comparative Pb isotope systematics and zircon U-Pb
commentary, in Dilles, J.H., Barton, M.D., Johnson, D.A., Proffett, J.M., geochronology for the Coastal, San Nicols and Cordillera Blanca
and Einaudi, M.T., eds., Contrasting styles of intrusion-associated hy- batholiths, Peru: Unpublished Ph.D. dissertation, University of California,
drothermal systems: Society of Economic Geologists Guidebook Series, v. Santa Barbara, 362 p.
32, pt. 1, p. 101125. Myers, J.S., 1974, Cretaceous stratigraphy and structure, western Andes of
Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits: Peru between latitudes 10 and 10 30': American Association of Petroleum
ECONOMIC GEOLOGY SEVENTY-FIFTH ANNIVERSARY VOLUME, p. 317391. Geologists Bulletin, v. 58, p. 474487.
Eskola, P., 1951, Around Pitkranta: Suomalaisen Tiedeakatemian Toimituk- 1975, Vertical crustal movements of the Andes in Peru: Nature, v. 254,
sia Sarja A (Annales Academi Scientiarum Fennic Series A), v. III, n. 27, p. 672674.
p. 196. 1976, Erosion surfaces and ignimbrite eruptions, measures of Andean
Gutscher, M.-A., Olivet, J.-L., Aslanian, D., Eissen, J.-P., and Maury, R., uplift in northern Peru: Geological Journal, v. 11, p. 2944.
1999, The lost Inca Plateau: Cause of flat subduction beneath Peru?: 1980, Geologa de los cuadrngulos de Huarmey (21-g) y Huayllapampa
Earth and Planetary Science Letters, v. 171, p. 335341. (21-h), Boletn 33, Serie A: Carta Geolgica Nacional: Instituto Geolgico,
Hampel, A., 2002, The migration history of the Nazca Ridge along the Peru- Minero y Metalrgico, Sector Energa y Minas, Per, 153 p., 1 map,
vian active margin: A re-evaluation: Earth and Planetary Science Letters, v. 1:100,000 scale.
203, p. 665679. Noble, D.C., McKee, E.H., and Mgard, F., 1979, Early Tertiary Incaic
Harrison, J.V., 1940, The geology of the central Andes in part of the Province tectonism, uplift and volcanic activity, Andes of central Peru: Geological
of Junn, Peru: Quarterly Journal of the Geological Society, London, v. 99, Society of America Bulletin, v. 90, p. 903907.
p. 136. Noble, D.C., McKee, E.H., Eyzaguirre, V.R., and Marocco, R., 1984, Age
INGEMMET, 1999, Mapa geolgico del Per: Instituto Geolgico, Minero and regional tectonic and metallogenetic implications of igneous activity
y Metalrgico, Sector Energa y Minas, Per 1:1,000,000, Digital Version and mineralization in the Andahuaylas-Yauri belt of southern Peru: ECO-
CD-ROM. NOMIC GEOLOGY, v. 79, 172176.
Jacay, J., 1996, Geologa del cuadrngulo de Singa (19-j), Boletn 67, Serie A: Noble, D.C., Wise, J.M., Vidal, C.E., 1999, Episodes of Cenozoic extension
Carta Geolgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sec- in the Andean orogen of Peru and their relation to compression, magmatic
tor Energa y Minas, Per, 215 p., 1 map, 1:100,000 scale. activity and mineralization, in Machar, J. Benavides, V., and Rosas, S.,
Jaillard, E., 1987, Sedimentary evolution of an active margin during middle eds., 75 Anniversario de la Sociedad Geologica del Peru, Julio, 1999: So-
and upper Cretaceous times: The north Peruvian margin from late Aptian ciedad Geologica del Peru, Volumen Jubilar 5, p. 4566.
up to Senonian: Geologische Rundschau, v. 76, p. 677697. Norabuena, E., Dixon, T., Stein, S., and Harrison, C.G.A., 1999, Decelerat-
Knopf, A., 1908, Geology of the Seward Peninsula tin deposits, Alaska: ing Nazca-South America and Nazca-Pacific motions: Geophysical Re-
United States Geological Survey Bulletin 358, 71 p. search Letters, v. 26, p. 34053408.
Kwak, T.A.P., and Askins, P.W., 1981, The nomenclature of carbonate re- OConnor, K., 2000, Yacimiento polimetlico Antamina: historia, exploracin
placement deposits, with emphasis on Sn-F(-Be-Zn) wrigglite skarns: y geologa, in primer volumen de monografas de yacimientos minerales
Journal of the Geological Society of Australia, v. 28, p. 123136. PeruanosHistoria, exploracin y geologa, volumen Luis Hochschild
Love, D.A., Clark, A.H., and Schwarz, F.P., 2000, The Antamina deposit, An- Plaut: Lima, Peru, Instituto de Ingenieros de Minas del Peru, p. 231244.
cash, Peru: Anatomy and petrology of a giant copper skarn [abs.]: Geologi- Pardo-Casas, F., and Molnar, P., 1987, Relative motion of the Nazca (Faral-
cal Society of America Abstracts with Programs, v. 32, no. 7, p. A137. lon) and South American plates since Late Cretaceous time: Tectonics, v. 6,
Love, D.A., Clark, A.H., Strusievicz, O.R., and Lee, J.K.W., 2001, The re- p. 233248.
gional tectonic setting of the giant Antamina Cu-Zn skarn deposit, north- Petersen, U., 1965, Regional geology and major ore deposits of central Peru:
central Peru [abs.]: Geological Society of America Abstracts with Programs, ECONOMIC GEOLOGY, v. 60, p. 407476.
v. 33, no. 6, p. A358. Petersen, U., and Vidal C., C.E., 1996, Magmatic and tectonic controls on
Love, D.A., Clark, A.H., Ullrich, T.D., Archibald, D.A., and Lee, J.K.W., the nature and distribution of copper deposits in Peru, in Camus, F., Silli-
2003, 40Ar-39Ar evidence for the age and duration of magmatic-hydrother- toe, R.S., and Peterson, R., eds., Andean copper deposits: New discoveries,
mal activity in the giant Antamina Cu-Zn skarn deposit, Ancash, north-cen- mineralization, styles and metallogeny: Society of Economic Geologists
tral Peru [abs.]: Geological Association of Canada/Mineralogical Associa- Special Publication 5, p. 118.
tion of Canada/Society of Economic Geologists Abstracts, v. 28, Abstract Pilger, R.H., 1981, Plate reconstructions, aseismic ridges, and low-angle sub-
no. 396 CD-ROM. duction beneath the Andes: Geological Society of America Bulletin, v. 92,
Manrique, A.I., 1998, Promocin de la minera del carbn en El Peru: Ge- p. 448456.
ologa economica de las cuencas de Alto Chicama, Santa, Oyn y Jatun- Pitcher, W.S., Atherton, M.P., Cobbing, E.J., and Beckinsale, R.D., eds.,
huasi: PROCARBON, Instituto Geolgico, Minero y Metalrgico, Sector 1985, Magmatism at a plate edge: The Peruvian Andes: Glasgow, Blackie
Energa y Minas, Per, 66 p. and Son Ltd., 328 p.
McKee, E.H., Noble, D.C., Scherkenbach, D.A., Drexler, J.W., Mendoza, J., Pohn, H.A., 2000, Lateral ramps in the folded Appalachians and in overthrust
and Eyzaguirre, V.R., 1979, Age of porphyry intrusion, potassic alteration, belts worldwideA fundamental element of thrust-belt architecture:
and related Cu-Zn skarn mineralization, Antamina district, northern Peru: United States Geological Survey Bulletin 2163, 71 p.
ECONOMIC GEOLOGY, v. 74, p. 928930. Redwood, S.D., 1998, The Antamina copper-zinc skarn deposit, northern
McLaughlin, D.H., 1924, Geology and physiography of the Peruvian Peru [abs.]: Geological Association of Canada/Mineralogical Association
cordillera, Departments of Junn and Lima: Geological Society of America of Canada/Canadian Geophysical Union Program with Abstracts, v. 23, p.
Bulletin, v. 35, p. 591632. 153.
McNulty, B.A., Farber, D.L., Wallace, G.S., Lopez, R., and Palacios, O., 1999, The geology of the Antamina copper-zinc skarn deposit, Peru: The
1998, Role of plate kinematics and plate-slip-vector partitioning in conti- Gangue, Newsletter of the Mineral Deposits Division, Geological Associa-
nental magmatic arcs: Evidence from the Cordillera Blanca, Peru: Geology, tion of Canada, issue 60, p. 17.
v. 26, p. 827830. Reyes, L., 1980, Geologa de los cuadrngulos de Cajamarca (15-f), San Mar-
Mgard, F., 1984, The Andean orogenic period and its major structures in cos (15-g) y Cajabamba (16-g), Boletn 31, Serie A: Carta Geolgica Na-
central and northern Peru: Journal of the Geological Society, London, v. cional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas,
141, p. 893900. Per, 70 p., 3 maps, 1:100,000 scale.
1987, Structure and evolution of the Peruvian Andes, in Schaer, J.-P., Richards, J.P., 2000, Lineaments revisited: Society of Economic Geologists
and Rodgers, J., eds., The anatomy of mountain ranges: Princeton, New Newsletter, 42, p. 1 and 1420.
Jersey, Princeton University Press, p. 179210. Rivera, J.N., 1996, El megafracturamiento pre-Mendaa de Casma-Pasto
Mourier, T., Bengtson, P., Bonhomme, M., Buge, E., Cappetta, H., Crochet, Bueno y su influencia en la metalognia Andina: Instituto de Ingenieros de
J.-Y., Feist, M., Hirsch, K.F., Jaillard, E., Laubacher, G., Lefranc, J.P., Moul- Minas del Per Revista Minera, v. 240, p. 612.
lade, M., Noblet, C., Pons, D., Rey, J., Sig, B., Tambareau, Y., and Taquet, Samam-Boggio, M., 1980, ed., El Per Minero: Instituto Geolgico, Minero
P., 1988, The Upper Cretaceous-lower Tertiary marine to continental y Metalrgico, Sector Energa y Minas, Lima, Per, 18 volumes.

0361-0128/98/000/000-00 $6.00 915


916 LOVE ET AL.

Snchez, A., 1995, Geologa de los cuadrngulos de Culebras (20-g), Casma Vidal, C.E., and Noble, D.C., 1994, Yacimientos hidrotermales controlados
(19-g), Chimbote (19-f), Boletn 59, Serie A: Carta Geolgica Nacional: In- por estructura y magmatismo en la region central del Per: Instituto de In-
stituto Geolgico, Minero y Metalrgico, Sector Energa y Minas, Per, genieros de Minas del Per Revista Minera, v. 230, p. 1619.
263 p., 3 maps, 1:100,000 scale. Wheeler, R.L., 1978, Cross-strike structural discontinuities, possible explo-
Snchez, J., lvarez, D., and Lagos, A., 1998, Geologa de los cuadrngulos ration tool in detached forelands [abs.]: Geological Society of America,
de Juscusbamba (16-i) y Plvora (16-j), Boletn 119, Serie A: Carta Ge- Southeastern Section Abstracts with Programs, v. 10, no. 4, p. 201.
olgica Nacional: Instituto Geolgico, Minero y Metalrgico, Sector En- Wilson, J.J., 1963, Cretaceous stratigraphy of central Andes of Peru: Ameri-
erga y Minas, Per, 262 p., 2 maps, 1:100,000 scale. can Association of Petroleum Geologists Bulletin, v. 47, p. 133.
Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe, Wilson, J.J., and Reyes, L., 1964, Geologa del cuadrngulo de Pataz, Boletn
O., Neeraudeau, D., Crdenas, J., Rosas, S., and Jimnez, N., 2002, Late 09, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y Met-
Permian-Middle Jurassic lithospheric thinning in Peru and Bolivia, and its alrgico, Sector Energa y Minas, Per, 91 p., 1 map, 1:100,000 scale.
bearing on Andean-age tectonics: Tectonophysics, v. 345, p. 153181. Wilson, J.J., Reyes, L., and Garayar, J., 1967, Geologa de los cuadrngulos de
Soler, P., Grandin, G., and Fornari, M., 1986, Essai de synthse sur la mtal- Pallasca (17-h), Tayabamba (17-i), Corongo (18-h), Pomabamba (18-i),
logenie du Prou: Godynamique, v. 1, p. 3368. Carhuaz (19-h) y Huari (19-i), Boletn 16, Serie A: Carta Geolgica Na-
Somoza, R., 1998, Updated Nazca (Farallon)-South America relative motions cional: Instituto Geolgico, Minero y Metalrgico, Sector Energa y Minas,
during the last 40 m.y.: Implications for mountain building in the central Per, 95 p., 6 maps, 1:100,000 scale.
Andean region: Journal of South American Earth Sciences, v. 11, p. 1995, Geologa de los cuadrngulos de Pallasca (17-h), Tayabamba (17-
211215. i), Corongo (18-h), Pomabamba (18-i), Carhuaz (19-h) y Huari (19-i), Bo-
Strusievicz, O.R., Clark, A.H., Lee, J.K.W., Farrar, E., Slauenwhite, M., and letn 60, Serie A: Carta Geolgica Nacional: Instituto Geolgico, Minero y
Hodgson, C.J., 2000, Metallogenetic relationships of the Huaraz, Ancash, Metalrgico, Sector Energa y Minas, Per, 63 p., 6 maps, 1:100,000 scale.
segment of the precious-base metal subprovince of northern Peru [abs.]: Zentilli, M., and Maksaev, V., 1995, Metallogenetic model for the late
Geological Society of America Abstracts with Programs, v. 32, no. 7, p. Eocene-early Oligocene supergiant porphyry event, northern Chile, in
A504. Clark, A.H., ed., Giant ore deposits II: Controls on the scale of orogenic
Szekely, T.S., 1967, Geology near Huallacocha Lakes, central high Andes, magmatic-hydrothermal mineralization, April 1995, Queens University,
Peru: American Association of Petroleum Geologists Bulletin, v. 51, p. Kingston, Ont., Giant Ore Deposits Workshop, 2nd, Proceedingssecond
13461353. (corrected) printing, 1996, p. 152165.
Teal, L., Harvey, B., Williams, C., and Goldie, M., 2002, Geology of the Yana- Zentilli, M., Doe, B.R., Hedge, C.E., Alverez, O., Tidy, E., and Daroca, J.A.,
cocha gold deposits, northern Peru [extended abs.], in Marsh, E.E., Gold- 1988, Istopos de plomo en yacimientos de tipo prfido cuprfero com-
farb, R.J., and Day, W.C., eds., Global exploration 2002: Integrated meth- parados con otros depsitos metalferos en los Andes del norte de Chile y
ods for discovery: Denver, Colorado, Society of Economic Geologists, Argentina (English abstract): Actas, V Congreso Geolgico Chileno, Santi-
Abstracts, p. 4344. ago, v. 1, p. B331369.
Terrones, A.J., 1958, Structural control of contact metasomatic deposits in Zharikov, V.A., 1970, Skarns: International Geology Review, v. 12, p. 541559,
the Peruvian cordillera: Mining Engineering, Transactions of the American 619647, 760775.
Institute of Mining Engineers, v. 11, p. 365372. Zuzunaga, A., 2003, Closing the loopUsing actual concentrator perfor-
Thomas, W.A., 1977, Evolution of Appalachian-Ouachita salients and re- mance to determine the true value of ore sources [abs.]: Canadian Institute
cesses from reentrants and promontories in the continental margin: Amer- of Mining, Metallurgy and Petroleum, Annual Meeting, 105th, Montreal,
ican Journal of Science, v. 277, p. 12331278. Quebec, May 57, 2003, Program with Abstracts, p. 170.
Tracy, R.J., and Frost, B.R., 1991, Phase equilibria and thermobarometry of
calcareous, ultramafic and mafic rocks, and iron formations, in Kerrick,
D.M., ed., Contact metamorphism: Reviews in Mineralogy, v. 26, p.
207289.

0361-0128/98/000/000-00 $6.00 916

You might also like