Anda di halaman 1dari 30

2010 Society of Economic Geologists, Inc.

Economic Geology, v. 105, pp. 12711299

The Hypogene Iron Oxide Copper-Gold Mineralization in


the Mantoverde District, Northern Chile
ANA A. RIEGER,1, ROBERT MARSCHIK,1, MANUEL DAZ,2,* STEFAN HLZL,3 MASSIMO CHIARADIA,4 BUR AKKER,1
AND JORGE E. SPANGENBERG5
1Resource

Geology Research Group, Department of Earth and Environmental Sciences, Ludwig-Maximilians Universitt,
Luisenstrasse 37, 80333 Munich, Germany
2AngloAmerican

3Bayerische
4Section

Chile, Mantoverde Division, Santiago de Chile, Chile

Staatssammlung fr Palontologie und Geologie, Richard-Wagner-Str. 10, 80333 Munich, Germany

des Sciences de la Terre, Universit de Genve, Rue des Marachers 13, 1205 Genve, Switzerland

5Institut

de Minralogie et Gochimie, Universit de Lausanne, Anthropole, 1015 Lausanne, Switzerland

Abstract
The Mantoverde iron oxide copper-gold (IOCG) district, northern Chile, is known for its Cu production
from supergene ores. Recently, exploration outlined an additional hypogene ore resource of 440 Mt with 0.56
percent Cu, and 0.12 g/t Au. The hypogene sulfide mineralization occurs mainly as chalcopyrite and pyrite, typically in specularite or magnetite-cemented breccias and associated stockworks. The host rocks underwent variably intense K feldspar alteration, chloritization, sericitization, silicification, and/or carbonatization. A district
scale Na(-Ca) alteration is absent. The IOCG mineralization in the district shows a strong tectonic control by
northwest- to north-northwesttrending brittle structures. Large Cu sulfide-rich veins or Cu sulfide-cemented
breccias are absent. Therefore, head grades of 4 percent Cu are an exception. There is a positive correlation
between Cu and Au grades. Gold is probably contained mostly in chalcopyrite and pyrite. Elevated concentrations of light rare-earth elements (LREE) occur locally but are attributed to redistribution of LREE within the
deposits rather than to derivation from external sources. The Cu-Au ores in the Mantoverde district are low in
U and have relatively low contents in heavy metals that are potentially hazardous to the environment, such as
As (avg 14 ppm), Hg (<5 ppm), or Cd (<0.2 ppm). The sulfur isotope ratios of chalcopyrite from the IOCG deposits lie between 5.6 and 8.9 per mil 34SVCDT. They show systematic variations within the district, which are
interpreted to reflect relative distance to inferred fluid conduits and the level of deposition within the hydrothermal system. Most initial 87Sr/86Sr values of altered volcanic rocks and hydrothermal calcite from the
Mantoverde district are between 0.7031 and 0.7060 and are similar to those of the igneous rocks of the region.
Lead isotope ratios of chalcopyrite are consistent with Pb (and by inference Cu) derived from Early Cretaceous
magmatism. The sulfur, strontium, and lead isotope data of chalcopyrite, calcite gangue, or altered host rocks,
respectively, are compatible with a genetic model that involves cooling of metal and sulfur-bearing magmatichydrothermal fluids that mix with meteoric waters or seawater at relatively shallow crustal levels. An additional
exotic sulfur input is likely, though not required, for the copper mineralization. Apart from the IOCG deposits,
there are a number of smaller magnetite(-apatite) bodies in the district. These are geologically similar to the
Cu-Aubearing magnetite bodies, but are related to splays of the north-southtrending Atacama fault zone and
differ in alteration and texture.

Introduction
THE MANTOVERDE iron oxide Cu-Au (IOCG) district is located about 50 km southeast of the Chaaral harbor and 100
km north of the city of Copiap, northern Chile (Fig. 1). It
has become one of the most important IOCG districts worldwide. The first documented workings date back to the 18th
century. After a period of mining between 1906 and 1938 and
diverse exploration campaigns from the mid 1950s to early
the 1980s by various companies, the Empresa Minera de
Mantos Blancos, S.A. (AngloAmerican) started exploring the
site in 1989 and commenced mining in 1996. Reported initial
reserves were 85 Mt at 0.82 percent Cu of supergene Cu-Au
ores (Vila et al., 1996). Since then, district exploration has
identified additional resources that have expanded the mine
life to 2014 (Fig. 2). Currently, the annual production is 14.5
Corresponding authors: e-mail, ARieger@lmu.de; marschik@lmu.de
*Present address: AngloAmerican Chile, El Soldado Division, Santiago de
Chile, Chile.

0361-0128/10/3921/1271-29

Mt of ore (oxide heap and dump) with an output of 60,000 t


cathode copper. The geologic resources estimated from production (19962007), proven ore reserves, and resources add
up to 410 Mt with 0.58 percent total Cu for the oxide heap
and dump operation (AngloAmerican Mantoverde Division,
unpub. internal report, 2007). Preliminary estimates of the
hypogene copper sulfide and gold ores suggest resources of
440 Mt with 0.56 percent Cu, and 0.12 g/t Au at a 0.20 percent Cu cutoff (AngloAmerican Mantoverde Division internal
report, 2007). The feasibility of these ores is currently being
investigated.
The Mantoverde district has been subject to several studies. Orrego and Zamora (1991) described characteristics of
the Manto Ruso ore zone. Vila et al. (1996) gave the first
overview about the geology of the district and presented a genetic model for mineralization in the Mantoverde Norte area.
Their work summarizes microthermometric and geochronologic data of unpublished company reports. Investigations by
Cornejo et al. (2000) focused on the ore-related alteration

1271

Submitted: August 10, 2009


Accepted: June 30, 2010

1272

RIEGER ET AL.
70

26

69

a Fault Zone
Atacam

71

25

Chanaral

Copiap
CandelariaPunta del Cobre

28

Argenti
na

27

Pacific
Ocean

Mantoverde

50 km

Pliocene to Recent

Lower Cretaceous

Upper Cretaceous to Miocene

Pre-Cretaceous

FIG. 1. Schematic geologic map of the area around Copiap and Chaaral,
northern Chile. The location of the two major IOCG districts, Mantoverde
and Candelaria-Punta del Cobre, are indicated.
= mine.

and paragenetic relationships, whereas Sanhueza et al. (2000)


studied the relationships of magmatism, tectonism, and mineralization. Zamora and Castillo (2001) established a geologic
model integrating new findings in the Mantoverde Sur and
Manto Ruso areas. More recently, Benavides et al. (2007)
summarized the geology of the Mantoverde district, compiled
a paragenetic sequence, and discussed new stable isotope
data. They emphasized the roles of nonmagmatic hydrothermal fluids and evaporite-derived sulfur in the formation of
IOCG-type mineralization. Their conclusions seem to be biased by observations and data from the more distal, hematiterich portions of the hydrothermal system.
In this contribution, we give an update on the newest developments in the Mantoverde district. Extensive exploration
drilling in the mine areas during recent years also enables investigations of the hypogene, magnetite-rich ores, which allow
conclusions relevant to exploration, mining, and for genetic
interpretations. We describe the distribution and characteristics of the hypogene IOCG mineralization based mainly on
the geochemical composition of representative composite
core samples from within the orebodies and present a new
paragenetic sequence. We infer fluid, metal, and sulfur
sources based on new strontium, lead, and complementary
sulfur isotope data. Although we agree on a fluid-mixing
model, which is widely accepted for many larger IOCG deposits (e.g., Marschik and Fontbot, 2001; Oliver et al., 2004;
Chiaradia et al., 2006; Simard et al., 2006; Baker et al., 2008;
Kendrick et al., 2008; Monteiro et al., 2008; De Haller and
Fontbot, 2009; Kolb et al., 2009) and implied by the work of
Benavides et al. (2007), we provide an alternate interpretation as favored by the latter authors regarding the role of nonmagmatic hydrothermal fluid and sulfur components.
0361-0128/98/000/000-00 $6.00

District Geology
The Mantoverde IOCG district forms part of the Coastal
Cordillera of northern Chile (Figs. 1, 3). The area represents
a Jurassic-Early Cretaceous continental magmatic arc environment related to the subduction of the Aluk plate under the
South American continent (e.g., Scheuber and Andriessen,
1990). The arc and an associated back-arc basin developed on
a late Paleozoic to Triassic basement (e.g., Mpodozis and
Ramos, 1990). The arc-derived rocks in the Mantoverde district are mainly andesitic volcaniclastic conglomerates, breccias, or flows of probable Late Jurassic age, which are correlated with the Jurassic La Negra Formation (Lara and Godoy,
1998). The volcanic rocks are intruded by Cretaceous granitoids of the Chilean Coastal batholith. Granodiorites and
monzonites of the 125 to 130 Ma Las Tazas complex (Berg
and Breitkreuz, 1983; Brown et al., 1993; Wilson et al., 2000)
are exposed in the west of the Mantoverde district, whereas
granodioritic and quartz dioritic rocks of the 120 to 127 Ma
Sierra Dieciocho, and diorites, monzodiorites, granodiorites,
and tonalites of the 90 to 110 Ma Sierra Merceditas-Remolino
plutonic complexes occur to the east (Zentilli, 1974; Berg and
Breitkreuz, 1983; Dallmeyer et al., 1996; Lara and Godoy,
1998; Fig. 3). To the south, quartz-monzodiorites, pyroxene
granodiorites, and amphibole-biotite-granodiorites of the 130
to135 Ma Cerro Morado complex crop out (Lara and Godoy,
1998). The Cerro Morado complex includes lavas of the La
Negra Formation as hornfelsed roof pendants (Gelcich et al.,
2005). Intrusive stocks and dikes are common at depth in the
mineralized areas. The latter crop out in many places of the
district.
The Mantoverde district represents a structural block that
is limited to the east and west by north-southtrending
branches of the Atacama fault zone (Figs. 2, 3), which is a
subduction-related, arc-parallel, strike-slip fault system that
stretches over 1,000 km along the Chilean coast (e.g.,
Scheuber and Andriessen, 1990). The district is divided into
a northeastern and a southwestern tectonic wedge by an approximately 12-km-long, N 15 to 20 W-trending, 40 to 50
E-dipping major brittle fault, the Mantoverde fault, which
connects the two branches of the Atacama fault zone (Figs.
24). The Mantoverde fault and several minor northwest
faults initially developed as left lateral strike-slip structures
related to a duplex evolution of the Atacama fault zone
(Brown et al., 1993; Sanhueza and Robles, unpub. internal report, 1999). Later extension caused (1) the reactivation of the
Mantoverde fault as a normal scissor fault; (2) the development of associated, curved north-northwest to north-south
normal faults (Sanhueza and Robles, unpub. internal report,
1999); and (3) tilting of the northeastern part of the district
with a northeastern-down sense (Orrego and Zamora, 1991;
Sanhueza and Robles, unpub. internal report, 1999; Fig. 4).
Ore Distribution
The Cu-Au orebodies within the district are controlled
mainly by north-northwest to northwest brittle faults. The
two largest ore zones, Mantoverde Norte and Mantoverde
Sur, occur along the central part of the Mantoverde fault
(Figs. 2, 4), and new exploration drilling has shown they may
be physically continuous with one another. The Laura and
Kuroki orebodies are located along the Mantoverde fault

1272

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

Fault
La

sA

nim
as

Gravel deposit (Miocene)


Magmatic rocks (Jurassic to Lower Cretaceous)

Celso (Fe)

Va
lley

Celso
(Cu)
Manto Ruso
Punto 62

Kuroki

Sierra Dieciocho pluton (120-127 Ma)

AFZ oriental branch

N 7066

Alluvium (Recent)

rde
tove
Ma n

AFZ central branch

Legend

Las Tazas pluton (125-130 Ma)


Cerro Morado pluton (130-135 Ma)
La Negra Formation (Late Jurassic)
Alteration and Mineralization units
Breccia Verde (Breccia Temprana)

Rebosadero
Sur (Fe)

Laura

K feldspar-quartz rock
Argillic alteration
Calcite vein or breccia

N 7064

Tectonic breccia (Mantoverde Breccia)


FerrferaSanta Clara
(Fe)

Specularite breccia (Manto Atacama)


Specularite stockwork (Transition Zone)
Magnetite-rich rock
Fault, observed, inferred

Mantoverde Norte

Mine in operation
Project
Mine closed

N 7062

500 m
ley
Val
de
Ver
a
L

Mantoverde Sur
Gu

ista

or

sect

ga

an

am

Altav

lle
Va

N 7060

risto

tec
Mon

or

sect

UTM

inate
Coord

)
s (km

Franko

N 7058

San Juan
E 368

E 370

UTM Coordinates (km)

E 372

FIG. 2. Geologic map of the Mantoverde district (note: the Manto Monstruo mine mentioned in the text is located outside the area covered by this map). AFZ = Atacama fault zone.
0361-0128/98/000/000-00 $6.00

1273

1273

1274

RIEGER ET AL.
7015W Geographic coordinate

7025W
AFZ central bra
nc

7080 DCce

2625S
Geographic coordinate

AFZ oriental branch

7070
Jgla

Jln

Kgsm

MVF

130.1+1.9
130.2+0.6

Legend

Jigf

Sedimentary deposit

130.3+1.3
128.7+2.8

153+1.0

Plutonic Complexes
Sierra Merceditas (Kgsm) ca. 90-110 Ma

126.8+1.3
123.7+2.7
126.8+0.5

7060

Remolino (Kgr) ca. 90-110 Ma


Sierra Dieciocho (Kgsd) ca. 120-127 Ma
Las Tazas (Kglt) ca. 125-130 Ma

120+4
115+3

Kglt

Cerro Morado (Kgm) ca. 130-135 Ma


Cerro Moradito (JKgm) ca. 140-145 Ma

125.7+0.6

Las Animas (Jgla) ca. 150-160 Ma


Flamenco (Jigf) ca. 190-200 Ma

7050

UTM Coordinates (km)

La Negra Fm (Jln) Jurassic


Chaaral Metamorphic Complex (DCce)
Devon-Carbon

Kgsd
Kgr

Fault, observed, inferred


130.1+1.9

Kgm

25 km

102.3+1.5
2645S

JKgm

7040

360

Radiogenic age (Ma)

UTM Coordinates (km)

370

380

FIG. 3. Geologic map of the larger Mantoverde area (modified after Lara and Godoy, 1998). Locations and radiometric
ages of major plutonic complexes (see text for references) and position of the Mantoverde district, as shown in Figure 2, are
indicated. AFZ = Atacama fault zone; MVF = Mantoverde fault.

AF

ch

VF

ran

al b

tr
cen

M
Kuroki (Cu)

Mantoverde
Norte (Cu)
Mantoverde
Sur (Cu)
Franko (Cu)

Z h
AF anc
br
tal
n
ie

Manto Ruso
(Cu)
Celso
(Cu)
or

Celso
(Fe)
Rebosadero
Sur (Fe)
FerrferaSanta Clara
(Fe)

Legend
Mine in operation
Project
Mine closed

FIG. 4. Structural model of the Mantoverde district (based on Orrego and


Zamora, 1991, and Sanhueza and Robles, unpub. internal report, 1999). AFZ
= Atacama fault zone; MVF = Mantoverde fault. Gray shades and white horizon are used to show the displacement of the northeastern part of the district.
There is no lithostratigraphic meaning.
0361-0128/98/000/000-00 $6.00

toward the north, whereas the Franko ore zone is found to


the south of the two Mantoverde pits (Fig. 2). Other Cu-Au
bodies, unrelated to the Mantoverde fault, are the Punto 62
ore zone, the old Manto Ruso mine, the Celso project, and
Manto Monstruo mine in the northeast of the district. A small
orebody in the southwest of the district, the San Juan Cu
mine, is situated to the west of the Mantoverde fault (Fig. 2).
The Cu production in the district came from near surface,
supergene, leachable ores. The oxidation level commonly
reaches between 200 and 250 m below the surface but may
lie deeper close to major faults (Fig. 5). Beneath this level, a
poorly developed enrichment zone is found, characterized
mainly by chalcocite and covellite. Hypogene Cu mineralization at depth is present mainly as chalcopyrite. The associated alteration is very similar in all the Cu-Au orebodies and
characterized by variably intense K feldspar alteration, chloritization, sericitization, silicification, and/or carbonatisation.
Particularly, K alteration is a characteristic feature of the
IOCG deposits. Massive K feldspar-quartz altered zones are
locally exposed in the Mantoverde Norte-Sur and Manto
Ruso sectors. These zones commonly contain sulfides and, in

1274

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

1275

1000
900

(masl)

900
800

800
700

700
600

Manto Ruso: Section 103140 N

100 m

100 m

Celso: Section 103550 N

c
1100

1000
M
VF

Legend

900

Alluvium (Recent)
Dioritic dike

(masl)

800

Sierra Dieciocho pluton


(120-127 Ma)
700

Las Tazas pluton


(125-130 Ma)
Cerro Morado pluton
(130-135 Ma)

600

Andesite, andesitic
breccia (Late Jurassic)
500

Alteration and Mineralization unit


Breccia Verde
(Breccia Temprana)

400

K feldspar-quartz rock
Argillic alteration

300

Calcite vein or breccia


100 m

Mantoverde Norte: Section 100100 N

Tectonic breccia
(Mantoverde Breccia)
Specularite breccia
(Manto Atacama)

d
M
VF

Specularite stockwork
(Transition Zone)

1000

Magnetite Zone

900

Sulfide ore body

Fault, observed, inferred


Oxidation level
Drill hole
800

(masl)

Composite sample

700

600

500

100 m

Mantoverde Sur: Section 98450 N

FIG. 5. Representative geologic sections through four major ore zones in the Mantoverde district (from north to south):
a) Manto Ruso; b) the Celso project; c) Mantoverde Norte; and d) Mantoverde Sur. MVF = Mantoverde fault; asl = above
sea level.
0361-0128/98/000/000-00 $6.00

1275

1276

RIEGER ET AL.

places, have experienced brecciation, forming Cu-Aubearing specularite breccias with pervasively K feldspar-quartz
rich fragments.
In addition to the IOCG ores, the district hosts magnetite
deposits such as Rebosadero Sur and Ferrfera (Santa Clara
of Vila et al., 1996) which have been mined in the past (e.g.,
Espinoza et al., 1999; Fig. 2). These iron orebodies are hosted
in splays of the eastern branch of the Atacama fault zone and
are correlated with similar magnetite(-apatite) deposits of the
Chilean iron belt, which occur along the Atacama fault zone
between La Serena and Taltal. The hydrothermal alteration
associated with the magnetite(-apatite) bodies in the Mantoverde district is manifested mainly as variable proportions
of sericite, chlorite, carbonate, and quartz. Copper sulfides
are absent or scarce, whereas minor pyrite is present.
The Mantoverde Norte deposit
The Mantoverde Norte orebody was the first from which
supergene Cu-Au ores were extracted (total reserves of 101.6
Mt leachable copper ores). It is located in the middle part of
the Mantoverde fault and the ores show a close spatial relationship to the fault (Figs. 2, 5c; see also Vila et al., 1996).
Three main ore-bearing geologic units were distinguished at
the beginning of production: (1) a specularite-cemented, hydrothermal breccia, termed Manto Atacama, in the hanging
wall of the Mantoverde fault, that grades into; (2) a specularite stockwork zone, called the Transition zone, to the east; and
(3) a tectonic breccia (Mantoverde Breccia), which is located
in the footwall of the main fault plane of the Mantoverde fault
(Fig. 5c). The Manto Atacama, Mantoverde Breccia, and
Transition zone are parallel to the Mantoverde fault. To the
east of the Transition zone and to the west of the Mantoverde
Breccia, another hydrothermal breccia, the Breccia Verde, occurs in altered andesitic or granitoid rocks (Fig. 5c; Vila et al.,
1996). Vila et al. (1996) note that an increase in cataclastic deformation, along with an increase in Cu contents, occurs as the
Mantoverde fault is approached. Exploration drilling in 2000
and 2001 penetrated a sulfide-bearing, magnetite-rich zone in
the footwall of the Mantoverde fault, below the current bottom of the open pit. Copper-bearing magnetite near the Mantoverde fault was previously only known to occur in the Mantoverde Sur area (Montecristo and Altavista: Vila et al., 1996;
Zamora and Castillo, 2001).
The Manto Atacama: The Manto Atacama is a roughly tabular shaped hydrothermal breccia with a thickness that averages 80 m but reach up to 200 m. The breccia is composed of
subangular to subrounded fragments, mainly of andesite or
granitoid igneous rocks in a mineralized, calcite-bearing,
coarse-grained specularite matrix. The rock fragments range
from a few millimeters up to about 35 cm in diameter. They
are commonly affected by variable degrees of pervasive K
feldspar alteration with more or less intense chloritization,
sericitization, silicification, and/or carbonatization. Locally,
tourmaline, minor sphene, or scapolite may be present.
Below the oxidation zone (below mining level 750), the specularite-rich matrix contains pyrite-chalcopyrite. Digenite and
bornite have replaced chalcopyrite locally. The rock is cut
mainly by veinlets of K feldspar quartz, tourmaline, or
sericite. Postdating these, calcite veinlets and specularite
veinlets are manifestations of the latest hydrothermal activity.
0361-0128/98/000/000-00 $6.00

The Transition zone: The specularite stockwork of the Transition zone contains supergene copper oxide and hypogene
sulfides (Fig. 5c). The stockwork developed in both La Negra
volcanic rocks and the Breccia Verde. The wall-rock alteration
is essentially the same as in the Manto Atacama. Minor
phases include tourmaline, sphene, and apatite.
The Mantoverde Breccia: The Mantoverde Breccia is a cataclastite rock with a rock flour matrix (Vila et al., 1996; Fig. 5c).
Its thickness is commonly between 20 and 40 m, but can reach
up to 150 m in La Verde Valley, which lies between the Mantoverde Norte and Mantoverde Sur areas. In the Mantoverde
Norte pit, the rock is composed of angular andesite and/or
diorite fragments that commonly are a few mm to about 10 cm
in diameter, and show silicification, chlorite, and minor to moderate K feldspar alteration. The alteration mineralogy in the
matrix includes chlorite, quartz, calcite, K feldspar, and clay
minerals. The rock is cut by specularite, K feldspar quartz,
and calcite veins and veinlets. The calcite veins may contain
fragments of subangular magnetite or chalcopyrite.
The Breccia Verde: The Breccia Verde is composed of silicified volcanic and dioritic fragments in a chlorite-quartzrich
matrix, which also contains calcite, subordinate sericite, and,
locally, zeolite group minerals (chabasite-phillipsite) and
small amounts of tourmaline (Fig. 5c). The rock is cut by low
to moderately dense sets of K feldspar quartz, calcite
siderite, specularite quartz, sericite veinlets. Rock fragments show pervasive K feldspar alteration. Toward the surrounding country rocks, the Breccia Verde grades into a chlorite-quartzrich stockwork. The Breccia Verde has been
considered mostly barren, though it can locally contain significant grades of Cu-Au. Mineralization is normally related to
specularite veinlets.
The Magnetite zone: At Mantoverde Norte, the Magnetite
zone occurs below 680 m above sea level (a.s.l) and is open toward depth (Fig. 5c). It developed between the Breccia Verde,
an intrusion of the 120 to 127 Ma Sierra Dieciocho pluton to
the west, and the Mantoverde Breccia to the east (Fig. 5c). The
rocks of the Magnetite zone are magnetite-chlorite-sericite-K
feldsparcemented breccias, with igneous rock fragments that
are altered mainly by magnetite, K feldspar, and quartz. Magnetite and mushketovite (the pseudomorphous replacement of
an earlier specular hematite by magnetite) are observed. The
rocks are cut by K feldspar quartz, calcite, sericite, and late
specularite-calcite veinlets. Pyrite and chalcopyrite are present
as small patches, disseminations, or in discontinuous veinlets.
The Mantoverde Sur deposit
The Mantoverde Sur sector (including the former Altavista
and Montecristo mines) contains reserves of 46.8 Mt Cu of
leachable copper ores. Overall, the mineralization does not
show such a strong relationship with the Mantoverde fault,
and some rock types differ from those in Mantoverde Norte.
The Magnetite zone is characteristic of Mantoverde Sur (Fig.
5d). Two mineralization styles are distinguished: (1) magnetite stockworks and disseminations; and (2) elongate magnetite-cemented breccia or massive magnetite bodies. The
magnetite-rich rocks may be barren or may contain chalcopyrite (and/or pyrite), mainly as disseminations and discontinuous veinlets. The magnetite zone is overlain or enveloped by
pervasive argillic alteration, respectively, in the Altavista and

1276

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

Montecristo sectors (Fig. 2). The argillic zone stretches from


the present surface to the bottom of the oxidation zone. Astudillo (unpub. internal report, 2006) notes that the Cu
grades (Cu oxides) are lower in the argillic zone than in adjacent magnetite-rich rocks.
In the northern, shallow part of the Mantoverde Sur area
are rock types sufficiently similar to those of Mantoverde
Norte that they can be included with the previously defined
units (Manto Atacama, Mantoverde Breccia, and Breccia
Verde). The Manto Atacama is only poorly developed and
gives way to magnetite-rich rocks in the hanging wall, footwall
and toward depth (Fig. 5d). The Transition zone is essentially
absent. The style of mineralization changes from the northern
part of Mantoverde Sur southward to a point where magnetite-rich rocks (Magnetite zone; Altavista and Montecristo
sectors) dominate over specularite-rich rocks. In the Altavista
sector, located about 2 km south of Mantoverde Norte, the
magnetite-rich orebody (Cu oxides) is 400 m long and 300 m
thick and controlled by northwest-trending structures (Fig. 2;
Astudillo, unpub. internal report, 2006). The orebody in the
Montecristo sector is 500 m long and 200 m wide. It follows
a northeast structural trend and dips 50 to the north. The
thickness of the oxidation zone varies between 120 and 250 m
(bottom of oxide ores roughly at 800 m a.s.l.). The Mantoverde Breccia and the Breccia Verde are both present in the
entire Mantoverde Sur area. The Mantoverde Breccia in this
area is composed of fragments of silicified and/or chloritized
andesite or diorite rocks or magnetite in a quartz-chlorite-calcite sericite-rich matrix. In the intermediate levels (~800
650 m), the Mantoverde Breccia is interspersed with massive
magnetite bodies (Fig. 5d). At deeper levels, quartz plus
traces of scapolite and patchy K feldspar alteration are observed locally (640630 m). The Breccia Verde occurs in the
footwall and hanging wall of the Mantoverde fault (Fig. 5d).
It contains fragments of andesite or diorite, showing a pervasive chlorite-quartz, K feldspar and/or sericite tourmaline
alteration, in a quartz-chlorite-K feldsparrich matrix. Parts
of the Breccia Verde are mineralized. Chalcopyrite and/or
pyrite are present as aggregates in the matrix or in discontinuous veinlets. Accessory allanite and sphene occur locally.
The alteration minerals in veinlets of all units at Mantoverde
Sur, except the argillic zone, are very similar to those described above for the Mantoverde Norte deposit.
The Manto Ruso deposit and the Celso project
Manto Ruso was mined in the past on a small scale. The
area was reconsidered for mining in the 1980s (Orrego and
Zamora, 1991) and subsequent exploration activities identified reserves of 15.1 Mt of supergene, leachable copper ores.
They occur as a 400-m-long, 100-m-wide, N70 to 80E
trending body, which is covered by unconsolidated material to
the south (Fig. 5a). The thickness of the oxidation zone is variable and lies between 40 to 150 m below the surface level
(800850 m a.s.l.). Hypogene Cu ore is present at depth. The
Manto Ruso deposit is hosted in andesitic volcanic-volcaniclastic rocks and diorite of the Sierra Dieciocho complex; the
latter crop out immediately to the east of the deposit (Fig.
5a). The orebody is a subvertical to vertical, specularite-cemented hydrothermal breccia, which is similar to the Manto
Atacama. It contains angular andesite and diorite fragments,
0361-0128/98/000/000-00 $6.00

1277

which are affected by strong K feldspar alteration and silicification chloritization. Pyrite, chalcopyrite, and, locally, bornite and digenite occur interstitial to specularite. The specularite-cemented breccia grades into a specularite stockwork
zone (Transition zone), which also contains chalcopyritepyrite (the specularite breccia and stockwork envelope have
been termed Manto Ruso Breccia by Orrego and Zamora,
1991). The host rocks are characterized by strong pervasive
quartz, K feldspar, or sericite alteration. Moderate chlorite alteration may also be present. The rocks at Manto Ruso host
veinlets similar to those observed at Mantoverde.
The Manto Ruso Breccia developed over a preexisting
breccia that is correlated with the Breccia Verde (Breccia
Temprana of Orrego and Zamora, 1991; Fig. 5a, c). The fragments of the Breccia Verde-Breccia Temprana show a strong
quartz-chlorite-sericite alteration, accompanied by variably
intense K feldspar alteration. The matrix is formed of chlorite
and quartz, plus locally abundant K feldspar and/or epidote,
specularite, pyrite, and chalcopyrite.
Magnetite-rich rocks are also present at Manto Ruso. Their
dimensions vary between some decimeters to a few meters.
They may occur within the igneous host rocks, or in the Transition zone, or the Breccia Temprana (Fig. 5a). Magnetite occurs in massive form in veins and veinlets, as replacements, as
fragments, or in the matrix of breccias. Large magnetite veins
(up to several meters wide) are typically surrounded by brecciated magnetite-rich rocks or occasionally by magnetite
stockwork zones. The magnetite may show a mushketovite
texture. Magnetite-rich rocks are cut by specularite-calcite
veinlets. Pyrite chalcopyrite occur as disseminations and
veinlets in magnetite. Allanite is commonly associated with
magnetite and present as allanite-magnetite microveinlets in
the magnetite stockwork zones.
The Celso area, to the east of Manto Ruso (Fig. 5b), was previously mined for iron (Ferrfera Celso). The martitized magnetite bodies seem to be controlled by a northwest-trending
structural corridor and are located mainly at the contact of volcanic rocks and diorites (C. Astudillo, 2007, pers. commun.).
Exploration activities during 2005 and 2006 outlined an orebody with 5.2 Mt of leachable copper ores immediately to the
west of the old iron ore workings. The copper mineralization is
related to a 20 to 60 m thick, N 40 to 55 W-trending and 55
to 70 W-dipping, specularite-cemented hydrothermal breccia
body, the Celso Breccia. The breccia is enclosed by a specularite stockwork hosted predominantly in diorites. The oxidation
level reaches down to 100 to 150 m depth (820920 m a.s.l.).
Beneath these levels, hypogene sulfides occur over a vertical
distance of at least 100 m (Fig. 5b). The rock units in the Celso
area are similar to those at Manto Ruso. The Celso Breccia is
lithologically similar to the Manto Atacama at Mantoverde or
the Manto Ruso Breccia. It is best developed at depth, where
it is composed of fragments of silicified diorite and minor andesite in a specularite matrix. The breccia contains chalcopyrite-pyrite disseminated in the matrix and is cut by chalcopyrite-pyritebearing veinlets (Astudillo et al., unpub. internal
report, 2006). The surrounding Cu-sulfidebearing specularite
stockwork (Transition zone) is hosted in pervasive silicified and
chloritized rocks. Rocks similar to the Breccia Verde are also
recognized in the Celso area. They consist of silicified dioritic
and andesitic fragments in a chlorite-quartz sericiterich

1277

1278

RIEGER ET AL.

matrix. Massive magnetite bodies developed in the northeast


part of the Celso sector but they were completely converted to
hematite (martite). The dimensions of these bodies can be inferred from the abandoned workings in Ferrfera Celso.
Mineralogy and Paragenetic Sequence
The hypogene mineralization and related alteration in the
Mantoverde district show many similarities with those of the
IOCG deposits in the Punta del Cobre district (e.g., Marschik
and Fontbot, 1996, 2001). The paragenetic sequence at
Mantoverde can be subdivided into three hydrothermal
stages: an early iron oxide stage, a sulfide stage, and a late hydrothermal stage (Fig. 6). The iron oxide stage is characterized by intense iron metasomatism associated with pervasive
potassic alteration, silicification, and hydrolitic alteration of
the host rocks. During this stage, an early generation of specular hematite (hm I) was followed by magnetite (mt I, mt II)
and pyrite (Figs. 6, 7). Hematite (hm I) developed in fractures and in the matrix of breccias. During the subsequent
formation of magnetite, pseudomorphous replacement of the
early hematite (hm I) by magnetite (mushketovitization) took
place within the Magnetite zone (Fig. 7a). Outside the Magnetite zone (i.e., at shallower levels) hematite (hm I) remained unaffected. The bulk of the specularite at shallow
levels is hematite I. K metasomatism caused widespread pervasive K feldspar alteration. In areas in which it was less intense, it caused halos around the fluid conduits, and converted
magmatic feldspars into K feldspar. It also formed biotite,
which is mostly observed as relics. Biotite may have been

more abundant but its former presence is probably masked


by later chloritization. Abundant biotite, as reported from
deep drill holes by Cornejo et al. (2000), has not been observed. Growth textures and crosscutting relationships suggest that tourmaline, scapolite, white mica, and apatite
formed before widespread chloritization and that magnetite
(mt II) formation (locally with growth of allanite and/or epidote) took place and marked the end of the iron oxide stage
(Figs. 6, 7b-d). Locally, specularite may have developed instead of (or in addition to?) magnetite (mt II).
The sulfide stage is characterized by the formation of chalcopyrite, silicification, and quartz, K feldspar-quartz, and
sericite quartz veining (Figs. 6, 7e). Bornite and digenite are
only observed around the oxidation front of Manto Ruso and
Mantoverde Norte mines or in the deeper part of Mantoverde
Norte, near the Mantoverde fault. They replace earlier formed
chalcopyrite to variable degrees, most commonly along grain
boundaries (Fig. 7f-g). Therefore, it is assumed that both minerals are not hypogene but supergene in origin. Gold mineralization seems closely linked to Cu (chalcopyrite), as suggested
by the positive correlation of Au and Cu concentrations in the
ore (see below) and the scarce occurrence of native gold or
gold alloys. The bulk of the pyrite (py II) formed during the
sulfide stage and pyrite II texturally predates chalcopyrite
(Figs. 6, 7h). Small amounts of an earlier pyrite generation developed during the iron oxide stage, and minor pyrite chalcopyrite also formed locally during the late hydrothermal
stage. However, this late stage is characterized by pervasive
carbonatization, the formation of calcite specularite (hm II)

Iron Oxide Stage


Hematite
Magnetite
Pyrite
Gold
Chalcopyrite
Biotite
K feldspar
Tourmaline
Scapolite
Quartz
White mica
Apatite
Allanite
Epidote
Chlorite
Rutile
Sphene
Zeolite
Calcite

Sulfide Stage

Late Stage

at shallow levels

hm I

hm II

mt I

mt II
py II

py I

at deep levels

py III
?
cp I

cp II

?
P

V
V

V
V

V
?

V+D
Loc
V+D
D
V

V+P

Time
FIG. 6. Paragenetic sequence of the iron oxide Cu-Au mineralization in the Mantoverde district. The bulk of the specularite at shallow levels is hematite I. At depth, hematite I gives way to, and was replaced by, magnetite I (mushketovitization)
as the hydrothermal system fully developed. Abbreviations: cp = chalcopyrite, D = disseminations, hm = hematite, Loc =
local, mt = magnetite, P = pervasive alteration, py = pyrite, V = veinlets.
0361-0128/98/000/000-00 $6.00

1278

1279

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

chl-mt II

mt II
scp
cte

ser

mushketovite

py I

100

250

100

all-mt II

cp I
Kfs

bn
Kfs-qtz

chl
Kfs-qtz

cp I
qtz

all-mt II
100

py II

250

100

cte

cp I
dg

ser
hm II
cp I
zeo

py II
100

mt

100

cte
250

FIG. 7. Photomicrographs of thin sections (ts) and polished slabs (ps) documenting paragenetic relationships in the Mantoverde district. a) Magnetite (mt I) pseudomorphously replacing bladed hematite (hm I; ps, parallel nicols). b) Scapolite
crystals cut by a chlorite-magnetite (mt II) veinlet. Calcite occurs as infill between scapolite crystals and in microveinlets cutting the scapolite (ts, cross nicols). c) Sericite and pyrite (py I) formation are followed by a second generation magnetite
(mt II; ps, parallel nicols). d) Allanite and second generation magnetite (mt II) cut by a chlorite veinlet (ts; parallel nicols).
e) Pervasive K feldspar alteration cut by quartz veinlets, which in turn are cut by K feldsparquartz veinlets. Chalcopyrite
(cp I) represents a late infill (ts, parallel nicols). f) Chalcopyrite (cp I) marginally replaced by bornite (ps, parallel nicols). g)
Chalcopyrite (cp I) marginally converted to digenite (ps, parallel nicols). h) Pyrite (py II) veinlets cut magnetite (locally
weakly martitized). Pyrite in turn is cut by thin chalcopyrite (cp I) veinlets. Late hematite (hm II) cuts the aforementioned
minerals (ps; parallel nicols). i) Pervasive zeolite (chabasite-phillipsite) alteration with interstitial calcite that is cut by a late
thin sericite veinlet (ts; cross nicols). Mineral abbreviations: all = allanite, bn = bornite, cte = calcite, chl = chlorite, cp = chalcopyrite, dg = digenite, hm = hematite, Kfs = K feldspar, mt = magnetite, py = pyrite, qtz = quartz, scp = scapolite, ser =
sericite, zeo = zeolite.

veins and veinlets, specularite (hm II) calcite veinlets (Figs.


6, 7h), and minor quartz or sericite veinlets. The amount of
hematite II is by far smaller than that of hematite I, and
mainly limited to small veinlets. Zeolites (chabasite plus traces
of phillipsite), representing low-temperature hydrothermal alteration, occur locally at Mantoverde Norte (Fig. 7i).
Mineralization Age
A loosely constrained Early Cretaceous mineralization age is
based on K-Ar geochronology on hydrothermal sericite yielding 121 3 and 117 3 Ma (Vila et al., 1996) and a two-point
Re-Os isochron age derived from magnetite yielding 116 Ma
(Mathur et al., 2002). The range of alteration-mineralization
ages overlaps within error with that of granitoids of the Sierra
0361-0128/98/000/000-00 $6.00

Dieciocho plutonic complex (120 4 to 126.8 1.3 Ma; Zentilli, 1974; Berg and Breitkreuz, 1983; Berg and Baumann,
1985; Dallmeyer et al., 1996). This is in permissive agreement
with a hydrothermal system that is coeval with or younger than
the constraints for the magmatic rocks. In the mined areas, intrusive rocks assigned to the Sierra Dieciocho complex are
mineralized, indicating that they were already emplaced when
mineralization occurred. A 116 to 121 Ma mineralization age
is compatible with magnetic properties of the orebodies in the
district, which show normal (Manto Ruso) or inverse polarities
(Mantoverde Norte; Tassara et al., 2000). The mineralization
in the Mantoverde district, therefore, may have occurred coeval with a change from an interval with inverse into normal
polarity at about 120 Ma (Vila et al., 1996; Tassara et al., 2000),

1279

1280

RIEGER ET AL.

or alternatively with a change to a short interval of reverse polarity between 115 to 116 Ma (Chronozone M-1r; high-resolution Mesozoic timescale of Gradstein et al., 1995).
Analytical Methods
Whole-rock and mineral geochemistry
Analysis of 10-m composite samples of drill cores from different units from Mantoverde Norte-Sur (Magnetite zone,
Breccia Verde, and Manto Atacama), Manto Ruso, (Transition zone, Breccia Verde-Breccia Temprana, and Manto Ruso
Breccia) and Celso (Transition zone, Breccia Verde-Breccia
Temprana, Celso Breccia, and magnetite-rich rocks) has been
carried out. The samples represent the hypogene ores, although they locally may contain minor supergene overprint.
Concentrations of total Cu, leachable Cu (in H2SO4 and CN
solutions), residual Cu (AAS), and S (LECO) were determined on 121 samples. From these, 96 were analyzed for
their Fe content and 27 were selected for multi-element
analysis (ICP). Additionally, 25 hand specimens of altered
rocks from drill cores from the Mantoverde district were collected for whole-rock major, trace, and rare-earth element
(REE) analysis by the lithium metaborate-tetraborate fusion
ICP-MS method. Although these samples cannot be used to
determine potential REE grades, they facilitate interpretations regarding REE mobility and distribution.
Hydrothermal calcite and sulfides targeted for isotope analysis were handpicked under a binocular microscope. The trace

element contents of hydrothermal calcite were determined to


calculate initial Sr ratios using a total digestion hydrochloric,
nitric, perchloric, and hydrofluoric acid attack and induced
coupled plasma-mass spectrometry (ICP-MS) analytical
method. The calcite contained fluid inclusions (mainly three
phase liquid-vapor-salt up to 10 m in size and liquid-vapor
inclusions typically about 3 m). Seven pyrite and twelve
chalcopyrite samples were analyzed for gold, Pb, Zn, As, Ni,
and Co (aqua regia digestion, ICP-MS). All chemical analyses
were carried out in certified commercial laboratories using
internal standards. Magnetite from Mantoverde Norte was
analyzed at the Ludwig-Maximilians University, Munich,
using a Cameca SX100 electron microprobe calibrated with
natural and synthetic standards. Operating conditions were
15 keV acceleration voltage at 60 nA intensity.
Sulfur isotope analytical methods
Sulfur isotope ratios of 30 samples of handpicked chalcopyrite and pyrite from the Manto Ruso (n = 10), Mantoverde
Norte (n = 11), Mantoverde Sur (n = 7), and San Juan orebodies (n = 1) were determined at the Stable Isotopes Laboratory of the University of Lausanne, Switzerland, using an
online elemental analyzer-continuous flow-isotope ratio mass
spectrometer. Details on the procedure are given in Marschik
et al. (2008). The results are reported in the notation as per
mil () deviations relative to Vienna Caon Diablo Troilite
standard (34SVCDT; Table 1). Replicate analyses show reproducibility of 0.2 per mil.
TABLE 1. Sulfur and Lead Isotope Analyses and Geochemical Data

Sample no.

Location

Drill core or
Altitude

MV07156
MV07178
MV07209
MV07214
MV07221
MV07232
MV07277
MV07286
MV07184
MV99900
MV07421
MV07425
MV07454
MV07460
MV07483
MV07520
MV08001
MV08002
MV08004
MV08009
MV08011
MV08015
MV07323
MV07337
MV07356
MV07356
MV07368
MV07409
MV08005
MV09001
SJ09002

Manto Ruso
Manto Ruso
Manto Ruso
Manto Ruso
Manto Ruso
Manto Ruso
Manto Ruso
Manto Ruso
Manto Ruso
Manto Ruso
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Sur
Mantoverde Sur
Mantoverde Sur
Mantoverde Sur
Mantoverde Sur
Mantoverde Sur
Mantoverde Sur
Mantoverde Sur
San Juan

DDH01GT04
DDH01GT04
DDH01GT07
DDH01GT07
DDH01GT07
DDH01GT07
DDH00MR37
DDH00MR37
902 m a.s.l. (pit)
930 m a.s.l. (pit)
DDH01DS10
DDH01DS10
DDH01DS10
DDH01DS10
DDH00DS05
DDH00MV03
DDH07DS21
DDH07DS21
DDH07DS21
DDH07DS09
DDH07DS09
DDH07DS12
DDH00DS04
DDH00DS04
DDH01DS21
DDH01DS21
DDH01DS21
DDH01DS12
DDH07DS06
RCH09MC14
Stockpile

Depth
(m)
196
330
151
196
230
333
284
381
109
165
546
585
282
667
543
555
620
269
275
450
202
461
231
231
414
380
335
348

Alteration/rock type

Mineral

Specularite breccia
Specularite breccia
Specularite breccia
Specularite breccia
Magnetite-rich rocks
Qtz-chl
Specularite breccia
Qtz-chl
Specularite breccia
Specularite breccia
Specularite breccia
Specularite breccia
Magnetite-rich rocks
Magnetite-rich rocks
Specularite breccia
Magnetite-rich rocks
Magnetite-rich rocks (cte vein)
Magnetite-rich rocks
Qtz-chl
Specularite breccia
Specularite breccia
Magnetite-rich rocks (cte vein)
Calcite vein (Mantoverde Fault)
Magnetite-rich rocks
Specularite breccia
Specularite breccia
Magnetite-rich rocks
Kfs-qtz
Magnetite-rich rocks
Magnetite-rich rocks
Magnetite-rich rocks

Pyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Pyrite
Pyrite
Chalcopyrite
Chalcopyrite
Pyrite
Chalcopyrite
Pyrite
Pyrite
Pyrite
Chalcopyrite
Chalcopyrite
Pyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite
Pyrite
Chalcopyrite
Pyrite
Chalcopyrite
Chalcopyrite
Chalcopyrite

Abbreviations: a.s.l.= above sea level, chl = chlorite, cte = calcite, Kfs = K feldspar, qtz = quartz
0361-0128/98/000/000-00 $6.00

1280

34S ()
9.3
8.9
7.7
7.5
6.3
3.2
7.2
3.8
9.4
7.2
1.4
1.1
1.9
1.5
1.4
0.2
1.1
0.8
2.6
2.6
3.0
3.0
2.5
5.0
5.8
2.5
1.6
5.6
3.6
4.7

1281

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

Radiogenic isotope analytical methods


Variably altered rocks (n = 13) and hydrothermal calcite (n
= 4) from the Manto Ruso, Mantoverde Norte, Mantoverde
Sur orebodies, and outside these mineralized zones have
been collected for Sr isotope analysis. The rock samples were
finely crushed and ground in a tungsten carbide ball mill. The
calcite samples were ground in an agate mortar. Approximately 100 mg of powdered whole-rock were dissolved in
Savillex PFA vessels in a mixture of hydrofluoric acid (HF)
and nitric acid (HNO3) for 36 hours at 120C. The samples
were then cooled down and centrifuged. Solutions were
dried, picked up with concentrated nitrogenic acid (HNO3),
and again dried to assure nitritic form of interesting elements.
After dissolution in 6N HnO3, an appropriate amount of the
solution was introduced in preconditioned PTFE columns
filled with 0.025 ml pre-cleaned Sr-specific resin (Sr-Spec,
Eichrom industries). Strontium separation was achieved following a modified recipe of Horwitz et al. (1992). After elemental separation, ~50 to 200 ng Sr were loaded on precleaned tungsten filaments using an emitter solution
(modified after Birck, 1986). Procedures for calcite were
largely the same but concentrated hydrochloric acid (HCl)
was used for mineral decomposition of ~50 mg powdered
sample. Measurements were performed with a thermal ion
mass spectrometer (Finnigan MAT 261.5) at the laboratory
for radiogenic isotopes of Bayerische Staatsammulung fr
Palontologie und Geologie, Munich, Germany, in static

mode using faraday cups. Results were corrected for instrumental mass fractionation to 88Sr/86Sr = 8.375209. Isobaric interference of 87Rb was controlled by monitoring 85Rb, but no
significant corrections were necessary. Total procedural
blanks were less than 100 pg, which is negligible concerning
sample Sr contents. The mean 88Sr/87Sr for NIST 987 was
0.710248 with a standard deviation of 0.000023 (2008: n = 60,
nominal value NIST 987= 0.710240). No corrections were
made. External reproducibility of 88Sr/87Sr in natural samples
usually is 0.000040 or better.
The lead isotope compositions of 12 samples of chalcopyrite and 7 samples of pyrite were analyzed in static mode
using Finnigan MAT-262 thermal ionization multicollector
mass spectrometer at the Department of Mineralogy of the
University of Geneva, Switzerland (Table 1). The sulfide samples represent different units within the Mantoverde Norte (n
= 11), Mantoverde Sur (n = 3), and Manto Ruso (n = 4) deposits. Sample preparation was carried out following the procedures described in Chiaradia and Fontbot (2003) and
Marschik et al. (2003b). The lead isotope ratios have been
corrected for mass fractionation by a 0.10 percent factor per
atomic mass unit, calculated from replicate analyses of the
SRM981 international standard. An external reproducibility
of lead isotope ratios of 0.07 percent for 206Pb/204Pb, 0.11 percent for 207Pb/204Pb, and 0.15 percent for 208Pb/204Pb has been
demonstrated at the 2 confidence level through multiple
analyses of SRM981 standard.

for Sulfides from the Main Orebodies, Mantoverde IOCG District


206Pb/204Pb

Error
(at 1)

207Pb/204Pb

Error
(at 1)

208Pb/204Pb

Error
(at 1)

Co
(ppm)

Ni
(ppm)

Zn
(ppm)

As
(ppm)

18.511
20.329
22.566

0.002
0.001
0.002

15.561
15.692
15.792

0.002
0.001
0.001

38.269
39.114
42.119

0.004
0.003
0.004

1.2
244.0
3200.0

2.4
49.2
151.0

5.9
5.5
9.0

1.3
23.0
54.1

19.227
18.816

0.001
0.007

15.609
15.634

0.001
0.006

39.160
38.549

0.002
0.015

70.3
223.0

137.0
68.6

8.9
4.0

18.598
19.189
19.500
18.713
18.502

0.001
0.005
0.001
0.000
0.000

15.595
15.642
15.633
15.565
15.575

0.001
0.004
0.001
0.000
0.000

38.706
40.131
38.429
38.529
38.275

0.003
0.010
0.003
0.001
0.001

2810.0
> 5000.0
2950.0
6.0
12.7

157.0
494.0
428.0
2.6
3.2

18.533
19.540
18.637
18.514
18.528
18.565
18.505
18.882

0.001
0.000
0.000
0.001
0.000
0.000
0.000
0.001

15.555
15.604
15.568
15.568
15.600
15.559
15.572
15.590

0.001
0.000
0.000
0.001
0.000
0.000
0.000
0.001

38.260
38.306
38.255
38.230
38.367
38.245
38.289
39.154

0.003
0.001
0.001
0.002
0.001
0.000
0.000
0.001

62.6
54.5
493.0
724.0
888.0
538.0
4.2
109.0

19.200

0.002

15.607

0.002

38.611

0.004

2700.0

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

1281

Au
(ppb)

Pb
(ppm)

Th
(ppm)

U
(ppm)

26.5
< 0.5
< 0.5

1.4
2.2
2.5

< 0.1
< 0.1
4.5

< 0.1
< 0.1
0.8

6.7
23.8

5560.0
< 0.5

5.7
0.7

< 0.1
< 0.1

< 0.1
< 0.1

17.0
6.0
7.5
17.2
51.3

58.7
220.0
26.8
5.0
3.5

27.9
1.2
< 0.5
2720.0
2200.0

2.9
1.3
2.3
19.3
44.8

0.5
< 0.1
0.1
< 0.1
< 0.1

< 0.1
< 0.1
0.3
< 0.1
0.1

10.8
14.3
60.9
88.2
85.9
60.3
3.4
20.7

17.8
9.9
17.6
43.9
69.8
15.7
20.5
28.5

8.3
5.2
18.7
18.5
27.1
25.1
1.7
4.6

1040.0
208.0
675.0
33.5
3840.0
1410.0
3030.0
> 10000.0

4.2
7.7
14.2
22.1
51.0
28.6
25.2
4.4

< 0.1
< 0.1
< 0.1
< 0.1
0.1
0.1
< 0.1
< 0.1

< 0.1
< 0.1
< 0.1
< 0.1
0.1
0.1
< 0.1
< 0.1

618.0

3.7

43.7

112.0

1.1

< 0.1

< 0.1

1282

RIEGER ET AL.

Analytical Results
Composite sample geochemistry
The range of total Cu (soluble and residual) and residual
Cu values in composite samples are similar in all three deposits, with highest values in samples from Mantoverde Sur.
Total Cu and Au concentrations show a good positive correlation in the district (Fig. 8). The distribution of Cutotal and
Au in the various defined units is shown in Figure 8. Zinc
concentrations are generally low (33254 ppm) and present
at a similar range of values in the Mantoverde and Manto
Ruso deposits. Silver, Pb, and Mo are typically below detection limit, although some measureable Mo concentrations
(<32 ppm) occur locally in the Magnetite zone of Mantoverde. Cobalt concentrations in composite samples range
from 11 to 803 ppm. The Co concentrations in the samples
from Manto Ruso are near or below the crustal average of
29 ppm (crustal values of Taylor and McLennan, 1985),
whereas at Mantoverde, Co values are higher by almost one
order of magnitude. Mercury and Cd values are commonly
low (<5 ppm or <0.2 ppm, respectively). Titanium contents
are generally below detection limit (i.e., <0.01%), whereas V

concentrations are between 7 and 203 ppm, and therefore


below or near the average crustal value of 230 ppm. Arsenic
values (up to 42.4 ppm) are enriched, and Ba values (6-60
ppm) depleted compared to their average crustal concentrations of 1 and 250 ppm, respectively. The average K content
in the composite samples is 0.3 percent and is one order of
magnitude higher than that of Na.
Results of rare earth element geochemistry
Rare earth elements are mobile in many IOCG systems and
may reach potentially economic concentrations locally in
these deposits. We analyzed REE at Mantoverde to locate
areas or units enriched in REE and assess possible REE
sources. Chondrite-normalized REE patterns (Nakamura,
1974) of hydrothermally altered rocks in the Mantoverde district are compared to reference basaltic andesite (Ewart,
1982), andesite (GERM database), and to Jurassic volcanic
rocks from Cerro Blanco, Cerro Plomo, Cerro del Difunto,
Sierra Minillas between Taltal and Chaaral (Lucassen et al.,
2006), Jurassic volcanic rocks from Sierra Fraga and Quebrada la Tranquita east of Mantoverde (Lucassen et al., 2006),
and Early Cretaceous plutons from the Candelaria-Punta del

Cu vs (Au x1000) per sector

Cu vs (Au x1000) in Mantoverde Sur

0.7

0.7

Mantoverde Sur
Manto Ruso
Celso

Magnetite Zone
Breccia Verde
Manto Atacama

0.6

0.5

0.5

0.4

0.4

Au ppm

Au ppm

0.6

0.3

0.3

0.2

0.2

0.1

0.1

0.0

0.0
0.1

1.0

10.0

0.1

1.0

Cu vs (Au x1000) in Manto Ruso

Cu vs (Au x1000) in Celso

0.7

0.7

Manto Ruso Breccia


Transition Zone
Breccia Verde

0.6

Celso Breccia
Transition Zone
Breccia Verde
Magnetite Zone

0.6

0.5

0.5

Au ppm

Au ppm

10.0

Cutotal wt.%

Cutotal wt.%

0.4

0.3

0.4

0.3

0.2

0.2

0.1

0.1

0.0

0.0
0.1

1.0

10.0

0.1

1.0

Cutotal wt.%

Cutotal wt.%

FIG. 8. Representative Cutotal and Au contents of composite samples from drill cores in the Mantoverde district. Overall,
there is a good positive correlation between Cu and Au.
0361-0128/98/000/000-00 $6.00

1282

10.0

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

Cobre district (Marschik et al., 2003a; Fig. 9). These kinds of


rocks are commonly enriched in REE relative to chondrite,
and their LREE may be moderately enriched (commonly up
to ~10 times) relative to the HREE (Fig. 9j).
The host rocks in the Mantoverde district show a variety of
REE patterns, including relatively flat (Fig. 9b, f, h, i), slightly
to moderately enriched character (Fig. 9a-f); slightly Ushaped patterns (Fig. 9a, b); HREE-enriched relative to
LREE (Fig. 9f, g) or vice versa; patterns with a more or less
pronounced positive (Fig. 9f) or negative Eu anomaly (Fig.
9g); or without an Eu anomaly (Fig. 9a-i). Three samples contain significant LREE concentrations. The highest combined
concentrations of La, Ce, Pr, and Nd of 1,893 ppm were detected in altered dioritic rocks at Mantoverde Norte. Another
LREE-enriched sample (MV96-5) stems from the Manto Atacama breccia of Mantoverde Norte. The volcanic breccia
clasts (MV96-5-1) and the specularite matrix in this rock were
separately analyzed (MV96-5-2; Fig. 9b). The whole-rock
breccia contains combined La, Ce, and Nd concentrations of
401 ppm, the breccia clasts have 974 ppm, and the specularite matrix contains 502 ppm.
Isocon plots according to Grant (2005) have been established
using the most conservative behaving elements for drawing the
chord (gain-loss reference line; Fig. 10). These diagrams suggest that REE were redistributed within the deposits. Enrichments in LREE (Fig. 10a) or HREE (Fig. 10b) are detected
and rocks with variably depleted REE are the likely source for
these enrichments (Fig. 10c, d). The depletions and enrichments may have occurred in close spatial proximity.
Results of mineral geochemical analysis
Pyrite from the investigated IOGC deposits has <17 ppm
Zn, <3 ppm Pb, up to 618 ppm Ni, and up to >5,000 ppm Co.
Arsenic values are typically below 60 ppm, although a value of
220 ppm has been obtained from one sample. Chalcopyrite
contains up to 27 ppm As, up to 51 ppm Pb, <70 ppm Zn,
<140 ppm Ni, and up to 888 ppm of Co. Gold contents in
pyrite are up to 112 ppb, whereas Au in chalcopyrite may
reach >10,000 ppb (Fig. 11). Full analytical results are given
in Table 1.
Magnetites from Mantoverde Norte and Sur deposits show
similar median contents of trace elements (Ni = 0.01 wt %,
Zn = 0.01 wt %, Cr = 0.01 wt %, Ti = 0.01 wt %, V = 0.04 wt
%, Si = 0.08 wt %, Al = 0.02 wt %, K = 0.01 wt %, Ca = 0.01
wt %, and Cu and Mn below detection limit). Compared with
Mantoverde, magnetites from the barren magnetite deposits
Rebosadero Sur and Ferrfera-Santa Clara in the district
show higher median values in Ti (0.16 wt %), V (0.43 wt %),
and Al (0.15 wt %) and lower values in Si (0.03 wt %; Fig. 12).
Negligible differences were observed in the other elements.
Results of sulfur isotope geochemistry
Chalcopyrite yields 34S values from 5.6 to 8.9 per mil,
whereas 34SVCDT values of pyrite range from 0.2 to 9.4 per
mil (Fig. 13; Table 1). Chalcopyrite from the Manto Ruso orebody (3.88.9 34S) tends to be more enriched in 34S compared to chalcopyrite from the Mantoverde Norte (3.0 to
+2.6 34S) and the Mantoverde Sur orebodies. Four out of
five chalcopyrites from Mantoverde Sur show 34S between
5.6 and 2.5 per mil, and the outlier has a 34S value of 5.0
0361-0128/98/000/000-00 $6.00

1283

per mil. The single chalcopyrite sample from the magnetiterich ore of the San Juan mine has a 34S value of 4.7 per mil.
The sulfur isotope signature of pyrite shows a similar though
less pronounced spatial variation, with 34S from 3.2 to 9.4
per mil at Manto Ruso and 0.2 to 1.9 per mil at Mantoverde
Norte. Two pyrites from the Mantoverde Sur orebody have
34S values of 1.6 and 5.8 per mil.
Results of radiogenic isotope analyses
Rubidium and Sr contents in the altered host rocks range
from 17 to 204 ppm, and from 12 to 198 ppm, respectively
(Table 2). The Rb/Sr ratio in all but two samples (MV07450
and MV07658) is >1. Calcite has Rb contents of 0.2 to 0.5
ppm and Sr contents of 62 to 103 ppm (Table 2). The initial
87Sr/86Sr (87Sr/86Sr ) is based on an average alteration age of
i
119 Ma (Vila et al., 1996). The altered rocks show a relatively
wide spread in 87Sr/86Sri values from 0.703054 to 0.710053. In
contrast, the 87Sr/86Sri values of hydrothermal calcite are fairly
homogeneous, ranging from 0.703628 to 0.703708 (Table 2;
Fig. 14). Strontium contents tend to correlate negatively with
measured 87Sr/86Sr, whereas Sr contents and 87Sr/86Sri do not
show any correlation.
Lead isotope ratios of chalcopyrite have 206Pb/204Pb between 18.502 and 19.540, 207Pb/204Pb between 15.555 and
15.609, and 208Pb/204Pb between 38.230 and 39.160 (Table 1).
Those from pyrite show a wider range with 206Pb/204Pb from
18.598 to 22.566, 207Pb/204Pb from 15.595 to 15.792, and
208Pb/204Pb from 38.429 to 42.119. Most of the Pb isotope ratios of chalcopyrite plot close to the Orogene Pb isotope evolution line of Zartman and Doe (1981; Fig. 15). They show
signatures that mainly plot into the compositional field defined by arc igneous rocks in a 206Pb/204Pb versus 207Pb/204Pb
diagram (Zartman and Doe, 1981). The sample provenance,
Pb isotope analytical data, and Pb, U, and Th concentrations
in chalcopyrite and pyrite are given in Table 1.
Discussion
District alteration
The various rock, alteration, and ore types are similar among
the IOCG deposits in the Mantoverde district, as are the
mineral associations in veinlets and their relative temporal occurrence. Isocon diagrams suggest substantial mobility of elements (Fig. 10). Most notable is the widespread and variably
intense potassium metasomatism, which is expressed by
abundant K feldspar (Fig. 10). Hydrothermal biotite occurs
only in minor quantities. However, a significant early biotite
alteration may have taken place, but could have been masked
by later chloritization. Unlike other IOCG districts (e.g., Candelaria-Punta del Cobre: Marschik and Fontbot, 2001;
Ernest Henry-Cloncurry: Mark and Forster, 2000; Norbotten
region in Sweden: Edfelt et al., 2005), at Mantoverde there is
no evidence of a district- or regional-scale sodic or sodic-calcic alteration event that is commonly manifested as variable
proportions of secondary albite, marialitic scapolite, epidote, calcic-amphibole, and/or diopside. Scapolite alteration
is locally present in the Mantoverde Norte-Sur deposits,
whereas calcic-amphibole is essentially absent within the district at the present levels of exposure. In general, however,
the alteration and mineralization styles and processes appear

1283

1284

RIEGER ET AL.

similar to those of the IOCG deposits in the Punta del Cobre


district (Marschik and Fontbot, 1996, 2001) to the south of
Mantoverde. These deposits are characterized by pervasive K
feldspar-quartz-chlorite/biotite plus abundant calcite veining,
and are dominated by specularite-chalcopyrite-pyrite at shallow
levels and by magnetite-chalcopyrite-pyrite at depth. By analogy with the Candelaria-Punta del Cobre IOCG system and in
Mantoverde Norte: Tectonic breccia - Mantoverde Breccia

10000

Ore characteristics
The Cu and Au grade distribution in the orebodies Mantoverde Norte-Sur and Manto Ruso is fairly homogeneous

1000

100

10

10

0.1
La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

La

Lu

Mantoverde Norte: Specularite stockwork - Transition Zone

100

10

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

Lu

10000

MV96-5 (WR)
MV96-5-1 (F)
MV96-5-2 (matrix)

MV96-4 (WR)
MV07524 (F)

1000

Ce

Mantoverde Norte: parts of Specularite breccia - Manto Atacama

Abundance/chondrite

Abundance/chondrite

100

0.1

1000

100

10

0.1

0.1
La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

La

Lu

Mantoverde Norte: altered volcanic rock

10000

10000
MV07464 (WR)
MV07540 (WR)

1000

Abundance/chondrite

Abundance/chondrite

MV96-5 (WR)
MV07485 (WR)
MV07427 (F)
MV07514 (F)
MV07526 (F)

1000

10000

Mantoverde Norte: Specularite breccia - Manto Atacama

10000

MV96-1 (WR)
MV07450 (WR)

Abundance/chondrite

Abundance/chondrite

accordance with the geological model of Hitzman et al. (1992),


the Mantoverde district represents a more distal or shallow
facies of the general zonation seen in many IOCG systems.

100

10

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

Lu

Mantoverde Sur

MV07335 (WR)
MV07337 (WR)
MV07356 (F)
MV07368 (F)
MV07577 (WR)

1000

100

10

0.1

0.1
La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

Lu

La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

FIG. 9. Chondrite-normalized REE patterns of altered rocks from various ore zones and units in the Mantoverde district
showing local enrichment of LREE (chondrite of Nakamura, 1974). a) Results of whole-rock (WR) analyses of samples from
the tectonic breccia (Mantoverde Breccia) at Mantoverde Norte. b) Cu-Aubearing specularite breccia (WR) and its rock
fragments (F) at Mantoverde Norte (Manto Atacama). c) Altered rocks from the Transition zone of Mantoverde Norte. d)
Sample MV96-5 of the Manto Atacama breccia of Mantoverde Norte. The patterns of the whole-rock fraction (MV96-5 WR),
the volcanic rock breccia clasts (MV96-5-1 F), the specularite matrix (MV96-5-2 matrix) are shown. e) Altered volcanic rocks
from the Mantoverde Norte sector showing near original to slightly elevated LREE. f) Samples of altered whole-rock and
breccia fragments from the Mantoverde Sur sector suggesting locally elevated LREE or relative depletion of LREE compared to HREE in other areas. Relatively flat REE patterns occur in rocks from g) Manto Ruso; h) altered plutonic rocks;
and i) volcanic rocks outside of the mineralized zones. j) REE pattern of reference basaltic andesite (Ewart, 1982), Jurassic
volcanic rocks of Cerro Blanco, Cerro Plomo, Cerro del Difunto, Sierra Minillas, Sierra Fraga, and Quebrada la Tranquita
(Lucassen et al., 2006), andesite (GERM database), and Early Cretaceous plutons from the Candelaria-Punta del Cobre district. These rock types are similar to those in the Mantoverde district (Marschik et al., 2003a).
0361-0128/98/000/000-00 $6.00

1284

Tm

Yb

Lu

1285

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE


Manto Ruso

10000

MV07176 (WR)
MV07206 (WR)
MV07135 (F)

1000

100

10

10

0.1

0.1
La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

La

Lu

Ce

Pr

Altered volcanic rock outside the mineralized zone

MV07076 (WR)
MV07658 (WR)
MV07058 (WR)

1000

10000

Abundance/chondrite

Abundance/chondrite

MV07498 (WR)
MV07641 (WR)
MV07646 (WR)

100

10000

Altered intrusive rock

1000

Abundance/chondrite

Abundance/chondrite

10000

100

10

Nd

Sm

Eu

Gd

Tb

Ho

Er

Tm

Yb

Lu

Reference analyses of magmatic rock

Basaltic andesite
Jurassic volcanic rock
Andesite
Copiap Batholith, diorite
Copiap Batholith, hornblende-diorite
Copiap Batholith, tonalite

1000

Dy

100

10

0.1

0.1
La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

Lu

La

Ce

Pr

Nd

Sm

Eu

Gd

Tb

Dy

Ho

Er

Tm

Yb

Lu

FIG. 9. (Cont.)

owing to the fact that mineralization is mostly disseminated or


in veinlets. Large Cu sulfide-rich veins or Cu sulfide cemented breccias as described in other Chilean IOCG deposits (e.g., Marschik and Fontbot, 1996, 2001) are absent.
However, high-grade ores of up to 4.2 percent Cu are present
locally (e.g., in the Magnetite zone at Mantoverde NorteSur). Overall, a positive correlation of Cutotal and Au concentrations is recognized (Fig. 8), whereas Cu versus Fe or Au
versus Fe do not show any kind of trend. Plotting leachable or
residual Cu versus Au or Fe does not provide any additional
information. These relationships are similar to those reported
in the Candelaria IOCG deposit (Marschik et al., 2000) and
are consistent with the paragenetic position of the iron oxide
and the Cu-Au mineralization (Fig. 6). The positive correlation of Cu and Au, the fact that chalcopyrite is Au-bearing
(Fig. 11), and the scarce occurrence of native gold or electrum are compatible with the hypothesis that Au was transported together with Cu in chlorine complexes (e.g., Seward,
1991). Despite the paragenetic position of hematite-magnetite and chalcopyrite and the nonexisting correlation between Fe and Cu-Au contents in the ores, as mentioned
above, there is a close spatial association of Cu with iron oxides. Bornite, which forms part of the hypogene ore in some
IOCG deposits such as Salobo (Requia et al., 2003) or in Andean porphyry copper deposits (e.g., Gustafson and Hunt,
1975), is interpreted as a product of supergene processes at
Mantoverde based on replacement textures and its limited
spatial occurrence near the supergene oxidation front. Sphalerite-rich zones, as at the Candelaria or the Carola IOCG
deposits in the Punta del Cobre district (Marschik and Fontbot, 2001), have not yet been observed at Mantoverde. Zinc
0361-0128/98/000/000-00 $6.00

concentrations in most composite samples cluster around the


average crustal value of ~80 ppm (Taylor and McLennan,
1985). Also noteworthy is the relatively low Ag contents (typically <0.5 ppm) in the ores as compared to the CandelariaPunta del Cobre IOCG system, where Ag is typically 3 to 7
ppm (Marschik and Fontbot, 2001).
Elevated LREE contents are detected locally in IOCG deposits (e.g., Hitzman et al., 1992; Williams et al., 2005). Although Mantoverde does not show the extreme enrichment
observed in deposits such as Olympic Dam (Oreskes and Einaudi, 1990), some significant REE concentrations are detected. In composite samples from the mineralized Breccia
Verde at Mantoverde Norte-Sur, La is enriched up to 128
times over chondrite and up to 174 times in the equivalent
unit at Manto Ruso. The Mantoverde Breccia, Manto Atacama Breccia, and Transition zone at Mantoverde Norte are
also locally enriched in LREE (Fig. 9a-c). In the Manto Atacama Breccia (sample MV96-5) both the altered rock fragments (MV96-5-1; Table 3) and the specularite matrix (MV965-2) contain these elements, although the rock fragments host
higher concentrations (Fig. 9d). The fact that the sample representing the entire breccia contains less LREE than the matrix fraction may result from the total Fe content, which is 15
percent higher in the breccia than in the matrix. The matrix
fraction contains abundant, minute, altered volcanic rock
fragments, which may host the bulk of the LREE detected in
the matrix fraction. Most samples of the Manto Atacama from
Mantoverde Norte have low concentrations in LREE (~10
times chondrite; Fig. 9b). They are lower than those of the
reference rocks (Fig. 9j). Overall, the REE patterns of the altered rocks of the Mantoverde district are heterogeneous.

1285

1286

RIEGER ET AL.

120

50

gain
loss
25 %

Gd
Al2O3
Fe2O3
Ta Zr Hf
SiO2
Cs Nb
TiO2 MnO

60
Tb

40

Tm

Dy

20

Er Yb Lu

0
0

10

Zn

30

40

60

70

40

K2O

50 %

80

LOI Rb

20

Sr

90

20
0

30

500
%
400
%

0%

Mantoverde Norte: Specularite breccia - MV07526


%

10

15

Ni

%
75
Ta

0%

50

25

100

50

25

gain
loss
25 %

%
0

Tb
Zr Hf

Al2O3

Nb
Zn Sc

Fe2O3
MgO

20

30

40

50 %

Cr

Sm
Gd
Nd
Eu

CaO

Ni

50

60

70

75 %
V

80

90

10

15

0%

75

Ni

50

25

100
gain
s
los

K2O

80

gain
loss

80

25 %

60
Hf

CiA

Ta

Ba

CiA

CiO
Mantoverde Norte: Specularite breccia - MV07485

120

TiO2
MnO

P2O5
Ba
Cs

10

CiO
120

Ta

Dy
La
Ce
Pr
SiO

Na2O

Na2O

50

75

10

80

20

0%

Lu

60

75 %

P2O5

Sr CaO

Ho

0%

Ho

80

15

100

20
0

25 %

Sc

MgO
Eu

Tm
Yb

0%

Sm

100

LOI

20

Er

30

Ba

30

75

50
0%
40
0%

140

Nd

120

CiA

Rb

140

100

400

500

%
0

K2O

160

20

30

Pr

Manto Ruso: Specularite stockwork - MV07176

160

%
50

CiA

La Ce

40
0

180

50
0%

Mantoverde Norte: Specularite stockwork - MV07524

200

Zr

Zr

40

MnO

Rb

P2O5

20
LOI

Sr

Cs

Al2O3SiO2

Lu
MgO
Ho TmYb
Pr Sm Gd
Fe2O3
TiO2
Dy Er
La Ce Nd Eu Tb
Na2O CaO

50 %

Zn

Nb

Sc V

40

75 %

20

LOI Rb

Sr

0
0

10

20

30

40

50

60

70

80

25 %

K2O

60

90

Yb
Ba
Er
Tb Ho Tm Lu
Fe2O3
SiO2
Dy
Na2O
Gd
Al2O3 Pr Sm
MnO
Nb
Eu
La Ce Nd
TiO2
MgO
P2O5
CaO
Cs

10

20

30

40

50

60

Hf

Y
50 %

Cr

Zn Sc V

70

75 %

80

90

CiO

CiO
FIG. 10. Isocon analysis of selected rock samples from Mantoverde Norte and Manto Ruso (Grant, 2005). The composition of the least altered reference rock is an average of whole-rock data of Jurassic basaltic andesite and andesite (5255 wt
% SiO2) from Cerro del Difunto, Sierra Fraga, and Quebrada la Tranquita of Lucassen et al. (2006). Note that the scales of
the diagrams are different to accommodate all data. The oxides or elements used to establish the chords are: a) SiO2, Fe2O3,
and Dy; b) Tb; c) SiO2 and Al2O3; d) SiO2, and Ta. Dashed lines indicate element or element oxide percentage of enrichment
or depletion relative to the reference rocks. Abbreviations: CiA = scaled element concentration of the altered rock; CiO =
scaled element concentration of the reference rock.

The isocon diagrams (Fig. 10) suggest that REE were variably
mobilized and deposited by the hydrothermal fluids. There is
no compelling evidence for a significant LREE introduction

from outside the deposit by the mineralizing fluids. Therefore,


significant amounts of recoverable LREE concentrations in
1.2
Mantoverde
Norte/Sur deposit
Magnetite-apatite
deposit

S vs Au per sector
100

1.0

10

Concentration in wt. %

Mantoverde Norte
Mantoverde Sur
Manto Ruso

Au ppm

chalcopyrite
0.1

0.01

0.8

0.6

0.4

pyrite
0.2

0.001

0.0001
0

10

12

14

16

18

0.0

20

Ti

S wt.%

FIG. 11. Gold content of pyrite and chalcopyrite from the Mantoverde
IOCG ores.
0361-0128/98/000/000-00 $6.00

Ti

Al

Al

Si

Si

FIG. 12. Box plots of Ti, V, Al, and Si concentrations in magnetite from
IOCG and magnetite(-apatite) bodies in the Mantoverde district.

1286

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

1287

: chalcopyrite
: pyrite

MVF

h
ntal branc
AFZ orie

Manto Ruso

Mantoverde
Norte

Mantoverde
Sur

San Juan
-6

-4

-2

10

34SCDT ()

FIG. 13. Location map of the Mantoverde district showing the main ore zones and distribution of sulfur isotope signatures of chalcopyrite and pyrite. A general trend in enrichment of 34S in chalcopyrite from south to north is recognized
(see arrow).

the deposit are unlikely. The identification of the LREEbearing units and the host minerals to these elements is still
in progress. Parts of the LREE are hosted by allanite and apatite. Both minerals, however, are present only in minor
quantities. In general, the Cu-Au ores in the Mantoverde district are low in U and have relatively low contents in heavy
metals that are potentially hazardous to the environment,
such as As (avg 14 ppm), Hg (<5 ppm), or Cd (<2 ppm).
The Cu and Au grades in supergene and hypogene zones
are similar, suggesting insitu supergene alteration. An enrichment blanket as, for instance, in the El Salvador porphyry

copper deposit (Gustafson and Hunt, 1975) has not developed. The relative immobility of the supergene Cu seems unfavorable for development of related exotica-type deposits in
the vicinity (e.g., Mote et al., 2001).
Iron oxide distribution
A zonation of iron oxide species exists within the individual
IOCG deposits in the Mantoverde district. Specularite is characteristic of the upper levels of the deposits, whereas magnetite is present at depth. This zonation, which is common in
IOCG deposits, is obscured at district-scale by postore tilting

Early Cretaceous seawater


Early Cretaceous granitoids
(magmatic-hydrothermal fluids)
Early Cretaceous volcanic rocks

Mantoverde district

Jurassic volcanic and intrusive rocks


Altered volcanic rocks
Manto Ruso altered rocks
Mantoverde Norte altered rocks
Mantoverde Norte calcites
Mantoverde Sur altered rocks
Mantoverde Sur calcites
0.703

0.704

0.705

0.706

0.707

0.708

0.709

0.710

initial 87Sr/86Sr

FIG. 14. Strontium isotope initial ratios of altered rocks (sample labels refer to Fig. 9) and calcite from the Mantoverde
district compared to Sr isotope signatures of relevant Sr reservoirs. Early Cretaceous seawater (Jones et al., 1994); Early Cretaceous granitoids (Chilean Coastal Batholith near Copiap, Marschik et al., 2003a; Las Tazas complex, Berg and Baumann,
1985); Early Cretaceous volcanic rocks (Morata and Aguirre, 2003); Jurassic igneous rocks (Lucassen et al., 2006). The strontium isotope signature of Early Cretaceous granitoids corresponds to the inferred composition of magmatic hydrothermal
fluids.
0361-0128/98/000/000-00 $6.00

1287

1288
0.703688
0.000028
0.703707
0.004
0.4
365
Mantoverde Sur
MV07366

DDH01DS21

194
Mantoverde Sur
MV07322

DDH00DS04

452
Mantoverde Norte
MV07490

DDH00DS05

226

114
380
461

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

Abbreviations: a.s.l. = above sea level, chl = chlorite, cte = calcite, hm = hematite, Kfs = K feldspar, qtz = quartz, ser = sericite

119
Calcite

101.0

0.703708
0.000033
0.703735
0.005
0.5
119
Calcite

91.2

0.703681
0.000017
0.703700
0.004
0.4
119
Calcite

103.0

0.705629
0.705568
0.704323
0.705812
0.703054
0.704527
0.703886
0.704894
0.704771
0.706081
0.704260
0.704014
0.710053
0.703628
0.000016
0.000018
0.000053
0.000084
0.000059
0.000036
0.000076
0.000043
0.000032
0.000019
0.000036
0.000034
0.000390
0.000022
0.726872
0.722766
0.708080
0.723974
0.718668
0.709703
0.712725
0.714687
0.709937
0.718777
0.710165
0.705185
0.733354
0.703644
4.333
3.509
0.768
3.706
3.188
1.058
1.806
2.000
1.056
2.592
1.207
0.239
4.750
0.003
104.0
186.0
152.0
126.0
204.0
55.0
65.0
54.0
76.0
127.0
111.0
17.0
57.0
0.2
24.0
53.0
198.0
34.0
64.0
52.0
36.0
27.0
72.0
49.0
92.0
71.0
12.0
62.0
119
119
119
119
119
119
119
119
119
119
119
119
119
119
Ser-qtzchl
Kfsser-cal
Chl-serqtz
Kfs-ser-qtzcal
Kfs-qtzser
Chl-qtz-hm
Chl-qtz
Chl-qtz-cte
Kfs-qtz-chl
Kfs-qtz
Kfs-chl-hm
Chl-cte
Qtz-ser
Manto Ruso
Manto Ruso
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Norte
Mantoverde Sur
Mantoverde Sur
Mantoverde Sur
East from Laura
Northwest from Celso
Quebrada Guamanga
Mantoverde Norte
MV07176
MV07206
MV07450
MV07485
MV07514
MV07464
MV07540
MV07335
MV07337
MV07577
MV07076
MV07658
MV07058
MV07477

DDH01GT04
DDH01GT07
DDH01DS10
DDH00DS05
DDH99MV03
1001 m a.s.l. (pit)
DDH07N04
DDH00DS04
DDH00DS04
910 m a.s.l. (pit)
971 m a.s.l.
995 m a.s.l.
999 m a.s.l.
DDH00DS05

317
106
519
314
520

Specularite stockwork
Specularite breccia
Tectonic breccia
Specularite breccia
Specularite breccia
Volcanic rock
Volcanic rock
Breccia Verde
Breccia Verde
Volcanic rock (?)
Volcanic rock
Volcanic rock
Volcanic rock
Calcite vein
(Breccia Verde)
Calcite vein
(Breccia Verde)
Calcite vein
(Mantoverde fault)
Calcite vein
(Magnetite zone)

Rock
Rock
Rock
Rock
Rock
Rock
Rock
Rock
Rock
Rock
Rock
Rock
Rock
Calcite

Rb/Sr
Rb
(ppm)
Sr
(ppm)
Age
(Ma)
Rock/
mineral
Alteration
Unit/rock type
Depth
(m)
Drill core or
Altitude
Location
Sample
no.

TABLE 2. Strontium Isotope and Geochemical Data of Altered Rocks and Calcite from the Mantoverde IOCG District

(87Sr/86Sr)

Error
(at 2)

(87Sr/86Sr)i

RIEGER ET AL.

of the structural block east of the Mantoverde fault (Fig. 4).


The structural configuration explains the predominance of
near-surface magnetite-rich rocks (i.e., the exposure of
deeper portions of the system) in the Mantoverde Sur sector,
whereas toward the north (Mantoverde Norte and Manto
Ruso), the upper portions of the system are preserved and
hematite is the predominant near-surface iron oxide phase
(Orrego and Zamora, 1991; Zamora and Castillo, 2001; Sanhueza and Robles, unpub. internal report, 1999).
Apart from near-surface magnetite found in the Mantoverde Sur sector, there are magnetite(-apatite) bodies barren in Cu and Au along the eastern branch of the Atacama
fault zone (e.g., Ferrfera Celso, Rebosadero Sur, FerrferaSanta Clara). Apart from their lack of Cu-Au, these iron deposits show some differences to the nearby IOCG bodies.
They are controlled by north-south structures, whereas the
IOCG deposits occur in northwest to north-northwest faults
(Fig. 4). The alteration associated with the magnetite(-apatite) bodies is similar to that of the IOCG deposits, yet the
intense K alteration is absent in the former. Calcite occurs as
veinlets and pervasive carbonatization, although the thick calcite veins that characterize the IOCG deposits along the
Mantoverde fault are lacking. A common feature of the CuAubearing magnetite in the Mantoverde district is abundant
mushketovite. Magnetite in the iron orebodies along the Atacama fault zone is typically massive and mushketovite has not
been identified. The iron ores at Rebosadero Sur and Ferrfera-Santa Clara are not spatially associated with Breccia
Verde-type rocks. The relationships between Breccia Verde
and the iron mineralization at Ferrfera Celso remain unclear
because of the close proximity of the Celso Cu-Au ores.
Magnetite composition has been used to discriminate between various magnetite-bearing mineral deposit types (e.g.,
Espinoza, 1985; Beaudoin et al., 2007). Magnetite from the
magnetite(-apatite) ores in the district tends to have higher Ti
and V contents than magnetite from the IOCG deposits (Fig.
12). However, the analytical data scatter widely so that they
may not be of any practical use.
Paragenesis and ore formation
Cooling-mixing models in which magmatic-hydrothermal
metal-bearing fluids interact with nonmagmatic hydrothermal fluids have been envisaged for South American IOCG
systems (e.g., Marschik et al., 2000, 2003a; Marschik and
Fontbot, 2001; Chiaradia et al., 2006) and this process is favored here for the Mantoverde district (e.g., Vila et al., 1996;
Cornejo et al., 2000; Williams et al., 2005). The physicochemical conditions of hydrothermal evolution have not been
rigorously established, but the following constraints have
emerged from our study. The new paragenetic scheme is
compatible with such a hypothesis. Barren hydrothermal
brecciation (Breccia Verde) preceded the IOCG mineralization in the district. The metallic mineralization started developing under oxidizing conditions and/or low temperatures, as
documented by the formation of early specularite (hm I; Figs.
6, 7a). A change to magnetite (mt I) and the transformation of
early specularite into mushketovite in the deeper parts of the
system record an increase in temperature as the system became fully established and/or a drop in the redox state (see
De Haller and Fontbot, 2009). The Fe metasomatism was

1288

1289

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

40.0

39.0
39.5

208Pb/204Pb

208Pb/204Pb

B
C
39.0
B
38.5

38.5

38.0
38.0

MORB
A
37.5
17.5

18.0

18.5

19.0

19.5

20.0

20.5

18.5

15.7

204

Pb/

19.5

Pb

C
207Pb/204Pb

207Pb/204Pb

15.7

19.0
206

206Pb/204Pb

15.6

15.5

15.6

MORB
A
15.4
17.5

18.0

18.5

19.0

19.5

20.0

20.5

18.5

206Pb/204Pb
Chalcopyrite
Mantoverde Norte

Pyrite
Mantoverde Norte

Mantoverde Sur

Mantoverde Sur

Manto Ruso

Manto Ruso

19.0

19.5

206Pb/204Pb
Volcanic rocks

Coastal Batholith
Copiap (Early Cretaceous)1

North and east of Chaaral (Jurassic)2

South of Antofagasta (Jurassic - Early Cretaceous)2


North of Taltal (Jurassic)2

FIG. 15. Lead isotope signatures of chalcopyrite and pyrite from IOCG bodies in the Mantoverde district. Pb isotope evolution curves after Zartman and Doe (1981). Letters within charts: A = mantle, B = orogene, and C = upper crust contributed
to the orogene. Reference fields of Marschik et al. (2003b)1 and Lucassen et al. (2006)2. Note that sample MV07232 has an
anomalous Pb isotope signature and is not plotted on the diagrams.

accompanied by intense K alteration and silicification, which


are typical of cooling systems (e.g., Dilles and Einaudi, 1992).
The marialitic scapolite (Benavides et al., 2007) observed locally may be a manifestation of local influx of saline nonmagmatic hydrothermal fluids during the earlier, high temperature
(>400C) magnetite mineralization (mt I). The main Cu-Au
mineralization followed widespread quartz-K feldspar veining
and probably overlapped with a hydrolysis event expressed by
sericitization. The occurrence of Au in chalcopyrite and mineralogical evidence of relatively high salinities are consistent
with Cu and Au transport as Cl complexes. The late hydrothermal stage is characterized by abundant calcite veining
and pervasive carbonatization, with hematite-martite and
locally minor sulfide. It occurred at lower temperatures
(<250C), required by the local presence of zeolites like chabasite and phillipsite (Fig. 7i; Tschernich, 1992). Apart from
the small amounts of late-stage pyrite (py III) or chalcopyrite
0361-0128/98/000/000-00 $6.00

(cp II) that formed roughly coeval with calcite, calcite veins
may enclose coarse-grained, rounded to subrounded, earlier
pyrite and/or chalcopyrite (cp I) fragments, which may have
been erroneously interpreted as intergrowth textures in the
past (e.g., Benavides et al., 2007). The intense and widespread hydrothermal brecciation, the abundant open spacefilling, and the large calcite veins suggest that the mineralization occurred at relatively shallow crustal levels. This would
have potentially allowed a significant influx of basin-derived
nonmagmatic fluids that may have been relatively cool and/or
oxidized. The stable isotope data of Benavides et al. (2007)
are permissive for involvement of nonmagmatic fluids, although they are also compatible with a significant magmatic
fluid component throughout the ore-related hydrothermal
activity, based on the temperatures inferred here. A fluid in
equilibrium with magnetite at 500C has a calculated 18O
composition between 7.6 and 9.6 per mil 18O (fractionation

1289

1290

RIEGER ET AL.
TABLE 3. Major and Trace Element Concentrations in Rocks of the Mantoverde District

Sample
Location
Alteration/rock type
UTM North
UTM East
Altitude/
drill core-depth
Major elements (wt %)
SiO2
TiO2
Al2O3
Fe2O3 total
MnO
MgO
CaO
Na2O
K2O
P2O5
LOI
Total
Trace elements (ppm)
Cs
Tl
Ba
Th
U
Hf
Ta
Nb
Y
Zr
Rb
Sr
Sc
V
Cr
Co
Ni
Cu
Zn
Mo
Ag
Pb
Bi
Be
In
Ga
Ge
As
Sn
Sb
La
Ce
Pr
Nd
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu

MV96-1 (WR)
Mantoverde
Norte

MV07450 (WR)
Mantoverde
Norte

Tectonic breccia
7062500
369000
970 m a.s.l.

Tectonic breccia
7062195
369717
DDH01DS10522 m

31.71
0.60
10.44
39.53
0.04
5.61
0.23
0.16
3.92
0.23
3.23
95.71

49.73
0.99
18.32
6.67
0.11
4.28
2.58
1.37
10.45
0.17
4.80
99.48

<1.0
740.0
7.8
<0.5
2.0
<0.5
8.0
63.0
87.0
40.0
20.1
303.0
74.0
44.0
90.0
2037.0
2.0
<2.0
<0.4
<5.0
<5.0
<2.0

8.0
2.0
434.0
692.0
171.0
21.3
3.3
<0.5

0.8
0.1

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

<0.5
0.4
4319.0
0.9
1.4
2.1
0.3
3.0
11.0
94.0
152.0
198.0
19.0
196.0
<20.0
14.0
30.0
20.0
40.0
<2.0
<0.5
<5.0
<0.4
<1.0
<0.2
18.0
1.0
<5.0
1.0
0.9
9.4
12.4
1.4
5.2
1.2
0.3
1.6
0.3
1.8
0.4
1.4
0.3
1.7
0.3

MV96-5 (WR)
Mantoverde
Norte
Specularite
breccia
7062500
369000

MV96-5-1 (F)
Mantoverde
Norte
Specularite
breccia
7062500
369000

970 m a.s.l.

970 m a.s.l.

28.66
0.55
8.73
51.44
0.03
2.31
0.04
0.12
4.42
0.04
1.97
98.31

54.80
0.57
17.10
8.16
0.05
3.00
0.07
0.26
11.35
0.07
2.53
97.96

32.60
0.57
10.23
45.17
0.04
2.55
0.04
0.12
5.48
0.05
2.03
98.86

<1.0

<1.0

<1.0

740.0
2.4
1.5
1.0
<0.5

2100.0
4.8
3.0
2.0
<0.5

1000.0
2.7
1.7
1.0
<0.5

4.0
50.0
86.0
40.0
24.9
139.0
75.0
40.0
34.0
6615.0
2.0
<2.0
0.4
<5.0
<5.0
<2.0

8.0
133.0
194.0
110.0
20.6
57.0
63.0
67.0
46.0
7249.0
16.0
<2.0
0.8
5.0
<5.0
<2.0

4.0
57.0
91.0
51.0
24.3
118.0
66.0
31.0
38.0
6224.0
2.0
<2.0
0.4
<5.0
<5.0
<2.0

5.3

5.0

5.3

2.2
116.0
217.0

1.3
260.0
528.0

1.9
145.0
271.0

68.0
10.5
1.8

186.0
25.0
4.4

86.0
12.7
2.2

<0.5

<0.5

<0.5

1.3
0.2

2.6
0.5

1.1
0.3

1290

MV96-5-2 (Matrix) MV07485 (WR)


Mantoverde
Mantoverde
Norte
Norte
Specularite
Specularite
breccia
breccia
7062500
7062396
369000
369544
DDH00DS05970 m a.s.l.
310 m
56.90
0.18
12.43
8.09
0.23
1.76
2.06
1.34
8.67
0.06
7.75
99.47
0.5
0.3
1049.0
5.2
1.8
2.8
0.4
2.0
25.0
129.0
126.0
34.0
5.0
36.0
<20.0
46.0
30.0
310.0
<30.0
<2.0
<0.5
<5.0
<0.4
<1.0
<0.2
8.0
1.0
14.0
1.0
3.3
5.8
13.1
2.0
8.9
2.5
0.8
3.7
0.6
4.1
0.9
2.7
0.4
2.5
0.4

MV07427 (F)
Mantoverde
Norte
Specularite
breccia
7062195
369717
DDH01DS10221 m
69.51
0.28
13.19
1.66
0.07
0.44
1.30
0.20
10.27
0.03
2.54
99.49
0.8
0.2
1524.0
10.8
1.7
5.1
1.1
5.0
5.0
188.0
151.0
90.0
4.0
33.0
<20.0
44.0
<20.0
130.0
<30.0
10.0
<0.5
6.0
<0.4
<1.0
<0.2
10.0
1.0
<5.0
1.0
2.5
3.6
6.3
0.7
2.7
0.6
0.2
0.5
<0.1
0.5
0.1
0.5
0.1
0.7
0.1

1291

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE


TABLE 3. (Cont.)
Sample
Location
Alteration/rock type
UTM North
UTM East
Altitude/
drill core-depth
Major elements (wt %)
SiO2
TiO2
Al2O3
Fe2O3 total
MnO
MgO
CaO
Na2O
K2O
P2O5
LOI
Total
Trace elements (ppm)
Cs
Tl
Ba
Th
U
Hf
Ta
Nb
Y
Zr
Rb
Sr
Sc
V
Cr
Co
Ni
Cu
Zn
Mo
Ag
Pb
Bi
Be
In
Ga
Ge
As
Sn
Sb
La
Ce
Pr
Nd
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu

MV07514 (F)
Mantoverde
Norte
Specularite
breccia
7062449
369769
DDH99MV03520 m

MV07526 (F)
Mantoverde
Norte
Specularite
breccia
7062545
369133
DDH07N05155 m

MV96-4 (WR)
Mantoverde
Norte
Specularite
stockwork
7062500
369000

54.30
0.58
17.08
5.35
0.11
1.14
2.01
0.18
13.42
0.22
5.45
99.84

56.94
0.65
16.35
6.45
0.02
1.93
0.06
0.26
12.20
0.03
1.93
96.82

1.0
0.3
1931.0
1.9
3.9
4.1
0.6
6.0
7.0
159.0
204.0
64.0
4.0
34.0
< 20.0
196.0
< 20.0
< 10.0
< 30.0
< 2.0
< 0.5
< 5.0
< 0.4
< 1.0
< 0.2
10.0
1.0
9.0
3.0
1.8
4.3
9.3
1.1
4.4
1.0
0.3
1.1
0.2
1.1
0.2
0.9
0.2
1.3
0.3

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

0.6
0.3
1814.0
2.5
1.8
2.9
0.6
3.0
4.0
104.0
199.0
61.0
9.0
75.0
60.0
30.0
30.0
5110.0
50.0
< 2.0
< 0.5
6.0
< 0.4
< 1.0
< 0.2
12.0
1.0
< 5.0
6.0
1.4
2.9
5.6
0.8
2.9
0.6
0.2
0.7
0.1
0.9
0.2
0.7
0.1
1.0
0.2

MV07464 (WR)
Mantoverde
Norte

MV07540 (WR)
Mantoverde
Norte

MV07335 (WR)
Mantoverde
Sur

Chl-qtz-hm
7063109
368781

970 m a.s.l.

MV07524 (F)
Mantoverde
Norte
Specularite
stockwork
7062545
369133
DDH07N0575 m

1001 m a.s.l.

Chl-qtz
7062566
369218
DDH07N04115 m

Chl-qtz-cte
7061600
369653
DDH00DS04380 m

51.06
0.76
16.16
11.18
0.10
5.05
0.05
0.24
10.13
0.03
2.80
97.55

50.56
0.81
19.70
10.37
0.11
5.38
0.06
0.09
8.11
0.12
4.34
99.65

45.73
0.83
13.29
18.76
0.11
8.34
1.71
0.14
3.64
0.14
6.22
98.90

56.73
0.68
13.90
14.93
0.04
2.40
0.37
3.90
2.60
0.15
2.73
98.41

56.26
0.58
12.44
9.22
0.08
5.21
3.71
0.15
3.21
0.12
6.61
97.58

< 1.0

1.1
0.2
767.0
1.6
1.8
2.3
0.2
3.0
17.0
87.0
197.0
33.0
45.0
219.0
110.0
22.0
50.0
7000.0
50.0
< 2.0
< 0.5
< 5.0
< 0.4
1.0
< 0.2
33.0
2.0
8.0
5.0
2.5
574.0
937.0
103.0
279.0
45.3
9.2
26.4
2.1
6.3
0.8
1.8
0.2
1.3
0.2

2100.0
1.5
< 0.5
3.0
< 0.5
9.0
79.0
155.0
108.0
30.3
150.0
96.0
43.0
46.0
1349.0
38.0
< 2.0
< 0.4
< 5.0
< 5.0
< 2.0

7.9
2.5
42.9
75.0
23.0
5.3
1.5
< 0.5

2.1
0.4

1291

< 0.5
0.3
760.0
3.3
0.7
2.2
0.2
2.0
14.0
103.0
55.0
52.0
34.0
222.0
370.0
19.0
70.0
20.0
40.0
< 2.0
< 0.5
< 5.0
< 0.4
1.0
< 0.2
26.0
2.0
6.0
2.0
1.2
57.1
114.0
14.1
45.3
8.5
2.0
5.2
0.6
2.9
0.6
1.6
0.2
1.4
0.2

1.3
< 0.1
386.0
5.4
1.2
3.4
0.8
5.0
9.0
103.0
65.0
36.0
18.0
188.0
20.0
24.0
< 20.0
290.0
< 30.0
< 2.0
< 0.5
< 5.0
< 0.4
1.0
< 0.2
14.0
1.0
5.0
2.0
4.4
9.2
21.3
2.7
9.3
2.1
0.5
2.1
0.3
1.9
0.4
1.3
0.2
1.3
0.2

< 0.5
< 0.1
363.0
10.8
1.4
5.1
1.1
9.0
18.0
182.0
54.0
27.0
13.0
90.0
50.0
28.0
40.0
800.0
< 30.0
5.0
< 0.5
< 5.0
< 0.4
< 1.0
< 0.2
20.0
1.0
55.0
1.0
< 0.5
83.8
139.0
14.2
40.3
6.0
1.1
4.1
0.6
3.1
0.6
1.9
0.3
1.6
0.2

1292

RIEGER ET AL.
TABLE 3. (Cont.)

Sample
Location
Alteration/rock type
UTM North
UTM East
Altitude/
drill core-depth
Major elements (wt %)
SiO2
TiO2
Al2O3
Fe2O3 total
MnO
MgO
CaO
Na2O
K2O
P2O5
LOI
Total
Trace elements (ppm)
Cs
Tl
Ba
Th
U
Hf
Ta
Nb
Y
Zr
Rb
Sr
Sc
V
Cr
Co
Ni
Cu
Zn
Mo
Ag
Pb
Bi
Be
In
Ga
Ge
As
Sn
Sb
La
Ce
Pr
Nd
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu

MV07337 (WR)
Mantoverde Sur
Kfs-qtz-chl
7061600
369653
DDH00DS04461 m
80.25
0.28
6.99
1.49
0.02
0.61
0.74
0.20
5.28
0.04
1.25
97.14
< 0.5
0.2
1219.0
3.5
2.8
8.0
2.6
6.0
4.0
244.0
76.0
72.0
3.0
24.0
< 20.0
129.0
< 20.0
< 10.0
< 30.0
< 2.0
< 0.5
< 5.0
0.5
< 1.0
< 0.2
6.0
1.0
< 5.0
1.0
4.2
3.6
8.1
0.9
3.0
0.6
0.5
0.8
0.1
0.8
0.2
0.8
0.1
1.0
0.2

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

MV07356 (F)
MV07368 (F)
MV07577 (WR)
Mantoverde Sur Mantoverde Sur Mantoverde Sur
Specularite
breccia
Magnetite rich rock
Kfs-qtz
7061620
7061620
7060466
369776
369776
369360
DDH01DS21DDH01DS21231 m
414 m
910 m a.s.l.
38.13
0.41
13.88
24.81
0.15
3.48
0.57
0.26
6.23
0.01
10.59
98.52
1.6
0.2
967.0
1.5
1.5
2.0
0.3
1.0
12.0
86.0
99.0
76.0
37.0
183.0
60.0
321.0
60.0
150.0
< 30.0
98.0
< 0.5
12.0
< 0.4
1.0
< 0.2
21.0
2.0
6.0
6.0
< 0.5
2.5
5.2
0.6
2.5
0.6
0.2
0.8
0.2
1.6
0.4
1.5
0.3
1.9
0.3

82.12
0.31
4.03
1.02
0.02
0.33
2.41
0.08
1.55
0.13
2.72
94.72

74.87
0.06
11.38
1.41
0.06
0.33
1.65
0.20
8.08
0.15
2.16
100.40

< 0.5
< 0.1
43.0
1.9
1.0
7.0
1.3
4.0
6.0
248.0
47.0
8.0
6.0
44.0
20.0
50.0
< 20.0
540.0
< 30.0
3.0
< 0.5
6.0
< 0.4
< 1.0
< 0.2
11.0
1.0
< 5.0
1.0
< 0.5
1.8
3.7
0.5
2.2
0.6
0.2
0.8
0.2
1.1
0.2
0.8
0.1
0.9
0.2

0.8
0.3
1037.0
1.1
1.0
1.7
2.3
4.0
6.0
30.0
127.0
49.0
3.0
7.0
< 20.0
35.0
< 20.0
120.0
< 30.0
< 2.0
30.7
< 5.0
< 0.4
< 1.0
< 0.2
6.0
1.0
< 5.0
1.0
< 0.5
2.0
3.6
0.5
2.3
0.7
0.2
1.0
0.2
1.3
0.3
0.7
0.1
0.6
0.1

1292

MV07176 (WR)
Manto Ruso
Specularite
stockwork
7065659
369567
DDH01GT04315 m

MV07206 (WR)
Manto Ruso
Specularite
breccia
7065614
369351
DDH01GT07106 m

MV07135 (F)
Manto Ruso
Specularite
stockwork
7065659
369567
DDH01GT04129 m

62.23
0.11
10.33
6.14
0.08
0.96
7.25
0.69
3.67
0.10
7.53
99.10

46.70
0.73
15.86
0.94
0.15
0.44
12.26
0.19
10.32
< 0.01
11.15
98.74

76.09
0.29
12.46
1.67
0.00
0.29
0.06
1.00
5.73
0.01
1.91
99.51

0.5
0.4
213.0
8.9
1.2
2.2
0.5
2.0
67.0
88.0
104.0
24.0
4.0
17.0
< 20.0
43.0
< 20.0
2540.0
< 30.0
< 2.0
< 0.5
< 5.0
< 0.4
< 1.0
< 0.2
12.0
< 1.0
< 5.0
< 1.0
< 0.5
23.0
45.7
5.4
18.2
3.8
1.0
4.8
1.0
7.8
2.0
7.3
1.2
7.0
0.9

0.7
0.3
1101.0
3.0
0.6
1.8
0.4
4.0
20.0
76.0
186.0
53.0
18.0
98.0
50.0
10.0
< 20
90.0
< 30.0
< 2.0
< 0.5
< 5.0
< 0.4
< 1.0
< 0.2
14.0
< 1.0
< 5.0
2.0
< 0.5
14.3
27.9
3.9
15.2
3.8
1.3
3.8
0.6
3.2
0.7
2.0
0.3
2.2
0.3

0.5
0.2
669.0
1.8
1.2
5.1
0.8
3.0
4.0
202.0
132.0
20.0
4.0
11.0
< 20.0
24.0
< 20.0
160.0
< 30.0
< 2.0
< 0.5
< 5.0
< 0.4
< 1.0
< 0.2
16.0
1.0
< 5.0
2.0
3.1
0.6
1.4
0.2
0.7
0.2
< 0.05
0.2
< 0.1
0.4
0.1
0.4
0.1
0.7
0.2

1293

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE


TABLE 3. (Cont.)
Sample

MV07498 (WR)

MV07641 (WR)

MV07646 (WR)

MV07076 (WR)

Mantoverde Norte
Cte-chl
7062449
369769

Mantoverde Norte
Kfs-chl-cte
7061965
370085

Mantoverde Norte
ab-ep
7061713
370784

DDH99MV03-189 m

1100 m a.s.l.

Major elements (wt %)


SiO2
TiO2
Al2O3
Fe2O3 total
MnO
MgO
CaO
Na2O
K2O
P2O5
LOI
Total

51.65
0.81
17.64
8.35
0.33
5.44
7.59
4.11
1.78
0.16
2.87
100.73

50.10
1.12
14.48
8.15
0.16
5.98
5.41
0.72
6.64
0.72
6.34
99.82

Trace elements (ppm)


Cs
Tl
Ba
Th
U
Hf
Ta
Nb
Y
Zr
Rb
Sr
Sc
V
Cr
Co
Ni
Cu
Zn
Mo
Ag
Pb
Bi
Be
In
Ga
Ge
As
Sn
Sb
La
Ce
Pr
Nd
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu

0.8
0.2
597.0
3.4
1.0
2.7
0.4
4.0
16.0
107.0
38.0
466.0
28.0
243.0
40.0
37.0
20.0
110.0
280.0
< 2.0
< 0.5
< 5.0
< 0.4
1.0
< 0.2
20.0
2.0
14.0
< 1.0
2.6
13.0
28.5
4.0
16.3
3.9
1.2
3.3
0.5
3.1
0.6
1.7
0.3
1.6
0.3

Location
Alteration/rock type
UTM North
UTM East
Altitude/
drill core-depth

< 0.5
0.2
908.0
5.8
1.0
4.8
0.6
6.0
34.0
171.0
167.0
64.0
37.0
107.0
70.0
29.0
< 20.0
480.0
70.0
< 2.0
< 0.5
< 5.0
< 0.4
< 1.0
< 0.2
15.0
1.0
16.0
1.0
3.8
19.9
46.7
6.2
24.0
5.8
1.4
5.9
1.1
7.0
1.3
3.7
0.6
3.4
0.5

East from Laura


Kfs-chl-hm
7063983
368711

MV07658 (WR)
Northwest
from Celso
Chl-cte
7066314
369573

MV07058 (WR)
Quebrada
Guamanga
Qtz-ser
7058510
369756

1169 m a.s.l.

971 m a.s.l.

995 m a.s.l.

999 m a.s.l.

55.63
1.03
18.77
2.94
0.11
3.98
10.65
5.20
0.57
0.02
0.92
99.82

50.39
0.96
14.43
15.98
0.24
5.43
0.50
0.13
5.99
0.29
4.14
98.47

54.32
2.27
12.66
11.29
0.12
3.48
4.18
3.54
0.98
0.73
4.73
98.29

85.79
0.26
4.74
2.14
0.14
0.28
1.73
0.10
1.71
0.05
2.54
99.47

< 0.5
< 0.1
175.0
3.4
0.5
2.7
0.8
5.0
16.0
99.0
11.0
407.0
33.0
101.0
40.0
31.0
< 20.0
10.0
70.0
< 2.0
< 0.5
< 5.0
< 0.4
1.0
< 0.2
19.0
1.0
< 5.0
< 1.0
2.1
5.0
9.9
1.5
7.1
2.1
0.9
2.2
0.4
2.8
0.6
1.6
0.3
1.6
0.3

0.8
0.2
1635.0
2.9
1.0
2.4
0.3
2.0
10.0
89.0
111.0
92.0
13.0
189.0
130.0
33.0
< 20.0
< 10.0
30.0
< 2.0
< 0.5
< 5.0
< 0.4
< 1.0
< 0.2
15.0
1.0
41.0
1.0
2.4
4.0
7.1
0.9
4.1
1.1
0.3
1.5
0.2
1.5
0.3
1.2
0.2
1.3
0.2

< 0.5
0.1
142.0
4.2
1.4
7.5
1.0
12.0
39.0
275.0
17.0
71.0
18.0
256.0
< 20.0
22.0
30.0
670.0
60.0
< 2.0
0.8
< 5.0
< 0.4
2.0
< 0.2
18.0
1.0
47.0
3.0
5.6
7.2
17.2
2.7
13.4
4.2
1.3
5.6
1.2
8.1
1.7
4.9
0.7
4.5
0.7

0.6
< 0.1
85.0
3.0
0.9
10.3
1.3
4.0
4.0
380.0
57.0
12.0
3.0
23.0
< 20.0
42.0
< 20.0
< 10.0
< 30.0
< 2.0
< 0.5
< 5.0
1.5
< 1.0
< 0.2
7.0
< 1.0
< 5.0
1.0
3.4
0.6
1.6
0.2
1.1
0.3
0.1
0.6
0.1
0.7
0.2
0.6
0.1
0.8
0.2

Abbreviations: WR = whole rock, F = breccia fragments, a.s.l. = above sea level, ab = albite, chl = chlorite, cte = calcite, ep = epidote, hm = hematite, Kfs
= K feldspar, qtz = quartz, ser = sericite; blank spaces = not analyzed
0361-0128/98/000/000-00 $6.00

1293

1294

RIEGER ET AL.

coefficients of Bottinga and Javoy, 1973). A fluid in equilibrium with hematite at 350C has a calculated 18O between
8.0 and 11.6 per mil (fractionation coefficients of Zheng and
Simon, 1991). A fluid in equilibrium with chlorite at 350C
has a calculated 18O composition from 6.0 to 11.6 per mil
(fractionation coefficients of Cole and Ripley, 1998) and D
composition from 12.5 to 37.5 per mil (Graham et al.,
1984). A fluid in equilibrium with calcite at 220C has a calculated 18O composition from 7.2 to 8.4 per mil (Kim and
ONeil, 1997).
The new paragenetic scheme and relationships discussed
above differ from the recently published compilation by Benavides et al. (2007), particularly in the temporal relationships of hematite and magnetite, or the timing of carbonatization, calcite veining, and sulfide mineralization. Benavides
et al. (2007) distinguish four hydrothermal stages (stages I, II,
ore stage III, and stage IV). They state that magnetite and
hematite formed almost entirely at different paragenetic
stages, assuming that all of the hematite postdates magnetite
and disregarding the zonation of IOCG systems. Our observations suggest that magnetite at depth formed coeval with
the bulk of the hematite at shallow levels (Fig. 6). Hematite
crosscutting magnetite, as mentioned by Benavides et al.
(2007), is a late and minor event (hematite II). It also cuts the
bulk of the Cu mineralization, although it may contain minor
amounts of sulfides (e.g., Fig. 7h). The occurrence of chalcopyrite is considered to be relatively late (Ore stage III of
Benavides et al., 2007 or sulfide stage, this study). However,
Benavides et al. (2007) recognize an early chalcopyrite phase
co-genetic with early magnetite (stage I chalcopyrite of Benavides et al., 2007, cited in their Table 2, samples 2425 and
2425A). They also see evidence for the presence of another
chalcopyrite generation during their stage II mineralization
(Benavides et al., 2007, see Table 2, e.g., sample 2427), which
predates the main Cu mineralization. These chalcopyrite generations are not indicated in the paragenetic scheme of Benavides et al. (2007). Therefore, pyrite and chalcopyrite form together with the iron oxides during their stages I, II, and III.
The observations may have led Benavides et al. (2007) to conclude that the sulfides formed in equilibrium with the iron oxides (see below).
Implications of sulfur isotope data
The 34S values of chalcopyrite (5.6 to 8.9) and pyrite
(0.2 and 9.4) presented here are similar to those of Benavides et al. (2007; chalcopyrite 6.6 to 10.0; pyrite 6.8 to
11.2). A compilation of the available IOCG-related sulfur
isotope data in the Mantoverde district is given in Fig. 16.
Sulfides in association with magnetite have 34S values from
6 to 3 per mil (Fig. 16b), whereas sulfides associated with
specularite vary from 1 to 11 per mil (Fig. 16c). The data suggest that there is a variation in sulfur isotope signatures of
chalcopyrite relative to the inferred fluid conduits (e.g., the
Mantoverde fault) and the level of deposition within the hydrothermal system. Chalcopyrite with negative 34S values or
values close to 0 per mil is characteristic of magnetite-rich
ores at deeper levels or specularite-rich ores in the central
parts of the orebodies (i.e., the more internal or upflow zone),
whereas chalcopyrite enriched in 34S typically occurs outside
these fluid conduits or is developed at shallower levels. In
0361-0128/98/000/000-00 $6.00

contrast to chalcopyrite, the distribution of sulfur isotopes in


pyrite within the district is more heterogeneous. The 34S values of chalcopyrite tend to decrease from north to south in
the district. This is a consequence of the tilting of the structural block east of the Mantoverde fault (Fig. 4), resulting in
downthrow of the northeastern part of the district (Orrego
and Zamora, 1991; Sanhueza and Robles, unpub. internal report, 1999) thereby exposing deeper, magnetite-rich portions
of the system in the south (Mantoverde Sur area) and shallower environments with hematite-rich ores in the northeast
(Manto Ruso area). Alternatively, the southern part of the
Mantoverde district (Mantoverde Sur) could be interpreted
as the proximal upflow zone, whereas the northern part (shallow levels of Mantoverde Norte; Laura, Kuroki, Punto 62,
Manto Ruso, Celso) represents the distal outflow zone. Sulfides in calcite veins or associated with calcite show a similar
range of sulfur isotopic compositions to those associated with
magnetite (Fig. 16d). This is consistent with the observation
that much of the chalcopyrite in calcite veins represents
transported fragments of the main chalcopyrite mineralization (cp I) from deeper levels in the fluid conduits (see
above).
Benavides et al. (2007) emphasize the role of a nonmagmatic (seawater) source for sulfur. They conclude that exotic
sulfur input is a prerequisite for economic or even subeconomic copper mineralization in Andean IOCG deposits, referring also to the nearby La Candelaria deposit and the RalCondestable deposit in Peru. This conclusion is based on the
fact that most of their sulfur isotope data is from district exploration drill holes from the more lateral or shallow
(hematite-rich) portions of the system and a consequence of
the assumption that pyrite (and chalcopyrite) formed in equilibrium with iron oxides. Benavides et al. (2007) explicitly
state in their conclusions that hematite breccias and stockworks host all significant chalcopyrite. Only three of their
samples are from the magnetite-rich ores (stage I pyrite and
chalcopyrite with 34S values from 0.5 to 2; see Benavides
et al., 2007, Table 2) that represent the bulk of the known hypogene 440 Mt Cu-Au ore resource. They propose that the
incursion of brines with a major seawater component, possibly mediated by evaporites, occurred after high temperature
Fe-K metasomatism by magmatic-hydrothermal fluids and at
the outset of a period of intense hydrolytic alteration. Benavides et al. (2007) calculate that the 34S signature of their
main Cu-stage fluid (stage III) ranges from 26.4 to 36.2 per
mil, which is much higher than the 34S value of Aptian seawater (~14; Claypool et al., 1980). Although they do not explicitly state that they invoke fluid mixing, the fact that the
dominant metal reservoir is magmatic (excluding derivation
from volcanic rocks) and that sulfur is exotic, such a model is
implied. Mixing of a metal-bearing magmatic fluid with a seawater or evaporite sulfate-bearing nonmagmatic fluid would
require sulfate reduction and should lead to low 34S values in
the more oxidized (hematite-rich assemblage) and relatively
higher 34S in more reduced (magnetite-rich assemblage) of
the system. This, however, is not the case.
We like to emphasize the fact that the main-stage Cu mineralization texturally postdates the main iron oxide formation
(Fig. 6) and sulfides in association with magnetite or in the
proximal portions of the system show relatively low 34S values,

1294

1295

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE


Sulfur isotope data of Mantoverde IOCG district
14
13

data from this work


data from Benavides
et al. (2007)

12
11
10

Frequency

9
8
7
6
5
4
3
2
1
0
-8

-7

-6

-5

-4

-3

-2

-1

34SCDT

Magnetite-dominated zone

chalcopyrite
pyrite

10

11

12

chalcopyrite
pyrite

Frequency

Frequency

Specularite-dominated zone

5
4
3

5
4
3

1
0

0
-8 -7 -6 -5 -4 -3 -2 -1 0

-8 -7 -6 -5 -4 -3 -2 -1 0

9 10 11 12

9 10 11 12

34SCDT (%0) (n=35)

34SCDT (%0) (n=20)


Calcite veins

Indistinctive altered rocks

chalcopyrite
pyrite

chalcopyrite
pyrite

Frequency

Frequency

(%0) (n= 90)

8
7

5
4
3

5
4
3

0
-8 -7 -6 -5 -4 -3 -2 -1 0

9 10 11 12

-8 -7 -6 -5 -4 -3 -2 -1 0

34SCDT (%0) (n=17)

9 10 11 12

34SCDT (%0) (n=18)

FIG. 16. a) Compilation of sulfur isotope compositions of sulfides from IOCG ores in the Mantoverde district (Benavides
et al., 2007; this study). b) Sulfides associated with magnetite have 34S values from 6 to 3 per mil, whereas c) sulfides associated with specularite vary from 1 to 11 per mil (note that the total number of analyses of sulfides associated with hematite
is almost double the amount of analyses of sulfides associated with magnetite). d) Sulfides in calcite veins or associated with
calcite show a similar range of sulfur isotopic compositions to those associated with magnetite. e) Sulfides in altered host
rocks (with no association to magnetite, hematite, or calcite) span a similar 34S compositional range as all sulfides of the previously mentioned three groups.
0361-0128/98/000/000-00 $6.00

1295

1296

RIEGER ET AL.

as compared to sulfides in the more distal, hematite-rich


zones (Fig. 16c). Magnetite does not show martitization that
could be related to reduction processes leading to sulfide formation (e.g., Bastrakov et al., 2007). Martitization, where present, instead developed during Late stage hematite veining.
The conditions at Mantoverde are broadly similar to those
described and modeled by De Haller and Fontbot (2009) for
Ral-Condestable, Peru, although the presently exposed
parts of Mantoverde represent the more distal parts of a
zoned IOCG system. Chalcopyrite and pyrite associated with
magnetite formed under relatively more reduced, weakly
acidic conditions, at temperatures below 350C, due to cooling and wall-rock reactions. The 34S composition of these
sulfides is near the sulfur isotopic composition of the fluid
(i.e., ~0 34S; Ohmoto and Rye, 1979; Ohmoto and Goldhaber, 1998), although some fractionation has occurred due
to redox reactions, resulting in strongly negative 34S values
locally at Mantoverde (Fig. 16b). Magnetite was not replaced
by sulfides because of relatively low sulfur fugacities. The
34S compositional range observed in the district may be interpreted to result from mixing of magmatic sulfur with nonmagmatic (possibly seawater evaporitic) sulfur. The exotic,
sulfur-bearing, nonmagmatic hydrothermal fluids likely predominate at shallow levels and in the lateral, typically
hematite-rich parts of the ore system.
A spatial variation in sulfur isotope ratios similar to that at
Mantoverde is recognized in the Candelaria-Punta del Cobre
district, where the sulfide mineralogy includes pyrrhotite,
suggesting that locally all sulfur was in the form of H2S
(Ohmoto, 1972). Essentially all of the sulfides at CandelariaPunta del Cobre have a restricted 34S signature, consistent
with an igneous source (3.2 to 3.1; e.g., Rabbia et al.,
1996; Marschik and Fontbot, 2001; Mathur et al., 2002, and
references therein). Sulfides with relatively high 34S ratios
(up to 7.2) occur only locally, generally at shallow levels
and/or at the periphery of the orebodies (Rabbia et al., 1996;
Ullrich and Clark, 1999). Even at Ral-Condestable, where
34S values from 32.7 to 26.3 per mil have been reported
(Ripley and Ohmoto, 1977; De Haller and Fontbot, 2009),
the presence of external seawater or evaporitic (i.e., exotic)
sulfur is not seen as a requirement for IOCG deposit formation (De Haller and Fontbot, 2009).
Implications of strontium and lead isotope data
The range of initial Sr isotope ratios of altered volcanic
rocks from the Mantoverde district largely overlaps with
those of equivalent Jurassic volcanic and coeval intrusive
rocks in the Taltal-Chaaral-Mantoverde area (Berg and Baumann, 1985; Lucassen et al., 2006), Early Cretaceous volcanic
rocks north of La Serena (Morata and Aguirre, 2003), and
Early Cretaceous plutons of the Coastal Batholith in the
northern Atacama Region (Berg and Baumann, 1985;
Marschik et al., 2003a; Fig. 14). Therefore, the hydrothermal
fluid that caused the alteration at Mantoverde may have variably equilibrated with the host rocks. However, most of the
Jurassic to Early Cretaceous igneous rocks show 87Sr/86Sri values from 0.703 to 0.705. Relatively higher 87Sr/86Sri values in
plutonic or volcanic rocks are typically explained with magma
contamination by crustal rocks or by seawater alteration (e.g.,
Lucassen et al., 2006). If the latter interpretation is valid, all
0361-0128/98/000/000-00 $6.00

but one of our samples (MV 07058) show initial Sr isotope


ratios between those of igneous and seawater Sr reservoirs
(87Sr/86Sr of Early Cretaceous seawater: 0.707235 to 0.707464;
Jones et al., 1994; Fig. 14).
The whole-rock sample MV07058 has the highest initial Sr
ratio of 0.710053, which exceeds that of Early Cretaceous seawater (Table 2). Contamination by, or direct derivation from,
a radiogenic crustal source (Triassic rocks north of Chaaral
with 87Sr/86Sri > 0.710 are reported by Berg and Baumann,
1985) seems unlikely since the analyzed rock experienced intense silicification and sericitization and consists almost entirely of secondary quartz-sericite-hematite and minor tourmaline. It has a SiO2 content of >85 wt percent, a K2O/ Na2O
value of 17, the lowest Sr content (12 ppm), and the highest
Rb/Sr (4.75) of the sample set (Table 2). Decay of Rb introduced during K metasomatism does not explain the radiogenic Sr isotope signature, given the fact that other K metasomatized rocks with much higher Rb contents or a similarly
high Rb/Sr of 4.33 do not show such a radiogenic signature.
Furthermore, a sample (MV07514) that experienced intense
pervasive K feldspar alteration, causing very high K2O and Rb
contents, has the lowest 87Sr/86Sri of 0.703054 (i.e., an igneous
Sr signature; Table 2). Although some formation waters may
show similarly high Sr isotope ratios as sample MV07058
(Moldovanyi et al., 1993), we conclude that this Sr isotope
value is an outlier.
The Sr isotope signature of hydrothermal calcite (87Sr/86Sri
= 0.703628 to 0.703708) is fairly homogeneous and consistent
with an igneous Sr source or fluids equilibrated with such a
source. The O and C isotope characteristics of this calcite
(18O from 11.3 to 12.5, and 13C of 5 and 5.3; Benavides et al., 2007) and of similar calcite at Carola mine in the
Punta del Cobre IOCG district are permissive for calcite derivation from magmatic-hydrothermal fluids or fluids equilibrated with magmatic silicates (Rabbia et al., 1996). Overall,
the Sr isotope data are compatible with a cooling magmatichydrothermal fluid that caused K metasomatism and mixed
with meteoric fluids or with fluids with more radiogenic Sr,
such as coeval seawater. The latter interpretation is based on
the assumption that the pristine igneous Sr initial isotope signature was between 0.703 and 0.705 and that the hydrothermal fluid caused Sr initial ratios greater than 0.705 in the altered rocks.
All but three chalcopyrite samples (MV07337, MV07286
and MV08002) from the Manto Ruso, Mantoverde Norte,
and Mantoverde Sur orebodies show a relatively narrow range
of Pb isotope compositions (206Pb/204Pb = 18.50218.713;
207Pb/204Pb =15.55515.597; 208Pb/204Pb = 38.23038.529).
They define fields in 206Pb/204Pb versus 207Pb/204 Pb and
206Pb/204Pb versus 208Pb/204Pb diagrams that largely overlap
with those of the Early Cretaceous plutonic rocks in the La
Candelaria-Punta del Cobre district near Copiap and are
similar to those of sulfides from this district (Marschik et al.,
2003b; Fig. 15). Although they tend to plot close to, or overlap with, the fields of Jurassic volcanic and intrusive rocks in
the Chaaral and Taltal area (Lucassen et al., 2006), the Pb
isotope signatures clearly differ from these rocks, which supports the hypothesis that the Pb (and by inference the Cu) is
derived from Early Cretaceous magmatism and not from the
Jurassic country rocks.

1296

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE

Three chalcopyrite and most of the pyrite samples show


anomalous radiogenic Pb signatures. The reason for this remains to be investigated. It may be related to the fact that
IOCG fluids are variably enriched in U and Th. Although U
and Th concentrations are generally low at Mantoverde and
may only occasionally exceed measurable amounts (>0.1
ppm) in the sulfides (Table 1), minute inclusions of U- and
Th-bearing phases could affect the Pb isotope systematics of
pyrite in the district, since pyrite typically has low Pb contents
(<6 ppm) compared to chalcopyrite (measured Pb up to 55
ppm). However, a systematic relationship between a radiogenic Pb component as a function of U and/or Th contents
in pyrite could not be recognized in our sample set.
Conclusions and a Revised Ore Genesis Model
The Mantoverde district, in its current exposures and drill
holes, shows features typical of the more distal or shallow facies of zoned IOCG systems. The IOCG deposits in the district are characterized by a variety of barren or mineralized
breccias, in the latter case with Cu sulfides mostly disseminated or in veinlets. They formed at a relatively shallow
crustal level under brittle conditions. The orebodies lack
large massive sulfide veins or Cu sulfide-cemented breccias.
Therefore, ore grades are fairly homogeneous around the
average grade and head grades of 4 percent Cu are an exception. Copper shows a close spatial association with magnetite and hematite, although the bulk of the hypogene
chalcopyrite is associated with magnetite. Hypogene Cu-Au
ores without associated iron oxides are not observed. The
district shows a north to south and vertical zonation with regard to the distribution of iron oxide minerals. Specularite
breccias and associated specularite stockworks pass into
magnetite bodies at depth, surrounded by magnetite stockworks. Low U and locally elevated REE concentrations are
characteristic of the IOCG deposits in the Mantoverde district. The REE enrichments may be explained by mobilization and redistribution of these elements within the deposits
but are not sufficiently large to require an external REE
source.
The IOCG mineralization developed under relatively oxidized conditions at the very beginning (early specularite) and
it evolved toward high temperature (>400C) and/or more reduced conditions (magnetite-mushketovite) as the system became fully established. Early specularite gave way to magnetite and mushketovite formation only at depth. At shallow
levels, the hematite prevailed and hosts the bulk of the oxidized supergene and parts of the hypogene ores.
The main Cu-Au mineralization took place after the formation of the bulk of the iron oxides (hematite and magnetite)
and before widespread calcite veining. The Cu-Au association
is consistent with metal transport as chlorine complexes (e.g.,
Seward, 1991). The mineralization occurred under relatively
low sulfur fugacities and likely at temperatures below 350C.
Chalcopyrite and pyrite in the Magnetite zone formed under
relatively reducing, weakly acidic conditions and the 34S values of chalcopyrite and pyrite are compatible with a magmatic sulfur source. A north to south and vertical zonation,
similar to that observed with respect to the iron oxides, is
also recognized in the distribution of sulfur isotope values of
chalcopyrite, which are a function of the relative distance to
0361-0128/98/000/000-00 $6.00

1297

the inferred fluid conduits and the level of deposition. The


34S compositional range observed in the district may be interpreted to result from mixing of magmatic with nonmagmatic hydrothermal, sulfur-bearing fluids. Assuming that unaltered Jurassic volcanic rocks originally had 87Sr/86Sri from
0.703 to 0.705, the nonmagmatic fluid component could have
had a seawater Sr isotope signature. However, an intense and
voluminous sodic scapolite alteration, which may be taken as
evidence for large-scale dissolution of evaporites, is not present in the Mantoverde area. Alternatively, a nonmagmatic
aqueous fluid component, which likely predominates at shallow levels and in the lateral, typically hematite-rich parts of
the ore system, could be of meteoric origin, having a Sr isotope signature close to that of the igneous host rocks. The Pb
isotope ratios of chalcopyrite are consistent with the hypothesis that the Pb (and by inference the Cu) is derived from
Early Cretaceous magmatism.
Beside Cu-Aubearing magnetite bodies, barren magnetite
(-apatite) bodies are also present in the district. Whereas the
Cu-Au ores typically show a relationship with north-northwest to northwest structures, the barren magnetite(-apatite)
bodies are controlled by splays of the north-southtrending
Atacama fault zone. The alteration related to the magnetite(apatite) bodies is similar to that associated with the Cu-Au
ores, however, the intense potassic component to alteration is
absent. This study has shown that the magnetite is texturally
distinct from that in the main Mantoverde Cu-Au bodies. The
IOCG orebodies are commonly spatially related to an early
hydrothermal breccia (Breccia Verde-Breccia Temprana), in
contrast to the former magnetite(-apatite) iron ores.
Acknowledgments
This study was financed by the German Research Foundation (DFG) and supported by AngloAmerican Chile, Mantoverde Division. Constructive discussions and contributions
by Astudillo, M. Cortz, A. Sanguinetti, M. Zuleta, and the
friendly help of L. Guerrero, L. Ramos, C. Robledo, and D.
Vliz are thanked. Financial support by the Bayerische Gleichstellungsfrderung to A.A. Rieger is gratefully acknowledged. The authors thank G. Davidson, D. Kontak, R. Xavier,
and L. Meinert for their constructive comments.
REFERENCES
Baker, T., Mustard, R., Fu, B., Williams, P.J., Dong, G., Fisher, L., Mark, G.,
and Ryan, C.G., 2008, Mixed messages in iron oxide copper gold systems
of the Cloncurry district, Australia: Insights from PIXE analysis of halogens
and copper in fluid inclusions: Mineralium Deposita, v. 43, p. 599608.
Bastrakov, E.N., Skirrow, R.G., and Davidson, G.J., 2007, Fluid evolution
and origins of iron oxide Cu-Au prospects in the Olympic Dam district,
Gawler craton, South Australia: ECONOMIC GEOLOGY, v. 102, p. 14151440.
Beaudoin, G., Dupuis, C., Gosselin, P., and Jbrak, M., 2007, Mineral chemistry of iron oxides: application to mineral exploration: Biennial SGA Meeting, 9th, Dublin, 2023 August 2007, Proceedings, p. 497500.
Benavides, J., Kyser, T. K., Clark, A.H., Oates, C.J., Zamora, R., Tarnovschi,
R., and Castillo, B., 2007, The Mantoverde iron oxide-copper-gold district,
III Regin, Chile: The role of regionally derived, nonmagmatic fluids in
chalcopyrite mineralization: ECONOMIC GEOLOGY, v. 102, p. 415440.
Berg, K., and Baumann, A., 1985, Plutonic and metasedimentary rocks from
the Coastal Range of northern Chile: Rb-Sr and U-Pb isotopic systematics:
Earth and Planetary Science Letters, v. 75, p. 101115.
Berg, K., and Breitkreuz, C., 1983, Mesozoische Plutone in der nordchilenischen Kstenkordillere: Petrogenese, Geochronologie, Geochemie und
Geodynamik mantelbetonter Magmatite: Geotektonische Forschung, v. 66,
p. 1107.

1297

1298

RIEGER ET AL.

Birck, J.L., 1986, Precision K-Rb-Sr isotopic analysis: Application to Rb-Sr


chronology: Chemical Geology, v. 56, p. 7383.
Bottinga, Y., and Javoy, M., 1973, Comments on oxygen isotope geothermometry: Earth and Planetary Science Letters, v. 20, p. 250265.
Brown, M., Diaz, F., and Grocott, J., 1993, Displacement history of the Atacama fault system 2500'S-2700'S, northern Chile: Geological Society of
America Bulletin, v. 105, p. 11651174.
Chiaradia, M., and Fontbot, L., 2003, Separate lead isotope analyses of
leachate and residue rock fractions: Implications for metal source tracing in
ore deposit studies: Mineralium Deposita, v. 38, p. 185195.
Chiaradia, M., Banks, D., Cliff, R., Marschik, R., and De Haller, A., 2006,
Origin of fluids in iron oxide-copper-gold deposits: Constraints from 37Cl,
87Sr/86Sr and Cl/Br: Mineralium Deposita, v. 41, p. 565573.
i
Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, H., and Zak, I., 1980, The
age curves of sulfur and oxygen isotopes in marine sulfate and their mutual
interpretation: Chemical Geology, v. 28, p. 199260.
Cole, D.R., and Ripley, E.M., 1998, Oxygen isotope fractionation between
chlorite and water from 170-350C: A preliminary assessment based on
partial exchange and fluid/rock experiments: Geochimica et Cosmochimica
Acta, v. 63, p. 449457.
Cornejo, P., Matthews, S., Orrego, M., and Robles, W., 2000, Etapas de mineralizacin asociadas a alteracin potsica en un sistema de Fe-Cu-Au: Yacimiento Mantoverde, III. Regin de Atacama, Chile: Congreso Geolgico
Chileno, 9th, Puerto Varas, 31 July 4 August 2000, Actas, p. 97101.
Dallmeyer, R.D., Grocott, J., Brown, M., Taylor, G.K., and Treloar, P.J., 1996,
Mesozoic magmatic and tectonic events within the Andean plate boundary
zone, 26-2730'S, North Chile: Constraints from 40Ar/39Ar mineral ages:
Journal of Geology, v. 104, p. 1940.
De Haller, A., and Fontbot, L., 2009, The Ral-Condestable iron oxide
copper-gold deposit, central coast of Peru: Ore and related hydrothermal
alteration, sulfur isotopes, and thermodynamic constraints: ECONOMIC GEOLOGY, v. 104, p. 365384.
Dilles, J.H., and Einaudi, M.T., 1992, Wall-rock alteration and hydrothermal
flow paths about the Ann Mason porphyry copper deposit, Nevada: A 6-km
vertical reconstruction: ECONOMIC GEOLOGY, v. 87, p. 10632001.
Edfelt, ., Armstrong, R.N., Smith, M., and Martinsson, O., 2005, Alteration
paragenesis and mineral chemistry of the Tjarrojakka apatite-iron and Cu
(-Au) occurrences, Kiruna area, northern Sweden: Mineralium Deposita, v.
40, p. 409434.
Espinoza, C., Gelcich, S., Vivallo, W., 1999, Yacimientos metalferos de la
Hoja Quebrada Salitrosa, Regin de Atacama: Santiago, Chile, Servicio Nacional de Geologa y Minera, Mapa de Recursos Minerales de Chile, scale
1:100.000.
Espinoza, R.S., 1985, Distribucin de elementos de transicin en magnetitas
de yacimientos de hierro Chilenos y su significado en la tipologa y gnesis
probable de los yacimientos: Congreso Geolgico Chileno, 4th, Antofagasta,
Actas, p. 3/8363/853.
Ewart, E., 1982, The mineralogy and petrology of Tertiary-Recent orogenic
volcanic rocks: With special reference to the andesitic-basaltic compositional range, in Thorpe, R.S., ed., Andesites: orogenic andesites and related
rocks: Chichester, U.K., John Wiley and Sons, p. 2587.
Gelcich, S., Davis, D.W., and Spooner, E.T.C., 2005, Testing the apatitemagnetite geochronometer: U-Pb and 40Ar/39Ar geochronology of plutonic rocks, massive magnetite-apatite tabular bodies, and IOCG mineralization in Northern Chile: Geochimica et Cosmochimica Acta, v. 69, p.
33673384.
Gradstein, F.M., Agterberg, F.P., Ogg, J.G., Hardenbol, J., Van Veen, P.,
Thierry, J., and Huang, Z., 1995, A Triassic, Jurassic and Cretaceous time
scale: Society for Sedimentary Geology Special Publication 54, p. 95126.
Graham, C.M., Atkinson, J., and Harmon, R.S., 1984, Hydrogen isotope fractionation in the system chlorite-water: NERC 6th Progress Report of Research 1981-1984, NERC Publication Series D, n. 25, p. 139.
Grant, J.A., 2005, Isocon analysis: A brief review of the method and applications: Physics and Chemistry of the Earth, v. 30, p. 9971004.
Gustafson, L.B., and Hunt, J.P., 1975, The porphyry copper deposit at El Salvador, Chile: ECONOMIC GEOLOGY, v. 70, p. 875912.
Hitzman, M.W., Oreskes, N., and Einaudi, M.T., 1992, Geological characteristics and tectonic setting of Proterozoic iron oxide (Cu-U-Au-REE) deposits: Precambrian Research, v. 58, p. 241287.
Horwitz, E.P., Chiarizia, R., and Dietz, M.L., 1992, A novel strontium-selective extraction chromatographic resin: Solvent Extraction and Ion Exchange, v. 10, p. 313336.
0361-0128/98/000/000-00 $6.00

Jones, C.E., Jenkyns, H.C., Coe, A.L., and Hasselbo, S.P., 1994, Strontium
isotopic variations in Jurassic and Cretaceous seawater: Geochimica et Cosmochimica Acta, v. 58, p. 30613074.
Kendrick, M.A., Honda, M., Gillen, D., Baker, T., and Phillips D., 2008, New
constraints on regional brecciation in the Wernecke Mountains, Canada,
from He, Ne, Ar, Kr, Xe, Cl, Br and I in fluid inclusions: Chemical Geology, v. 255, p. 3346.
Kim, S.T., and ONeil, J.R., 1997, Equilibrium and nonequilibrium oxygen
isotope effects in synthetic carbonates: Geochimica et Cosmochimica Acta,
v. 61, p. 34613475.
Kolb, J., Meyer, M., Prantl, S., Sindern, S., Sakellaris, G.A., Vennemann, T.,
Bttcher, M.E., 2009, Characteristics of hydrothermal fluids forming the
Guelb Moghrein Fe Oxide-Cu-Au-Co deposit, Mauritania: Ore mineral
chemistry, fluid inclusions and isotope geochemistry, in Porter, T.M., ed.,
Hydrothermal iron-oxide copper-gold & related deposits: A global perspective: Adelaide, Australian Mineral Foundation, in press.
Lara, L., and Godoy, E., 1998, Hoja Quebrada Salitrosa, III Regin de Atacama: Santiago, Chile, Servicio Nacional de Geologa y Minera, Mapas Geolgicos 4, scale 1:100.000.
Lucassen, F., Kramer, W., Bartsch, V., Wilke, H.G., Franz, G., Romer, R.L.,
and Dulski, P., 2006, Nd, Pb, and Sr isotope composition of juvenile magmatism in the Mesozoic large magmatic province of northern Chile
(1827S): Indications for a uniform subarc mantle: Contributions to Mineralogy and Petrology, v. 152, p. 571589.
Mark, G., and Forster, D.R.W., 2000, Magmatic-hydrothermal albite-actinolite-apatite-rich rocks from the Cloncurry district, NW Queensland, Australia: Lithos, v. 51, p. 223245.
Marschik, R., and Fontbot, L., 1996, Copper(-iron) mineralization and
superposition of alteration events in the Punta del Cobre belt, northern
Chile: Society of Economic Geologists, Special Publication 5, p.
171189.
2001, The Candelaria-Punta del Cobre iron oxide Cu-Au(-Zn-Ag) deposits, Chile: ECONOMIC GEOLOGY, v. 96, p. 17991826.
Marschik, R., Leveille, R.A., and Martin, W., 2000, La Candelaria and the
Punta del Cobre district, Chile: Early Cretaceous iron oxide Cu-Au(-ZnAg) mineralization, in Porter, T.M., ed., Hydrothermal iron-oxide coppergold & related deposits: A global perspective, Adelaide, Australian Mineral
Foundation, p. 163175.
Marschik, R., Fontignie, D., Chiaradia, M., and Voldet, P., 2003a, Geochemical and Nd-Sr-Pb-O isotope characteristics of granitoids of the Early Cretaceous Copiap plutonic complex (2730'S), Chile: Journal of South
American Earth Sciences, v. 16, p. 381398.
Marschik, R., Chiaradia, M., and Fontbot, L., 2003b, Implications of Pb isotope signatures of rocks and iron oxide Cu-Au ores in the Candelaria-Punta
del Cobre district, Chile: Mineralium Deposita, v. 38, p. 900912.
Marschik, R., Spikings, R., and Kuscu, I., 2008, Geochronology and stable
isotope signature of alteration related to hydrothermal magnetite ores in
Central Anatolia, Turkey: Mineralium Deposita, v. 43, p. 111124.
Mathur, R.D., Marschik, R., Ruiz, J., Munizaga, F., Leveille, R.A., and Martin, W., 2002, Age of mineralization of the Candelaria iron oxide Cu-Au deposit and the origin of the Chilean iron belt based on Re-Os isotopes: ECONOMIC GEOLOGY, v. 97, p. 5971.
Moldovanyi, E.P., Walter, L.M., and Land, L.S., 1993, Strontium, boron, oxygen, and hydrogen isotope geochemistry of brines from basal strata of the
Gulf Coast sedimentary basin, USA: Geochimica et Cosmochimica Acta, v.
57, p. 20832099.
Monteiro, L., Xavier, R., de Carvalho, E., Hitzman, M.W., Johnson C.A., de
Souza Filho C.R., and Torresi, I., 2008, Spatial and temporal zoning of hydrothermal alteration and mineralization in the Sossego iron oxide-coppergold deposit, Carajs Mineral Province, Brazil: Paragenesis and stable isotope constraints: Mineralium Deposita, v. 43, p. 129159.
Morata, D., and Aguirre, L., 2003, Extensional Lower Cretaceous volcanism
in the Coastal Range (2920'30S), Chile: Geochemistry and petrogenesis:
Journal of American Earth Sciences, v. 16, p. 459476.
Mote, T.I., Becker, T.A., Renne, P., and Brimhall, G.H., 2001, Chronology of
exotic mineralization at El Salvador, Chile, by 40Ar/39Ar dating of copper
wad and supergene alunite: ECONOMIC GEOLOGY, v. 96, p. 351366.
Mpodozis, C., and Ramos, V., 1990, The Andes of Chile and Argentina, in Ericksen, G.E., Caas-Pinochet, M.T., and Reinemund, J.A., eds., Geology of
the Andes and its relation to hydrocarbon and mineral resources: Houston,
Texas, Circum-Pacific Council for Energy and Mineral Resources, Earth
Sciences Series, v. 11, p. 5991.

1298

HYPOGENE IOCG MINERALIZATION IN THE MANTOVERDE DISTRICT, NORTHERN CHILE


Nakamura, N., 1974, Determination of REE, Ba, Fe, Mg, Na, and K in carbonaceous and ordinary chondrites: Geochimica et Cosmochimica Acta, v.
38, p. 757775.
Ohmoto, H., 1972, Systematics of sulfur and carbon isotopes in hydrothermal
ore deposits: ECONOMIC GEOLOGY, v. 67, p. 551578.
Ohmoto, H., and Rye, R.O., 1979, Isotopes of sulfur and carbon, in Barnes,
H.L., ed., Geochemistry of hydrothermal ore deposits, 2nd ed.: New York,
Wiley and Sons, p. 509567.
Ohmoto, H., and Goldhaber, M.B., 1997, Sulfur and carbon isotopes, in
Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed.: New
York, Wiley and Sons, p. 517611.
Oliver, N.H.S., Cleverley, J. S., Mark, G., Pollard, P. J., Fu, B., Marshall, L.
J., Rubenach, M. J., Williams , P. J., and Baker, T., 2004, Modeling the role
of sodic alteration in the genesis of iron oxide-copper-gold deposits, eastern
Mount Isa block, Australia: ECONOMIC GEOLOGY, v. 99, p. 11451176.
Oreskes, N., and Einaudi, M.T., 1990, Origin of rare earth element-enriched
hematite breccias at the Olympic Dam Cu-U-Au-Ag deposit, Roxby
Downs, South Australia: ECONOMIC GEOLOGY, v. 85, p. 128.
Orrego, M., and Zamora, R., 1991, Manto Ruso: Un yacimiento de cobre ligado a la Falla de Atacama, Norte de Chile: Congreso Geolgico Chileno,
6th, Via del Mar, Extended Abstracts, 1991, p. 174178.
Rabbia, O.M., Frutos, J., Pop, N., Isache, C., Sanhueza, V., and Edelstein, O.,
1996, Caractersticas isotpicas de la mineralizacin de Cu(-Fe) de Mina
Carola, distrito minero Punta del Cobre, norte de Chile: 8th Congreso Geolgico Argentino, Buenos Aires, Actas, p. 241254.
Requia, K., Stein, H., Fontbot, L., and Chiaradia, M., 2003, Re-Os and PbPb geochronology of the Archean Salobo iron oxide copper-gold deposit,
Carajs mineral province, northern Brazil: Mineralium Deposita, v. 38, p.
727738.
Ripley, E.M., and Ohmoto, H., 1977, Mineralogic, sulfur isotope and fluid inclusion studies of the stratabound copper deposits at Raul mine, Peru:
ECONOMIC GEOLOGY, v. 72, p. 10171041.
Sanhueza, A., Robles, W., Cembrano, J., Orrego, M., and Pavez, A., 2000,
Origen tardimagmtico de la magnetita en el distrito Mantoverde y su asociacin a magmatismo sintectnico y exhumacin rpida en un dplex extensional de la Zona de Falla Atacama: Congreso Geolgico Chileno, 9th,
Puerto Varas, 31 July4 August 2000, Actas, p. 161165.
Scheuber, E., and Andriessen, P.A.M., 1990, The kinematic and geodynamic
significance of the Atacama fault zone, northern Chile: Journal of Structural Geology, v. 12, p. 243257.
Seward, T.M., 1991, The hydrothermal chemistry of gold, in Foster, R.P., ed.,
Gold metallogeny and exploration: Blackie, p. 3762.

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

1299

Simard, M., Beaudoin, G., Bernard, J., and Hup A., 2006, Metallogeny of
the Mont-de-lAigle IOCG deposit, Gasp Peninsula, Qubec, Canada:
Mineralium Deposita, v. 41, p. 607636.
Tassara, A., Roperch, P., and Pavez, A., 2000, Paleomagnetsmo de los yacimientos Mantos Blancos y Manto Verde: Implicancias tectnicas y cronolgicas: Congreso Geolgico Chileno, 9th, Puerto Varas, 31 July4 August
2000, Actas, p. 166170.
Taylor, S.R., and McLennan, S.M., 1985, The Continental Crust; Its composition and evolution; An examination of the geochemical record preserved
in sedimentary rocks: Oxford, Blackwell, 312 p.
Tschernich, R., 1992, Zeolites of the world: Phoenix, Arizona, Geoscience
Press Inc., 563 p.
Ullrich, T.D., and Clark, A.H., 1999, The Candelaria copper-gold deposit,
Regin III, Chile: Paragenesis, geochronology and fluid composition, in
Stanley, C.J. et al., eds., Mineral deposits: Processes to processing: Rotterdam, Balkema, p. 201204.
Vila, T., Lindsay, N., and Zamora, R., 1996, Geology of the Manto Verde Copper Deposit, northern Chile: A specularite-rich, hydrothermal-tectonic
breccia related to the Atacama fault zone: Society of Economic Geologists,
Special Publication 5, p. 157170.
Williams, P.J., Barton, M.D., Johnson, D.A., Fontbot, L., De Haller, A.,
Mark, G., Oliver, N.H.S., and Marschik, R., 2005, Iron oxide-copper-gold
deposits: Geology, space-time distribution, and possible modes of origin:
ECONOMIC GEOLOGY 100th ANNIVERSARY VOLUME, p. 371405.
Wilson, J., Dallmeyer, R.D., and Grocott, J., 2000, New 40Ar/39Ar dates from
the Las Tazas Complex, northern Chile: Tectonic significance: Journal of
South American Earth Sciences, v. 13, p. 115122.
Zamora, R., and Castillo, B., 2001, Mineralizacin de Fe-Cu-Au en el distrito
Mantoverde, Cordillera de la Costa, III Regin de Atacama, Chile: Congreso Internacional de Prospectores y Exploradores (ProExplo 2001), 2nd,
Lima, Per, Abstracts. (CD-ROM).
Zartman, R.E., and Doe, B.R., 1981, Plumbotectonics. The model: Tectonophysics, v. 75, p. 135162.
Zentilli, M., 1974, Geological evolution and metallogenetic relationships in
the Andes of northern Chile between 26 and 29 S: Unpublished Ph.D.
thesis, Kingston, Ontario, Queens University, 394 p.
Zheng, Y.F., and Simon, K., 1991, Oxygen isotope fractionation in hematite
and magnetite: A theoretical calculation and application to geothermometry of metamorphic iron-formation: European Journal of Mineralogy, v. 3,
p. 877886.

1299

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

1300

Anda mungkin juga menyukai