Sie sind auf Seite 1von 26

2007 Society of Economic Geologists, Inc.

Economic Geology, v. 102, pp. 415440

The Mantoverde Iron Oxide-Copper-Gold District, III Regin, Chile:


The Role of Regionally Derived, Nonmagmatic Fluids in Chalcopyrite Mineralization
JORGE BENAVIDES,,* T. K. KYSER, ALAN H. CLARK,
Department of Geological Sciences and Geological Engineering, Queens University, Kingston, Ontario, Canada K7L 3N6

CHRISTOPHER J. OATES,
Geochemistry Division, Anglo American plc, 20 Carlton House Terrace, London, United Kingdom SW1Y 5AN

RICHARD ZAMORA, RAL TARNOVSCHI, AND BORIS CASTILLO**


Anglo American Chile Ltda, Avenida Pedro de Valdivia 291, Santiago, Chile

Abstract
Located in the Cordillera de la Costa of northern Chile, the mines of the Mantoverde district exploit supergene oxide ore developed over several Lower Cretaceous, hematite-rich, iron oxide-Cu-Au (IOCG) deposits
with an average protore grade of 0.52 percent Cu and 0.11 g/t Au (e.g., Mantoverde proper, Manto Ruso). The
geologic setting and genesis of this productive IOCG district are clarified herein through regional petrologic and
lithogeochemical study and light stable isotope analysis of paragenetically constrained samples from Mantoverde
and its satellite deposits. Together with chalcopyrite-bearing, but subeconomic, bodies of metasomatic magnetite (e.g., Montecristo and Franco) and Cu-barren magnetite-fluorapatite-pyrite bodies (e.g., Ferrfera), the
deposits of the Mantoverde district were emplaced along the main and, more commonly, subsidiary segments of
the plate boundary-parallel Atacama fault system. They are hosted by Middle to Upper Jurassic andesites of the
La Negra Formation and diorites and monzodiorites assigned to the Lower Cretaceous Sierra Dieciocho plutonic complex. Prior to mineralization, the Jurassic and Neocomian igneous rocks of this Andean transect were
subjected to moderate albitization (spilitization) and hydrolytic alteration and, subsequently, to regional, nondeformational metamorphism, which locally attained the lower greenschist facies. Both processes, however, were
focused along the western margin of a Neocomian marginal basin, 25 to 30 km east of the Atacam fault system,
and there is no evidence of widespread albitization in the vicinity of the major IOCG centers.
An extensively revised paragenetic model for Mantoverde and its satellite deposits incorporates four stages.
Stage I was dominated by widespread potassium and iron metasomatism which converted granitoid and volcanic rocks to orthoclase and magnetite, respectively. Stage II comprises chloritic and sericitic alteration and
veining. The deposition, early in stage II, of marialitic scapolite, subsequently largely replaced by chlorite, was
probably contemporaneous with regional scapolitization in the area between the Atacama fault system and the
marginal basin. Chalcopyrite deposition was restricted to the ensuing stage III, hosted by calcite veins and, particularly, specular hematite-dominated hydrothermal breccias and stockworks. Stage IV barren calcite-quartz
vein swarms record the terminal hydrothermal activity. Stable isotope fractionation relationships and published
fluid inclusion microthermometry define a retrograde thermal evolution, from above ~460C in stage I,
through ~350C in stage II, to ~210 to 280C in ore stage III, and ~110 to 240C in stage IV.
The 34S values of chalcopyrite and pyrite from Mantoverde and its associated orebodies and prospects range
overall from 6.8 to +11.2 per mil, overlapping extensively. However, the narrow range, 0.6 to +2 per mil, of
34S values of pyrite associated with stage I magnetite contrasts with the much wider range, 1.2 to +9.1 per
mil, of that deposited in stage II. The compositional variability increases from +1.4 to 11.2 per mil in the mineralized assemblages of stage III, chalcopyrite generally having higher values than pyrite. The iron oxides in
the district have 18O values that vary overall from 1.9 to +4.1 per mil, the highest values, +1.4 to +4.1 per
mil, occurring in stage I metasomatic magnetite, whereas stage III hematite has lower values of 2.0 to +1.7
per mil. Estimated equilibrium 34Sfluid values increased dramatically with time, from +0.4 to +4 per mil in
stage I, through +9.1 to +14.9 per mil during stage II, to +26.4 to +36.2 per mil for the most richly mineralized hematitic breccias. Stage III hematite equilibrated with a fluid with 18O values of +3.0 to +8.0 per mil,
significantly lower than those of fluids from which stage I magnetite crystallized (i.e., +7.3 to +9.9).
The fluids responsible for barren stage I magnetite-pyrite assemblages, with 34S and 18O values close to 0
and +8 per mil, respectively, may have been products of the second boiling of granitoid magmas, possibly of
the Sierra Dieciocho complex. Markedly higher 34S and lower 18O values in stages II and, particularly, stage
III, in which all significant chalcopyrite and gold were deposited, are interpreted as evidence for the incursion
of modified seawater, possibly via evaporitic sediments. Such externally derived fluids, probably mobilized by
marginal basin inversion and recorded by the district-wide scapolitization (Na-Cl metasomatism), may have
been a prerequisite for hypogene Cu(-Au) mineralization in the Mantoverde district.

Corresponding author: e-mail, jbenavides@cambriageosciences.com


*Present address: Cambria Geosciences Inc., 3035455 West Boulevard, Vancouver, Canada V6M 3W5.
**Present address: Rmulo J. Pea no. 170, Departamento 21-B, Condominio Las Palmas, Copiap, Chile.

0361-0128/07/3662/415-26

415

BENAVIDES ET AL.

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

IQUIQUE

72

70

68

SYMBOLS
Fe oxide Cu-Au deposits
Volcanic-hosted Cu-(Ag) dep.

INA
ENT
ARG
CHILE

AFS

Magnetite-apatite deposits
Fault
Town/City
O

22 S

TOCOPILLA
Buena Esperanza (Cu-Ag)
Mantos de Luna
Cu-Ag
Michilla District
Cu-(Ag)
Naguayan Cu-(Au)
Mantos Blancos (Cu-Ag)
ANTOFAGASTA
0

OCEAN

100 km
SCALE

AFS

24 S

Julia (Cu)

AFS

Santo Domingo
Cu-(Ag)
TALTAL

26 S
CHANARAL

Cerro Negro Cu-(Au)

28 S

Punta del Cobre


(Cu-Au)
Candelaria
(Cu-Au)

ILE
AR
GE
NT
IN

COPIAPO

Fig. 2

CH

Mantoverde
Cu-(Au)

AFS

Boqueron Chanar (Fe)

Los Colorados (Fe)


VALLENAR
Algarrobo (Fe)
SOUTH AMERICA

N
EA

TIC

CHILE

AN

30 S

Romeral (Fe)

OC

Santiago

LA SERENA

FIG. 1
ARGENTINA

AT
L

Los Cristales (Fe)

PACIFIC OCEAN

Introduction
THE CENTRAL Andean orogen provides a unique context for
the clarification of the genetic factors responsible for the economic concentration of copper sulfides and gold in deposits of
the iron oxide-copper-gold (IOCG) clan. The majority of such
systems in Chile and Per are of Mesozoic age, and the geologic record provides better constraints on their geodynamic
setting and on the petrogenesis of the spatially and temporally
associated granitoid rocks than are afforded by most Precambrian IOCG provinces. Moreover, the deposits embrace the
entire spectrum of IOCG composition, from centers with negligible Cu and Au (e.g., the magnetite deposits of the Chilean
iron belt, such as El Romeral: Bookstrom, 1977), through
large magnetite deposits with proportionately minor associated Cu-Au mineralization (e.g., Carmen, Chile: Espinoza,
1990; Marcona and Pampa de Pongo, Per: Hawkes et al.,
2002), to major copper sulfide deposits in which the associated
Fe oxide mineralization is uneconomic (e.g., La Candelaria:
Ryan et al., 1995; Mantoverde: Vila et al., 1996).
A fundamental uncertainty in the evolving genetic model
for IOCG mineralization is whether it is generated by metalbearing brines exsolved from crystallizing granitoid magmas,
and therefore controlled primarily by melt-aqueous fluid
equilibria, or alternatively, whether the intervention of nonmagmatic waters is a prerequisite for chalcopyrite and gold
enrichment (Williams et al., 2005, and references therein).
Numerous authors (e.g., Sillitoe, 2003; Sillitoe and Perell,
2005) interpreted the geological and geochemical relationships of the IOCG deposits of northern Chile in entirely magmatic-hydrothermal terms and argued that their characteristic metal association (i.e., Cu, Au, Co, Ni, As, Mo, and U)
reflects the basic, dioritic to gabbroic, nature of the inferred
parental magmas. Pollard (2006) proposed that IOCG systems in the Andes and elsewhere differ from porphyry copper
deposits in that vapor saturation in parental magmas occurred
at higher pressures, owing to the abundance of CO2, and that
the evolution of the hydrothermal fluids was controlled by unmixing of the carbonic phase.
In contrast, Ullrich and Clark (1999) and Ullrich et al.
(2001) concluded that temporal changes in the sulfur and oxygen isotope compositions of the hydrothermal fluids at La
Candelaria resulted from incursion of water from contiguous
evaporitic strata of the Chaarcillo Group during emplacement of the chalcopyrite-gold ore. Their findings were consistent with the conclusion of Barton and Johnson (1996, 2000)
that many salient features of IOCG deposits are difficult to
reconcile with straightforward magmatic-hydrothermal models. Fluid mixing has been advocated by Haynes et al. (1995)
and Johnson and McCulloch (1995) for Olympic Dam, the
most Cu- and Au-rich large IOCG deposit, whereas Menuge
et al. (2002) recorded late-stage, saline, oxidized fluids of possible evaporite origin in the Pea Ridge magnetite-hematite deposit in Missouri. However, Marschik and Fontbot (2001a)
discounted the stable isotope data presented by Ullrich and
Clark (1999), a decision later supported by Pollard (2006).
In this study, we document sulfur and oxygen isotope data
for representative, paragenetically constrained, samples from
the Mantoverde mining district, located in the Coastal
Cordillera, III Regin, northern Chile (2630'40"2636'03"
S, 7017'39"7020'05" W; Figs. 1, 2). The Coastal Cordillera

PACIFI
C

416

FIG. 1. Location map of the Mantoverde district. Numerous Fe oxide (i.e.,


magnetite-apatite), Fe oxide-copper-gold, and volcanic-hosted, strata-bound
Cu (Ag) deposits, located between latitudes 22 and 27 S, are controlled by
the main or subsidiary structures in the Atacama fault system. Area of Figure
2 is also shown. After Sillitoe and Perell (2005).

416

26 45 S
7.040

7.050

7.060

7.070

26 25 S

Jigf

Q.

Jgla

tern bran

ch

Jkgm

Kglt

Jln

Kgm

Jln

Jln

Jln

Kgm

Kgm

Mantoverde

Kglt

Manto Ruso

Kglt

Kgsd

Fig. 3

Kgsd

Q. d
e

Jln

380

Kgr

Pirula

Kgr

390

Kpc

Chivato

400
o

Kpc

10 km

Kgch

Berta

Jln

Kch

Sierra Santo
Domingo district

Rodados Negros

Kpc

Palmira

Kgsm

UTM Coordinates (x 1.000)


UTM 19S PSAD 56

nga

Jln

Santa Rosa

Gua
ma

as

im

An

Kgsm

Kgsm

Q.
La
s

Ferrifera

Kgsd

Jln

Q. del Salado

Jln

70 00 W

26 45 S
7.040

Magnetic
Declination

7.050

7.060

7.070

26 25 S

7.080

LEGEND

IOCG Mine/Prospect

NW-SE lineament
Contact
Mafic dike

Reverse fault

Fault, observed; covered

SYMBOLS

Metasedimentary rocks
(Devonian-Carboniferous, DCce)

Metamorphic Rocks

Lower Jurassic Jigf (Flamenco, ca. 190-200 Ma)

Middle-Upper Jurassic
Jgla (Las Animas, ca. 150-160 Ma)

Jurassic-Cretaceous
JKgm (Cerro Moradito, ca. 140-145 Ma)

Lower Cretaceous
Kgsd (Sierra Dieciocho, 120-126 Ma)
Kglt (Las Tazas, 125-130 Ma)
Kgm (Cerro Morado, 130-135 Ma)

mid-Cretaceous
Kgsm (Sierra Merceditas, ca. 90-110 Ma)
Kgr (Remolino, ca. 90-110 Ma)
Kgch (Chivato, ca. 111-114 Ma)

Plutonic Complexes

La Negra Fm., Jln (Middle-Upper Jurassic)

Punta del Cobre Fm., Kpc (L. Cretaceous)

Chanarcillo Group, Kch (L. Cretaceous)

Alluvial deposits (Neogene-Quaternary)

FIG. 2. Regional geologic setting and location of the Mantoverde district, modified from Lara and Godoy (1998) and Godoy and Lara (1998). Black stars indicate the
main IOCG deposits and prospects in the area. AFS, MVF, and ChF refer to the Atacama fault system, Mantoverde fault, and Chivato fault, respectively. For sources
of geochronologic data for plutonic complexes, see Lara and Godoy (1998) and Godoy and Lara (1998). The age of the Punta del Cobre Formation is based on Pop et
al. (2000) and Marschik and Fontbot (2001b). Area of Figure 3 is also shown.

Jgla

Jgla

Jgla

Jgla

os

litr

Sa

DCce

DCce

DCce

entral

branch
AFS, c

AFS, wes

Salado district

AFS, eastern branch

7.080

70 25 W
360

360

Ch

370
370

417
380

MVF
390

0361-0128/98/000/000-00 $6.00
400

Sedimentary/Volcanic Rocks

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

417

418

BENAVIDES ET AL.

between latitudes 22 and 27 S hosts numerous Fe oxide


(e.g., magnetite-apatite), Fe oxide-Cu-(Au), and volcanichosted strata-bound Cu-(Ag) deposits and constitutes a distinctive metallogenic subprovince of the Central Andes (Ruiz
et al., 1965; Boric et al., 1990; Sillitoe, 2003, and references
therein; Fig. 1). The Mantoverde district is one of only three
Andean IOCG camps (including the La Candelaria-Punta del
Cobre district in Chile and Ral-Condestable in Per), which
have supported significant Cu production in recent years.
This research is a part of a regional study of a 3,500-km2 area
surrounding Mantoverde (Benavides, 2006), with the objective of establishing lithogeochemical vectors to, specifically,
Cu-rich IOCG centers.
Mining in the district is now focused in three open pits,
Manto Ruso, Mantoverde, and Mantoverde Sur (Fig. 3), operated by Anglo American plc. The district has measured reserves of 140 million metric tons (Mt) at 0.63 percent of Cu,
at a cutoff grade of 0.38 percent and an annual production of
~60,000 t of Cu, entirely from supergene oxide ore. Gold,
with an average grade of 0.4 ppm, is not recovered (C. Astudillo, pers. commun., 2005). In the Mantoverde deposit itself the geologic resource of hypogene protore is 400 Mt at
0.52 percent Cu (at a cutoff grade of 0.2%) and 0.11 ppm Au
(C. Astudillo, pers. commun., 2005).
Regional Geology
The Mantoverde district is located in an ensialic calc-alkaline volcanoplutonic arc terrane of Mesozoic age (Dallmeyer
et al., 1996; Grocott and Taylor, 2002), hosted by Devonian to
Carboniferous metasedimentary strata (Lara and Godoy,
1998; Fig. 2) and Permo-Triassic plutonic and volcaniclastic
rocks (Brown, 1991; Lara and Godoy, 1998). Both the arc and
basement are transected by the regionally extensive Atacama
fault system and widely covered by Neogene to Quaternary
alluvial and coluvial deposits (Fig. 2).
The most voluminous volcanism in the wider Mantoverde
area has been assigned to the La Negra and Punta del Cobre
Formations (Fig. 2). The former is a Middle to Upper Jurassic succession (Garca, 1967) of basaltic andesitic to andesitic
lava flows with subordinate volcaniclastic and marine sedimentary units (Lara and Godoy, 1998; Vivallo and Henrquez,
1998). This formation constitutes either fault-bounded blocks
separating the central and eastern branches of the Atacama
fault system or roof pendants in Neocomian plutons (Fig. 2).
The younger Punta del Cobre Formation comprises a thick
package of andesitic flows with intercalations of tuffs, tuffaceous sandstones, and welded tuffs, and thin beds of lithic
arenites and limestones (Lara and Godoy, 1998). A Neocomian U-Pb zircon date of 131.3 1.4 Ma (i.e., Valanginian/Hauterivian boundary age) has been reported by Pop et
al. (2000) for a near-basal andesitic member of the type section in the La Candelaria-Punta del Cobre mining district,
~120 km south of Mantoverde. In the study area, the Punta
del Cobre Formation is restricted to the eastern part of the
district (Fig. 2), where it concordantly overlies the La Negra
Formation and exhibits gradational contacts with the sedimentary Chaarcillo Group (Segerstrom and Parker, 1959;
Lara and Godoy, 1998). Exposures of the latter, preserved
east of the Mantoverde mining district proper (Fig. 2), comprise mudstones, calcareous sandstones and siltstones, chert
0361-0128/98/000/000-00 $6.00

and fossiliferous limestones with intercalations of tuffs and


conglomerates (Lara and Godoy, 1998), recording marine
sedimentation in a marginal back-arc basin. Naranjo (1978)
reported a Valanginian age (i.e., 132137 Ma) for limestones
in the Sierra Santo Domingo district area, 30 km north-northeast of the Mantoverde mining district (Fig. 2). Faunal assemblages documented by Moraga (1977) indicate a range of
ages from Berriasian to Barremian (i.e., ca. 121144 Ma) for
Chaarcillo Group strata cropping out 60 km southeast of the
Mantoverde district (location not shown in Fig. 2).
In the wider Mantoverde district (Fig. 2), the intrusive
rocks range in age from Late Triassic to mid-Cretaceous
(Dallmeyer et al., 1996; Lara and Godoy, 1998; Gelcich et al.,
2002). On the basis of Rb-Sr whole-rock isochron and U-Pb
zircon dates (Berg and Breitkreuz, 1983) and, more extensively, 40Ar/39Ar hornblende age spectra (Dallmeyer et al.,
1996), Lara and Godoy (1998) delimit the following pre-Albian intrusive complexes (Fig. 2): Flamenco monzogranites,
granodiorites, and tonalites at 190 to 200 Ma; Las Animas pyroxene quartz diorites at 150 to 160 Ma; Cerro Moradito
hornblende quartz diorites and hornblende granodiorites at
140 to 145 Ma; Cerro Morado quartz monzodiorites and granodiorites with hornblende, biotite, and pyroxene at 130 to
135 Ma; Las Tazas biotite-hornblende granodiorites at 125 to
130 Ma; Sierra Dieciocho hornblende-biotite quartz diorites
at 120 to 126 Ma. Although the main intrusive focus was displaced eastward from the Early Jurassic to the mid-Cretaceous, there was considerable areal overlap throughout the
Cretaceous (Fig. 2). The granitoid rocks are I-type (Chappell
and White, 1974), members of the magnetite series (Ishihara,
1977) and characteristic of volcanic arcs (Pearce et al., 1984).
In the western parts of the area, the La Negra Formation, the
plutonic bodies, and basement units are cut by swarms of
dioritic and/or andesitic dikes with north-south, northwest,
and northeast strikes (Fig. 2; Lara and Godoy, 1998).
Tectonic relationships
The major tectonic features in the area are assigned to the
north-south Atacama fault system (Fig. 2), which extends for
more than 1,000 km from Iquique to La Serena (Naranjo,
1987; Thiele and Pincheira, 1987; Brown et al., 1993, and references therein). Initiated in the Early Jurassic, this arc-parallel structure (Scheuber and Andriessen, 1990; Brown et al.,
1993) records a complex kinematic evolution, but dip-slip and
left-lateral strike-slip displacements predominated during the
Early Cretaceous (Brown et al., 1993, and references therein;
Dallmeyer et al., 1996). The Atacama fault system controlled
both the emplacement of the Upper Jurassic and Lower Cretaceous plutons (Grocott et al., 1994; Wilson and Grocott,
2001; Grocott and Taylor, 2002) and the development of iron
oxide-copper-gold deposits, including Mantoverde (Fig. 2)
and the sulfide-poor magnetite deposits of the Chilean iron
belt (Espinoza, 1990; Grocott and Taylor, 2002). The Chivato
fault in the southeast part of the area (Fig. 2) exhibits a reverse displacement, with a dominant northeast to northnortheast strike and northwest-directed tectonic transport
that translated the La Negra Formation over the Punta del
Cobre Formation (Fig. 2; Lara and Godoy, 1998). During the
late Neocomian, this structure behaved as a ductile shear
zone, with both dip- and strike-slip displacements, that

418

419

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

70 1739W

AFS, central branch

VOLCANIC/SEDIMENTARY ROCKS

01GT11

Neogene/Quaternary alluvial deposits

01DS07

RCH04MR66

99CN12

Manto Ruso

La Negra Formation, andesites (Middle-Upper Jurassic)

RCH04MR68

PLUTONIC ROCKS

MVF
B
01CN03

Mafic dikes
Sierra Dieciocho complex (ca. 120-126 Ma)

B
N 7.064.000

LEGEND
LITHOLOGIC UNITS

99MS01

N 7.065.000

N 7.066.000

E 371.000
26 3040S

E 370.000

E 369.000

E 368.000

E 367.000

Laura

Las Tazas complex (ca. 125-130 Ma)

01DS22

Ferrifera

00CN02
01CN04

Cerro Morado complex (ca. 130-135 Ma)

N 7.063.000

ALTERATION UNITS
Tectonic Breccia (Mantoverde body, Vila et al., 1996;
mineralized cataclastic rock along the Mantoverde Fault)

99MV06
99MS03

Green Breccia (Chlorite/quartz- bearing hydrothermal


breccia)

01DS13

00DS05

N 7.062.000

Felsic Body (Fine- grained aggregate of potassium feldspar


and quartz)

Mantoverde
main open pit C

MINERALIZATION UNITS

N 7.061.000

Sulfide-bearing stockwork of specular hematite veins


(Transition Zone, Vila et al., 1996)
Sulfide-bearing specular hematite-cemented hydrothermal
breccias (Manto Atacama, Vila et al., 1996)

01GT01
00DS02

Bodies of magnetite-pyrite (Stage I, this study)


Mantoverde Sur
N 7.060.000

01GT02

Calcite vein system

N 7.059.000
N 7.058.000

Franco

SYMBOLS
Fault: observed; covered

AFS, easter

n branch

Montecristo

96GM21

NW - SE lineaments
Drill-hole collar

01DS18

Geological cross section (Fig. 8)

MVF
Trillizos

70 2005W

E 367.000

Bodies of massive magnetite-apatite-pyrite

01DS18

01DS16

E 368.000

E 369.000

E 370.000

MAGNETIC
DECLINATION
4 59

26 3603S
E 371.000

SCALE

1,000 m

Datum: PSAD56
Projection: UTM 19S

FIG. 3. Geology of the Mantoverde mining district. Black stars indicate the locations of the collars of drill holes selected
for sulfide and iron oxide sampling. Modified from Zamora and Castillo (2001). See text for details. Note: Reverse circulation drill holes 04DT01, 04DT04, and 04DT02, not shown in the map, are located, respectively, 1.5, 2.2, and 2.3 km south
of Trillizos. AFS and MVF refer to Atacama fault system and Mantoverde fault, respectively.
0361-0128/98/000/000-00 $6.00

419

420

BENAVIDES ET AL.

controlled the emplacement of the Remolino pluton (Grocott


and Taylor, 2002: Fig. 2) as well as the small Berta and Chivato Cu-Au deposits (Fig. 2). In addition to the above structures, southwest-northeast to north-southstriking, eastverging, reverse faults to the east of the main Mantoverde
district exerted a primary control on the location of Palmira
and several other IOCG centers (Fig. 2). Finally, a series of
northwest-southeast lineaments may be shallow expressions
of episodically reactivated older structures rooted in the basement (Bonson, 1998).
In the Andes of northern Chile between 22 to 27 S, extensional tectonism during the Jurassic led to the formation of
the Tarapac basin in the foreland of the magmatic arc
(Mpodozis and Ramos, 1990). In the Early Cretaceous, however, changes in both the rate and obliquity of convergence at
the plate boundary and steepening of the subducting oceanic
slab generated a transtensional tectonic regime that resulted
in the formation of the Atacama fault system and an aborted
marginal back-arc basin (berg et al., 1984; Aguirre et al.,
1989; Mpodozis and Ramos, 1990; Grocott and Taylor, 2002).
In this setting, arc magmatic activity and sedimentation were
broadly contemporaneous (Grocott and Taylor, 2002). The
former is represented by both the Las Tazas and Sierra
Dieciocho plutonic complexes and the Punta del Cobre Formation, and the sedimentary filling by the Chaarcillo Group.
The present areal separation (Fig. 2) of Neocomian granitoid
rocks to the west and volcanic strata to the east may be an effect of erosion but probably reflects an inherent separation of
large-scale intrusive and eruptive activity. The tectonic inversion of the Neocomian basin took place in the mid-Cretaceous (Mpodozis and Ramos, 1990).
Regional alteration and metamorphism
Igneous rock compositions and alteration features have
been documented by Benavides (2006) for 460 outcrop samples from an area extending 32 km east, 50 km north, and 33
km south of the Mantoverde mine. The major alteration and
metamorphic domains are shown in Figure 4, subdivided
into early albitization and argillic alteration and later lowgrade metamorphism. Neither is associated with penetrative
deformation.
Albitization: Weak to moderate replacement of magmatic
plagioclase by albite is widespread in the andesites of the
Punta del Cobre Formation, in the eastern part of the wider
Mantoverde area (Fig. 4), whereas the La Negra Formation
volcanic rocks only locally exhibit such effects. Secondary albite initially developed as patches in plagioclase phenocrysts
(Fig. 5A), with concomitant destruction of albite twin lamellae, whereas more intense albitization resulted in complete
pseudomorphism of plagioclase crystals (Fig. 5B). However,
most associated augite phenocrysts are unaltered, with only
local marginal replacement by probably magmatic magnesiohornblende. Such selective alteration is interpreted as subocean floor spilitization. Complete replacement of the rocks by
albite, such as occurred in the Upper Andesite at La Candelaria (Ullrich and Clark, 1999), is not observed.
Weak to moderate albitization is focused in a 10- to 15-kmwide, north-south zone extending for 55 km from the Rodados Negros area to the latitude of the Cerro Negro deposit
(Figs. 2, 4), where it affects units of the La Negra Formation
0361-0128/98/000/000-00 $6.00

as well as the overlying Punta del Cobre Formation (Fig. 4).


The zone of Na metasomatism therefore rims the Neocomian
basin and lies immediately west of the outliers of Chaarcillo
Group sedimentary rocks. There is no evidence of extensive
secondary albite development in either volcanic or plutonic
rocks in the vicinity of the Atacama fault system and the major
IOCG centers along that structure (see below).
Hydrolytic alteration: The regionally developed albitic alteration in the north-south domain between Rodados Negros
and Sierra Aspera (Fig. 4) is extensively overprinted by both
intermediate argillic assemblages (illite, chlorite, and smectite:
Fig. 5C) and, particularly in the Mina Berta and Diego del Almagro areas (Fig. 4), by muscovite and quartz (Fig. 5D).
Low-grade metamorphism: As documented by Palacios
(1977), the Jurassic volcanic strata of the region are widely affected by low-grade metamorphism, comparable in all salient
aspects to that described for central Chile by Aguirre et al.
(1989). In this study, similar effects have been confirmed in
the Neocomian Punta del Cobre Formation. The most widespread metamorphic assemblages may be assigned to the
prehnite-pumpellyite facies (Alt, 1999) but locally both
prehnite-actinolite and pumpellyite-actinolite facies are represented. Characteristic mineralogical relationships in andesites of the La Negra and Punta del Cobre Formations are
shown in Figure 6. As emphasized by Frey et al. (1991) and
Alt (1999), this range of assemblages, although extending
from the sub- to the lower greenschist facies, may not record
a significant temperature range.
The metamorphic assemblages clearly developed after the
formation of secondary albite and even the superimposed hydrolytic alteration (Fig. 6). As with those processes, however,
metamorphism was most intense in the eastern part of the
study area along the western margin of the marginal basin
(Fig. 4) and only weakly affected areas adjacent to the Atacama fault system. The replacement of metamorphic actinolite by magnetite in the Santa Rosa Cu-Au prospect (Figs. 4,
7) shows that the regional metamorphism predated the initial
stages of hydrothermal alteration, at least in the minor IOCG
centers located east of the main Mantoverde district.
Geology of the Mantoverde Mining District
Host-rock units
The least altered andesites of the La Negra Formation in
the Mantoverde district proper contain plagioclase (An6070)
phenocrysts in a microcrystalline to aphanitic groundmass.
The main mafic mineral is augite, and titaniferous magnetite
is finely disseminated in the groundmass. In the northern part
of the district, amygdules are filled with calcite, chlorite, and
chalcedony (Lpez, 2002). Numerous beds of fine-grained
lithic and crystal tuffs are intercalated with the andesites. The
major plutonic rocks of the immediate district are equigranular, medium-grained, hornblende and biotite-bearing diorites
and monzodiorites assigned to the Hauterivian-Barremian
(i.e., ca. 121132 Ma) Sierra Dieciocho complex (Lara and
Godoy, 1998).
Structural relationships
Mineralization in the Mantoverde district developed within
an intensely fractured structural block delimited by the

420

400

70 00W

390

380

370

70 25W

26 06 20S

360

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

LEGEND

26 06 20

7.110

7.110

Trgc
Jln

Alluvial deposits (Neogene-Quaternary)

Jln

Cerro Negro

Kpc

Chanarcillo Group, Kch (L. Cretaceous)


Punta del Cobre Fm., Jln (L. Cretaceous)

Jln

7.100

Jln

7.090

Jln

mid-Cretaceous complexes
Kgsm (Sierra Merceditas, ca. 90-110 Ma)
Kgr (Remolino, ca. 90-110 Ma)
Kgch (Chivato, ca. 111-114 Ma)
Kglb (La Borracha, ca. 105-110 Ma)

Jln

7.090

Sierra Aspera
district

Kgsa

Carmen
Kgsa

Jln

Jgla

Plutonic Complexes

Jln

AFS, centra

Kglt

Jgla

La Negra Fm., Kpc (Middle-Upper Jurassic)


Kgsa

l branch

Jgla

DCce

AFS,
weste
rn bra

nch

7.100

Kpc

Ester

7.080

Diego de 7.080
Almagro

Jln

Salado district

Sierra Santo
Domingo district

Santa
Rosa

Q. del Salado

Jgla

Lower Cretaceous complexes


Kgsd (Sierra Dieciocho, ca.120-126 Ma)
Kgsa (Sierra Aspera, ca. 125-130 Ma)
Kglt (Las Tazas, ca. 125-130 Ma)
Kgm (Cerro Morado, ca.130-135 Ma)

Jln

Jln

Kglt

Jurassic-Cretaceous complexes
JKgm (Cerro Moradito, ca. 140-145 Ma)

Kpc
Jln

Kglt

Kpc

Kch

7.070

7.070

Middle-Upper Jurassic complexes


Jgas (Agua de Sol, ca. 150 Ma)
Jgla (Las Animas, ca. 150-160 Ma)

Kgsm
Jln

Manto Ruso

Kgsm

Kglt

Ferrifera
Mantoverde
7.060

La
sA
nim

Kgm

7.050

AFS, eastern branch

Kgsd

Kglt

Triassic plutons, Trgc (Capitana, ca. 215 Ma)

Pirula

Q.

Kpc

7.060

Metamorphic Rocks

Palmira

as

Metasediments, DCce (DevonianCarboniferous)

Kgsm
Jln

Kpc

Ch
F

Jgla

Rodados
Negros

Q. d
e

SYMBOLS

Kgr

7.050

Gua
ma

Western limit of moderate albitization

nga

Remolino
Kgsd

Jln

Kgr

Western limit of pervasive low-grade


metamorphism

Chivato

Kgr
Kgm

Berta

Zones of scapolitization: intense; moderate

Kgch

7.040

Q.
S

Jkgm

7.040

Fault, observed; covered

ali
tr

Kgr

os

Kgm

Jln

Reverse fault; NW-SE lineament

Kgr

Sierra Desiertito

390

380

370

360

70 25W

Jgas

10 km

Magnetic Declinatio n

IOCG Mine/Prospect

7.030
400

Kglb

7.030

Contact
70 00W

Ch
FS

?
26 50 35S

Sedimentary/Volcanic Rocks

26 50 35S

UTM Coordinates (x 1.000)


UTM 19S PSAD 56

FIG. 4. Geologic map showing the western boundaries of regional albitization (thick dashed line) and low-grade metamorphism (thick dashed-dotted line) in the wider Mantoverde district. Based on regional petrographic studies by Benavides
(2006). These earlier alteration events affected predominantly Neocomian and Jurassic volcanic rocks in the vicinity of the
western margin of the back-arc basin. Areas of volcanic rocks with moderate and strong development of marialitic scapolite
are also shown. Symbols for lithologic units as in Figure 2, with the addition of the following plutonic complexes: La Capitana (Trgc, Triassic), Agua de Sol (Jgas, Jurassic), Sierra Aspera and La Borracha (Kgsa and Kglb, respectively, Lower Cretaceous). Base map modified from Lara and Godoy (1998) and Godoy and Lara (1998). The area of the Mantoverde mining
district is also shown.
0361-0128/98/000/000-00 $6.00

421

421

422

BENAVIDES ET AL.

FIG. 5. Regional albitization. (A). Patchy albitization of plagioclase phenocryst in Punta del Cobre Formation andesite.
Sample 124633, Diego de Almagro sector (coordinates: x = 398,684, y = 7,083,747; transmitted light, crossed nicols). (B).
More intense albitization, accompanied by the destruction of twinning. Sample 9795, Rodados Negros sector (coordinates:
x = 391,529, y = 7,051,142; transmitted light, crossed nicols). (C). Regional hydrolysis. Na-metasomatized plagioclase phenocrysts are overprinted by assemblages dominated by clays and illite. Sample 124682, Santa Rosa mine area (coordinates: x
= 386,991, y = 7,075,031; plane-polarized transmitted light). (D). Hydrolytic assemblage dominated by muscovite and quartz
(Ms-Qtz), overprinting a moderately albitized plagioclase phenocryst. Sample 124633, Diego de Almagro sector (coordinates:
x = 398,684, y = 7,083,747; transmitted light, crossed nicols).

subvertical central and eastern branches of the Atacama fault


system (Figs. 2, 3). These are connected by the Mantoverde
fault (MVF, Fig. 3), which strikes N15 to 20 W and dips 40
to 50 E (Figs. 3, 8, 9). With their subsidiary structures, the
Mantoverde fault and the eastern branch of the Atacama fault
system exerted a strong control on the location and morphology of both barren magnetite-apatite-pyrite bodies and sulfide-bearing specular hematite-cemented breccias (Fig. 8).
The Mantoverde deposit itself has been interpreted as being
hosted by a releasing strike-slip duplex located in a transfer
zone between the central and eastern branches of the Atacama fault system (A. Sanhueza and W. Robles, 1999, Estudio
Estructural del Distrito Mantoverde, unpublished report for
Anglo American Chile, Santiago, 1999, 25 p.).
Distribution of mineralization
The Mantoverde mining district extends for 10 km in a
north-south direction and comprises several mineralized
0361-0128/98/000/000-00 $6.00

centers (Figs. 3, 8), which range from chalcopyrite- and


specular hematite-rich breccias and stockworks to massive
magnetite-pyrite and magnetite-apatite pyrite bodies. The
mineralized breccias and mineralogically identical stockworks are commonly located in the hanging wall of the Mantoverde fault and in the northern half of the district, whereas
crudely tabular, massive bodies of magnetite-pyrite occur
predominantly in the footwall of the Mantoverde fault and in
the southern half of the district (Fig. 3). Massive and irregular bodies of magnetite-apatite pyrite are developed along
the eastern branch of the Atacama fault system (Figs. 3, 8).
From north to south, the most important deposits and prospects
are Manto Ruso, a sulfide-bearing hematite-cemented
breccia pipe and associated stockwork (Fig. 8A); Laura, a sulfide-bearing specularite-cemented breccia (Fig. 8B); Ferrfera, a magnetite-apatite pyrite body (Fig. 3); Mantoverde
proper, the largest known deposit in the district, comprising
mineralized tectonic breccias, tabular chalcopyrite-rich,

422

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

423

FIG. 6. Mineral associations and textural relationships of regional, nondeformational, low-grade metamorphism. (A). Vesicle filled with prehnite (Prh) and calcite (Cal) in a fine-grained andesitic volcanic rock of the La Negra Formation, east of
Cerro Negro. Sample 124579 (coordinates: x = 382,696, y = 7,104,867; transmitted light, crossed nicols). (B). Actinolite disseminated in the matrix of an andesite (upper-right corner) and as thin veinlets cutting moderately albitized plagioclase phenocrysts. Sample 124670 (coordinates: x = 381,720, y = 7,075,744; transmitted light, crossed nicols). (C). Irregular patches
of metamorphic chlorite (Chl) in the matrix of a fine-grained andesite of the Punta del Cobre Formation. Note the moderately albitized plagioclase phenocryst altered to clays (right side), prior to the development of chlorite veinlets. Sample
124699, Sierra Santo Domingo district (coordinates: x = 394,593, y = 7,074,909; plane-polarized transmitted light). (D). Aggregate of epidote (Ep), chlorite (Chl), and actinolite (Act) in fine-grained andesite of the Punta del Cobre Formation. Sample 9769, Palmira mine area (coordinates: x = 388,499, y = 7,060,987; transmitted light, crossed nicols).

hematite-cemented, hydrothermal breccias, and stockworks


(Fig. 8C); and Montecristo and Franco-Trillizos, chalcopyritebearing subvertical veins and lenses of magnetite (Fig. 8D).
Chalcopyrite and pyrite, the dominant hypogene sulfide
minerals, are associated with minor bornite and pyrrhotite.
Hematite, chlorite, apatite, calcite, quartz, potassium feldspar, sericite, and tourmaline are the major associated minerals. Vila et al. (1996) defined three main alteration-mineralization units in the Mantoverde deposit itself (Fig. 8C).

FIG. 7. Metasomatic magnetite (Mag) of hydrothermal stage I forms irregular patches overprinting metamorphic actinolite (Act) and epidote (Ep)
in an andesite of the Punta del Cobre Formation. Sample 9815, west of the
Sierra Santo Domingo district (coordinates: x = 392,394, y = 7,067,006;
plane-polarized transmitted light).
0361-0128/98/000/000-00 $6.00

423

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

1368

1378

2469

1373

424
100

200 m

SCALE H:V
0

2447

DDH00DS05

2357
2358
2359

2470

2465

DDH01DS13

200 m

50

150

250

350

450

550

650

750

850

950

1050

m.a.s.l.
1150

50

150

250

350

1361
1362B
1363

1358

SCALE H:V

BLSP

100

2384

200 m

100 m

SCALE H:V

DDH96GM21

Mantoverde Fault

SECTION D-D (Montecristo)

700

750

800

850

900

950

m.a.s.l.
1000

150

250

350

450

550

1356

750

850

950

450

DDH01CN03

1050

650

BLSP

Mantoverde Fault

m.a.s.l.
1150

550

650

750

850

m.a.s.l.
950

SECTION B-B (Laura)

FIG. 8. Representative east-west geologic profiles in the Mantoverde district. Selected drill holes have been projected to show the locations of some analyzed samples. The bodies of metasomatic magnetite deepen in a northward direction, and the Montecristo sector (Section D-D) is inferred to represent the deepest parts of the
hydrothermal system. Most specular hematite-cemented breccias are located in the hanging wall of the Mantoverde fault. Massive bodies of magnetite-apatite pyrite
along the eastern branch of the Atacama fault system are not shown. See text for details. BLSP records the base of the supergene profile. Legend as in Figure 3. Redrawn with permission from sections prepared by the Mantoverde Geology Division, Anglo American-Chile.

BLSP

Mantoverde Fault

100

SCALE H:V

DDH01GT11

SECTION C-C, (Mantoverde deposit)

BLSP

DDH01DS07

Mantoverde Fault

SECTION A-A (West Manto Ruso)

424
BENAVIDES ET AL.

425

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

FIG. 9. Photograph of supergene copper oxide zone, showing the structural relationships of the Mantoverde and Manto Atacama breccias with respect to the Mantoverde fault, which dips moderately to the east. View to
north, Mantoverde main open pit, 960 m bench (for location see Fig. 3).

From west to east, these are (1) Mantoverde, a mineralized


tectonic breccia in the footwall of the Mantoverde fault (Fig.
9); (2) Manto Atacama, a specularite-cemented, sulfide-bearing hydrothermal breccia in the hanging wall of, and elongated parallel to, the Mantoverde fault and grading eastward
into a sulfide-bearing specular hematite vein system cutting
moderately altered andesites and diorites (the Transition zone
of Vila et al., 1996; Figs. 8C, 9); and (3) Brecha Verde, a chlorite-quartz-sericitecemented hydrothermal breccia body developed on both sides of the Mantoverde fault (Fig. 8C).

STAGE I
K and Fe
Metasomatism

Paragenetic Relationships
The paragenetic relationships of hydrothermal alteration and
hypogene mineralization in the district are herein modified
from the observations of Vila et al. (1996), Cornejo at al. (2000),
and Lpez (2002). The sequence of events is based primarily
on drill core samples and exposures in the Mantoverde mine
(Figs. 3, 8, 9), but very similar associations and crosscutting relationships are observed at Manto Ruso, Laura, Montecristo,
and Franco (Fig. 3). Four stages are distinguished (Fig. 10). Of
these, stage I, dominated by widespread potassium and iron
metasomatism of both plutonic and volcanic rocks, and stage
II, chlorite-rich hydrolytic alteration and veining (Fig. 11A-B),
preceded the emplacement of stage III chalcopyrite-bearing,
specular hematite-cemented hydrothermal breccias and stockworks (Fig. 11C-D). Barren calcite-quartz veining represents
the latest paragenetic stage at Mantoverde (stage IV; Figs. 10,
11D). The magnetite-apatite pyrite bodies cropping out
along the eastern branch of the Atacama fault system (Fig. 3)
have somewhat different mineral compositions and textural relationships and are considered separately.
Albitization: There is no evidence that intense Na metasomatism occurred along the Atacama fault system, and in this
respect Mantoverde differs markedly from other major central Andean IOCG districts (e.g., Ullrich and Clark, 1999;
Hawkes et al., 2002). Secondary albite has, however, been observed in a single clast in a chloritic hydrothermal breccia
from the Franco area (Figs. 3, 12), where it is extensively replaced by orthoclase, presumably representing stage I of the
hydrothermal paragenesis (Fig. 10).
Stage I potassium and iron metasomatism: Intense potassium metasomatism is recorded by the assemblage K-feldspar

STAGE II
Chlorite-Sericite
Quartz

Magnetite
Orthoclase
Biotite
Tourmaline
Titanite
Quartz
Muscovite
Chlorite
Anhydrite
Scapolite
Rutile
Epidote
Hematite
Pyrite
Chalcopyrite
Gold
Siderite
Calcite

ORE STAGE III


Specular HematiteChalcopyrite

STAGE IV
Late calcite
veins

(v)

FIG. 10. Paragenetic sequence of hypogene mineralization and alteration in the Mantoverde deposit (modified after Vila
et al., 1996, Cornejo et al., 2000, and Lpez, 2002). See text for details (thick line = major mineral, thin line = minor mineral). (v) = veins.
0361-0128/98/000/000-00 $6.00

425

426

BENAVIDES ET AL.

FIG. 11. Samples showing the crosscutting relationships of alteration and mineralization units in the Mantoverde district.
(A). Stage II, chlorite-bearing hydrothermal breccia (i.e., Brecha Verde: Vila et al., 1996) containing fragments of Kfeldspathized host rock. Sample 2397, drill hole 00DS18, 190 m. (B). Subangular fragments of stage I metasomatic magnetite
are enclosed in a pyrite-bearing, chlorite-dominated stage II assemblage, which in turn is cut by a stage III hematite vein
(left side). Sample 2357, drill hole 01DS13, 967.7 m. (C). Hand sample showing moderately K-feldspathized and chloritized
granitoid rock cut by stage III hematite veins and assigned to the Transition zone of Vila et al. (1996). Note the aggregates
of fine-grained chrysocolla (Ccl) in the hematitic veins. Mantoverde main open pit. (D). Stage III chalcopyrite intergrown
with hematite fills fractures in host rock altered to K-feldspar and chlorite of stages I and II, respectively. Note the thin, irregular, stage IV calcite veins. Sample 1400, drill hole 99CN12, 388 m.

+ quartz, with subordinate magnetite, titanite, and tourmaline. Although Lpez (2002) reported the presence of microcline, X-ray powder diffraction analyses indicate that orthoclase is the dominant secondary feldspar. K-feldspar
selectively replaced plagioclase, with preservation of the original textures of volcanic and plutonic host rocks. More intensely altered rocks were transformed into fine-grained,

FIG. 12. Fragment of an altered host rock included in a stage II hydrothermal breccia with chlorite-rich matrix. In the fragment, secondary albite (Ab), possibly associated with early regional albitization, is partially replaced by orthoclase (Kfs) of metasomatic stage I. A planar stage IV quartz
vein is visible in the upper right corner. Sample 124720, northwest of the
Franco sector (Fig. 3; coordinates: x = 369,480, y = 7,059,145; plane-polarized transmitted light).
0361-0128/98/000/000-00 $6.00

426

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

pink, granular aggregates (i.e., the felsic bodies of Vila et al.,


1996, and Zamora and Castillo, 2001; Figs. 3, 8, 13). Hydrothermal biotite developed more erratically in the groundmass of the andesites.
Although there is little textural evidence that magnetite
and K-feldspar formed contemporaneously, oxygen isotope
data, discussed below, and crosscutting relationships indicate
that, as proposed by Lpez (2002), both are the products of
high-temperature metasomatism by magmatic fluids. Fe
metasomatism is manifested as the progressive replacement
of the groundmass of andesites and tuffs by magnetite (Fig.
14A-C). Plagioclase phenocrysts, quartz, and lithic fragments
are generally less altered (Fig. 14A-B, D). In the Mantoverde
mine, fine-grained massive bodies of magnetite and pyrite
occur mainly in the deepest parts of the deposit, in the footwall of the Mantoverde fault (Fig. 8), although in the Franco
and Montecristo sectors, tabular and subvertical bodies of

FIG. 13. Stage I, texturally destructive, K-feldspathization in the Mantoverde district. (A). Andesitic host rock southeast of Manto Ruso replaced
by fine-grained orthoclase (Kfs). A stage II chlorite vein and patches (Chl)
and stage IV calcite (Cal) developed along the chlorite vein. Sample 1399,
drill hole 99CN12, 340 m. (B). Intense K-feldspathization of a volcanic host
rock located east of the Mantoverde main open pit (see Fig. 3). Orthoclase
(Kfs) is cut by stage II muscovite veins (Ms). Sample 2464, drill hole 01DS13,
317 m. Both in plane-polarized transmitted light.
0361-0128/98/000/000-00 $6.00

427

magnetite occur in both the footwall and hanging wall of the


Mantoverde fault (Figs. 3, 8D). Small irregular cavities in
magnetite are widely filled with chlorite, calcite, and quartz,
and magnetite bodies are cut by thin veins of chlorite (Fig.
14D), quartz, orthoclase, and calcite. Magnetite is locally replaced by hematite (Fig. 14C). The major magnetite bodies
lie progressively deeper from south to north, suggesting that
the southern part of the district has been uplifted relative to
the northern part since ore formation through the northward
tilting of the hanging-wall block of the Mantoverde fault
(Zamora and Castillo, 2001; Lpez, 2002; Figs. 3, 8).
The pyrite associated with stage I magnetite (Fig. 10) forms
<1-mm, irregular grains (Fig. 14B-C), erratically distributed
and rarely exceeding 3 vol percent. Both pyrite and magnetite
are cut by stage III chalcopyrite (Fig. 14C).
Microthermometric data provided by S. Collao and R. Ortega (1999, Mineralizacin-alteracin y Inclusiones Fluidas
en Muestras del Distrito Minero Mantoverde, Copiap, III
Regin, Chile, unpub. report for Empresa Minera Mantos
Blancos, Divisin Mantoverde, 1999, in Lpez, 2002) indicate
that Fe metasomatism (i.e., stage I, Fig. 10) was the product
of high-temperature (e.g., homogenization temperatures
from 460550C), highly saline fluids (Table 1). Because the
hydrothermal system at Mantoverde evolved at shallow levels,
indicated by hydraulic fracturing, corrections of homogenization temperatures for pressure are expected to be minimal
(Table 1).
The second type of magnetite-rich mineralization in the
Mantoverde district is represented by Ferrfera and other
centers along the eastern branch of the Atacama fault system
(Fig. 3). These bodies are texturally and compositionally distinctive (Fig. 15), comprising fine-grained, equigranular intergrowths of magnetite (6070 modal %), fluorapatite (25
35%), and pyrite (5%). Apatite occurs as equant, <0.5-mm
grains and the pyrite forms irregular and massive aggregates
with diameters of up to a few millimeters (Fig. 15) with subhedral to euhedral magnetite inclusions. Chlorite, quartz, and
calcite occur as thin veins and fill small cavities. The textural
relationships, including smooth, cuspate contacts, suggest that
magnetite, fluorapatite, and pyrite formed in equilibrium (Fig.
15). The columnar and dendritic magnetite textures which
have been interpreted by Nystrm and Henrquez (1994) as
evidence for a magmatic origin for several Cretaceous iron
deposits in northern Chile are not developed at Ferrfera.
Chalcopyrite occurs only in trace amounts at Ferrfera and in
other apatite-rich bodies, mainly replacing pyrite.
Stage II scapolite, chlorite-quartz-hematite-pyrite, and muscovite-calcite-hematite-pyrite: Crosscutting relationships indicate that K-Fe metasomatism was followed by hydrolytic alteration. The major minerals of this stage are chlorite, quartz,
and fine-grained 2M1 muscovite (sericite), with minor pyrite,
rutile, and hematite (Fig. 10). However, widespread remnants
of marialitic scapolite in chlorite aggregates (Fig. 16A) indicate
that it was originally a major phase. Scapolite deposition,
widely replacing plagioclase phenocrysts in wall-rock andesites, postdated stage I magnetite formation and we assign it
to early stage II (Fig. 10). On a wider scale (Fig. 4), marialite
is moderately to intensely developed in several areas both
along the Atacama fault system and between that structure
and the marginal basin, again superimposed on hydrothermal

427

428

BENAVIDES ET AL.

FIG. 14. Stage I iron metasomatism of volcanic rocks. (A). The matrix of a silicic volcaniclastic rock has been replaced by
magnetite (Mag), but the original texture has been preserved. Note the quartz remnants (Qtz) and stage II chlorite veins
(Chl). Sample 2358, drill hole 01DS13, 967 m (plane-polarized transmitted light). (B). Same field, in plane-polarized reflected light; the small, irregular grains of pyrite on the right are associated with magnetite. (C). Stage III chalcopyrite (Cpy)
veins cut pyrite (Py) associated with stage I magnetite (Mag). Irregular patches of hematite (Hem) replace magnetite. Sample 2425, drill hole 00DS02, 216 m (plane-polarized reflected light). (D). Fe-metasomatized andesite cut by stage II chlorite-quartz veins. Sample 2426, drill hole 00DS02, 302 m (plane-polarized transmitted light).

magnetite (Fig. 16B-C). Stage II scapolite in the Mantoverde


district may therefore represent a facies of a single, regional,
Na-Cl metasomatic event.
In addition to replacing scapolite, stage II chlorite initially
replaced mafic minerals and feldspars in plutonic rocks (Fig.
11C) and both the matrix and plagioclase phenocrysts in volcanic rocks, forming irregular patches and thin veins (Fig.
13A). The so-called Brecha Verde (Vila et al., 1996; Zamora
and Castillo, 2001; Fig. 11A), herein assigned to stage II,
comprises subangular to subrounded fragments of Kfeldspathized and silicified host rock in a pyrite and hematitebearing chlorite-quartz matrix with minor titanite, rutile,
tourmaline, and magnetite (Fig. 17A-B). Lpez (2002)
demonstrated that penninite was superseded by daphnite
with increasing alteration intensity. In this study, anhydrite
was observed at depth, filling fractures in pyrite and magnetite in association with chlorite (sample 2357, 954m, drill
0361-0128/98/000/000-00 $6.00

hole 01DS13; Figs. 3, 8C). Chloritized host rocks are cut by


thin orthoclase veins, some of which were later invaded by
specular hematite. These reopened veins are, in turn, cut by
thin, hematite-bearing quartz veins.
Crosscutting relationships indicate that sericitization occurred early in stage II, in association with quartz, chlorite,
and calcite pyrite hematite (Figs. 10, 17C-D) and replaced both relict plagioclase grains and hydrothermal orthoclase. Locally, sericite veins cut K-feldspathized host rocks
(Fig. 13B). The most intense development of sericite occurred in the Trillizos-Franco sector (Fig. 3).
Stage II pyrite is largely idiomorphic, forming either isolated grains, 0.5 to 6 mm in diameter (Fig. 11B), and with numerous inclusions of chlorite and tourmaline (Fig. 17B, D),
or irregular and massive aggregates up to a few centimeters
across, intergrown with chlorite, sericite, and quartz. It also
occurs as fracture fillings in altered host rocks and in quartz

428

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

(2)
210280
(3)
(3)
460550
(1)
Estimated temperature
References

(2)

180250
(3)

1421
Boiling (?)
3256
Boiling (?)

Notes: Fluid inclusion type: L-V = liquid-vapor, V.R.= vapor-rich, L-V-S = liquid-vapor-solid (halite-bearing)
References: (1) Collao and Ortega (1999, in Lpez, 2002), (2) Benavides (2006), (3) Vila et al. (1996), (4) Collao and Campos (2003)
1 Oxygen isotope geothermometry: muscovite-calcite pair according to the fractionation factors of Taylor (1997), sample 2394; calcite-hematite by integrating:
6 2
hem - water = 2.78 10 /T - 3.39
(ONeil et al., 1969) and hem - water = 1.63 106/T2 - 12.3 (Yapp, 1990), sample 1361 (for sample location see Table 3)

110
3040
Boiling and mixing
with meteoric waters
110240
(4)

L-V-S
140240
350
Not stated
460550

Calculated temperatures (C)1


Fluid inclusion type
Homogenization temperature (C)
Mean Th
Salinity (wt % NaCl equiv)
Inferred conditions

Stable isotope geothermometry


18O values ()

Quartz coexisting
with magnetite
Sample description

Calcite: + 12.5
Muscovite: + 9.3

Fragments in hydrothermal breccia

L-V (V.R.)
L-V-S
180260
215340
237
26
3256
Boiling, hydrostatic pressure

L-V-S
160360
240
3240

Coarse,
crystalline
calcite

230250
L-V
210280

Sulfidebearing
calcite vein

Calcite: +11.3
Hematite: 2.0

L-V
112260

Veins cutting
Cu mineralization

Microthermometry

O isotope geothermometry
Intergrown in
hydrothermal
breccia
Microthermometry
Microthermometry
Microthermometry
Method

O isotope geothermometry
Intergrown

Calcite-hematite
III
Calcite FI
Quartz FI
II
Muscovite-calcite
I
Quartz FI
Stage

TABLE 1. Synthesis of Microthermometric and Oxygen Isotope Geothermometric Data, Mantoverde District

IV
Calcite FI

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

429

FIG. 15. Textural relationships of the Ferrfera magnetite-apatite-pyrite


mineralization on the eastern branch of the Atacama fault system (Fig. 3).
These bodies are composed of, in decreasing abundance, magnetite (Mag),
fluorapatite (Ap), and pyrite (Py). Note the irregular to planar contacts, suggesting that these minerals formed in equilibrium. See text for details. Sample 2389, drill hole 99MS03, 264 m (plane-polarized reflected light).

veins cutting host rocks that are intensely altered to magnetite


overprinted by chlorite (Fig. 10). Textural relationships indicate that the minor chalcopyrite spatially associated with
stage II chlorite and quartz replaced pyrite, forming very fine
blebs and, at more advanced stages of replacement, irregular
fillings (i.e., interstitial) or thin veinlets in either K-feldsparquartz veins or altered host rocks of stage II (Fig. 17E-F).
The larger chalcopyrite aggregates are either massive or
porous, with rounded relics of pyrite (Fig. 17F). There are no
textural relationships unambiguously supporting the deposition of chalcopyrite during stage II (Fig. 10).
On the basis of the 18O values of coexisting muscovite and
calcite (Table 1), the temperature during stage II is estimated
to have been ~350C, in agreement with chlorite crystallinity
determined from X-ray powder diffraction spectra, which indicate crystallization at ~300 to 350C (rkai, 1991).
Stage III Cu (-Au) mineralization: Sulfide-rich, hematitecemented, hydrothermal breccias and mineralogically identical veins represent the main ore stage in the Mantoverde district (Fig. 10). Irregular to crudely tabular breccia bodies
(e.g., Manto Atacama: Vila et al., 1996) and breccia pipes
(e.g., Manto Ruso) are located preferentially in the hanging
wall of the Mantoverde fault (Figs. 3, 8, 9). The matrix of the
breccias is dominated by bladed specular hematite, enclosing
angular to subangular fragments of K-feldspathized, silicified,
or chloritized host rocks with diameters of up to a few centimeters (Fig. 18A). Locally, magnetite is pseudomorphous
after hematite (i.e., mushketovite). Breccia bodies grade eastward into networks of specular hematite veins cutting moderately altered host rocks (the Transition zone of Vila et al.,
1996; Fig. 8). Textural relationships in both breccias and veins
(Fig. 18B-C) show that chalcopyrite and pyrite were deposited at the same time as the hematite.
The development of hydrothermal breccias was locally preceded by the formation of sulfide-bearing calcite veins (Fig.
10). These occur predominantly in the northern half of the
district, in the hanging wall of the Mantoverde fault (Fig. 3).

429

430

BENAVIDES ET AL.

FIG. 16. Scapolite formation. (A). Scapolite (Scp) is overprinted by later


stage II chlorite (Chl). Stage I titanite (Ttn) is enclosed by scapolite. Sample
2395, drill hole 01DS18, 169 m (see Fig. 3; plane-polarized transmitted
light). (B). Scapolite (Scp) associated with quartz (Qtz) and titanite (Ttn).
Sample 5770 from unmineralized area located 15 km south of the Mantoverde district (see Fig. 3; coordinates: x = 367,005, y = 7,044,994; transmitted light; crossed nicols). (C). Scapolite postdates stage I magnetite
(Mag). Note the alteration of scapolite to calcite (turbid patching), the albitized plagioclase in the upper-left corner (AbPl), and the quartz (Qtz) aggregate at the right side. Sample 9826, barren sector, district-scale sampling
line west of Palmira (coordinates: x =384,779, y = 7,059,141; transmitted
light, crossed nicols).
0361-0128/98/000/000-00 $6.00

Locally, sulfides and calcite are intergrown with quartz and


specular hematite (samples 1378 and 1363, drill holes
01GT11, 876 m and 01CN03, 798 m, respectively; Figs. 3,
8B).
The pyrite associated with hematite-cemented breccias occurs as small, subangular to subrounded, generally fractured
grains (Fig. 18B) and, less commonly, grain aggregates with
diameters of up to a few millimeters. Chalcopyrite forms either irregular open-space fillings with widths of up to a few
centimeters or thin, discontinuous veinlets. The chalcopyrite
aggregates are massive or porous, enclosing relics of pyrite as
isolated grains, clusters (Fig. 18C) or discontinuous trails.
The pyrite associated with the early stage III calcite veins
forms subangular to subrounded and generally fractured
grains up to 5 mm in diameter. The associated chalcopyrite
occurs as irregular and rounded aggregates up to 1 cm in diameter, generally replacing pyrite and intergrown with calcite
and, locally, specular hematite. Although Zamora and Castillo
(2001) argued that significant chalcopyrite was introduced in
stage I along with magnetite and pyrite, we interpret microscopic textural relationships as evidence that chalcopyrite was
largely deposited during stage III (Fig. 10), in agreement with
Lpez (2002). Vila et al. (1996) reported increasing gold
grades with proximity to the Mantoverde fault (Figs. 3, 8),
and Lpez (2002) reported that gold is commonly intergrown
with chalcopyrite associated with specular hematite (Fig. 10).
Vila et al. (1996) reported homogenization temperatures
and salinities of fluid inclusions in what are interpreted to be
early stage III quartz and stage III calcite. They accepted the
coexistence of liquid- and vapor-rich inclusions as evidence
that stage III fluids were boiling and that mineralization occurred at ~180 to 250C. This is supported by a temperature
of ~240C calculated from the oxygen isotope fractionation
between coexisting calcite and hematite associated with stage
III chalcopyrite and pyrite (see below: Table 1).
Stage IV late calcite quartz veins: On the basis of crosscutting relationships, Vila et al. (1996), Cornejo et al. (2000),
Zamora and Castillo (2001), and Lpez (2002) concluded that
the development of calcite veins with variable amounts of sulfides and minor siderite (Astudillo, 2001) was a terminal hydrothermal event in the district. We assign these weakly mineralized veins to stage III (Fig. 10). In contrast, stage IV
calcite veins are generally barren and variably accompanied
by quartz (Figs. 10, 11D). They are observed throughout the
district but are more abundant in the northern centers, where
veins up to a few meters wide are located in the vicinity of the
Mantoverde fault (i.e., at Laura and to the west of Manto
Ruso, Figs. 3, 8A).
Minimum temperatures and fluid salinities during the terminal stage IV calcite veining are defined by microthermometric studies by S. Collao and E. Campos (Inclusiones Fluidas en Muestras de la Mina Mantoverde, IIIe Regin, Chile,
unpub. report for Empresa Minera Mantos Blancos, Divisin
Mantoverde, 2003, 19 p.; Table 1).
Supergene oxidation: The supergene oxidation profile in
the Mantoverde district, has a consistent thickness of 200 m
(Fig. 8) and is dominated by chrysocolla, brochantite, malachite, and Cu-bearing hematite, jarosite, and goethite (Astudillo, 2001; Lpez, 2002). These minerals occur as fine disseminations and fracture fillings in both tectonic and

430

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

FIG. 17. Textural relationships of the chlorite and sericite (muscovite)-dominated hydrolytic stage II and the replacement
textures associated with stage III chalcopyrite. (A). Advanced stages of hydrolytic alteration led to the formation of a chlorite-cemented breccia (Brecha Verde: Vila et al., 1996), containing fragments of K-feldspathized host rock (Kfs). The opaque
mineral in the matrix is hematite. Sample 2448, drill hole 00DS05, 403 m (plane-polarized transmitted light). (B). Stage II
chlorite (Chl) generally encloses fine, euhedral pyrite grains (Py) and very small, subhedral tourmaline crystals (Tur). Note
the tourmaline inclusion in the pyrite grain. Sample 2397, drill hole 01DS18, 190 m (plane-polarized transmitted light).
(C). Host rock completely replaced by muscovite (Ms), which is in turn cut by thin chlorite veins (Chl), both of stage II. Sample 2500, drill hole 01CN03, 8 m (plane-polarized transmitted light). (D). Idiomorphic crystals of pyrite (Py) coexisting with
stage II muscovite (Ms) and hematite (Hem). Sample 2394, drill hole 01DS16, 296 m (plane-polarized reflected light).
(E). At advanced stages of replacement, very small rounded relics of pyrite (Py) survive in a chalcopyrite-bearing matrix
(Cpy). Sample 1391, drill hole 99MS01, 576 m (plane-polarized reflected light). (F). Chalcopyrite (Cpy) enclosing very fine,
rounded pyrite relics (Py) occurs as irregular aggregates or veinlets in host rocks altered to K feldspar (Kfs) and chlorite (Chl).
Sample 2470, drill hole 01DS13, 830 m (plane-polarized reflected light).
0361-0128/98/000/000-00 $6.00

431

431

432

BENAVIDES ET AL.

hydrothermal breccias (Vila et al., 1996; Figs. 9, 11C), locally


increasing the Cu grades to 5 percent (Astudillo, 2001). Less
commonly, chalcocite replaces both pyrite and chalcopyrite
(e.g., sample 2430, drill hole 01GT11, 290 m, Mantoverde,
Southern pit; Fig. 3).
Age of mineralization
Vila et al. (1996) reported K-Ar ages of 117 3 and 121
3 Ma for two Mantoverde sericite (muscovite) samples from,
respectively, the Transition zone and an altered granitoid
dike. These data would suggest that at least stage II alteration
may have been associated with the emplacement of the Sierra
Dieciocho pluton (Fig. 3), for which Lara and Godoy (1998)
proposed an age of 120 to 126 Ma, based on dates of 120 4
Ma (K-Ar on hornblende: Haynes, 1975, recalculated using
the constants of Steiger and Jger, 1977), 125.7 0.6 and
126.8 0.5 Ma (40Ar/39Ar on hornblende: Dallmeyer et al.,
1996), and 123.7 2.7 and 126.8 1.3 Ma (Rb-Sr whole-rock
isochron and U-Pb zircon ages, respectively: Berg and Breitkreuz, 1983). However, Gelcich et al. (2002) reported significantly older U-Pb zircon ages of 129.6 0.3 and 128.9
0.3 Ma for, respectively, a pluton of the Sierra Dieciocho
complex and a potassically altered quartz monzodiorite dike
from the mine area. These dates are, moreover, significantly
older than a U-Pb age of 126.4 0.3 Ma for hydrothermal titanite associated with stage I orthoclase, probably from the
main Mantoverde orebody (Gelcich et al., 2002). The U-Pb
data suggest that K-feldspathization and, by implication, Fe
metasomatism significantly postdated crystallization of the
Sierra Dieciocho pluton. Further, although micas are susceptible to Ar outgassing at relatively low temperatures, the temperatures during stages III and IV (i.e., 250C) were probably insufficient to reset muscovite 2M1 argon systematics.
Robust geochronologic data are required for stage III Cu-Au
mineralization in this district, but an age in the range 117 to
121 Ma is favored here.
Stable Isotopes
We report the results of 34S analyses of 73 high-quality sulfide separates (36 of chalcopyrite and 37 of pyrite) from the
Mantoverde deposit and its satellite orebodies and prospects.
Additionally, 18O values were determined for 10 samples of
magnetite and 11 of hematite. The sulfur isotope data, including sample descriptions and locations, are summarized in
Table 2. The 18O values and relevant information on the iron
oxide samples are listed in Table 3, together with 18O, D,
and 13C data for selected associated minerals.

FIG. 18. Textural relationships of stage III hydrothermal breccias. (A).


The breccias comprise a matrix of bladed hematite (Hem) and subangular to
subrounded fragments of altered host rock. Sample 1404, drill hole 125 N2,
89 m, 1 km south of Ferrfera (see Fig. 3). (B). Subangular to subrounded
and fractured grains of pyrite are common in the matrix of hematitic breccias. Note the incipient development of chalcopyrite in pyrite grains located
left-center. Sample 2496, drill hole 00CN02, 300 m. (C). Chalcopyrite (Cpy)
vein cuts hematite and pyrite of stage III. Sample 2488, drill hole 99MV06,
167 m. All views in plane-polarized reflected light.
0361-0128/98/000/000-00 $6.00

Analytical methods
Selection of samples for 34S analysis was based on detailed
petrographic study. Pyrite and chalcopyrite separates of 0.2
and 0.3 mg, respectively, were converted to SO2 in a Carlo
Erba element analyzer NCS 2500, with CuO as oxidant. The
sulfur isotope composition was measured with a Finnigan
MAT 252 mass spectrometer at the Queens Facility for Isotope Research. Data are reported using the notation in units
of per mil, relative to the CDT standard (Can Diablo
troilite). The analytical precision for 34S values is 0.3 per
mil. Using the above procedure, the 34S values of NIST-123
and NIST-127 are 17.1 and 20.0 per mil, respectively.

432

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

433

TABLE 2. 34S Values of Sulfides, Mantoverde District


Sample no. Drill hole

Depth
(m)

Manto Ruso
2469
DDH01GT11
155008
RCH04MR68
1378A
DDH01GT11
154856
RCH04MR66
1378B
DDH01GT11
154856-1 RCH04MR66
154866
RCH04MR66
1373
DDH01GT11
1378
DDH01GT11
1375
DDH01GT11
1391
DDH99MS01
West Manto Ruso
1368
DDH01DS07
SE Manto Ruso
1400-1
DDH99CN12
1398
DDH99CN12
1399
DDH99CN12
1399B
DDH99CN12
1400
DDH99CN12
1400-2
DDH99CN12
1400-C-2 DDH99CN12
1400-C-1 DDH99CN12
East Laura sector
1362-B
DDH01CN03
1361
DDH01CN03
1362
DDH01CN03
1363-02
DDH01CN03
1363
DDH01CN03
1362
DDH01CN03
1356
DDH01CN03
1358
DDH01CN03
NNE-Mantoverde main pit
2492
DDH99MV06
2488B
DDH99MV06
2488
DDH99MV06
2475C
DDH01CN04
2475A
DDH01CN04
2477
DDH01CN04
2477B
DDH01CN04
2473
DDH01CN04
2475B
DDH01CN04
NW Ferrifera
2496
DDH00CN02
2498
DDH00CN02
2498
DDH00CN02
2496
DDH00CN02
2483
DDH01DS22
AFS-East Mantoverde main pit
2457
DDH99MS03
2459
DDH99MS03
2459B
DDH99MS03
2459
DDH99MS03
2459B
DDH99MS03
SE-Mantoverde main pit
2470A
DDH01DS13
2470
DDH01DS13
2470B
DDH01DS13
2359
DDH01DS13
2465
DDH01DS13
2357
DDH01DS13
2358
DDH01DS13
2447
DDH00DS05
North-Mantoverde pit south
2430
DDH01GT01
NW-M.V. pit south
2425
DDH00DS02
2427
DDH00DS02
0361-0128/98/000/000-00 $6.00

UTM_E

UTM_N

Stage

Sulfide

Host mineral/rock

Habit/texture

34S

653
218
876
234
876
236
254
351
876
501
576

369,200
369,475
369,200
369,135
369,200
369,135
369,135
369,200
369,200
369,200
370,170

7,065,700
7,065,390
7,065,700
7,065,580
7,065,700
7,065,580
7,065,580
7,065,700
7,065,700
7,065,700
7,066,140

III
III
III
III
III
III
III
III
III
III
II

Cpy>>Py
Cpy
Cpy>>Py
Cpy
Cpy
Cpy>>Py
Py
Py>>Cpy
Py
Py>>Cpy
Py>>Cpy

Specular hematite
Specular hematite
Quartz
Specular hematite
Quartz
Specular hematite
Specular hematite
Specular hematite
Quartz
Specular hematite
Potassic alteration

Irregular aggregate
Irregular aggregate
Irregular and fine aggregates
Fine aggregate
Irregular and fine aggregates
Irregular aggregate
Rounded crystals
Subhedral grains
Irregular and fine aggregates
Irregular aggregates
Vein

8.9
10.0
6.6
7.2
5.1
9.3
1.9
4.6
6.8
5.5
6.8

294

368,560

7,065,570

III

Cpy

Calcite vein

Fine/rounded aggregates

4.7

386
255
340
340
388
386
386
386

370,065
370,065
370,065
370,065
370,065
370,065
370,065
370,065

7,065,490
7,065,490
7,065,490
7,065,490
7,065,490
7,065,490
7,065,490
7,065,490

III
III
III
III
III
III
III
III

Cpy
Cpy>>Py
Cpy
Cpy
Cpy
Cpy
Cpy
Py

Calcite vein
Specular hematite
Specular hematite
Specular hematite
Specular hematite
Calcite vein
Calcite vein
Calcite vein

Irregular aggregate
Irregular filling
Irregular filling
Irregular filling
Open space fillings
Aggregate
Irregular and massive aggregates
Irregular aggregates

2.2
2.5
6.1
5.0
3.7
0.9
0.4
2.8

758
732
757
798
798
752
448
591

369,300
369,300
369,300
369,300
369,300
369,300
369,300
369,300

7,064,660
7,064,660
7,064,660
7,064,660
7,064,660
7,064,660
7,064,660
7,064,660

III
III
III
III
III
III
II
III

Cpy
Cpy
Cpy>>Py
Cpy
Cpy
Py>>Cpy
Py>>Cpy
Py>>Cpy

Calcite vein
Calcite vein
Calcite vein
Calcite/Fe oxide
Calcite/Fe oxide
Calcite vein
Chlorite
Specular hematite

Irregular aggregate
Irregular aggregate
Aggregate
Irregular aggregate
Irregular vein
Irregular aggregate
Irregular fillings
Irregular aggregates

0.4
0.1
0.1
2.5
0.9
5.1
4.9
5.1

549
167
167
404
404
404
404
311
404

369,620
369,620
369,620
369,360
369,360
369,360
369,360
369,360
369,360

7,062,950
7,062,950
7,062,950
7,063,785
7,063,785
7,063,785
7,063,785
7,063,785
7,063,785

III
III
III
III
III
III
III
III
III

Cpy>>Py
Cpy
Py>>Cpy
Cpy>>Py
Cpy>>Py
Cpy>>Py
Cpy>>Py
Cpy>>Py
Cpy>>Py

Calcite vein
Specular hematite
Specular hematite
Specular hematite
Specular hematite
Specular hematite
Specular hematite
Specular hematite
Specular hematite

Fine aggregates
Open space fillings
Irregular filling
Irregular aggregate
Irregular aggregate
Irregular fracture fillings
Irregular fracture fillings
Irregular fracture fillings
Irregular aggregate

5.4
4.0
11.2
8.7
5.3
3.8
3.9
1.4
4.5

302
448
448
301
236

370,090
370,090
370,090
370,090
370,180

7,063,930
7,063,930
7,063,930
7,063,930
7,063,950

III
III
III
III
III

Cpy>>Py
Py
Py
Py
Py

Specular hematite
Specular hematite
Specular hematite
Specular hematite
Quartz vein

Irregular aggregate
Irregular aggregate
Irregular aggregate
Subhedral cumulates
Subhedral cumulates

3.9
3.6
3.1
5.0
8.0

324
453
453
453
453

370,525
370,525
370,525
370,525
370,525

7,062,700
7,062,700
7,062,700
7,062,700
7,062,700

II
II
II
II
II

Py
Py>>Cpy
Py
Py>>Cpy
Py>>Cpy

Chl/sericite
Chl/sericite
Chl/sericite
Chlorite
Chlorite

Disseminations
Subhedral cumulates
Subhedral cumulates
Subrounded grains
Subrounded grains

1.5
1.7
0.9
0.7
1.2

830
830
830
970
427
954
968
290

369,930
369,930
369,930
369,930
369,930
369,930
369,930
369,535

7,062,510
7,062,510
7,062,510
7,062,510
7,062,510
7,062,510
7,062,510
7,062,490

III
III
III
I
II
II
II
III

Cpy>>Py
Cpy>>Py
Cpy>>Py
Py
Py
Py>>Cpy
Py
Cpy

Calcite vein
Calcite vein
Calcite vein
Magnetite
Sericite
Chlorite
Chl/sericite
Specular hematite

Irregular aggregate
Irregular aggregate
Irregular aggregate
Subhedral cumulates
Fracture filling
Irregular aggregates
Rounded disseminations
Irregular fillings/veins

2
1.8
1.4
0.5
9.1
1.0
0.9
4.5

290

369,610

7,061,130

II

Py

Altered host rock

Matrix of breccia

1.2

216
357

369,880
369,880

7,061,050
7,061,050

I
II

Cpy
Cpy

Magnetite
Altered host rock

Irregular aggregate
Discontinous veinlet

2.0
0.9

433

434

BENAVIDES ET AL.
TABLE 2. (Cont.)

Sample no. Drill hole

Depth
(m)

UTM_E

UTM_N

Stage

Sulfide

Host mineral/rock

Habit/texture

34S

216
302

369,880
369,880

7,061,050
7,061,050

I
II

Cpy
Py>>Cpy

Magnetite
Chlorite

Irregular fillings/veins
Subhedral grains

0.6
0.5

169
220

370,445
370,445

7,060,290
7,060,290

II
II

Py>>Cpy
Py

Chlorite
Chlorite

Subhedral cumulates
Subhedral grains

0.8
0.0

194
296
296

370,860
370,860
370,860

7,057,050
7,057,050
7,057,050

III
II
II

Cpy
Py
Py

Calcite vein
Sericite
Sericite

Irregular and fine aggregates


Subrounded grains
Subrounded grains

242
210
238
232
110

371,040
371,040
371,040
371,040
371,230

7,056,550
7,056,550
7,056,550
7,056,550
7,055,650

II
II
II
II
II

Py>>Cpy
Py
Py
Py
Py

Chl-qtz/magnetite
Chl-qtz/magnetite
Chl-qtz/magnetite
Chl-qtz/magnetite
Chl-qtz/magnetite

Irregular aggregate
Irregular aggregate
Dissemination/aggregates
Fine disseminations
Disseminations

370,180

7,063,950

Py

Magnetite/apatite

Fine disseminations

1.7

370,525
370,525

7,062,700
7,062,700

Py
Py

Magnetite/apatite
Magnetite/apatite

Irregular aggregates
Irregular aggregates

0.6
0.9

2425A
DDH00DS02
2426
DDH00DS02
SE Mantoverde pit south
2395
DDH01DS18
2398
DDH01DS18
Trillizos
2386
DDH01DS16
2394
DDH01DS16
2394-1
DDH01DS16
S-SSW of Franco
155687
RCH04DT01
155671
RCH04DT01
155685
RCH04DT01
155682
RCH04DT01
155762
RCH04DT02

Magnetite-apatite-pyrite bodies
NW Ferrifera
2483A
DDH01DS22
236
Ferrifera
2389
DDH99MS03
264
2389-1
DDH99MS03
264

4.1
1.1
0.6
0.5
0.4
1.4
0.9
0.6

Samples are tabulated from north to south; for locations of the drill hole collars, see Figure 2

Extraction of oxygen from purified separates was carried


out at 600C, using the BrF5 technique of Clayton and
Mayeda (1963). The oxygen isotope compositions, measured
with a Finnigan MAT 252 mass spectrometer, are reported
using the notation in units of per mil relative to VSMOW
and have a precision of 0.2 per mil. Using these techniques,
analyses of NIST-28 gave a value of 9.6 per mil.
Sulfur isotope compositions
The 34S values of sulfides in the Mantoverde district range
from 6.6 to +11.2 per mil overall (Fig. 19). Values of pyrite
range from 6.8 to +11.2 per mil and those for chalcopyrite
from 6.6 to +10.0 per mil. Despite the overlap, the 34S values vary systematically with the paragenetic stage, with
markedly wider variations characterizing the younger hydrothermal activity (Figs. 10, 19).
Pyrite associated with the magnetite-apatite bodies along
the eastern branch of the Atacama fault system (Fig. 3) has
34S values of +0.6 to +1.7 per mil (Fig. 15, Table 2). A similar narrow range and relatively low values from 0.6 to +2 per
mil are exhibited by sulfides associated with the stage I magnetite bodies located in the footwall of the Mantoverde fault
(Figs. 3, 8, 14, 19). In contrast, pyrite and chalcopyrite spatially associated with stage II sericitic and chloritic alteration
have 34S values ranging from 0.9 to +4.9 per mil (Figs. 11B,
19; Table 2). Similarly, the values for stage II sulfides filling
fractures in altered host rocks range from 1.2 to as high as
+9.1 per mil (Fig. 19, Table 2).
The greatest isotopic variability is shown by chalcopyrite
and pyrite directly associated with stage III specular
hematite-cemented hydrothermal breccias, with 34S values
ranging from +1.4 to +11.2 per mil, chalcopyrite generally
having the higher values (Fig. 19, Table 2). In contrast, sul0361-0128/98/000/000-00 $6.00

fides associated with stage III calcite veins tend to have lower
values, 5.5 to +5.1 per mil, with chalcopyrite having lower
values than pyrite (Fig. 19, Table 2).
Sulfides from the central and northern parts of the district, (i.e., north of the Mantoverde main open pit; Fig. 3),
have the widest range of 34S values. In this sector, the lowest 34S value (i.e., 6.8 in sample 1378: Table 2) was
found in pyrite from a deep stage III quartz-calcite vein at
Manto Ruso, whereas the highest values (i.e., +7.2+10:
Table 2) occur in sulfides in shallow specular hematite-cemented breccias and veins in the vicinity of Manto Ruso
(Fig. 3). Pyrite associated with intensely chloritized volcanic host rocks and massive bodies of magnetite in the
southern part of the district (i.e., Montecristo, Franco, and
Trillizos sectors: Fig. 3) has 34S values close to 0 per mil
(Table 2).
Oxygen isotope compositions of iron oxides
A histogram of the 18O values of iron oxide phases is
shown in Figure 20. The magnetites include samples from
both the magnetite-apatite-pyrite bodies at Ferrfera and the
stage I magnetite bodies associated with the Cu-mineralized
orebodies contiguous with the Mantoverde fault. Hematite
samples are from chalcopyrite-rich hydrothermal breccias assigned to paragenetic stage III (Fig. 10, Table 3).
The iron oxides have 18O values which vary overall from
1.9 to +4.1 per mil (Fig. 20, Table 3). The highest values are
shown by magnetite and range from +1.4 to +4.1 per mil.
Magnetite in the sulfide-bearing orebodies has slightly lower
values, from +1.4 to +3.1 per mil, than those in the magnetite-apatite-pyrite bodies, which vary from +2.2 to +4.1 per
mil. Stage III hematite has lower isotopic values of 2.0 to
+1.7 per mil (Fig. 20).

434

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

435
164
296
296
296
164
128
218
236
264
180
252
180
264
236

01DS16
01DS16
01DS17
01DS18
01DS16
04DT041
04DT041
01DS22
99MS03
99MS03
99MS03
99MS03
99MS04
01DS22

180

99MS03

158

968
968

01DS13
01DS13

96GM21

404
500
167
223
223

01CN04
01CN04
99MV06
01DS05
01DS05

172
190

732
798
798
301

01CN03
01CN03
01CN03
00CN02

01DS18
01DS18

294

01DS07

216
302

387
254

99CN12
99CN12

00DS02
00DS02

351

Depth (m)

01GT11

Drill hole

370,180
370,525
370,525
370,525
370,525
370,525
370,180

371,100
371,100

370,860
370,860
370,860
370,860
370,860

370,110

370,445
370,445

369,880
369,880

370,525

369,930
369,930

369,360
369,360
369,620
369,540
369,540

369,300
369,300
369,300
370,090

368,570

370,065
370,065

369,200

UTM_E

7,063,950
7,062,700
7,062,700
7,062,700
7,062,700
7,062,700
7,063,950

7,055,800
7,055,800

7,057,050
7,057,050
7,057,050
7,057,050
7,057,050

7,059,370

7,60,290
7,60,290

7,061,050
7,061,050

7,062,700

7,062,510
7,062,510

7,063,785
7,063,785
7,062,950
7,062,500
7,062,500

7,064,660
7,064,660
7,064,660
7,063,930

7,065,570

7,065,490
7,065,490

7,065,700

UTM_N

II
II

I
II
II
II
II

II
II

I
I

II

I
III

III
III
III
II
II

III
III
III
III

II

III
III

III

Stage

Notes: for locations of the drill hole collars, see Figure 3; D and 13C values from Benavides (2006)
1 Reverse circulation drill holes

Manto Ruso
1375
SE Manto Ruso
1400
1398A
SW Manto Ruso
1368
NE-E Laura sector
1361
1363
1363
2496
NNE-Mantoverde main pit
2475
2477
2488
2446A
2446B
E-Mantoverde main pit
2358
2358A
South Ferrfera-E Mantoverde main pit
2454
NW-Mantoverde pit south
2425
2426
SE-Mantoverde pit south
2396
2397
Montecristo
2384
Trillizos
2387
2394
2394
2394
2387
South Trillizos
156613
156680
Eastern branch, AFS (Ferrfera)
2483
2389
2454
2456
2454A
2390
2483

Sample/location

Magnetite
Magnetite
Magnetite
Magnetite
Magnetite
Apatite
Apatite

Chlorite
Chlorite

Magnetite
Hematite
Muscovite
Calcite
Chlorite

Magnetite

Chlorite
Chlorite

Magnetite
Magnetite

Chlorite

Magnetite
Hematite

Hematite
Hematite
Hematite
Chlorite
Chlorite

Hematite
Hematite
Calcite
Hematite

Chlorite

Hematite
Hematite

Hematite

Mineral

Magnetite-apatite-pyrite bodies
Magnetite-apatite-pyrite bodies
Magnetite-apatite-pyrite bodies
Magnetite-apatite-pyrite bodies
Magnetite-apatite-pyrite bodies
Magnetite-apatite-pyrite bodies
Magnetite-apatite-pyrite bodies

Chloritized volcanic rock


Chloritized volcanic rock

Massive bodies
Breccia
Intergrown
Intergrown
Chloritized volcanic rock

Massive bodies

Chlorite-bearing breccia
Chlorite-bearing breccia

Massive bodies
Massive bodies

Fine grained

Fractured bodies
Vein

Breccia
Breccia
Breccia
Chloritized granitoid
Chloritized granitoid

Vein
Breccia
Breccia
Breccia

Chloritized volcanic rock

Breccia
Breccia

Breccia

Texture/host

TABLE 3. Oxygen, Hydrogen, and Carbon Isotope Compositions of Selected Minerals, Mantoverde District

4.1
2.2
3.5
2.5
3.1
7.2
8.2

8.4
7.8

2.5
1.5
9.3
12.5
9.5

3.1

11.2
9.0

1.4
3.4

5.6

1.4
1.9

1
0.5
0.0
9.7
8.7

1.7
1.9
11.3
1

7.6

1.7
1.0

1.7

18O

54
46

71

54

62
62

55

70
60

66

5.3

13C

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

435

436

BENAVIDES ET AL.
14

12

FREQUENCY

10

0
-7

-6

-5

-4

-3

-2

-1

10

11

12

13

S (, CDT)
34

LEGEND

Py with ironstones at Ferrifera


Cpy "associated" with Stage I-Mt
Cpy with Stage II-Chl/Ser
Py with Stage III-Cal
Cpy with Stage III-Cal/Hem
Cpy with Stage III-Qtz
Cpy with Stage III-Breccia

Py with Stage I-Mt


Py with Stage II-Chl/Ser
Sulfide veins/fillings-Stage II
Cpy with Stage III-Cal
Py with Stage III-Qtz
Py with Stage III-Breccia

FIG.19. Sulfur isotope composition of pyrite and chalcopyrite in the Mantoverde district, including sulfides associated
with the magnetite-apatite-pyrite bodies at Ferrfera. The 34S values range widely from 6.8 to +11.2 per mil (n = 73) and,
despite the extensive overlap, 34S values vary systematically with paragenetic stage. Sulfides associated with stage III hydrothermal breccias exhibit a wider range (+1.4 to +11.2, n = 25) and higher 34S values than those of stage I sulfides (0.6
to +2, n = 3). The isotopic composition of sulfides associated with both chlorite and sericite of stage II ranges from 0.9
to +4.9 per mil (n = 18), and that for stage III calcite veins is 5.0 to +5.1 per mil (n = 14). The 34S values of sulfides associated with bodies of magnetite-apatite at Ferrfera (n = 3) and stage I metasomatic magnetite in the Mantoverde district (n
= 3) are similar and close to 0 per mil. See text for details. Abbreviations: Ap = apatite, Cal = calcite, Chl = chlorite, Cpy =
chalcopyrite, Mag = magnetite, Py = pyrite, Qtz = quartz, Ser = sericite.

Isotopic characterization of the hydrothermal fluids


Estimation of the stable isotopic chemistry of the fluid requires knowledge of mineral and/or water isotope fractionation factors, temperature, and, for sulfur, oxidation state
(Ohmoto and Rye, 1979; Ohmoto and Goldhaber, 1997). The
parameters and information relevant to isotopic composition
and paragenetic stage are recorded in Table 4.
Stage I: The massive bodies of magnetite in the Mantoverde district are the product of high-temperature Fe
metasomatism (e.g., 460550C, see Table 1) and the coexistence of magnetite and pyrite suggests that oxygen fugacity was close to 1018 bars (see fig. 10.7, Ohmoto and Rye,
1979). These conditions imply that fluids with 34S values of
+0.4 to +4.0 per mil were responsible for preore magnetite
deposition, a range consistent with a magmatic origin
(Ohmoto and Goldhaber, 1997; Table 4). For the magnetiteapatite-pyrite bodies along the eastern branch of the Atacama fault system (Fig. 3), calculations based on the oxygen
isotope geothermometer for the pair apatite-magnetite
(Table 3) suggest a formation temperature of ~650C. The
0361-0128/98/000/000-00 $6.00

sulfur isotope composition of pyrite from Ferrfera, +0.6 to


+1.7 per mil, similarly is consistent with the involvement of a
magmatic fluid.
Crosscutting relationships indicate that magnetite and hypogene hematite formed almost entirely at different paragenetic stages in the Mantoverde district (Fig. 10). Using oxygen isotope fractionation factors for magnetite-water
(Bottinga and Javoy, 1973) and a temperature range of 460
to 550C (Table 1), stage I magnetite is inferred to have equilibrated with a fluid with an oxygen isotope composition of
+7.3 to +10 per mil. The whole-rock 18O value of +12 per
mil for a K-feldspathized host rock (i.e., sample 2464, drill
hole 01DS13, 371 m: Figs. 3, 13B) and the oxygen isotope
fractionation for the pair K-feldspar-water between 460 and
550C (ONeil and Taylor, 1967) indicate a fluid with 18O
value of about +10 per mil, indistinguishable from that in
equilibrium with magnetite. Similarly, the massive bodies of
magnetite-apatite-pyrite along the eastern branch of the Atacama fault system (e.g., Ferrfera: Fig. 3) are inferred to have
equilibrated with a fluid with 18O values of +8 to +9 per mil.

436

437

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE


6

FREQUENCY

0
-3

-2

-1

delta 18O

O (, SMOW)
18

LEGEND
Stage III-Hematite (Mantoverde)
Stage I- Magnetite (Mantoverde)
Magnetite-apatite-pyrite (Ferrifera)
FIG. 20. Frequency histogram of 18O values of magnetite and hematite
in the Mantoverde district (n = 20). The average value for magnetite is +3.5
per mil (n = 10), whereas hematite tends to have lower values, with an average of 0.6 per mil (n = 10). The isotopic composition of metasomatic magnetite (stage I, Figs. 11, 14) at Mantoverde is indistinguishable from that of
massive bodies of magnetite-apatite pyrite located along the eastern branch
of the Atacama fault system (i.e., Ferrfera, see Figs. 3, 15).

Stage II: The associations of muscovite-calcite hematite


pyrite and chlorite-quartz-pyrite rutile hematite indicate
that the hydrothermal fluids became more acidic in stage II,
following cooling to ~350C (see Table 1). The dominance of
pyrite and the incipient to moderate development of hematite
suggest that conditions fell between the hematite-magnetitepyrite and hematite-pyrite-FeSO4 oxygen buffers, with a O2
of ~1022.5 bars (Shi, 1992). The stage II fluids would therefore have had 34S values of +9.1 to +14.9 per mil (Table 4;
i.e., significantly heavier than those of fluids with a magmatic
derivation: Ohmoto and Goldhaber, 1997). Additionally, the
18O and D values of stage II chlorite range widely from

+5.6 to +11.2 and 71 to 46 per mil, respectively (Table 3).


The oxygen and hydrogen isotope fractionation factors defined for the pair chlorite-water (Graham et al., 1987; Cole
and Ripley, 1998) indicate that at 350C the fluid in equilibrium with stage II chlorite would have had 18O and D values of +5.7 to +11.2 and 38 to 13 per mil, respectively (Benavides, 2006). Such compositions indicate that the
hydrothermal system experienced a significant incursion of
nonmagmatic fluids during stage II (see Taylor, 1997).
Stage III: Megascopic and, particularly, microscopic crosscutting relationships indicate that stage III pyrite and chalcopyrite were not coprecipitated, precluding temperature estimation on the basis of sulfur isotope fractionation (Ohmoto
and Rye, 1979). However, the 18O values of coexisting calcite
and hematite (Table 1) define temperatures of ~230 to
250C. The coexistence of hematite and pyrite indicates that
at those temperatures, the oxygen fugacity would be between
the hematite-magnetite-pyrite and hematite-pyrite-FeSO4
buffers, with calculated fO2 values of ~1033 bars (Shi, 1992).
On this basis and Ohmoto and Rye (1979), Cu-rich stage III
assemblages with 34S values of +1.4 to +11.2 per mil (Tables
2, 4) are inferred to record deposition of fluids with 34S values ranging from +26.4 to as high as +36.2 per mil (Table 4).
Chalcopyrite and pyrite in the matrix of stage III hematite-cemented breccias equilibrated with fluids having 34S values of
+26.4 to +35 and of +26.9 to +36.2 per mil, respectively
(Table 4).
The oxygen isotope fractionation for the pair hematite-water
(Yapp, 1990) at 200 to 250C indicates that stage III hematite
equilibrated with a fluid with 18O values in the range +3.0 to
+8.0 per mil, overlapping with, but mainly lighter than, that of
the fluid in equilibrium with stage I magnetite.
Discussion and Conclusions
Role of exoticsulfur
The new light stable isotope data for the Mantoverde deposit and its satellites are directly germane to the ongoing debate concerning the genesis of central Andean Mesozoic iron
oxide-copper-gold mineralization and, more generally, to the
understanding of the IOCG clan globally. Despite the inherent uncertainties in both isotope fractionation factors and the
temperature and oxygen fugacity ranges for the various paragenetic stages, the data are sufficiently robust to conclude

TABLE 4. 34S Values of Sulfides from Paragenetic Stages I to III and Inferred Sulfur Isotopic Compositions of Ore Fluid
Paragenetic stage association
34Ssulfides
Estimated depositional
temperature1
fO2 indicator(s)
fO2 in bars 2
= 34Ssulfide - 34Sfluid
34Sfluid

I
(Mag-Py)

II
(Chl-Ms-Py Hem)

III
(Hem-Cpy-Py)

0.6 to +2.0

0.9 to +4.9

+ 1.4 to +11.2

460 to 550C
Mag-Py
1018
1 to 2
+ 0.4 to + 4

~350C
Py-(Hem)
1022.5
10
+ 9.1 to +14.9

~230250C
Hem-Py
1033
25
+ 26.4 to +36.2

Notes: The isotopic compositions of the fluid are calculated according to the procedures of Ohmoto and Rye (1979) and Ohmoto and Goldhaber (1997);
abbreviations: Py = Pyrite, Cpy = Chalcopyrite, Mag = Magnetite, Ms = Muscovite, Chl = Chlorite, Hem = Hematite, Cal = Calcite
1 For depositional temperature see Table 1
2 For stage I, based on Ohmoto and Rye (1979). For stages II and III, according to phase diagrams of Shi (1992)
0361-0128/98/000/000-00 $6.00

437

438

BENAVIDES ET AL.

that a major change in the isotopic chemistry, and hence predominant reservoir, occurred between stages I and II. Taken
in conjunction with the observations of Ullrich and Clark
(1999) and Ullrich et al. (2001) on the larger and higher grade
La Candelaria deposit, and those of Ripley and Ohmoto
(1977) on the small, high-grade, Ral-Condestable deposit,
the sulfur and oxygen isotope compositions for the different
paragenetic stages in the Mantoverde district strongly imply
that, at the least, exotic sulfur input was a prerequisite for
economic, or even subeconomic, copper mineralization in the
Andean IOCG centers. Moreover, the direct correlation between Cu and Au in these deposits implies that gold enrichment may also have been influenced by the incursion of such
nonmagmatic fluids.
The requirement for the introduction of sulfur from the
wider exocontact environment is indeed a satisfactory explanation for the fortuitous involvement of externally derived
brines in Cu-rich IOCG mineralization, in the central Andes
and elsewhere (Sillitoe, 2003). The geochemical relationships
documented in this study are consistent with the concepts of
Barton and Johnson (1996, 2000), in that seawater, plausibly
mediated through evaporitic processes, is the most likely
reservoir for the high 34S and low 18O fluids implicated in
the later stages of the Mantoverde and La Candelaria (Ullrich
et al., 2001) hydrothermal systems. It should be emphasized
that Cornejo et al. (2000) earlier proposed a fluid mixing
model for the Mantoverde deposit but envisaged meteoric
water rather than evaporite-sourced brines as the nonmagmatic component. However, the isotopic relationships preclude major meteoric water contributions, even in the terminal stages of hydrothermal activity.
Whereas the involvement of sulfur-bearing fluids with high
34S values (36.2 at Mantoverde; 20.2 at La Candelaria) is unambiguous in these two largest documented Andean IOCG deposits, there is no evidence that Cu and associated minor ore metals were similarly derived from either
the postulated evaporitic reservoir or from andesitic host
rocks. In this regard, the temporal relationships between initial chalcopyrite deposition and the introduction of exotic
sulfur are critical. At Mantoverde, marked changes in 34S
fluid occurred during the chlorite-dominated stage II (Fig.
10, Table 4), and therefore all significant Cu mineralization
took place from fluids with high 34S values. In contrast, Ullrich and Clark (1999) showed that chalcopyrite deposition at
La Candelaria was initiated by fluids with magmatic 34S values of 0.4 to +5.7 per mil, preceding marked increases in
34Sfluid values. Moreover, this polymetallic substage also involved the deposition of minor sphalerite and molybdenite.
These relationships suggest that at Mantoverde the dominant
ore metal reservoir may similarly have been magmatic.
Regional setting of mineralization
Sillitoe (2003) has emphasized that evaporitic successions
are either absent or minimally exposed in the vicinity of most
central Andean IOCG centers. This would constitute as a cogent objection to the involvement of evaporite-derived brines
in the genesis of these deposits. However, the Chaarcillo
Group is probably vestigially preserved in northern Chile as a
result of erosion during inversion of the back-arc basins, and
any evaporites also would have been eroded. In the case of
0361-0128/98/000/000-00 $6.00

Mantoverde, the widespread circulation of brines in the area


now separating the remnants of the Neocomian basin and the
Atacama fault system may be recorded by the numerous areas
of strong scapolitization (Fig. 4) and, as argued previously, the
marialitic scapolite deposited early in stage II in the Mantoverde deposit and its satellites may represent this event.
Such district-scale or even regional Na-Cl metasomatism is
characteristic of several Precambrian IOCG provinces (Frietsch et al., 1997) and has been ascribed to the large-scale dissolution of evaporites. Despite the relative youth of the Andean environment, erosion has certainly disguised the
original distribution of metasomatic scapolite within and behind the Neocomian arc, and there is no clear evidence for
the source and flow paths of the fluids. On the assumption,
however, that the back-arc basin, when developing, would
have extended farther west than the existing outcrops of the
Punta del Cobre Formation and Chaarcillo Group, a fluid
source along the basin margin is reasonable, possibly with
subsidiary flow within the Atacama fault system. Because
scapolite deposition occurred at the beginning of stage II, and
therefore coincided with the initial input of 34S-enriched fluids, it is plausible that the relict distribution of scapolite in the
wider district partially records the migration of the brines responsible for chalcopyrite deposition at Mantoverde and elsewhere.
The age of Cu mineralization (stage III) along the Atacama
fault system and in the vicinity of the Neocomian basin is
poorly constrained. However, as documented herein, the
131.3 1.4 Ma volcanic strata of the back-arc basin successively experienced spilitization and hydrolytic alteration and
low-grade metamorphism (Fig. 4) prior to the formation of
magnetite in the IOCG prospects of the Sierra Santo
Domingo district (Fig. 7). If we assume that all IOCG centers
in the district developed broadly simultaneously, the U-Pb
date (126.4 0.3 Ma) determined for stage I titanite from
Mantoverde by Gelcich et al. (2002) implies that the subocean-floor alteration and ensuing metamorphism were accomplished in no more than ca. 7 m.y. and, perhaps, in considerably less time. The metamorphism, therefore, was
probably diastathermal (i.e., extensional, see Robinson, 1987;
Alt, 1999), rather than due to burial. This would require extreme heat flow during the extensional opening of the backarc basin, and residual heat expelled subsequently during initial basin closure could have promoted destruction of
evaporitic strata and driven the scapolitizing brines westward.
Despite the absence of evaporites in the remnant outliers of
the Chaarcillo Group at this latitude, tentative correlations
between events in the back-arc basin and hydrothermal activity along the Atacama fault system may be proposed.
Concluding statement
The first comprehensive documentation of the stable isotope chemistry of hydrothermal minerals from the IOCG deposits and prospects of the Mantoverde district, in the framework of a refined paragenetic model, provides unambiguous
evidence that nonmagmatic sulfur played a critical role in the
formation of the chalcopyrite protore. We propose that the
incursion of brines with a major component of seawater, possibly mediated by evaporites, occurred after high-temperature Fe and K metasomatism by magmatic-hydrothermal

438

FLUIDS IN THE MANTOVERDE IOCG DISTRICT, III REGION, CHILE

fluids and at the outset of a period of intense hydrolytic alteration. Rather than representing merely a retrograde acidification of hydrothermal fluids, the chlorite-dominated stage II
assemblages are therefore envisaged as a key link between the
barren magnetite-rich mineralization and the stage III
hematitic breccias and stockworks, which host all significant
chalcopyrite. Mantoverde therefore adheres in fundamental
respects to the genetic model advocated by Barton and Johnson (1996, 2000), as do La Candelaria (Ullrich and Clark,
1999; Ullrich et al., 2001) and Ral-Condestable (Ripley and
Ohmoto, 1977; de Haller et al., 2002, 2006). Whereas the ore
metals in central Andean iron oxide-copper-gold deposits are
probably magma derived (e.g., Ullrich and Clark, 1999; Sillitoe, 2003), an influx of sulfur from a nonmagmatic source was
apparently essential for even subeconomic copper concentration in these hydrothermal centers. The paths followed by
such regionally derived brines may be partially revealed by
the district-scale distribution of scapolitic alteration.
Acknowledgments
This study is a component of the senior authors Ph.D. research and constitutes a contribution to the Queens University Central Andean Metallogenic Project (Q CAMP). Field
studies were generously facilitated by the geology staff of the
Mantoverde division of Anglo American-Chile, which has authorized publication of the paper. Kerry Klassen, Queens Facility for Isotopic Research, is greatly thanked for her valuable
support during isotopic analysis. Joan Charbonneau assisted
with the preparation of the revised manuscript. Critical and
constructive reviews by R.H. Sillitoe, J. Perell, S.J.
Matthews, and P.J. Pollard have considerably improved the
manuscript, as did Mark Hannington, editor. This project has
received generous financial and logistic support from AngloAmerican plc and was supported by the Natural Sciences and
Engineering Research Council of Canada (NSERC), the
Canada Foundation for Innovation (CFI), and the Ontario
Innovation Trust (OIT). Anglo American has also contributed
to the costs of color printing.
December 2, 2005; April 9, 2007
REFERENCES
berg, G., Aguirre, L., Levi, B., and Nystrm, J.O., 1984, Spreading-subsidence and generation of ensialic marginal basins: An example from the
Early Cretaceous of central Chile: Geological Society of London Special
Publication 16, p. 185193.
Aguirre, L., Levi, B., and Nystrm, J.O., 1989, The link between metamorphism, volcanism and geotectonic setting during the evolution of the
Andes: Geological Society of London Special Publication 43, p. 223232.
Alt, J.C., 1999, Very low-grade hydrothermal metamorphism of basic igneous
rocks, in Frey, M., and Robinson, D., eds., Low-grade metamorphism: Oxford, U.K., Blackwell Science, p. 169201.
rkai, P., 1991, Chlorite crystallinity: An empirical approach and correlation
with illite crystallinity, coal rank and mineral facies as exemplified by
Palaeozoic and Mesozoic rocks of northeast Hungary: Journal of Metamorphic Geology, v. 9, p. 723734.
Astudillo, C., 2001, Distribucin y Caracterizacin de Carbonatos en Mantoverde, Provincia de Chaaral, Tercera Regin, Chile: Unpublished B.Sc.
thesis, Antofagasta, Chile, Universidad Catlica del Norte, 103 p.
Barton, M.D., and Johnson, D.A., 1996, Evaporitic source model for igneous-related Fe oxide-(REE-Cu-Au-U) mineralization: Geology, v. 24, p.
259262.
2000, Alternative brine sources for Fe-oxide (-Cu-Au) systems: Implications for hydrothermal alteration and metals, in Porter, T.M., ed.,
0361-0128/98/000/000-00 $6.00

439

Hydrothermal iron-oxide copper-gold and related deposits: A global perspective: Adelaide, Porter Geoscience Consultancy Publishing, v. 2, p.
4360.
Benavides, J., 2006, Iron oxide-copper-gold deposits of the Mantoverde area,
northern Chile: Ore genesis and exploration guidelines: Unpublished
Ph.D. thesis, Kingston, Ontario, Queens University, 355 p.
Berg, K., and Breitkreuz, C., 1983, Mesozoiche Plutone in der Nordchilenischen Kstenkordillere: Petrogenese, Geochronologie, Geochemie und Geodynamik Mentelbetonter Magmatite: Geotektonische Forschungen 66,
107 p.
Bonson, C.G., 1998, Fracturing, fluid processes and mineralization in the
Cretaceous continental magmatic arc of northern Chile (2515'2715' S):
Unpublished Ph.D. thesis, U.K., Kingston University, 357 p.
Bookstrom, A.A., 1977, The magnetite deposits of El Romeral, Chile: ECONOMIC GEOLOGY, v. 72, p. 11011130.
Boric, R., Daz, F., and Maksaev, V., 1990, Geologa y yacimientos metalferos
de la Regin de Antofagasta: Santiago, Chile, Servicio Nacional de Geologa y Minera Boletn 40, 246 p.
Bottinga, Y., and Javoy, M., 1973, Comments on oxygen isotope geothermometry: Earth and Planetary Science Letters, v. 20, p. 250265.
Brown, M., 1991, Comparative geochemical interpretation of Permian-Triassic plutonic complexes of the Coastal Range and Altiplano (2530' to 2630'
S), northern Chile: Geological Society of America Special Paper 265, p.
157177.
Brown, M., Daz, 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.
Chappell, B.W., and White, A., 1974, Two contrasting granite types: Pacific
Geology, v. 8, p. 173174.
Clayton, R.N., and Mayeda, T.K., 1963, The use of bromine pentafluoride in
the extraction of oxygen from oxides and silicates for isotopic analysis:
Geochimica et Cosmochimica Acta, v. 27, p. 4352.
Cole, D.R., and Ripley, E.M., 1998, Oxygen isotope fractionation between
chlorite and water from 170 to 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 Fe-Cu-Au:
Yacimiento Mantoverde, III Regin de Atacama, Chile: IX Congreso Geolgico Chileno, Puerto Varas, Actas, p. 97101.
Dallmeyer, R.D., Brown, M., Grocott, J., Taylor, G.K., and Treolar, 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., Zuiga, A.J., Corfu, F., and Fontbot, L., 2002, The iron
oxideCu-Au deposit of Ral-Condestable, Mala, Per [abs.]: Congreso
Geolgico Peruano, 11th, Lima, Resmenes, p. 80.
de Haller, A., Corfur, F., Fontbot, L., Schaltegger, U., Barra, F., Chiaradia,
M., Frank, M., and Zuiga, A.J., 2006, Geology, geochronology, and Hf and
Pb isotope data of the Ral-Condestable iron oxide copper-gold deposit,
central coast of Per: ECONOMIC GEOLOGY, v. 101, p. 281310.
Espinoza, S., 1990, The Atacama-Coquimbo ferriferous belt, northern Chile,
in Fontbot, L., Amstutz, G., Cardozo, M., Cedillo, R., and Frutos, J., eds.,
Stratabound ore deposits in the Andes: Berlin, Springer-Verlag, p. 353364.
Frey, M., de Capitani, D., and Liou, J.G., 1991, A new petrogenetic grid
for low-grade metabasites: Journal of Metamorphic Geology, v. 9, p.
497509.
Frietsch, R., Tuiska, P., Martinsson, O., and Perdahl, J.-A., 1997, Early Proterozoic Cu(-Au) and Fe ore deposits associated with regional Na-Cl metasomatism in northern Fennoscandia: Ore Geology Reviews, v. 12, p. 134.
Garca, F., 1967, Geologa del Norte Grande de Chile [abs.]: Symposium
Sobre el Geosinclinal Andino, Sociedad Geolgica de Chile, 3rd, Santiago,
Abstracts, p. 138.
Gelcich, S., Davis, D.W., and Spooner, E.T.C., 2002, New U-Pb ages for host
rocks, mineralization and alteration of iron oxide (Cu-Au) deposits in the
Coastal Cordillera of northern Chile: South American Symposium on Isotope Geology, 4th, Salvador de Bahia, Brazil, Proceedings, p. 6365.
Godoy, E., and Lara, L., 1998, Hoja Chaaral-Diego de Almagro, III Regin
de Atacama: Santiago, Chile, Servicio Nacional de Geologa y Minera,
Mapas Geolgicos 3, escala 1:100,000.
Graham, C.M., Viglino, J.A., and Harmon, R.S., 1987, Experimental study of
hydrogen-isotope exchange between aluminous chlorite and water and of
hydrogen diffusion in chlorite: American Mineralogist, v. 72, p. 566578.

439

440

BENAVIDES ET AL.

Grocott, J., and Taylor, G., 2002, Magmatic arc fault systems, deformation
partitioning and emplacement of granitic complexes in the Coastal
Cordillera, north Chilean Andes (2530' to 2700' S): Journal of the Geological Society of London, v. 159, p. 425442.
Grocott, J., Brown, M., Dallmeyer, R.D., Taylor, G.K., and Treolar, P.J., 1994,
Mechanisms of continental growth in extensional arcs: An example from
the Andean plate-boundary zone: Geology, v. 22, p. 391394.
Hawkes, N., Clark, A.H., and Moody, T.C., 2002, Marcona and Pampa de
Pongo: Giant Mesozoic Fe-(Cu,Au) deposits in the Peruvian Coastal belt in
Porter, T.M., ed., Hydrothermal iron-oxide copper-gold and related deposits: A global perspective: Adelaide, Porter Geoscience Consultancy
Publishing, v. 2, p. 115130.
Haynes, D.W., Cross, K.C., Bills, R.T., and Reed, M.H., 1995, Olympic Dam
ore genesis: A fluid mixing model: ECONOMIC GEOLOGY, v. 90, p. 281307.
Haynes, S.J., 1975, Granitoid petrochemistry, metallogeny and lithospheric
plate tectonics, Atacama province, Chile: Unpublished Ph.D. thesis,
Kingston, Ontario, Canada, Queens University, 362 p.
Ishihara, S., 1977, The magnetite-series and ilmenite-series granitic rocks:
Mining Geology, v. 27, p. 293305.
Johnson, J.P., and McCulloch, M.T., 1995, Sources of mineralizing fluids for
the Olympic Dam deposit (South Australia): Sm-Nd isotopic constraints:
Chemical Geology, v. 121, p. 177199.
Lara, L., and Godoy, E., 1998, Hoja Quebrada Salitrosa, III Regin de Atacama: Santiago, Chile, Servicio Nacional de Geologa y Minera, Mapas Geolgicos 4, escala 1:100.000.
Lpez, E., 2002, Caracterizacin Petrogrfica y Estudio de la Alteracin
Hidrotermal y Mineralizacin del Distrito Mantoverde, Provincia de
Chaaral, Tercera Regin, Chile: Unpublished B.Sc. thesis, Antofagasta,
Chile, Universidad Catlica del Norte, 119 p.
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.
2001a, The Candelaria-Punta del Cobre iron oxide Cu-Au(-Zn-Ag) deposits, Chile: ECONOMIC GEOLOGY, v. 96, p. 17991826.
2001b, The Punta del Cobre Formation, Punta del Cobre-Candelaria
area, northern Chile: Journal of South American Earth Sciences, v. 14, p.
401433.
Menuge, J.F., Barton, M.D., Seeger, C.M., and Brewer, T.S., 2002, The relationship of Fe oxide-P-REE mineralization to magmatism: the Pea Ridge
deposit, Missouri [ext. abs.]: Transactions of the Institution of Mining and
Metallurgy, sec. B, v. 3, p. B155B156.
Moraga, A., 1977, Cuadrngulo Quebrada Desierto (Quebrada Salitrosa):
Santiago, Chile, Instituto de Investigaciones Geolgicas, Carta Geolgica
de Chile 25, escala 1:50.000.
Mpodozis, C., and Ramos, V., 1990, The Andes of Chile and Argentina: Circum-Pacific Council for Energy and Mineral Resources, Earth Science Series, v. 11, p. 5990.
Naranjo, J. A., 1978, Zona interior de la Cordillera de la Costa entre los
2600' y 2620' S, Regin de Atacama: Instituto de Investigaciones Geolgicas, Carta Geolgica de Chile 34, 48 p.
1987, Interpretacin de la actividad Cenozoica superior a lo largo de la
zona de la falla de Atacama, norte de Chile: Revista Geolgica de Chile, v.
31, p. 4355.
Nystrm, J.O., and Henrquez, F., 1994, Magmatic features of iron ores of
the Kiruna type in Chile and Sweden: Ore textures and magnetite geochemistry: ECONOMIC GEOLOGY, v. 89, p. 820839.
Ohmoto, H., and Goldhaber, 1997, Sulfur and carbon isotopes, in Barnes,
H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed.: New York,
Wiley Interscience, p. 517600.
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,
John Wiley and Sons, p. 509567.
ONeil, J.R., and Taylor, H.P., Jr., 1967, The oxygen isotope and cation exchange chemistry of feldspars: American Mineralogist, v. 52, p.14141437.
Palacios, C., 1977, Metamorfismo regional en rocas jurasicas en el norte de
Chile: Estudios Geolgicos, v. 33, p. 1116.
Pearce, J.A., Harris, N.B.W., and Tindle, A.G., 1984, Trace element discrimination diagrams for the tectonic interpretation of granitic rocks: Journal of
Petrology, v. 25, p. 956983.
Pollard, P.J., 2006, An intrusion-related origin for Cu-Au mineralization in
iron oxide-copper-gold (IOCG) provinces: Mineralium Deposita, v. 41, p.
179187.

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

Pop, N., Heaman, L., Edelstein, O., Isache, C., Zentilli, M., Pecskay, Z.,
Valdman, S., and Rusu, C., 2000, Geocronologa de las rocas gneas y los
productos de alteracin hidrotermal relacionados con la mineralizacin de
Cu-Fe(Au) del Sector Adriana-Carola-Cobriza (Parte Este del Distrito
Punta del Cobre-Candelaria), en base a dataciones U-Pb (en circn),
40Ar/39Ar y K-Ar: IX Congreso Geolgico Chileno, Puerto Varas, Chile,
Actas, p. 155160.
Ripley, E., and Ohmoto, H., 1977, Mineralogic, sulfur isotope, and fluid inclusion studies of the strata-bound copper deposits at the Ral mine, Peru:
ECONOMIC GEOLOGY, v. 72, p. 10171041.
Robinson, D., 1987, Transition from diagenesis to metamorphism in extensional and collision settings: Geology, v. 15, p. 866869.
Ruiz, C., Aguirre, L., Corvaln, J., Klohn, C., Klohn, E., and Levi, B., 1965,
Geologa y yacimientos metalferos de Chile: Santiago, Chile, Instituto de
Investigaciones Geolgicas, 302 p.
Ryan, P.J., Lawrence, A.L., Jenkins, R.A., Matthews, J.P., Zamora, G.,
Marino, W.E., and Urqueta, D.I., 1995, The Candelaria copper-gold deposit, Chile: Arizona Geological Society Digest, v. 20, p. 625645.
Scheuber, E., and Andriessen, P.A., 1990, The kinematic and geodynamic significance of the Atacama fault zone: Journal of Structural Geology, v. 12, p.
243257.
Segerstrom, K., and Parker, R., 1959, Cuadrngulo Cerrillos, Provincia de Atacama: Santiago, Chile, Instituto de Investigaciones Geolgicas, Carta Geolgica de Chile 1, 33 p.
Shi, P., 1992, Fluid fugacities and phase equilibria in the Fe-Si-O-H-S system: American Mineralogist, v. 77, p. 10501066.
Sillitoe, R.H., 2003, Iron oxide-copper gold deposits: An Andean view: Mineralium Deposita, v. 30, p. 787812.
Sillitoe, R., and Perell, J., 2005, Andean copper province: Tectonomagmatic
settings, deposit types, metallogeny, exploration, and discovery: ECONOMIC
GEOLOGY 100TH ANNIVERSARY VOLUME, p. 845890.
Steiger, R., and Jger, E., 1977, Subcommission on geochronology convention on the use of decay constants on geo- and cosmochronology: Earth and
Planetary Sciences Letters, v. 36, p. 359362.
Taylor, H.P., 1997, Oxygen and hydrogen isotope relationships in hydrothermal mineral deposits, in Barnes, H.L., ed., Geochemistry of hydrothermal
ore deposits, 3rd ed.: New York, John Wiley and Sons, p. 229302.
Thiele, R., and Pincheira, M., 1987, Tectonica transpresiva y movimiento de
desgarre en el segmento sur de la zona de Falla Atacama, Chile: Revista
Geolgica de Chile, v. 31, p. 7794.
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.
Ullrich, T.D., Clark, A.H., and Kyser, T.K., 2001, The Candelaria Cu-Au deposit, III regin, Chile: Product of long-term mixing of magmatic and evaporite-sourced fluids [abs.]: Geological Society of America Abstracts with
Programs, v. 33, no. 6, p. A3.
Vila, T., Lindsay, N., and Zamora, R., 1996, Geology of the Mantoverde 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.
Vivallo, W., and Henrquez, F., 1998, Gnesis comn de los yacimientos estratoligados y vetiformes de cobre del Jursico Medio a Superior en la
Cordillera de la Costa, Regin de Antofagasta, Chile: Revista Geolgica de
Chile, v. 25, p. 199228.
Williams, P.J., Barton, M.D., Johnson, D.A., Fontbot, L., de Haller, A.,
Mark, G., Oliver, N., 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., and Grocott, J., 2001, The emplacement of the granitic Las Tazas
complex, northern Chile: The relationships between local and regional
strain: Journal of Structural Geology, v. 21, p. 15131523.
Yapp, C., 1990, Oxygen isotopes in iron (III) oxides 1. Mineral-water fractionation factors: Chemical Geology, v. 85, p. 329335.
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, Lima, Conferencias,
2nd, Instituto de Ingenieros de Minas del Per, Lima, Actas, 13 p.

440

Das könnte Ihnen auch gefallen