You are on page 1of 27

Economic Geology Vol. 96, 2001, pp.

17991826

The Candelaria-Punta del Cobre Iron Oxide Cu-Au(-Zn-Ag) Deposits, Chile


ROBERT MARSCHIK,*
Lehrstuhl fr Lagerstttenlehre, Institut fr Mineralogie, TU Bergakademie Freiberg, Brennhausgasse 14, 09596 Freiberg/Sachsen, Germany
AND

LLUS FONTBOT

Section des Sciences de la Terre, Universit de Genve, Rue des Marachers 13, 1211 Genve 4, Switzerland

Abstract
Several iron oxide-rich Cu-Au(-Zn-Ag) deposits define an approximately 5-km-wide and at least 20-km-long belt along the eastern margin of the coastal batholith near Copiap, Chile. This belt includes the large Candelaria mine and a group of middle- and small-sized mines in the Punta del Cobre district, which is located about 3 km northeast of the Candelaria deposit. Estimated geologic resources of the belt are on the order of 700 to 800 million metric tons (Mt) at 1.0 percent Cu. The ore occurs in veins, breccia, and stringer bodies, and in replacement bodies that are roughly concordant to bedding. The orebodies are hosted mainly by volcanic and volcaniclastic rocks of the Punta del Cobre Formation and, in places, also occur in volcaniclastic intercalations in the lower part of the overlying Early Cretaceous Chaarcillo Group. Most of the larger orebodies in the belt are located where northwest-trending brittle faults intersect the contact between massive volcanic and volcaniclastic rocks. These northwest faults and a major northeast-trending ductile shear zone control portions of the ore of the Candelaria deposit. Chalcopyrite is the only hypogene Cu mineral. The Cu-Au ore is characterized by abundant magnetite and/or hematite and by locally elevated concentrations of Ag, Zn, Mo, and light rare earth elements. The ore is hosted mainly in zones with biotite-potassium feldspar calcic amphibole epidote alteration at Candelaria. In the Punta del Cobre district, ore in the deeper parts of the deposits is similarly associated, whereas at shallow levels it occurs in zones of biotite-potassium feldspar, or albite-chlorite calcite alteration. Mineralization at Candelaria-Punta del Cobre took place under relatively oxidized conditions manifested by the formation of specular hematite. In parts of the district, the pseudomorphic replacement of early specular hematite by magnetite during the main iron oxide formation marks a shift toward more reduced conditions or higher temperatures. The bulk of the magnetite probably formed at temperatures of about 500 to 600C. The main sulfide stage followed with formation of pyrite and chalcopyrite at temperatures of >470 to 328C. Subsequent martitization of the magnetite points to a temperature decrease. Cooling of the hydrothermal system is also indicated by the homogenization temperatures of 236C of saline fluid inclusions in late-stage calcite. Oxygen isotope combined with microthermometric data suggest that magmatic fluids or nonmagmatic fluids equilibrated with magmatic silicates were dominant during the main copper mineralization. Relatively light oxygen isotope signatures of fluids in equilibrium with late calcite suggest mixing with a nonmagmatic fluid (e.g., basinal brines or meteoric waters) during the late stages of hydrothermal activity. Sulfur isotope ratios of chalcopyrite, pyrite, pyrrhotite, and sphalerite from the Bronce, Candelaria, Las Pintadas, Santos, and Socavn Rampa deposits range from 34SCDT values of 0.7 to +3.1 per mil. This narrow range of sulfur isotope ratios near 0 per mil is consistent with sulfur of magmatic origin. Anhydrite from the Candelaria mine paragenetically overlaps with chalcopyrite. Fluid inclusions in this anhydrite homogenize between 340 and 470C and it has 34SCDT values between 14.5 and 17.5 per mil. A sulfate-sulfide value of 13.4 per mil for a sample with coexisting anhydrite and chalcopyrite is consistent with sulfide-sulfate fractionation at temperatures on the order of 400C. Ore lead isotope signatures are homogeneous and similar to those of least altered volcanic host rocks and nearby intrusive rocks. Radiometric ages, including new 40Ar/39Ar ages for hydrothermal alteration at Candelaria, point to a main Cu-Au mineralization event at Candelaria-Punta del Cobre at around 115 Ma. The ages indicate that ore formation was broadly coeval with batholithic granitoid intrusions and with regional uplift. They further imply that the Cu-Au(-Zn-Ag) deposits formed at shallow crustal levels (<3 km). The stable isotope data, the presence of previously reported hypersaline CO2-rich fluid inclusions in the main Cu ore stage and saline fluid inclusions in late-stage calcite, the oxidized character of the first ore-fluid pulses, and the mineralization age coeval with nearby intrusive activity are consistent with, but not unequivocally evidence of, magmatic fluid contribution into the hydrothermal system.

Introduction A GROUP of iron oxide-rich Cu-Au(-Zn-Ag) deposits defines a belt along the eastern margin of the composite coastal
Corresponding author: e-mail, robert.marschik@iaag.geo.uni-muenchen.de *Present address: Institut fr Allgemeine und Angewandte Geologie (IAAG), Ludwig-Maximilians Universitt, Luisenstrasse 37, D-80333 Mnchen, Germany.

batholith, southeast of Copiap, Chile (Fig. 1, Table 1). This belt, referred to as the Punta del Cobre belt, includes the Candelaria deposit with mineable reserves of 470 Mt at 0.95 percent Cu, 0.22 g/t Au, and 3.1 g/t Ag (Marschik et al., 2000), the Punta del Cobre district, and several middle- and smallsized mines with estimated combined reserves plus production of >120 Mt at 1.5 percent Cu, 0.2 to 0.6 g/t Au, and 2 to

0361-0128/01/3208/1799-28 $6.00

1799

1800
25

MARSCHIK AND FONTBOT

a)

71

70

69

b)

7015' W

Taltal
Atacama Fault Zone
26

N
Mine in operation
5-P an am eri ca na

Copiap
Providencia

Teresita

Chaeral

Carmen Manto Verde

Closed mine Prospect Road (paved) Road (unpaved) Copiap river

Q. Paipote

Paipote smelter

27

Cerro Negro Norte Cerro Imn Copiap Candelaria Bandurrias Boqern Chaar L. Colorados Vallenar

Ru ta

Marta-Venus Manto Monstruo Alcaparrosa Copiap Batholith

Mantos de Cobre Santos

28

Tierra Amarilla
Q. Melndez

2730' S

Algarrobo
29

City
Iron deposit
El Bronce

Candelaria

Cristales El Tofo

Trinidad Socavn Rampa Resguardo Carola

Iron oxide Cu-Au deposit


Pliocene to Recent Upper Cretaceous to Miocene Lower Cretaceous Pre-Cretaceous
0 100
Las Pintadas

Atacama-Kozan

Nantoco

La Serena
30

El Romeral

Q. Nantoco

Ovalle
31

Q. Cerrillos

0 km

km

Q. Los Toros

FIG. 1. (a). Geologic map of the Atacama region. Locations of major magnetite-apatite and iron oxide Cu-Au deposits are indicated (modified from Espinoza, 1990). (b). Location map of the Copiap area, showing selected iron oxide Cu-Au deposits of the Punta del Cobre belt.

8 g/t Ag. These deposits represent variations of essentially the same hydrothermal system but differ in size, intensity, and types of alteration, host rock, and position with respect to the contact metamorphic aureole of the Copiap batholith.

Historically, the mines of the Punta del Cobre belt have been grouped into several mining districts (Segerstrom and Ruiz, 1962). The mines Carola, Santos, and Socavn Rampa form part the Punta del Cobre district, which is located about

TABLE 1. Locations of Selected Mineral Deposits of the Punta del Cobre Belt Mine Alcaparrosa Bronce Candelaria Carola Lar Las Pintadas Manto Monstruo Manto Verde Mantos de Cobre Providencia Resguardo Santos Socavn Rampa Trinidad Venus-Marta
1No

UTM East 374000 372300 372600 377000 372900 366400 373750 377000 375900 376700 376900 375600 377000 376700 373500

UTM North 6962000 6955500 6956500 6956500 6956300 6947400 6962250 6957000 6962700 6972600 6957900 6961100 6958000 6959200 6963750

Characteristic hypogene ore mineralogy Mt-cpy-py Mt-cpy-py Mt-cpy-py (sl, po, mo) Mt-hm-cpy-py (sl) Mt-cpy-py Mt-cpy-py Mt-cpy-py Hm-cpy-py Mt-cpy-py Hm-cpy-py Cpy-py-hm-mt Mt-cpy-py-hm Cpy-hm-py Mt-hm-cpy-py Mt-cpy-py-hm

Main alteration Kspar, bio, act, qtz Scap, gt Bio, kspar, act, qtz Kspar, bio, chl, cte (act) Calc-silicate1 Gt, act, kspar Calc-silicate1 Kspar, chl, cte Kspar/ab-chl(-act) Ab, chl, cte, (kspar) Ab, chl, cte, (kspar) Kspar, bio, chl, (ab) Ab, chl, cte, (kspar) Kspar, bio, chl, (ab) Gt, act

Estimated reserves plus production 10 Mt @ 1.4% Cu 470 Mt @ 0.95 % Cu, 0.22 g/t Au, 3.1 g/t Ag 20 Mt @ 1.4% Cu 4.0 Mt @ 1.0-1.5% Cu 1.5 Mt @ 1.5% Cu 1.5 Mt @ 1.45% Cu 6 Mt @ 1.8-2.0% Cu, 0.4-0.5 g/t Au, 7.0 g/t Ag 20 Mt @ 1.5% Cu, 0.4-0.5 g/t Au, 7.0 g/t Ag 25 Mt @ 1.2-2% Cu, 0.2-0.3 g/t Au, 7.0 g/t Ag 15 Mt @ 1.5% Cu, 0.2-0.3 g/t Au, 7.0 g/t Ag

data; calc-silicate alteration in host sequence Abbreviations: ab = albite, act = actinolite, bio = biotite, chl = chlorite, cpy = chalcopyrite, cte = calcite, gt = garnet, hm = hematite, kspar = K feldspar, mo = molybdenum, mt = magnetite, po = pyrrhotite, py = pyrite, qtz = quartz, scap = scapolite, sl = sphalerite 0361-0128/98/000/000-00 $6.00

1800

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1801

3 km northeast of Candelaria (Figs. 1 and 2). These mines together with Candelaria display mineralization and associated alteration features that are similar to those found in deposits of the iron oxide (Cu-U-Au-REE) class as defined by Hitzman et al. (1992) based on Proterozoic examples. Genetic hypotheses proposed to explain the formation of the deposits of this class include hydrothermal models invoking magmatic fluid-dominated (e.g., Gow et al., 1994; Rotherham et al., 1998; Williams, 1998; Williams et al., 1999; Pollard, 2000) or saline nonmagmatic fluids (e.g., Battles and Barton, 1995; Haynes et al. 1995; Barton and Johnson, 1996, 2000; Barton et al., 1998). Metallogenic aspects and models for the ore formation at Candelaria and/or Punta del Cobre have been discussed, e.g., by Camus (1980), Hopf (1990), Ryan et al. (1995), Marschik and Fontbot (1996), Ullrich and Clark (1999), Marschik et al. (2000), and Mathur et al. (2002). We present a summary of field and analytical data on the Candelaria deposit and the Punta del Cobre district. We discuss the mineralogy,

paragenetic sequence, and the distribution of the principal alteration assemblages at the district scale and present a compilation of available data regarding the timing of ore formation and mineralization processes. Geologic Context The Candelaria-Punta del Cobre iron oxide Cu-Au(-Zn-Ag) deposits occur in the Chilean coastal cordillera (Fig. 1). The deposits are located to the east of the nearby main branches of the Atacama fault zone, which stretches over 1,000 km along the Chilean coast. The Atacama fault zone is a subduction-linked arc-parallel strike-slip fault system that has been active at least since Jurassic times (e.g., Scheuber and Andriessen, 1990; Brown et al., 1993; Scheuber et al., 1995; Dallmeyer et al., 1996). This fault system controlled mineralization of many of the iron deposits of the Chilean iron belt that also occur in the Chilean coastal cordillera (e.g., Bookstrom, 1977; Thiele and Pincheira, 1987; Espinoza, 1990;
7015' W

Tailing impoundments Mine area Open pit Mine in operation Mine closed Project
Kmd
40Ar/39Ar

Legend
Alluvium (Recent) Atacama Gravels (Miocene)

Co

pia

p riv

er

Paipote smelter

Punta Negra Viita Azul Marta-Venus Q. Florida Manto Monstruo Mantos de Cobre

Ladrillos district

Copiap Batholith (Early Cretaceous)


Tonalite (Kt) Monzodiorite (Kmd)
Q. Melndez

FSZ
Alcaparrosa

111.5 0.4 Ma
Santos

Quartz monzonite (Kqm)


Punta del Cobre district
Tierra Amarilla

Diorite (Khd) Intrusive rocks (undifferentiated)

Kqm

San Gregorio (Cu-Au)

Trinidad

2730' S San Francisco (Cu-Au)

Sacovn Rampa La Candelaria Carola

A'
Resguardo

Bandurrias Group
(late Valanginian to Aptian-Albian)

Khd

A
La Tigresa (Cu-Au) Transito (Au)

Chaarcillo Group
(late Valanginian to late Aptian)

OSZ

Bronce Atacama-Kozan

Punta del Cobre Formation


(Tithonian(?)-pre-late Valanginian)

CSZ
Ojancos Nuevo district
Nantoco

Shear zone CSZ FSZ OSZ Candelaria shear zone Florida shear zone Ojancos shear zone High-angle fault

40Ar/39Ar 109.9 0.4 Ma 109.9 1.7 Ma

Khd Las Pintadas district


Q. Los Toros

Q. Nantoco Q. Los Algarrobos

Low or medium angle fault

Kt
K-Ar 111 3 Ma

Q. Las Pintadas

N
2

Axis of the Tierra Amarilla Anticline

0 km

Las Pintadas

FIG. 2. Geologic map of the Candelaria-Punta del Cobre area. The Lar mine mentioned in the text was located where the Candelaria pit is shown. The position of the schematic cross section of Figure 6 is indicated. 0361-0128/98/000/000-00 $6.00

1801

1802

MARSCHIK AND FONTBOT

Mnard 1995). In the Copiap area, the Chilean iron belt is represented, e.g., by the Cerro Imn (Espinoza et al., 1994) and Cerro Negro Norte deposits (Vivallo et al., 1995; Fig. 1). The Candelaria-Punta del Cobre iron oxide Cu-Au(-Zn-Ag) deposits and most of the deposits of the Chilean iron belt are hosted in Early Cretaceous arc-derived volcanic and volcaniclastic rocks adjacent to intermediate plutons of the Chilean coastal batholith. Stratified rocks exposed in the Candelaria-Punta del Cobre area represent a facies transition of a continental volcanic arc to the west and northwest (Bandurrias Group) and a shallow marine back-arc basin to the east and southeast (Chaarcillo Group). Sedimentation in the back-arc basin commenced in Berriasian time (Early Cretaceous) with the deposition of the upper part of the volcanic-volcaniclastic Punta del Cobre Formation, which underlies the late Valanginian to Aptian (Early Cretaceous) carbonate rocks of the Chaarcillo Group (Abundancia-Nantoco, Totoralillo, and Pabelln formations; accumulated thickness 1,7002,000 m). Basin inversion, which started in late Aptian times (possibly at around 115 Ma), eventually resulted in the partial erosion of the back-arc sequence (Segerstrom and Parker, 1959; Zentilli 1974; Jurgan, 1977). Granitoid plutons of the Copiap batholith intruded the back-arc deposits in the western portion of the area, causing an extensive contact metamorphic aureole (Tilling, 1962, 1963, 1976). The Copiap batholith consists of several calc-alkaline intrusions ranging from diorite to quartz monzonite (SiO2 5068 wt %). These plutons are intruded, in places, by altered aplitic dikes (SiO2 7276 wt %). Dikes are abundant at the eastern margin of the batholith, near Candelaria. Hydrothermally altered dacite porphyry dikes and sills that locally contain minor sulfide mineralization and postore lamprophyric dikes occur at Candelaria and in the Punta del Cobre district. Portions of the Copiap batholith west of Candelaria are marginally affected mainly by intense sodic(-calcic) alteration that presumably is related to the ore formation (see below). Potassium-argon ages of batholith intrusions range between 119 to 97 Ma (Arvalo, 1994, 1995). The 40 Ar/39Ar ages of 111.5 Ma for a monzodiorite and 109.9 Ma for a granodiorite-tonalite near Candelaria have been determined (Fig. 2; Arvalo, 1999). District Stratigraphy The oldest rocks in the Candelaria-Punta del Cobre area belong to the Tithonian(?)-pre-late Valanginian Punta del Cobre Formation (Marschik and Fontbot, 2001), which hosts most of the iron oxide Cu-Au(-Zn-Ag) orebodies. The Punta del Cobre Formation (Fig. 3) is divided into the volcanic Geraldo-Negro Member (>500 m) and the overlying predominantly volcaniclastic Algarrobos Member (>800 m in drill cores; Marschik and Fontbot, 2001). The GeraldoNegro Member is further subdivided into the Lower Andesites (>300 m) that consist of altered massive andesitic volcanic rocks, and the Melndez Dacites (up to 200 m) comprising intensely alkali-metasomatized lava domes and flows of originally dacitic composition that overlie the Lower Andesites east of the Copiap River, in the Punta del Cobre district, sensu stricto (Fig. 4). The Algarrobos Member is formed by a succession of coarse, poorly bedded volcaniclastic conglomerates and breccias with centimetric to decimetric
0361-0128/98/000/000-00 $6.00

clasts that contain several intercalations of fine-grained sediments such as siltstones, arenites, coarse sandstones, and microconglomerates, commonly on the order of 10 to about 40 m thick, and also lenses of massive volcanic andesitic to basaltic rocks. The Algarrobos Member is characterized by marked lateral changes in thickness and facies. It passes vertically and laterally into the overlying calcareous Chaarcillo Group (Abundancia or Nantoco Formations) with a contact defined by the first continuous bed of massive limestone or its metamorphosed equivalent. The Algarrobos Member contains several horizons of economic or petrologic importance that were defined in the Punta del Cobre district and, to a certain extent, can be correlated within the Candelaria-Punta del Cobre area. These horizons include: the Basal Breccia, the Trinidad Siltstone, and the Upper Lavas. The Basal Breccia (up to 25 m) is a red sedimentary breccia, in places conglomeratic, that contains interdigitations of sandstone and grades into red sandstone toward the south. It is exposed in the Punta del Cobre district, where it overlies the Melndez Dacites. A similar horizon was identified in a drill core from Quebrada Los Algarrobos, where a hematite-bearing red sandstone rests on the Lower Andesites. The Basal Breccia hosts stratiform orebodies and the top of the Basal Breccia commonly marks the upper limit of the mineralization in the Punta del Cobre district. The Trinidad Siltstone (up to 60 m) that overlies the Basal Breccia is mainly composed of silt- and sandstone, chert, and tuffaceous sedimentary rocks, which locally contain elongate, decimetric to metric clasts of brecciated limestone. The Trinidad Siltstone is characterized by strong lateral changes in facies and thickness. In the Socavn Rampa mine, it covers a local erosional surface above the Basal Breccia. The Trinidad Siltstone can be correlated with biotitized originally fine-grained tuffaceous rocks at Candelaria. Whereas this unit is commonly barren in the Punta del Cobre district, it hosts high-grade ore at Candelaria. The Upper Lavas (up to 45 m) form a discontinuous horizon of lenses of basaltic to basalt-andesitic lavas and volcanic breccias with a primitive geochemical signature of less differentiated magmas. This magmatic event is interpreted to reflect the incipient opening of the Early Cretaceous marine backarc basin south of Copiap (Marschik and Fontbot, 2001). Structural Framework The dominating structural elements in the CandelariaPunta del Cobre area are a large northeast-trending antiform (Tierra Amarilla anticlinorium), a southeast-verging foldthrust system (El Bronce fold-thrust system; Arvalo and Grocott, 1997), and a dense set of north-northwest to northwest-trending high-angle sinistral transcurrent faults. The latter control parts of the mineralization (e.g., Camus, 1980; Marschik and Fontbot, 1996). Additionally, northeast- and east-northeasttrending high-angle and moderately (3050) west dipping faults and sinistral east-northeasttrending highangle faults are present. Mylonitic shear zones and cataclastic rocks locally form the contact between intrusive and Early Cretaceous country rocks (e.g., Ojancos shear zone). Ductile deformation is also recorded in the Candelaria-Florida shear zones, affecting volcanic and volcaniclastic rocks close to the batholith contact (Fig. 2). Such deformation is manifested in north-northeasttrending, 30 to 70 west-dipping zones of

1802

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1803

Regional stratigraphy
CHAARCILLO GROUP
Dacite dike/sill

Lithology
"Metasediments"

Main alteration assemblages


Scapolite pyroxene amphibole garnet skarn with garnetite horizons or quartz hornfels Quartz and pyroxene hornfels or pyroxene-scapolite-garnet skarn over biotite hornfels; amphibole veinlets Biotite hornfels or altered (K-feldspar or albite, biotite, amphibole) massive volcanic rocks; scapolite, and amphibole veinlets Intense biotitization (brown and/or green biotite) plus quartz and magnetite amphibole pink garnet, locally minor cordierite; overprinted by pervasive and fracture-controlled amphibole alteration Stratiform magnetite body plus amphibole alteration Volcaniclastic breccia with clasts intensely altered to K-feldspar in matrix of mainly magnetite commonly plus amphibole Albite and/or K-feldspar

Upper Lavas
up to 40 m

"Upper Andesites" up to 200 m (include biotite hornfelses)

Algarrobos Member

Trinidad Siltstone
40-100 m

PUNTA DEL COBRE FORMATION

"Tuffs (and volcaniclastic sediments)" 40-100 m


"Magnetite Manto" "K-feldspar Breccia"

Dacite dike

"Lower Andesites" (A) 200-300 m

Biotite-quartz-magnetite-albite/K-feldspar alteration, in places, overprinted by fracturecontrolled or pervasive amphibole

Geraldo-Negro Member

Lower Andesites
>350 m

"Lower Andesites" (B) >350 m Candelaria terminology (e.g., Ryan et al. 1995; Ullrich and Clark, 1999)

Biotite-quartz-magnetite-Na plagioclase alteration, locally minor K-feldspar and/or minor amphibole

FIG. 3. Schematic lithostratigraphic column of the Punta del Cobre Formation and the vertical distribution of the main alteration mineral assemblages at the Candelaria mine. The fracture-controlled calcic amphibole alteration is indicated. However, there are several other veining events (see paragenetic sequence, Fig. 9). The metasediment unit at Candelaria is roughly equivalent to the Abundancia Formation (Chaarcillo Group).

intensely foliated K-metasomatized (biotite) rocks (Fig. 5a). The Candelaria and Florida shear zones are possibly segments of a major shear-fault zone that may find its continuation in the Inca de Oro area (Sylvester and Palacios, 1992; Sylvester and Linke, 1993). The Candelaria shear zone predated the copper mineralization and is the oldest deformation recognized in the Candelaria mine to date. The CandelariaFlorida shear zones are cut and displaced by sinistral northnorthwest to northwest-trending high-angle faults, the eastnortheasttrending high-angle faults, and by the broadly northeast trending moderately west dipping faults. Shearing must have occurred between Berriasian (the age of the deformed rocks) and Aptian (age of mineralization) times, i.e., at a depth equivalent to or less than the maximum thickness of the late Valanginian to late Aptian Chaarcillo Group (2,000 m plus eroded material) that overlies the sheared rocks. Therefore, ductile deformation took place far above normal ductile-brittle transition and the Candelaria and Florida shear zones are interpreted to represent thermally
0361-0128/98/000/000-00 $6.00

moderated ductile deformation related to batholith emplacement, as suggested for other ductile shear zones associated with high-level intrusions now exposed in the Early Cretaceous magmatic arc of northern Chile (e.g., Grocott et al., 1993; Wilson et al., 2000). Mineralization and Alteration Ore occurrences In the Punta del Cobre belt, copper ore occurs as massive veins (Fig. 5b), in the matrix of hydrothermal breccias, as discontinuous veinlets or stringers in the altered host rocks (Fig. 5c) or superposed on massive magnetite replacement bodies (Fig. 5d), and as replacements and pore infilling of bodies roughly concordant with stratification (mantos, Fig. 5e). Larger orebodies are commonly located where northwest- to northnorthwesttrending faults intersect the contact between the massive volcanic rocks of the Geraldo-Negro Member and the overlying volcaniclastic Algarrobos Member (Figs. 3, 4, 6, and 7).

1803

1804

MARSCHIK AND FONTBOT

Regional stratigraphy

Lithology

Main alteration assemblages

CHAARCILLO GROUP

Limestone and volcaniclastic rocks Volcaniclastic breccia and conglomerate, siltstone, and chert Locally horizons of red chert magnetite

Algarrobos Member

Upper Lavas
up to 45 m

Altered massive volcanic rocks and volcanic breccia in volcaniclastic sediments Chlorite-carbonate sericite hematite alteration

Trinidad Siltstone
up to 60 m

PUNTA DEL COBRE FORMATION

Predominantly red and green chert, siltstone and tuffaceous rocks

Basal Breccia
up to 25 m

Red sedimentary breccia, in places conglomeratic, with sandy interdigitations

Chlorite and carbonate alteration, plus hematite and/or magnetite

Albite-chlorite calcite quartz

Melndez Dacites

Geraldo-Negro Member

up to >200 m

Altered massive volcanic rocks of originally dacitic composition

K-feldspar-chlorite/biotite calcite quartz

Albite-quartz-biotite/chlorite Biotite-quartz-K-feldspar/albite chlorite Altered massive volcanic rocks of originally andesitic composition Fracture-controlled amphibole

Lower Andesites
>300 m

FIG. 4. Schematic lithostratigraphic column of the Punta del Cobre Formation and vertical distribution of the main alteration mineral assemblages in the Punta del Cobre district.

In the Punta del Cobre district, sensu stricto, concordant stratiform bodies are hosted mainly by the Basal Breccia that overlies the Melndez Dacites. Subordinately, they occur also replacing small lenses of clastic sediments within or brecciated tops of volcanic flows. The mantos are commonly

underlain by veins or elongated breccia bodies. The latter may split up into veins at depth. The subvertical orebodies, which constitute the main portion of the mineralization, are emplaced along the northwest- to north-northwesttrending faults and are mainly confined to the volcanic Geraldo-Negro

FIG. 5. (a). Outcrop of the Candelaria shear zone, which is a north-northeasttrending, on average 50 west-dipping zone of intense foliated and K-metasomatized rocks. (b). North-northwesttrending vein of massive chalcopyrite plus minor pyrite in the Candelaria south pit. (c). Stringers of chalcopyrite-pyrite in intensely iron-metasomatized volcanic or volcaniclastic rocks, Candelaria north pit. (d). Massive magnetite with superposed amphibole and chalcopyrite-pyrite, Candelaria south pit. (e). Manto ore hosted in the lower part of the Basal Breccia at the Carola mine. (f). Chalcopyrite-pyrite veinlet cuts magnetite replacements, illustrating that the iron oxide formation preceded sulfide mineralization, Candelaria orebody. (g). Chalcopyrite-pyrite associated with amphibole cut and use postmagnetite albite veinlets, Candelaria orebody. (h) Amphibole containing interstitial chalcopyrite and small crosscutting chalcopyrite veinlets in previously K-metasomatized (K feldspar) rock. (i) Pyrite-chalcopyrite and K feldspar cut magnetite and the superposed pervasive amphibole alteration, Candelaria orebody. (j). Magnetite with superposed amphibole alteration both cut by a quartz plus K feldspar veinlet that was later used by chalcopyrite-pyrite; K feldspar commonly also uses the earlier quartz veinlets. (k) Epidote-pyrite cutting potassium-metasomatized (K feldspar) rock; the epidote-pyrite veinlet was later used by amphibole plus minor chalcopyrite; note the small amphibole veinlet that cuts the epidote-pyrite alteration. (l) Chalcopyrite-pyrite cutting an anhydrite veinlet. (m) Green biotite-quartz-magnetite alteration at Candelaria; chalcopyrite occurs similarly associated with green biotite. (n). Early specular hematite (hmI) from the Carola mine. (o) Pseudomorphic replacement of early specularite (hmI) by magnetite (mushketovite), relics of the original hematite are locally preserved. (p). Magnetite replacing and overgrowing hematite (hmI) during the main iron oxide mineralization. (q). Pyrrhotite as infilling between pyrite, chalcopyrite occurs in fractures of pyrite. (r). Late hematite (hmII) and martitized magnetite.

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

1804

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1805

a)
dike shear zone

b)

c)

d)

e)

f)
cpy-py mt mt
1 cm

mt plus cpy-py

1 cm

g)
ab

h)
amph kspar

i)
amph kspar cpy mt py-cpy

amph-cpy-py

1 cm

1 cm

1 cm

j)

k)
amph cuts epi-py kspar epi-py mt qtz +cpy-py amph cpy
1 cm 1 cm

l)
anh kspar py anh

amph

mt cut by anh cpy-py


1 cm

m)

n)

py

o)

hm I

hm I qtz mt bio
1 m 1 mm 1 m

cpy

mt mushketovite

p)
mt hm I

q)

cpy py po

r)
mt hm II

mt

cpy
1 m

FIG. 6. Schematic west-southwesteast-northeast section through the Candelaria-Punta del Cobre area. Orebodies at 1 m 1 mm Punta del Cobre are projected and are not to scale. The location of the cross section is shown in Figure 2. 0361-0128/98/000/000-00 $6.00

1805

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1807

%) and biotite-quartz-magnetite plus sodic plagioclase and/or K feldspar in the Lower Andesites show a strong spatial relationship with ore. Potassic assemblages pass at shallower levels into sodic assemblages of albite-chlorite calcite quartz (with Na2O up to 10 wt %). Sodium metasomatized rocks are considered to represent a peripheral part of the hydrothermal system. However, there is also ore in albitized rocks without evidence of related potassic alteration, e.g., in parts of the Socavn Rampa, Resguardo, and Mantos de Cobre mines (Fig. 7). Sodic alteration is best developed in the Melndez Dacites and particularly in the upper part of the lava dome south of Quebrada Melndez and north of the Manto Verde mine. Intensity of carbonatization and chloritization tends to increase higher in the stratigraphic section, i.e., toward the volcanicsedimentary rock contact and beyond, whereas the generally weak to moderate developed silicification diminishes. Chlorite-calcite hematite assemblages are typically found in the sedimentary rocks of the Basal Breccia and the Trinidad Siltstone and the overlying barren lavas and volcanic breccia of the Upper Lavas (Figs. 4 and 7). Dacite porphyry dikes are affected by various types of hydrothermal alteration. At Candelaria, these dikes commonly show sodic (albite) or potassic (K feldspar) alteration deeper in the section, and a sodium or potassium metasomatized core in an envelope of scapolite-garnet-pyroxene magnetite endoskarn with or without minor chalcopyrite-pyrite in the upper part. Dacite dikes in the Chaarcillo Group above the Candelaria orebody and in a similar level at Quebrada El Buitre are affected by sodic-calcic alteration. In the Punta del Cobre district, the dacite porphyries are intensely sodium or potassium metasomatized (e.g., the Trinidad and Carola mines) and, locally, host a sulfide-bearing stockwork (Marschik and Fontbot, 1996). Postore lamprophyric dikes that used similar tectonic structures as the Cu mineralization in the Punta del Cobre district are commonly affected by carbonatization and chloritization. Extensive areas of sodic or sodic-calcic alteration occur in plutonic, subvolcanic or volcanic, and sedimentary rocks along the eastern margin of the Copiap batholith (Fig. 8). Sodic alteration of igneous rocks commonly results in strong albitization, which leaves a bleached white- or light gray-colored rock. It is particularly well developed south and southwest of the Candelaria tailings pond. Locally observed albite-amphibole minor epidote assemblages may be caused by superposition of calcic amphibole on previous albitization or alternatively by a different sodic-calcic alteration event similar to that described by Dilles and Einaudi (1992) in the Ann Mason porphyry system. An early pervasive albitization that preceded potassic alteration is recognized in the Punta del Cobre district (Marschik and Fontbot, 1996). Andesitic host rocks at Candelaria were also affected by pervasive sodic prior to pervasive potassic alteration and mineralization (spilitization of Ullrich and Clark, 1999). We correlate this early albitization with the albitization found in igneous and sedimentary rocks west and southwest of the Candelaria deposit. Voluminous sodic scapolite-rich assemblages developed in rocks of the Abundancia Formation around Candelaria. Sodic scapolite commonly is associated with calcic amphibole and/or pyroxene, minor epidote andraditic garnet but may occur without significant amounts of these calc-silicates.
0361-0128/98/000/000-00 $6.00

These scapolite-rich beds may represent metamorphosed evaporitic horizons that are described to occur in the Lower Cretaceous rocks in the area (Segerstrom, 1962; Hopper and Correa, 2000). Albitization west and southwest of Candelaria is locally associated with minor disseminations of pyrite trace chalcopyrite and/or veinlets and disseminations of hematite, whereas rocks affected by scapolite amphibole and/or pyroxene alteration may locally host small magnetite chalcopyrite-pyrite mantos (e.g., the El Bronce mine; Daz, 1990) or contain traces of pyrite chalcopyrite, commonly in veinlets. Albitization occurs as discordant alteration within the district (e.g., Quebrada, Melndez; southwest of Candelaria; at the junction of Quebrada Nantoco and Quebrada Los Toros). In contrast, scapolite is largely confined to the layered rocks of the Chaarcillo Group and the uppermost part of the Punta del Cobre Formation near the batholith contact. It is found also in veinlets, commonly together with amphibole, in the altered margins of the batholith. Thermal contact metamorphism caused mineralogical changes without significant modification of the original geochemical composition in the affected andesitic volcanic rocks adjacent to the batholith contact. Calcic amphibole-rich assemblages developed in volcanic rocks close to the batholith contact grading into epidote-chlorite-rich assemblages farther outboard to the east (Marschik and Fontbot, 1996). These contact metamorphic assemblages locally grade into hydrothermal assemblages of the alkali metasomatized rocks that in places host the ore deposits. Ore mineralogy The ore consists essentially of magnetite and/or hematite, chalcopyrite, and pyrite, with local pyrrhotite, sphalerite, trace quantities of molybdenite, and arsenopyrite (see below and Fig. 9). Native gold occurs mainly as micron-sized inclusions in chalcopyrite. Ryan et al. (1995) described gold-filling microfractures in pyrite and Hopf (1990) an Hg-Au-Ag alloy. Minerals in the poorly developed supergene oxidation and enrichment zones include malachite, chrysocolla, massive and sooty chalcocite, and covellite (Sillitoe and Clark, 1969). Gangue mineralogy consists predominantly of quartz and anhydrite at Candelaria and calcite and/or quartz at Punta del Cobre. Tourmaline and traces of fluorite occur locally (Hopf, 1990). The Candelaria, Carola, and Trinidad deposits locally contain several hundreds parts per million of light rare earth elements (Fig. 10, Table 3; Marschik et al., 2000). Allanite is the only rare earth element-bearing mineral identified under the microscope so far. A detailed description of the hypogene ore minerals and their distribution is given in Hopf (1990) and Ryan et al. (1995). Paragenetic sequence The paragenetic sequences in the Candelaria deposit and in the Punta del Cobre district are similar (Fig. 11). Differences between the deposits arise from their relative position within the hydrothermal system, as detailed below. Intense iron metasomatism accompanied by early potassic alteration and silicification postdates large-scale early pervasive albitization. Both Ti-rich brown (TiO2 >2.0 wt %) and green hydrothermal biotite (TiO2 <1.8 wt %) are associated

1807

TABLE 2. Main Alteration Types and Their Occurrences

1808

Alteration type Host rocks Volcanic, volcaniclastic, and intrusive rocks Upper parts at P. del Cobre (e.g., Socavn Rampa), Trinidad area, Q. Melndez, Ladrillos district, at the batholith margin W and SW of Candelaria Beneath Candelaria; deeper parts of P. del Cobre Peripheral None to trace Na2O (up to 7 wt %) > K2O (2-3 wt %); CaO and MgO values about the same (2.5-4 wt %); L.O.I. low (<1 wt %) Post-Na pre-Ca associated with main Fe oxide formation; syn-Ca(?) Post early Na; pre-Ca; syn-Ca K2O up to 11 wt %; Na2O and CaO low (<1-2 wt %) Discordant; more than one event Peripheral None to strong Earliest hydrothermal alteration recognized Na2O up to 10 wt %; K and Ca low (<1-2 wt %); elevated CaO and L.O.I. values indicate carbonatization; high MgO, low CaO values indicate chloritization Location Comments Discordant; there are two sodic alteration events at and near Candelaria

Mineral assemblages1

Style

Relative position2

Spatially associated mineralization Relative time relationship

Typical whole-rock analysis3

0361-0128/98/000/000-00 $6.00 Volcanic rocks Volcaniclastic and volcanic rocks Candelaria; lower to intermediate parts at P. del Cobre (e.g., Santos, Carola) Intermediate parts at P. del Cobre (e.g., Carola) Internal, intermediate levels Strong Internal, intermediate levels Strong MARSCHIK AND FONTBOT Volcanic rocks K2O high (4-6 wt %) and low Na2O (<1 wt %); MgO high (2-4 wt %; chlorite); CaO and L.O.I. low (<1-2 wt %); in case of carbonatization higher CaO and L.O.I. Commonly strong Late In igneous rocks: CaO and MgO high and about the same (5-9 wt %); Na2O and K2O low (<1-3 wt %) Strong CaO and K2O high (4-5 wt %); Na2O low (<2 wt %) Peripheral to distal None to weak Pre- or syn-peak contact metamorphism; this event postdates main Fe metasomatism at Candelaria Not always associated with mineralization; typically present as veinlets in the P. del Cobre district In general Ca assemblages typically overprint potassic assemblages Concordant implying stratigraphic control; restricted to the conatct aureole Candelaria; lower parts Internal at Punta del Cobre (e.g., Santos, Carola) Volcaniclastic and volcanic rocks (Volcani)clastic, intrusive and marly rocks Above Candelaria, regional scale alteration in Ka Fm. between Q. Alcaparrosa and Q. Nantoco Candelaria orebody; deeper levels at P. del Cobre (e.g., Santos) Internal

Na

Ab or Na plag (chlcteqtz)

Commonly pervasive

Na plag-qtz-bio amph (kspar)

Pervasive

Bio ksparqtz-mt

Commonly pervasive; locally veinlets (e.g., Candelaria)

1808

Kspar-bio/chl qtz cte

Pervasive

Ca

Amphibole

Veins, veinlets, Volcanic, and pervasive volcaniclastic, carbonate, and intrusive rocks

Ca-K

Amph-bio qtz kspar epi

Veinlets, pervasive

Na-Ca

Sodic scap px amph epi

Pervasive, subordinate veinlets

TABLE 2. (Cont.)

Alteration type Host rocks Igneous rocks At the batholith margin W of Candelaria Peripheral to distal None to trace Na2O (6-9 wt %) > CaO (1.5-2.5 wt %); K2O > 1 wt % Location Comments

Mineral assemblages1

Style

Relative position2

Spatially associated mineralization Relative time relationship Discordant, transitional to the almost pure albitization SW of Candelaria; possibly a result of superposition of early albitization and late amph alteration or caused by a sodic-calcic event coeval with the scap-px amph formation above Candelaria Variable proportions of amph and bio; commonly without significant metasomatic changes in host-rock composition CaO 8-10 wt %; Na2O 3-4 wt %; K2O 1-3 wt % Without metasomatic changes in rock composition

Typical whole-rock analysis3

0361-0128/98/000/000-00 $6.00 Volcaniclastic and volcanic rocks Near the batholith contact; Punta Negra; in places at Q. La Pepita; near AtacamaKozan prospect; Upper Lavas at Candelaria Q. Nantoco, Q. Los Toros, Q Rivera and W of it Distal None Distal None Na2O (5-8 wt %) CaO (3-5 wt %) MgO high (up to 7 wt %); K2O variable (up to 3 wt %) Volcanic and volcaniclastic rocks CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

ab-amph epi ( sodic scap chl sericite)

Pervasive albitization; fracture controlled and disseminated amphibole

Upper Amph bio greenschist chl epi facies sericite

Pervasive

1809

Lower Epi-chl-cte greenschist qtz amph facies (propylitic alteration)

Pervasive

1 Abbreviations:

ab = albite, amph = amphibole, bio = biotite, chl = chlorite, cte = calcite, epi = epidote, kspar = K feldspar, plag = plagioclase, px = pyroxene, qtz = quartz, scap = scapolite Ka Fm. = Abundancia Formation, L.O.I. = loss of ignition Q. = Quebrada (valley or gorge) 2 Position relative to a postulated center of the ore system 3 Typical values of whole-rock analysis strongly depend on Fe contents Combined Ca-K or Na-Ca alteration mineral assemblages, in places, may be the result of superposition of Ca on preexisting K or Na alteration Data based on Marschik and Fontbot (1996)

1809

1810

MARSCHIK AND FONTBOT

W
El Bronce deposit 750 m

700 m
scap gt

650 m
to scale

0 m

50

Candelaria deposit

600 m 500 m 400 m 300 m


bio-qtz-mtkspar plus abundant Ca amph scap gt px Ca amph bio-qtz-alm gt cord commonly plus Ca amph kspar or Na plag-bio-Ca amph

200 m 100 m sea level


to scale

Na plag-qtz-bio-mt kspar minor Ca amph

100 m

Orebody, 0.4 % Cu contour

Punta del Cobre district Resguardo mine Socavn Rampa mine 800 m Hematite associated with ore
ab-chl-cte qtz kspar-chl/bio cte qtz chl-cte-hm

Carola mine

600 m

Magnetite-rich ore Massive magnetite

400 m Santos mine 200 m


bio-qtz + Na plag or kspar mt
not to scale

sea level
Ca-amph overprint

hm

Magnetite-hematite predominance

FIG. 7. Distribution of iron oxide minerals and main alteration types in the Candelaria-Punta del Cobre iron oxide CuAu(-Zn-Ag) deposits. Sections through the El Bronce deposit (after Daz, 1990), the Candelaria deposit (after Martin et al., 1997), and the Punta del Cobre district (modified from Flores, 1997) are shown. The El Bronce deposit was located above the Candelaria deposit to the west of the Bronce fault (Fig. 2). Abbreviations are given in Table 2. 0361-0128/98/000/000-00 $6.00

1810

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1811

CSZ FSZ OSZ

Candelaria shear zone Florida shear zone Ojancos shear zone


Viita Azul

7015' W
Co pia p rive r

Alteration type (at the surface)


Na or Na-Ca alteration Pervasive albitization (chlorite calcite quartz) Albite or Na-plagioclase amphibole epidote minor scapolite

Paipote smelter

Punta Negra

Q. Florida

Na-scapolite pyroxene amphibole skarn (garnetite horizons) K or Ca-K alteration


Q. Melndez

FSZ

Biotite and/or K-feldspar (chlorite calcite quartz) Calcic-potassic (mainly Ca-amphibole epidote, biotite, K-feldspar) Biotite Ca-amphibole (foliated rocks) Thermal contact metamorphic alteration Garnet pyroxene scapolite skarn Ca-amphibole biotite Epidote-chlorite calcite

Tierra Amarilla

2730' S

La Candelaria

OSZ CSZ
Nantoco

Q. Nantoco Q. Las Pintadas


Q. Los Toros

Q. Los Algarrobos

N
2

0 km

Las Pintadas

FIG. 8. Surface distribution of main alteration mineral assemblages in the Candelaria-Punta del Cobre area. See geologic map in Figure 2 for stratigraphic units.

with widespread pervasive magnetite alteration. Green-colored biotite is commonly the biotite variety that occurs in foliated (sheared) domains of the Candelaria deposit (Fig 5 a and m). In places, sheared massive green biotite cuts displaced fragments of massive magnetite replacements. Iron metasomatism resulted in the formation of specular hematite (hm I, Fig. 9), mainly in dilatational fractures and open spaces (Fig. 5n), and coevally, of massive magnetite replacement bodies during an early mineralization stage. Subsequent pseudomorphic replacement of fracture-controlled specularite (hm I) by magnetite (mushketovite; Ramdohr, 1980) and new formation of magnetite (mt I) records a shift toward more reduced conditions and/or higher temperatures (Figs. 5o and p, and 9). The main copper mineralization occurred after the early main iron oxide mineralization. Chalcopyrite pyrite crosscutting magnetite or early specularite with interstitial chalcopyrite pyrite are typically observed in the deposits of the Punta del Cobre belt (Fig. 5f). The existence of two temporally distinct major hydrothermal stages is demonstrated by the occurrence of physically separated veins
0361-0128/98/000/000-00 $6.00

of massive magnetite hematite without sulfides and massive chalcopyrite-pyrite without iron oxides, which in places occur in contact with each other (e.g., in the Carola mine). Chalcopyrite-pyrite texturally postdate a widespread albite, a calcic-amphibole, and a quartz K feldspar veining event at Candelaria. In the northern part of the Candelaria orebody, albite veinlets, occasionally plus minor scapolite, cut the early biotite-quartz-magnetite alteration. These albite veinlets in turn are cut and used by main copper mineralization. Albite veining in the Candelaria orebody is interpreted to occur coeval with sodic scapolite alteration in the overlying Abundancia Formation. Calcic-amphibole may largely overlap with the main-stage chalcopyrite-pyrite. Calcic amphibole associated with chalcopyrite-pyrite cutting or using previously formed albite veinlets is common (Fig. 5g). Similarly, pervasive K feldspar albite alteration is cut by calcic amphibole epidote. The association of calcic amphibole with interstitial chalcopyrite-pyrite and pre- or post-amphibole K feldspar replacements or veinlets is common (Fig. 5h and i). Crosscutting relationships in a few places indicate that there are at

1811

Punta del Cobre district

1812

MARSCHIK AND FONTBOT

Iron Oxide Stage CANDELARIA


Chalcopyrite Gold Hematite Magnetite Sphalerite Pyrite Pyrrhotite Molybdenite Albite K-feldspar Ca-amphibole Biotite Quartz Grunerite-cummingtonite Cordierite Almandine-rich garnet Epidote Chlorite Tourmaline Anhydrite Calcite

Sulfide Stage

Late Stage
Loc

I
P

P+V

II II
Loc Loc V

I
TEN

II

Loc

P P Pervasive biotitization V Pervasive silicification V V+P

V V

Widespread scapolite followed by andraditic garnet formation above the Candelaria orebody

V+P V V

V V

Punta del Cobre


Chalcopyrite Gold Hematite Magnetite Sphalerite Pyrite Albite K-feldspar Ca-amphibole Biotite Quartz Chlorite Tourmaline Calcite
Loc V

I
Upper part P P

V+P

II
TEN Loc

I I

II
TEN

Lower part Lower part Upper part TEN V P P+V

Ductile/brittle at Candelaria

Uplift

Brittle

Time

FIG. 9. Paragenetic sequences of main ore, alteration, and gangue minerals in the Candelaria orebody and the Punta del Cobre district. Peak contact metamorphic calc-silicate minerals, occurring outside the orebody, are not represented in this Figure (see text). The thickness of the lines roughly represents the relative abundance of the particular mineral but has no quantitative implications regarding other mineral species. The most relevant distinguishable veining events at Candelaria are shown. Peak contact metamorphism at Candelaria, indicated by the gray shaded field, is represented by the widespread sodic scapolite-pyroxene-andraditic garnet alteration above the orebody (see also Fig. 8). Peak contact metamorphism occurred at the end of the early potassic alteration, i.e., after the biotite-quartz-magnetite alteration. It was about the time at which deformation style changed from ductile-brittle to brittle at Candelaria and the available ages and the geologic context suggest that this was broadly coeval with regional uplift. The various hematite, magnetite, and pyrite generations mentioned in the text are indicated (see also Fig. 3). Abbreviations: P = pervasive alteration, TEN = tentatively, V = veinlets, ? = uncertain.

least two generations of calcic amphibole. Chalcopyrite of one and the same event may have various types of associated alteration minerals, depending on the position within the hydrothermal system. Veinlets of quartz-mushketovite with interstitial pyrite-chalcopyrite, veinlets of quartz K feldspar chalcopyrite-pyrite locally plus trace molybdenite cutting calcic amphibole (Fig. 5i), veinlets of chalcopyrite-pyrite with quartz-chlorite plus biotite cutting biotitized rocks, veinlets of anhydrite-chalcopyrite, veinlets of calcic amphibole chalcopyrite cutting and using epidote pyrite veinlets (Fig. 5j), veinlets of chalcopyrite pyrite sphalerite cutting anhydrite (Fig. 5k; see below), among others, are observed at Candelaria.
0361-0128/98/000/000-00 $6.00

Pyrite accompanies chalcopyrite but chalcopyrite as infilling in skeletal and cataclastic pyrite and other textural observations suggest that pyrite (py I) began to form earlier than chalcopyrite. A later minor phase of pyrite (py II) veinlets cut massive chalcopyrite veins. The fact that pyrrhotite commonly but not exclusively predates chalcopyrite, crosscutting relationships and other textural evidence suggest that pyrrhotite is roughly contemporaneous with the first stage of main chalcopyrite formation (Fig. 5q; see also Hopf, 1990; Ryan et al., 1995; Ullrich and Clark, 1999). Anhydrite cuts and is cut by chalcopyrite and/or pyrite and, in some instances, occurs intergrown with chalcopyrite (Fig. 5k). Post-anhydrite chalcopyrite

1812

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE


1000

1813

Candelaria mine

100

PC 1316 PC 1317 PC 1336 PC 1345 PC 1347 PC 1421 PC 1464 PC 1479 PC 1516

10

La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

1000

Carola mine

100

PC 148 PC 165 PC 188 PC 198 PC 207 PC 208 PC 217 PC 219 PC 221

10

1 La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

1000

rosettes of prismatic needles of about 2 mm intergrown with quartz at the Carola mine (Hopf, 1990) and is also common at Candelaria (Ryan et al., 1995). According to Hopf (1990), apatite formed after tourmaline. Hypogene metallic mineralization at a district scale ended with the deposition of hematite (hm II) and the martitization of the previously formed magnetite (mt I; Fig. 5r). The late hematite (hm II) event may correlate with specularite calcite veins recognized at a regional scale, which locally contain traces of magnetite, pyrite, and rarely chalcopyrite. Late calcite veins and veinlets cut the alteration and veinlet types previously mentioned. However, there are a few examples in which chalcopyrite and pyrite postdate calcite veining (e.g., in the upper part of the Carola mine). Minor chabazite and phillipsite have been microscopically identified in late veinlets. Peak of the contact metamorphic calc-silicate alteration occurs after the iron oxide formation and before the main Cu mineralization, because contact metamorphic andraditic garnet postdates sodic scapolite, which cuts biotite hornfelses and biotitized volcanic rocks of the Upper Lavas on top of the Candelaria orebody. Andraditic garnet in turn is cut by calcic-amphibole and by pyrite chalcopyrite veinlets that are correlated with the main mineralization stage (see above in the alteration section). The formation of biotite hornfels in the upper part of the Punta del Cobre Formation and lower part of the Chaarcillo Group is correlated with the barren brown biotite-almandine cordierite alteration in the Trinidad Siltstone and brown biotite-quartz-magnetite in the volcanic and volcaniclastic rocks at Candelaria. Metal and gangue mineral zonation A marked zonation is observed in the magnetite-hematite distribution at a district scale and within the ore deposits. Deposits lying closer to the batholith contact tend to have magnetite rather than hematite, even in relatively high stratigraphic positions (e.g., El Bronce or Las Pintadas, both hosted in rocks assigned to the Abundancia Formation). Virtually all of the early specularite (hm I) at Candelaria is now transformed into mushketovite (mt I). Early specular hematite (hm I) is preserved in the upper parts of the Carola and Socavn Rampa mines, i.e., in the marginal parts of the hydrothermal system. Second-generation hematite (hm II) and the martitization of the previously formed magnetite (mt I) can be best observed in the Punta del Cobre district (Fig. 5r). At Candelaria, veinlets of late specular hematite (hm II), with associated minor pyrite-chalcopyrite, are rare and martitization of magnetite is uncommon. Pyrrhotite is restricted to the upper central part of the northern Candelaria orebody (Ryan et al., 1995) and occurs locally in the Carola mine (Hopf, 1990). A redox boundary is observed at the upper limit of the ore system in the Punta del Cobre district, represented by ore mantos hosted in the Basal Breccia. The lower and internal parts of the mantos show the same greenish color as the underlying alkali metasomatized and chloritized volcanic rocks, but toward the top and margins, i.e., approaching the oxidized continental rocks of the Basal Breccia, the color changes to red due to an increase of hypogene hematite. In general, magnetite tends to be associated with potassic assemblages, whereas hematite is common in sodium

Socavn Rampa-Trinidad mines

PC 620 PC 704 PC 834 PC 1356

100

10

La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

FIG. 10. Chondrite-normalized rare earth element pattern of rocks from the Candelaria, Carola, and Socavn Rampa-Trinidad mines, illustrating local enrichments in light rare earth elements that occur with these deposits. Reference chondrite of Nakamura (1974).

locally contains sphalerite inclusions. There is a spatial association of sphalerite with epidote-allanite-chalcopyrite-pyriteanhydrite at Candelaria. Gold is present native as inclusions mainly in chalcopyrite and pyrite (Hopf, 1990; Ryan et al., 1995). The exact paragenetic position of molybdenite is poorly constrained due to its scarcity. Ullrich and Clark (1999) place the formation of molybdenite at the beginning of the main copper mineralization. In the few molybdenite-bearing samples available for this study, molybdenite appears to postdate calcic-amphibole and possibly occurs at the end of the main chalcopyrite formation. A textural relationship between anhydrite and molybdenite has not been observed. Apatite occurs as
0361-0128/98/000/000-00 $6.00

1813

1814

MARSCHIK AND FONTBOT TABLE 3. Whole-Rock Analysis of Ore Samples from Selected Ore Deposits of the Punta del Cobre Belt

Sample Location Unit Alteration SiO2 TiO2 Al2O3 Fe2O3* MnO MgO CaO Na2O K2O P2O5 L.O.I. Total Au (ppb) ppm Ag As Ba Br Cd Co Cr Cs Cu Hf Hg Mn Mo Ni Pb Rb Sb Sc Se Sr Ta Tb Th U V Y Zn La Ce Pr Nd Sm Eu Gd Dy Ho Er Tm Yb Lu

PC 1336 Candelaria Trinidad Siltstone Ca-K 41.69 0.62 12.00 21.98 0.16 6.36 4.39 0.17 4.51 0.47 0.91 93.25 474 10.9 11 400 <0.5 0.9 55 37 5 26614 3 <1 1268 <1 63 6 136 3 13 <3 105 <0.5 <0.5 2.5 3 137 20 61 19.7 35.7 3.2 13.9 2.8 0.57 1.6 2.1 0.46 1.3 0.24 1.6 0.26

PC 1345 Candelaria Algarrobos Member K 45.02 0.52 11.62 30.10 0.07 2.14 1.01 0.09 7.64 0.10 0.25 98.56 78 <0.4 5.8 1800 <0.5 <0.5 62 143 1 3787 3 <1 606 <1 49 17 175 0.9 11.4 <3 162 0.9 <0.5 3.2 <0.5 123 8 31 78.1 160.1 13.8 46.8 5.8 1.02 3 2.8 0.61 1.4 0.25 1.6 0.24

PC 188 Carola Melndez Dacites K 54.05 0.53 12.68 15.19 0.05 2.29 0.29 0.06 8.54 0.12 2.52 96.31 199 1 18 930 4.6 <0.5 37 <5 2 13104 4 <1 406 3 14 6 163 2 8 <3 42 0.8 <0.5 4.4 4.3 78 10 36 3.9 9.1 1.2 5 1.3 0.35 1.2 1.5 0.41 1.2 0.24 1.7 0.31

PC 198 Carola Melndez Dacites K-Na 56.41 0.51 13.73 11.36 0.07 2.58 1.13 3.10 4.42 0.16 3.29 96.77 152 0.9 10 1200 3.6 <0.5 45 <5 2 5059 4 <1 539 <1 14 <5 88 1.5 8 <3 85 1.1 <0.5 4.9 3.6 51 12 28 2.1 5.3 <1 3.8 1.2 0.28 1.4 1.7 0.46 1.6 0.27 1.9 0.34

PC 219 Carola Melndez Dacites K 46.05 0.46 11.01 19.40 0.10 6.21 1.34 0.00 4.98 0.15 6.06 95.76 686 2.6 35.2 980 4 <0.5 114 29 6 16249 3 <1 763 <1 63 12 111 2.9 12.6 <3 71 <0.5 0.5 6.8 2.4 94 11 101 69.8 134 10.6 33.4 4.8 0.69 2.2 1.8 0.4 1.1 0.18 1.1 0.17

PC 704 Socavn Rampa Melndez Dacites Carbonate 51.21 0.51 8.59 21.14 0.12 4.09 2.91 0.00 0.08 0.23 7.84 96.72 480 1.9 46.1 110 2 <0.5 54 155 <1 17845 3 <1 950 <1 47 <5 <15 3.9 9.1 <3 17 <0.5 <0.5 3.8 2.7 113 12 39 6.6 12.9 1.4 6.6 1.9 0.56 1.6 2 0.36 1.1 0.17 1.1 0.17

PC 834 Socavn Rampa Melndez Dacites K 46.71 0.72 15.02 15.81 0.10 2.76 1.28 0.04 7.68 0.15 4.74 95.01 499 1.4 3105 1200 1.6 <0.5 50 149 2 17518 3 <1 802 <1 17 5 175 2.4 14.7 <3 42 <0.5 <0.5 4.2 1 105 7 34 11 25.2 2.4 10.1 2.1 0.42 1.3 1.4 0.38 1.1 0.18 1.3 0.21

PC 1356 Trinidad Melndez Dacites K 53.33 0.46 12.48 17.61 0.09 4.41 1.05 0.00 6.09 0.14 3.29 98.95 15 <0.4 12.4 1900 <0.5 0.8 70 11 2 2316 4 <1 644 <1 71 7 91 1.6 6 <3 264 <0.5 0.7 5.2 2.4 91 13 47 115.1 196.7 14.8 57.1 6.4 1.15 2.9 1.9 0.41 <0.1 0.17 1.1 0.17

*Total iron as Fe2O3 Elements were analyzed by a combination of XRF, INAA, and ICP-AES methods Low totals are due to losses of sulfur and/or CO2

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

1814

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE


chalcopyrite chalcopyrite-pyrite pyrite pyrrhotite sphalerite anhydrite

1815

Las Pintadas

El Bronce

Socavn Rampa

Santos

stopes (N. Pop, pers. commun., 1997). In the upper levels of the Carola mine, sphalerite is virtually absent and massive ore usually contains below 100 ppm Zn (Hopf 1987). At Candelaria, two horizons have been identified in which Zn contents locally are about 1 wt percent (W. Martin, pers. commun., 2000). These horizons occur roughly at the upper contact of the Lower Andesites and at the lower contact of the Trinidad Siltstone. Highest concentrations of light rare earth elements occur locally near these same contacts (Marschik et al., 2000). Sulfur Isotope Geochemistry Sulfur isotope analyses on 47 samples from the Bronce (n = 2), Candelaria (n = 34), Las Pintadas (n = 1), Santos (n = 4), and Socavn Rampa mines (n = 6) were carried out at the Stable Isotopes Laboratory of the University of Lausanne, using an online elemental analyzer-continuous flow-isotope ratio mass spectrometer and in the G.G. Hatch Isotope Laboratories, University of Ottawa. Pyrrhotite, sphalerite, pyrite, anhydrite, and chalcopyrite that predates, is intergrown with, and postdates anhydrite, were analyzed. These samples represent the most relevant paragenetic positions (see Fig. 9). The 34S values of chalcopyrite, pyrite, pyrrhotite, and sphalerite from the deposits lie between 0.7 and +3.1 per mil relative to Caon Diablo Troilite (CDT; Table 4, Fig. 11). Sulfides from the Socavn Rampa mine have 34SCDT values between 0.7 and +0.5 per mil, and those from the Santos mine lie between 0.3 and +1.1 per mil. Sulfides from Candelaria have 34SCDT values of 0.3 and 3.1 per mil, whereas anhydrite yielded values between 14.5 and 17.5 per mil. We could not confirm relatively heavy values for sulfides reported by Ullrich and Clark (1999; up to 5.7 for the main-stage and 7.2 for the late-stage mineralization). A general shift toward lower 34SCDT values upsection in the Santos and Socavn Rampa mines is interpreted to reflect oxidation of the ore fluid approaching the Basal Breccia (Marschik et al., 1997a). This is in agreement with the abundance of magnetite in the lower parts and the predominance of hematite in the upper parts of the deposits of the Punta del Cobre district. A similar trend is also observed on a district scale where there is a general decrease in 34SCDT values from Candelaria, via Santos toward the Socavn Rampa mine (Fig. 11). Geochronology The Ar/ Ar analyses on biotite and amphibole separates from the Candelaria deposit were carried out by the Geochronological Research Laboratory of the New Mexico Bureau of Mines and Mineral Resources, Socorro, New Mexico, United States, using the furnace incremental-heating technique. These analyses were carried out to complement previously reported data, which are incorporated in the discussion below. The analytical results are given in Table 5. Analytical methods and parameters are detailed in the Appendix. Brown biotite (sample PC 99139) from the barren pervasive biotite-almandine-rich garnet cordierite alteration in the tuffaceous rocks that correlate with the Trinidad Siltstone has been analyzed. The age spectrum has two plateaus, one formed by heating steps D to F with 74.7 percent of the gas released and another plateau from steps F to H with 61 percent of the gas released (Fig. 12; Table 5). Each of the plateaus would conform the definition of a plateau age of at least three
40 39

Candelaria -2 0 2 4 6 34 S 8 10 12 14 16 18

(permil)

FIG. 11. Sulfur isotope variations within selected deposits of the Punta del Cobre belt. Sulfides from Candelaria show a 34S range from 0.3 to 3.1 per mil. Anhydrites have values between 14.5 and 17.5 per mil. Sulfides from all other mines studied are similar to those from Candelaria. However, a tendency toward lower 34S values from the internal (Candelaria mine) via intermediate (Santos mine) to the external portions of the system (Socavn Rampa mine) can be observed.

metasomatized rocks. These relationsips can be interpreted in terms of more internal and high-temperature vs. external and low-temperature portions of the ore-forming hydrothermal system. For example, the Socavn Rampa mine, which is characterized by sodic rather than potassic assemblages, is essentially devoid of magnetite, whereas hematite is widespread. In contrast, the Trinidad mine, which constitutes the northern continuation of the Socavn Rampa mine, has magnetite-rich ore, including massive magnetite bodies (J. Ponce, pers. commun., 1996), associated with potassic assemblages suggesting that the deposit represents another mineralization center between the Santos mine to the north and the Carola mine to the south. Within the Carola mine, massive magnetite bodies occur only below mining level 10 (362 m a.s.l) in the western part of the mine. In its eastern part, magnetite bodies are absent and the Cu/Fe is higher (Lino, 1984; Hopf, 1987). The distributions of quartz and calcite are similar to those of magnetite and hematite. In Candelaria and in deep parts of the Punta del Cobre district, quartz veinlets are widespread and silicification is common, whereas calcite occurs essentially in thin veinlets. In contrast, calcite veins and veinlets are abundant in the upper portions of the Punta del Cobre district, where pervasive carbonate alteration is common. Anhydrite is confined to the lower part of Candelaria orebody and locally occurs in the Trinidad and Carola mines. At Candelaria, highest Cu-Au grades occur in the Trinidad Siltstone in the uppermost part of the deposit, whereas most of the ore is hosted in the volcaniclastic and volcanic rocks below this horizon (Ryan et al., 1995). Gold concentrations show a good positive correlation with copper contents (Marschik et al., 2000). Gold grades in the deposits of the Punta del Cobre district are variable and appear to be slightly higher than at Candelaria (Table 3) and there may be a Cu/Au fractionation at a district scale. Local enrichments of Zn are recognized at the Carola and Candelaria mines. At deep levels of the Carola mine (below 335 m a.s.l.), Zn grades may exceed 2 wt percent in some
0361-0128/98/000/000-00 $6.00

1815

TABLE 4. Results of Sulfur Isotope Analysis

1816

Sample Abundancia Formation Abundancia Formation 17.5 14.6 15.2 15.3 14.5 PC 98193 Candelaria Chalcopyrite PC 98166 Candelaria Chalcopyrite PC 98148 Candelaria Chalcopyrite PC 98106 Candelaria Chalcopyrite PC 98014 Candelaria Chalcopyrite 0.0 PC 98009 Candelaria Chalcopyrite Massive chalcopyrite vein cut by pyrite veinlets Lower Andesites 0.2 PC 1513 Candelaria Chalcopyrite Chalcopyrite-quartzamphibole veinlet Lower Andesites 1.9

Mine

Mineral

Description

Unit

34S (CDT, ) Sample Mine Mineral Description Unit

34S (CDT, )

PC 98038

Bronce

Pyrite

0361-0128/98/000/000-00 $6.00
1.3 1.7 2.0 1.6 1.1 Algarrobos Member 2.9 Chalcopyrite-pyrite interstitial to quartz Massive chalcopyrite vein Algarrobos Member? Algarrobos Member Algarrobos Member Lower Andesites Chalcopyrite-pyrite interstitial to quartz Massive chalcopyrite vein cut by pyrite veinlets Algarrobos Member Lower Andesites Algarrobos Member Algarrobos Member Algarrobos Member Pyrite-rich chalcopyritemagnetite ore with associated with epidote and scapolite Chalcopyrite-rich magnetite-pyrite ore associated with epidote and scapolite Anhydrite cut by chalcopyrite Barren anhydrite veinlet cutting mineralized rock Anhydrite veinlet with clasts of chalcopyrite Anhydrite intergrown (cogenetic) with chalcopyrite Anhydrite cut by chalcopyrite Lower Andesites Trinidad Silstone Trinidad Silstone Algarrobos Member Lower Andesites 1.5 1.5 PC 99152 PC 98188 1.5 PC 99044 2.9 PC 98202 Candelaria 16.5 PC 98195 Candelaria Chalcopyrite Chalcopyrite Volcaniclastic microconglomerate, patchy epidote, amphibole veinlets, K-feldspar envelops and replacements, minor chalcopyrite-pyrite Chalcopyrite from andesitic wallrock Algarrobos Member Algarrobos Member Chalcopyrite Candelaria Candelaria Chalcopyrite Sphalerite Lower Andesites? Algarrobos Member Vein of pyrite-sphaleriteanhydrite Algarrobos Member? 0.3 1.0

PC 98039

Bronce

Chalcopyrite

PC 98150

Candelaria

Anhydrite

PC 98195

Candelaria

Anhydrite

PC 98196

Candelaria

Anhydrite

PC 99044

Candelaria

Anhydrite

PC 99152

Candelaria

Anhydrite

MARSCHIK AND FONTBOT

1816
Anhydrite as breccia infill with minor pyrite inclusions Chalcopyrite associated amphibole-quartz plus minor brown-green biotite Candelaria Chalcopyrite associated amphibole-epidote-Kfeldspar-quartz Chalcopyrite-pyrite-quartz vein cutting foliated biotitized rock Magnetite intergrown with and cut by finegrained biotite(-quartz) alteration and cut by chalcopyrite, minor pyrite Chalcopyrite-quartzamphibole veinlet Lower Andesites Lower Andesites 1.9 1.3 PC 1331 PC 635 Massive chalcopyrite vein cut by pyrite veinlets Las Pintadas Santos 1.7 2.0 1.6 PC 641 PC 939 PC 942 Santos Santos Santos Chalcopyrite-pyrite interstitial to quartz Massive chalcopyrite vein Chalcopyrite-pyrite interstitial to quartz Lower Andesites Algarrobos Member Algarrobos Member

PC 99158

Candelaria

Anhydrite

PC 1318

Candelaria

Chalcopyrite

PC 1336

Candelaria

Chalcopyrite

1.9 0.3 3.1

PC 1466

Candelaria

Chalcopyrite

Massive pyritechalcopyrite in the lowest part of the Algarrobos Member Chalcopyrite intergrown (cogenetic) with anhydrite Chalcopyrite-sphalerite veinlet cutting anhydrite

PC 1481

Candelaria

Chalcopyrite

PC 1513

Candelaria

Chalcopyrite

Chalcopyrite Chalcopyritepyrite Chalcopyrite Pyrite Pyrite

Nantoco Formation Lower Andesites

-0.3 0.6

PC 98009

Candelaria

Chalcopyrite

PC 98014

Candelaria

Chalcopyrite

0.3 1.1 -0.3

PC 98106

Candelaria

Chalcopyrite

PC 98148

Candelaria

Chalcopyrite

Magnetite-chalcopyrite associated with amphibole-epidote Massive magnetite replacements with disseminated chalcopyrite-pyrite Stockwork of chalcopyrite-pyrite Pyrite(-chalcopyrite) vein with secondary Cu oxides Layered chalcopyritepyrite in sedimentary

Lower Andesites Lower Andesites Lower Andesites

TABLE 4. (Cont.)
Description Unit 34S (CDT, ) Sample Mine Mineral Description Unit 34S (CDT, )

Sample

Mine

Mineral

PC 98166 Algarrobos Member 2.9 PC 790 Socavn Rampa Chalcopyrite Chalcopyrite-pyrite vein

Candelaria

Chalcopyrite

Massive chalcopyrite vein cut by pyrite veinlets

Algarrobos Member

1.1

PC 677

Socavn Rampa

Chalcopyrite

lense within andesitic rock Patchy replacements of chalcopyrite and pyrite

1000

FIG. 12. The 40Ar/39Ar age spectra of alteration minerals from the Candelaria deposit. (a). Age spectrum of brown biotite from the barren biotite-almandine-rich garnet cordierite alteration. (b). Age spectrum of green hydrothermal biotite from the biotite-quartz-magnetite alteration that hosts chalcopyrite pyrite veinlets and disseminations (see also Fig. 5m). (c). Age spectrum of amphibole intergrown with chalcopyrite-pyrrhotite.
% Radiogenic
100 120 130 140 90 40 80 130 140 0

1817

K/Ca

K/Ca

K/Ca

0361-0128/98/000/000-00 $6.00
Basal Breccia/ -0.7 Melndez Dacites contact Lower 0.5 Andesites 0.3 1.0 PC 834 Chalcopyrite PC 795 Chalcopyrite 0.4 0.0 Algarrobos Member Algarrobos Member Lower Andesites? 0.3 2.0 PC 795 Socavn Rampa Pyrite 1.9 PC 678 Socavn Rampa Chalcopyrite Socavn Rampa Socavn Rampa Chalcopyrite-pyrite as breccia infill Pyrite-chalcopyritespecularite (hm I) Calcite centered pyritechalcopyrite veinlet Chalcopyrite-pyrite as breccia infill Melndez Dacites Basal Breccia Volcaniclastic microconglomerate, patchy epidote, amphibole veinlets, K-feldspar envelops and replacements, minor chalcopyrite-pyrite Chalcopyrite from andesitic wallrock Massive pyritechalcopyrite in the lowest part of the Algarrobos Member Chalcopyrite intergrown (cogenetic) with anhydrite Basal Breccia/ -0.5 Melndez Dacites contact Melndez 0.4 Dacites Chalcopyrite-sphalerite veinlet cutting anhydrite Pyrrhotite-chalcopyrite veinlet Algarrobos Member Algarrobos Member
Apparent age (Ma) % Radiogenic
100 40 60 80 100

PC 98193

Candelaria

Chalcopyrite

PC 98195

Candelaria

Chalcopyrite

PC 98202

Candelaria

Chalcopyrite

PC 99044

Candelaria

Chalcopyrite

PC 99152

Candelaria

Chalcopyrite

PC 99125

Candelaria

Pyrrhotite

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1817
120 0 B C 750 110 850 D 1020 10 D 920 20 E 1000 E 1080 30 40 F 1075 50

Apparent age (Ma)

Apparent age (Ma)


110 120 130 140 90

% Radiogenic
100 80 60 40 20

100

110

90

B 750

C 850

10

10 D 920

20 60 F 1120 G 1110

20

E 1000

Cumulative percentage 39Ar released

Cumulative percentage 39Ar released

30

30

116.51 0.26 MSWD = 0.9

Total fusion age = 114.90 0.54

Total fusion age = 118.40 1.5


70

Total fusion age = 115.85 0.54

40 80

40

115.14 0.18 MSDW = 1.1

50

50

F 1075

60

60

70

70

(b) Age spectrum of sample PC 99131 biotite

(a) Age spectrum of sample PC 99139 biotite

Cumulative percentage 39Ar released


116.6 1.2 MSDW = 4.9**
90 J 1400 100
0.1 0.01

I G 1160 H 1300 1200

80

80

H 1180

(c) Age spectrum of sample PC 98137 amphibole


I 1210J 1250

H G 1110 1180

90

90

100

100

10

10

100

0.1

1000

1818

MARSCHIK AND FONTBOT TABLE 5. Results of 40Ar/39Ar Analysis

ID

Temp (C)

40Ar/39Ar

37Ar/39Ar

36Ar/39Ar

(103)

39Ar K (1018 mol)

K/Ca

40Ar*

39Ar

(%) 14.9 57.2 110.5 217.0 307.8 216.8 28.1 3.7 2.5 1.2 1.8 189.4 239.0 7.9 74.2 113.6 137.7 123.3 272.2 311.5 393.6 368.6 56.2 72.0 3.6 256.0 264.2 0.053 0.041 0.025 0.031 0.027 0.037 0.037 0.035 0.036 0.037 0.038 0.033 0.037 52.1 87.9 97.4 98.7 98.9 98.9 98.8 97.1 94.5 96.9 95.4

(%) 0.6 2.1 7.5 20.2 38.4 82.2 95.5 99.4 99.6 99.8 100.0 74.7

Age (Ma) 45.70 109.48 117.95 116.74 116.51 116.31 115.55 115.83 115.40 120.10 120.90 115.85 116.51 56.50 116.45 115.07 115.50 115.41 115.30 115.12 114.82 114.90 112.31 110.62 130.90 114.90 115.14 92 69.70 106.40 119.20 120.30 115.70 120.50 116.70 117.61 118.39 145.60 118.4 116.6 114.4

1 (Ma) 1.10 0.55 0.30 0.23 0.23 0.22 0.22 0.31 3.00 3.40 2.90 0.541 0.261 1.30 0.39 0.25 0.26 0.25 0.25 0.23 0.22 0.23 0.53 0.87 2.10 0.541 0.181 21.2 30.50 14.90 1.60 0.35 0.35 2.30 2.00 0.50 0.73 4.70 1.51 1.21 2.31

PC 99139, 3.69 mg biotite, J = 0.0039298, NM-131, Lab no. = 51807-01 A 650 12.53 0.0343 20.230 B 750 18.12 0.0089 7.357 C 850 17.66 0.0046 1.501 D 920 17.23 0.0024 0.666 E 1000 17.16 0.0017 0.535 F 1075 17.13 0.0024 0.542 G 1110 17.04 0.0181 0.639 H 1180 17.38 0.1392 1.669 I 1210 17.78 0.2021 3.263 J 1250 18.08 0.4306 1.962 K 1300 18.48 0.2846 2.852 Total gas age n = 11 Plateau MSWD = 0.9 n=3 steps D-F PC 99131, 4.60 mg biotite, J = 0.0039313, NM-131, Lab no. = 51809-01 A 650 16.63 0.0643 28.850 B 750 19.01 0.0069 6.840 C 850 17.69 0.0045 3.081 D 920 17.04 0.0037 0.661 E 1000 16.94 0.0041 0.393 F 1075 16.89 0.0019 0.282 G 1110 16.87 0.0016 0.278 H 1180 16.85 0.0013 0.378 I 1210 16.86 0.0014 0.377 J 1250 16.79 0.0091 1.447 K 1300 16.98 0.0071 2.943 L 1650 27.63 0.14 28.690 Total gas age n = 12 Plateau MSWD = 1.1 n=7 steps C-I

29.2 71.1 258.9 610.6 876.8 2105.9 642.1 185.2 10.0 9.3 10.5 4809.5 1746.3 39.9 165.2 536.4 837.2 1055.5 1924.0 1555.7 1550.2 442.8 66.8 36.1 19.2 8229.1 7901.9

48.6 89.2 94.7 98.7 99.2 99.4 99.4 99.2 99.2 97.3 94.7 69.3

0.5 2.5 9.0 19.2 32.0 55.4 74.3 93.1 98.5 99.3 99.8 100.0 96.0

PC 98137, 14.43 mg hornblende, J = 0.0039261, NM-131, Lab no. = 51805-01 A 800 209 9.573 665.3 6.28 B 850 61.71 12.45 178.700 1.1 C 950 49.73 20.07 122.400 2.2 D 1020 27.79 16.43 40.530 36.3 E 1080 19.31 19.01 11.880 376.2 F 1120 17.40 13.68 6.244 265.6 G 1160 19.95 13.96 12.550 15.9 H 1200 18.56 14.59 9.974 14.3 I 1300 17.84 14.15 6.943 98.0 J 1400 18.77 13.69 9.532 73.5 K 1650 39.74 13.47 66.540 9.1 Total gas age n = 11 898.3 Plateau MSWD = 4.92 n=5 steps F-J 467.3 Isochron MSWD = 6.52 n=5 steps F-J

6.3 16.1 30.7 61.9 89.8 96.0 87.3 90.7 95.1 91.1 53.4

0.7 0.8 1.1 5.1 47.0 76.5 78.3 79.9 90.8 99.0 100.0 32.9

Notes: Isotopic ratios corrected for blank, radioactive decay, and mass discrimination, not corrected for interferring reactions Individual analyses show analytical error only; plateau and total gas age errors include error in J and irradiation parameters n = number of heating steps K/Ca = molar ratio calculated from reactor produced 39ArK and 37ArCa 1 2 error 2 MSWD outside 95% confidence interval

consecutive steps that overlap within their errors and which contain 50 percent of the gas released (Fleck et al., 1977). However, there are marked decreases in the K/Ca from F to G, and from step G to H, which indicate mineral inclusions. We consider a 40Ar/39Ar weighted mean plateau age of 116.51 0.26 Ma (all errors at 2) calculated from heating steps D to F as representative. The total fusion age of this biotite sample is 115.85 0.54 Ma.
0361-0128/98/000/000-00 $6.00

Green hydrothermal biotite (sample PC 99131) from the foliated pervasive biotite-quartz-magnetite alteration with veinlets of quartz-chlorite green biotite, chalcopyrite, and minor pyrite, and disseminated chalcopyrite intergrown with the biotite gave a 40Ar/39Ar weighted mean plateau age of 115.14 0.18 Ma, based on heating steps C to I, which contain 96 percent of the gas released (Fig. 12; Table 5). The total fusion age of this biotite sample is 114.9 0.54 Ma.

1818

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1819

Euhedral amphibole (sample PC 98137) in a matrix of massive chalcopyrite-pyrrhotite has a somewhat disturbed age spectrum (Fig. 12; Table 5). Steps A to C are heterogeneous, giving ages younger than 110 Ma with large errors. Steps D to E yield an age of about 120 Ma, whereas steps F to J yield ages that increase from 115.70 0.70 to 118.39 1.4 Ma. The abrupt change in ages between steps E and F correlates with an increase in both K/Ca and radiogenic yield, which indicates the presence of inclusions within the amphibole grains. Steps F to J do not conform the definition of a plateau age. However, steps F to H contain 32.9 percent of the gas released and overlap within their errors. Steps I and J have a similar K/Ca ratio and comparable radiogenic yield to steps F to I, which leads us to believe that a pseudoplateau age of 116.6 1.2 Ma calculated from steps F to J is representative (Fig. 12; Table 5) and relevant. An isochron age of 114.4 2.3 Ma was calculated from these same steps. The total fusion age of this amphibole sample is 118.4 1.5 Ma. Discussion Alteration and mineralization patterns allow the establishment of a descriptive model for an idealized CandelariaPunta del Cobre-type iron oxide Cu-Au(-Zn-Ag) system. Most of the larger orebodies in the belt are located at the intersection of northwest- to north-northwesttrending brittle faults with the contact of massive volcanic and volcaniclastic rocks. These northwest- to north-northwesttrending brittle faults and a major north-northeasttrending ductile structure appear to be the main fluid conduits. In the internal parts of this system, represented by Candelaria and the deeper parts of the mines in the Punta del Cobre district, ore is associated with mainly calcic-potassic (calcic amphibole epidote-biotite, K feldspar) alteration (see also Table 2). Magnetite is abundant and occurs predominantly as massive replacement bodies that grade laterally into intense pervasive iron metasomatized (magnetite) volcanic and volcaniclastic rocks. Both early and late hematite are locally present in minor quantities. Quartz veins and veinlets are common and silicification is widespread. Calcite veinlets and carbonatization are minor or absent. Anhydrite locally occurs in veins and veinlets. The intermediate parts of the system are represented by the central upper portion of the Santos and Carola mines and are characterized by intense potassic (biotite and/or K feldspar) alteration, in places with subordinate calcic amphibole alteration. Pervasive biotitization of the host rock grades into intense chloritization toward more shallower levels, i.e., external parts. Magnetite occurs in veins and disseminated in the host rocks. Hematite is common and more abundant than in the internal parts of the system. In the external parts of the system, represented by the Socavn Rampa mine, mineralization is associated with intense sodic (albite) alteration and chloritization. Pervasive carbonatization and calcite veins are common and may be intense, whereas silicification and quartz veins are essentially absent. Hematite is the predominant iron oxide mineral, whereas magnetite is uncommon. However, iron oxides may be minor or virtually absent in some places with significant amounts of sulfides. The structural control and geometry of the ore, particularly that of the Punta del Cobre district, suggest that ascending
0361-0128/98/000/000-00 $6.00

fluids were focused by north-northwest to northwest-trending faults to form subvertical bodies in volcanic rocks. Where these fluids reached permeable horizons they spread out laterally to form concordant lens-shaped orebodies, e.g., in brecciated flow tops and intercalated volcaniclastic horizons, but mainly in the Basal Breccia, the lowermost part of the sediments of the Algarrobos Member that overlies the volcanic host rocks (Marschik and Fontbot, 1996). Fluid inclusions in postmagnetite quartz, which contains interstitial chalcopyrite from the Candelaria deposit homogenize at 370 to >440C (Marschik et al., 2000). Similar homogenization temperatures of 328C for hypersaline CO2rich fluid inclusions in quartz from Candelaria are reported by Ullrich and Clark (1999). Liquid-vapor inclusions in anhydrite that formed coevally with the main copper mineralization at Candelaria homogenize between 340 and 470C and liquid-vapor inclusions in calcite at temperatures of 180C (Marschik et al., 2000). Microthermometric measurements on fluid inclusions in postore calcite from the Punta del Cobre deposits indicate that this late-stage fluid was also saline, containing 12 to 24 wt percent NaCl equiv and 13 to 23 wt percent CaCl2 (Marschik et al., 1997b). Homogenization temperatures obtained are between 125 and 175C. Rabbia et al. (1996) reported homogenization temperatures of fluid inclusions in calcite from the Carola mine between 175 and 236C. Quartz associated with chalcopyrite from the Candelaria deposit has 18OSMOW values between 11.2 to 12.6 per mil (Marschik et al., 2000). A fluid in isotopic equilibrium with this quartz at temperatures between 370 and 440C has 18OSMOW values between 5.9 and 8.9 per mil (using isotopic fractionation factors of Friedman and ONeil, 1977). This isotopic composition is compatible with a fluid of magmatic origin or a nonmagmatic fluid equilibrated with silicates at high temperature. Oxygen isotope ratios of calcite from the Carola mine are between 15.4 and 15.9 per mil (Rabbia et al., 1996). The 18OSMOW values of calcite from the Santos and Socavn Rampa mines range from 14.3 to 15.3 per mil, and those of calcite from the Candelaria deposit are between 11.7 and 11.9 per mil (Marschik et al., 2000). A fluid in equilibrium with the calcite from the Carola mine would have 18OSMOW values approximately between 4.6 and 7.7 per mil for a temperature range of 175 to 235C. Accordingly, a fluid in equilibrium with the calcite from the Santos and Socavn Rampa mines would have an oxygen isotope composition between 2.8 and +4.7 per mil and with the calcite from the Candelaria deposit between 5.4 and +1.3 per mil at temperatures of 100 to 180C. The relatively low oxygen isotope ratios of a fluid in equilibrium with the late calcite at Santos and Socavn Rampa suggest that mixing between a magmatic fluid (or equilibrated with magmatic rocks) with a nonmagmatic fluid (e.g., basinal brine or meteoric waters) took place during the late stages of hydrothermal activity. Sulfur isotope ratios of sulfides from several mines of the Punta del Cobre belt determined for this study are fairly uniform. The narrow range of 34SCDT (0.7 and +3.1) near 0 per mil is consistent with a magmatic sulfur source and incompatible with significant input of sulfur derived from evaporites. The latter would imply more oxidized conditions than indicated by the paragenetic study and a larger spread in the

1819

1820

MARSCHIK AND FONTBOT

34SCDT values. The 34SCDT values of anhydrite from Candelaria between 14.5 and 17.5 per mil overlap with those suggested for Cretaceous seawater (Claypool et al., 1980). However, similar 34S values for sulfates are reported also from porphyry copper and high-sulfidation gold deposits (e.g., Ohmoto and Rye, 1979; Arribas, 1995). The sulfate-sulfide for a sample with coexisting anhydrite and chalcopyrite is 13.4 per mil, which is consistent sulfide-sulfate fractionation at temperatures on the order of 400C (Ohmoto and Rye, 1979; Ohmoto and Lasaga, 1982). Taking the paragenetic position of anhydrite within the main sulfide stage and the homogenization temperatures of fluid inclusions in anhydrite (340470C) into account (Fig. 9), the data are consistent with a predominantly cooling magmatic sulfur-bearing fluid. As discussed above, mineralization occurred under fairly oxidized conditions manifested in the formation of early specular hematite (hm I; Figs. 5n and 9). Pseudomorphic replacement of this early specular hematite by magnetite (mushketovite; Fig. 5o) and deposition of magnetite (mt I; Fig. 5p) could be explained by an increase in temperature as the hydrothermal system reached its climax at the site of ore formation (waxing stage) and/or by interaction of the oxidized fluid with Fe2+ abundantly contained in the andesitic host rock. Alteration mineral associations indicate that the deposits formed by near-pH neutral solutions. The oxidizing character of the fluid explains its metal transport capacity. The bulk of the iron oxide mineralization probably occurred at temperatures of about 500 to 600C. These temperature estimates are based on biotite-garnet Fe-Mg exchange geothermometry on early (pre-chalcopyrite) alteration in the tuffaceous rocks of the Algarrobos Member at Candelaria, determined by Ullrich and Clark (1997). The main sulfide stage follows with the formation of pyrite and chalcopyrite. Ore formation temperatures of about 400 to 500C at the Carola mine were estimated based on mineralogic evidence (Hopf, 1990). Homogenization temperatures of fluid inclusions in quartz and anhydrite from the Candelaria deposit suggest temperatures in the range of 328 to >470C for the main copper mineralization at Candelaria (Ullrich and Clark, 1999; Marschik et al., 2000; see above). Further cooling of the hydrothermal system is indicated by the homogenization temperatures of 236C of fluid inclusions in late-stage calcite (Rabbia et al., 1996; Marschik et al., 1997b, 2000). Martitization of magnetite and new formation of hematite (hm II) during the latest stage of mineralization are consistent with progressive cooling of the system. An increase in oxygen fugacity may also contribute to cause this mineralogical change. That the ore fluid had higher oxygen fugacities at the upper limit and distal portions of the ore system is also suggested by the above-mentioned variations toward lighter sulfur isotope composition of chalcopyrite and pyrite. Preliminary Pb isotope studies (Marschik et al., 1997a; Marschik and Chiaradia, 2000) show that the sulfide minerals and the least altered volcanic and intrusive rocks in the Candelaria-Punta del Cobre area have similar isotopic compositions. The relatively narrow isotopic compositional range of the ore minerals, with 208Pb/204Pb values between 38.2 and 38.4, 207Pb/204Pb values between 15.57 and 15.59, and 206Pb/ 204Pb values between 18.45 and 18.62, indicates that the hydrothermal fluid had a homogeneous Pb isotope composition.
0361-0128/98/000/000-00 $6.00

Initial 187Os/188Os values of 0.36 0.1 for an isochron calculated from data of hydrothermal magnetite and sulfide from Candelaria (Mathur, 2000; Mathur et al. (2002), and an initial 187 Os/188Os value of 0.33 for sulfide from the Bronce mine are similar to calculated initial 187Os/188Os values that range from 0.20 that 0.41 for magmatic magnetite in nearby batholithic rocks. The similarities in initial 187Os/188Os values of the ore and magmatic oxides and in the Pb isotope compositions of sulfides and granitoid plutons is consistent with the hypothesis that the batholithic magmas could be the metal source for the Candelaria-Punta del Cobre iron oxide Cu-Au deposits. The hydrothermally altered dacite porphyry dikes, although barren or only weakly mineralized, could represent an expression of an underlying igneous body that was responsible for mineralization. We underline the significance of the presence of specular hematite (later pseudomorphously replaced by magnetite, mushketovite) as an indicator for the oxidized character of the first mineralization stage. The presence of early hematite or of mushketovite is a typical feature of the deposits of the Candelaria-Punta del Cobre Cu-Au(-Zn-Ag) system and is probably a widespread phenomenon in iron oxide Cu-Au systems (e.g., Sossego, Brazil; (e.g., Sossego, Brazil, R. Marschik, unpub. data; Salobo, Brazil, Requia, 2002; Starra, Cloncurry district, Australia; Rotherham, 1997; and Ral-Condestable, Per, Haller, 2000). In other deposits it may have been overlooked. For instance, bladed magnetite from the Emmie Bluff deposit, Stuart Shelf, Australia (photo plate of fig. 2b, Gow et al., 1994), is likely to represent mushketovite. A variety of data are consistent with, but not unequivocally indicative of, magmatic fluid contribution to this hydrothermal system. This evidence includes: (1) oxygen isotope composition of quartz associated with chalcopyrite representing the main mineralization at Candelaria; (2) presence of hypersaline CO2-rich fluid inclusions (Ullrich and Clark, 1999) in the main ore stage, as well as of saline fluid inclusions in latestage calcite (Marschik et al., 1997b); (3) the 18OSMOW composition of 4.4 to 7.7 per mil of a fluid in equilibrium with calcite from the Carola mine (Rabbia et al., 1996); (4) the oxidized character of the ore fluid; (5) the narrow range of 34SCDT values of sulfides near 0 per mil; and (6) the age of the mineralization coeval with nearby intrusive activity. The similarities and transitions in terms of fluid composition and host-rock alteration of the Candelaria-Punta del Cobre deposits with porphyry skarn systems (e.g., Einaudi et al., 1981, Einaudi, 1982; Lang et al., 1995) provide additional empirical arguments that magmatic-sourced fluid is an essential part of the ore system. Cooling and fluid mixing are probably the main precipitation mechanisms at Candelaria-Punta del Cobre. Mineralization typically, but not exclusively, occurred near the contact of massive, relatively impermeable volcanic rocks and overlying porous and permeable volcaniclastic sediments. These sediments and the upper part of the faulted volcanic rocks were most likely saturated with either basinal brines, formation, or meteoric waters, which could contribute to the light oxygen isotope signatures mentioned above. An ascending saline magmatic(?) metal and sulfur-bearing fluid would have been diluted and cooled, causing saturation with respect to sulfides as it interacted with this external fluid, thereby precipitating

1820

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1821

the ore minerals. During the early stages of mineralization, reduction of the oxidized fluid through interaction with the volcanic host rocks could have been an additional precipitation mechanism. On the basis of the present data, however, it is not possible to dismiss a participation of evaporite-sourced brines as proposed by Barton and Johnson (2000) and Ullrich and Clark (1999). Alteration and mineralization ages Determination of alteration and mineralization ages at Candelaria-Punta del Cobre has been the objective of several
Age (Ma) 80
Regional cooling to 200-150C (closure temperature of K-feldspar)

contributions. Figure 13 is a compilation of the relevant age data, relating the absolute ages to the stratigraphic record. The mineralization age at Candelaria is best represented by two Re-Os molybdenite ages of 114.2 0.6 and 115.2 0.6 Ma (Mathur, 2000; Mathur et al., 2002). Previously published alteration ages cluster around 116 to 114 and 112 to 110 Ma (e.g., Marschik et al., 1997b; Arvalo, 1999; Ullrich and Clark, 1999). The new alteration ages presented in Table 5 and Figure 12 fall into the 116 to 114 Ma age range. The 40Ar/39Ar weighted mean plateau age of 115.14 0.18 Ma for green hydrothermal biotite associated with chalcopyrite-pyrite from
Stratigraphy Stage
Campanian Santonian Coniacian Cerrillos Turonian Cenomanian

90

100
Minimum age of mineralization (post-ore thermal events?) Hiatus (basin inversion, transpression) Unconformity Batholith emplacement

Albian

110

1700-2000 m

120

Mineralization age is 114.7 Ma (main mineralization) Overburden at the time of mineralization is less than 3 km

Pabelln

Aptian

Barremian Totoralillo NantocoAbundancia P. del Cobre Hauterivian Valanginian Berriasian

130
Stratigraphic position of the main orebodies

140
40Ar/39Ar 40Ar/39Ar 40Ar/39Ar

biotite plateau age (sample PC 99131, this study) biotite plateau age (sample PC 99139, this study) amphibole pseudo plateau age (sample PC 98137, this study)

Re-Os molybdenite ages (Mathur, 2000) Rb-Sr isochron (Marschik et al., 1997)
40Ar/39Ar 40Ar/39Ar 40Ar/39Ar 40Ar/39Ar 40Ar/39Ar 40Ar/39Ar 40Ar/39Ar 40Ar/39Ar

biotite inverse isochron age (Marschik et al., 1997) biotite plateau age (Ullrich and Clark, 1999) biotite total fusion age (Marschik et al., 1997) biotite isochron age (Arvalo, 1999) amphibole plateau age (Ullrich and Clark, 1999) whole rock plateau age (Marschik et al., 1997) whole rock inverse isochron age (Marschik et al., 1997) whole rock isochron age (Arvalo, 1999)

K-Ar whole rock age (Marschik et al., 1997)


FIG. 13. Summary of isotopic ages of rocks and minerals of the Candelaria-Punta del Cobre iron oxide Cu-Au deposits. Correlation of absolute and relative stratigraphic ages, according to Gradstein et al. (1995). Metallic mineralization at Candelaria-Punta del Cobre commenced as marine conditions ceased in the Atacama back-arc basin. 0361-0128/98/000/000-00 $6.00

1821

1822

MARSCHIK AND FONTBOT

Candelaria (sample PC 99131, Fig. 12, Table 5) is identical within errors with a 40Ar/39Ar inverse isochron age of 114.9 1.0 Ma for ore-related green hydrothermal biotite at the Santos mine and a less precise total fusion age of 114.6 1.4 Ma, similar to biotite from the Resguardo mine in the Punta del Cobre district (Marschik et al., 1997b). These biotite ages are similar to the Re-Os molybdenite ages and are interpreted to have recorded the same alteration event that predated and accompanied parts of the Cu-Au mineralization. The pseudoplateau age of 116.6 1.2 Ma of amphibole intergrown with chalcopyrite and pyrrhotite (sample PC 98137; Fig. 12, Table 5), although imprecise and only taken as an approximation, is compatible with the Re-Os molybdenite and the 40Ar/39Ar biotite ages mentioned above. Brown biotite from the barren biotite-almandine cordierite alteration in the Tuff unit (Trinidad Siltstone) at Candelaria with a weighted mean plateau age of 116.51 0.26 Ma (sample PC 99139, Fig. 12, Table 5) is somewhat older than the green biotite. Ullrich and Clark (1998, 1999) report 40Ar/39Ar plateau ages averaging 114.2 0.8 Ma for similar biotite from the barren biotitequartz-magnetite alteration and a 40Ar/39Ar plateau age of 111.7 0.8 Ma for amphibole associated with chalcopyrite from Candelaria. According to these authors, this later amphibole age represents the age of copper mineralization. The 40Ar/39Ar isochron ages of 110.7 1.6 Ma for biotite from deformed mineralized biotite-rich rocks and 111.0 1.4 Ma for a whole-rock chip of the same sample from Candelaria (Arvalo, 1999), and a 40Ar/39Ar total fusion weighted mean age (2 analyses) of green hydrothermal biotite from the Resguardo mine (Punta del Cobre district) yielding 111.6 1.4 Ma (Marschik et al., 1997b) are identical within errors with the amphibole plateau age of Ullrich and Clark (1998, 1999). These younger alteration ages of 110 to 112 Ma could indicate that there is another, possibly paragenetically similar alteration and mineralization event. This event could be related to nearby 110 to 112 Ma batholithic intrusions (Fig. 2). Implications for tectonic setting and depth of mineralization There is evidence for episodes of Middle to Late Cretaceous transpression (e.g., Sylvester and Palacios, 1992; Lara et al., 1996; Arvalo and Grocott, 1997; Taylor et al., 1998) as well as extension (Mpodozis and Allmendinger, 1993) at a regional scale. However, their exact timing is only poorly constrained. The age data summarized here have implications for the structural evolution of the area. Parts of the iron oxide and the bulk of the Cu ore at Punta del Cobre are hosted by north-northwest to northwest-trending high-angle sinistral transcurrent brittle faults, which were active by Aptian times (115 Ma). At Candelaria, mineralization occurred in these same north-northwest to northwest faults and in the northnortheasttrending Candelaria shear zone. The latter shows indication of minor movements during the early iron metasomatism but was essentially inactive as a ductile shear zone at the time of main Cu mineralization during which brittle conditions prevailed. The northwest-trending high-angle faults play an important role in structural models for Middle-Late Cretaceous transpressional conditions for the Atacama region. Age and field data presented here are compatible with the onset of generalized transpressional conditions in the region already at the
0361-0128/98/000/000-00 $6.00

time of mineralization. This is consistent with the geologic record indicating cessation of marine sedimentation in the back-arc and regional uplift in the late Aptian (e.g., Segerstrom and Parker, 1959; Zentilli, 1974; Jurgan, 1977, Prez et al., 1990). Taking into account that by around 115 Ma (Aptian) the Chaarcillo Group (late Valanginian to Aptian) should have reached its maximum thickness of 1,700 to 2,000 m, and that the ore deposits essentially occur in the upper part of the Punta del Cobre Formation, it is possible to estimate that mineralization took place at about a 2- to 3-km depth (Fig. 13). The change from ductile-brittle conditions during mineralization could be explained by a temperature decrease that went along with the regional uplift in the late Aptian. Related iron oxide-rich deposits Other Cretaceous ore deposits in the Andean Cordillera present similarities to the Candelaria-Punta del Cobre iron oxide Cu-Au(-Zn-Ag) system, although some have been interpreted using other genetic models. These deposits include the Manto Verde (K-Ar sericite ages are 117 3 and 121 3 Ma; Vila et al., 1996), Manto Ruso (Orrego and Zamora, 1991), and Teresa de Colmo iron oxide Cu-Au deposits (Hopper and Correa, 2000) in Chile, and the Ral-Condestable and Eliana iron oxide Cu-Au deposits in Per (K-Ar amphibole ages are 127 3.1 or 113 3.0 Ma, respectively; Vidal et al., 1990; see also Ripley and Ohmoto, 1977; Cardozo and Wauschkuhn, 1984; Cardozo, 1990; Haller, 2000). There are also similarities with the Early Cretaceous magnetite(-apatite) deposits of the Chilean iron belt (e.g., El Romeral, Bookstrom, 1977; Los Algarrobos, Mnard, 1995; Cerro Negro Norte, Vivallo et al., 1995) regarding setting, age, alteration, and host rocks. These iron deposits commonly contain minor amounts of sulfides, including chalcopyrite. Mineralization follows a similar paragenetic pattern with early magnetite and superposed sulfide formation. The Cerro Negro Norte deposit is a particularly good example. In the Cerro Negro Norte orebody, magnetite is cut by calcic amphibole veinlets that in turn are cut and used by later veinlets that contain pyrite chalcopyrite similar to those at Candelaria-Punta del Cobre (Fig. 9). The deposits of the iron belt occur in rocks that are affected by sodic(-calcic) alteration. However, a significant potassic component to alteration as in most iron oxide Cu-Au systems is absent. The common spatial and temporal association of iron and iron oxide Cu-Au deposits in Chile, Per, and elsewhere (e.g., Norbotten region, Sweden; Carlon, 2000) indicates that these deposits might also be genetically related. Conclusions The Candelaria-Punta del Cobre Cu-Au(-Zn-Ag) deposits have features that are characteristic of iron oxide (Cu-U-AuREE) deposits (Hitzman et al., 1992; Hitzman, 2000). These characteristics include abundant iron oxide as magnetite in the more internal and hematite in the more external portions of the system, and local enrichments in light rare earth elements. The deposits occur near a lithostratigraphic boundary that is intersected by a major structural zone. Brittle and ductile structures created localized dilatations that channeled the ore-forming fluids. Ore is associated with calcic-potassic

1822

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE

1823

alteration in the internal, potassic in the intermediate, and sodic alteration carbonatization in the marginal portions of the system. The radiometric ages point to a main Cu-Au mineralization event at Candelaria-Punta del Cobre at around 115 Ma. Ore formation was spatially and temporally associated with Early Cretaceous batholithic granitoid intrusions. The CandelariaPunta del Cobre Cu-Au(-Zn-Ag) deposits formed at a relatively shallow crustal level (<3 km) in rocks of an arc and/or back-arc environment. Although the host rocks are consistent with an extensional setting, as proposed for this deposit class, the available age data suggest that mineralization occurred during the onset of basin inversion, probably under transpressional conditions at a regional scale. The metal-bearing fluid was oxidized, saline, near-pH neutral, with oxygen and sulfur isotope compositions in equilibrium with magmatic rocks. This fluid had a homogeneous lead isotope signature similar to that of least altered volcanic host rocks and nearby intrusive rocks. The bulk of the magnetite probably formed at temperatures of about 500 to 600C. The main sulfide stage followed with formation of pyrite and chalcopyrite at temperatures of >470 to 328C. Cooling and mixing of the metal- and sulfur-bearing fluid with a fluid that has a light oxygen isotope signature (e.g., basinal brine or meteoric waters) are probably the main precipitation mechanisms at Candelaria-Punta del Cobre. During the early stages of mineralization, reduction of the oxidized fluid through interaction with the volcanic host rocks could have additionally caused ore mineral precipitation. The stable isotope data, the presence of previously reported hypersaline CO2-rich fluid inclusions in the main Cu-ore stage and saline fluid inclusions in late-stage calcite, the oxidized character of the ore fluid, and the fact that mineralization occurred coeval with nearby intrusive activity are consistent with, but not unequivocally evidence for, magmatic fluid contribution into the hydrothermal system. Finally, we would like to emphasize the preliminary nature of the conclusions drawn. In particular, the role of magmatic and the nature of nonmagmatic fluids in the Candelaria-Punta del Cobre iron oxide Cu-Au(-Zn-Ag) system is focus of the current research. Acknowledgments We thank Cominor, Compaa Contractual Minera Candelaria, Compaa Contractual Minera Carola, Compaa Minera Ojos del Salado, Empresa Minera Mantos Blancos, Phelps Dodge Exploration Corporation, Sali-Hochschild S.A., and Sociedad Punta del Cobre S.A. for access to their mines and properties. We are particularly grateful to L. Alvarez, P. Anguita, G. Arce, T. Balcu, C. Caldern, P. Caldern, B. Castillo, O. Edelstein, P. Flores, A. Gordillo, R. Leveille, W. Martin, E. Nuez, R. Nuez, R. Olivares, G. Oyarzn, J. Ponce, N. Pop, W. Robles, and R. Zamora for discussions and constructive comments. We thank M. Chiaradia (Univ. of Geneva) for assisting in the interpretation of the lead isotope data and Y. Haeberlin and L. Linares (Univ. of Geneva) for sulfur isotope analyses, some of which were carried out in the stable isotope laboratory of J. Hunzicker and J. Spangenberg, Section des Sciences de la Terre, University of Lausanne. We are grateful to J. Chesley (Univ. of Arizona) and L. Peters (NMGRL) for their help in interpreting the age data. This
0361-0128/98/000/000-00 $6.00

paper has greatly benefited from critical reviews and suggestions by M.D. Barton, J.W. Hedenquist, and P. Williams. This investigation was supported by the Deutsche Forschungsgemeinschaft (DFG) and the Swiss National Science Foundation. May 23, 2000; July 30, 2001
REFERENCES
Arvalo, C., 1994, Mapa Geolgico del cuadrngulo Los Loros: Santiago, Chile, SERNAGEOMIN, Documentos de Trabajo 6, scale 1:100.000. 1995, Mapa Geolgico de la Hoja Copiap, Regin de Atacama: Santiago, Chile, SERNAGEOMIN, Documentos de Trabajo 8, scale 1:100.000. 1999, The Coastal Cordillera-Precordillera boundary in the Copiap area, northern Chile and the structural setting of the Candelaria Cu-Au ore deposit: Unpublished Ph.D. thesis, Kingston-Upon-Thames, Kingston University, 244 p. Arvalo, C., and Grocott, J., 1997, The tectonic setting of the Chaarcillo Group and the Bandurrias Formation: An Early-Late Cretaceous sinistral transpressive belt between the Coastal Cordillera and the Precordillera, Atacama region, Chile: Congreso Geolgico Chileno, 8th, Antofagasta, 1997, Actas, p. 16041607. Arribas, A.J., 1995, Characteristics of high-sulfidation epithermal deposits, and their relation to magmatic fluid: Mineralogical Association of Canada Short Course Series v. 23, p. 419454. 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., Hydrothermal iron-oxide copper-gold & related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 4360. Barton, M.D., Johnson, D.A., and Hanson, R.B., 1998, Evaluation of possible roles of non-magmatic brines in igneous-related hydrothermal systems, especially Fe(-Cu-Au-REE) deposits [abs.]: Geological Society of America Abstracts with Programs, p. A-127. Battles, D.A., and Barton, M.D., 1995, Arc-related sodic hydrothermal alteration in the western United States: Geology, v. 23, p. 913916. Bookstrom, A.A., 1977, The magnetite deposits of El Romeral, Chile: ECONOMIC GEOLOGY, v. 72, p. 11011130. Brown, M., Diaz, F., and Grocott, J., 1993, Displacement history of the Atacama fault system 2500S-2700S, northern Chile: Geological Society of America Bulletin, v. 105, p. 11651174. Camus, F., 1980, Posible modelo gentico para los yacimientos de cobre del distrito minero Punta del Cobre: Revista Geolgica de Chile, v. 11, p. 5176. Cardozo, M., 1990, The Copara metallotect in central Peru: Geologic evolution and ore formation, in Fontbot, L., Amstutz, G.C., Cardozo, M., Cedillo, E., and Frutos, J., eds., Stratabound ore deposits in the Andes: Berlin-Heidelberg, Springer, p. 395-412. Cardozo, M., and Wauschkuhn, A., 1984, The Copara and the Patap metallotect on the western side of central Peru, in Wauschkuhn, A., Kluth, C., and Zimmermann, R.A., eds., Syngenesis and epigenesis in the formation of mineral deposits: Berlin-Heidelberg, Springer, p. 616646. Carlon, C.J., 2000, Iron oxide systems and base metal mineralization in northern Sweden, in Porter, T.M., ed., Hydrothermal iron-oxide coppergold & related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 283296. Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, Z., and Zak, I., 1980, The age curves of sulfur and oxygen isotopes in marine sulfate and their mutual interpretation: Chemical Geology, v. 28, p. 199260. Dallmeyer, R.D., Grocott, J., Brown, M., Taylor, G.K., and Treloar, P.J., 1996, Mesozoic magmatic and tectonic events within the Andean plate boundary zone, 26-2730S, north Chile: Constraints from 40Ar/39Ar mineral ages: Journal of Geology, v. 104, p. 1940. Deino, A., and Potts, R., 1990, Single-crystal 40Ar/39Ar dating of the Olorgesailie Formation, southern Kenya rift: Journal of Geophysical Research, v. 95, p. 84538470. Daz, R., 1990, Evolucin metasomatica del ara del depsito tipo skarn de Fe, Cu, Au El Bronce, Tierra Amarilla, III Region/Chile: Unpublished Memoria de Titulo, Antofagasta, Universidad Catolica del Norte, 181 p.

1823

1824

MARSCHIK AND FONTBOT Marschik, R., and Chiaradia, M., 2000, Lead isotope signatures of ore, volcanic, and batholithic rocks of the Candelaria-Punta del Cobre area, Chile [abs.]: International Geological Congress, 31st, Rio de Janeiro, Brazil, August 617, 2000, Abstracts.(CD-ROM) Marschik, R., and Fontbot, L., 1996, Copper(-iron) mineralization and superposition of alteration events in the Punta del Cobre belt, northern Chile: Society of Economic Geologists Special Publication 5, p. 171189. 2001, The Punta del Cobre Formation, Punta del Cobre-Candelaria area, Chile: Journal of South American Earth Sciences, v. 14, p. 401433. Marschik, R., Chiaradia, M., and Fontbot, L., 1997a, Intrusion-related Cu (-Fe)-Au mineralization of the Punta del Cobre belt, Chile: Lead and sulfur isotopic constraints, in Papunen, H., ed., Mineral deposits: Research and explorationwhere do they meet?: Rotterdam, Balkema, p. 655658. Marschik, R., Singer, B.S., Munizaga, F., Tassinari, C., Moritz, R., and Fontbot, L., 1997b, Age of Cu(-Fe)-Au mineralization and thermal evolution of the Punta del Cobre district, Chile: Mineralium Deposita, v. 32, p. 531546. Marschik, R., Leveille, R.A., and Martin, W., 2000, La Candelaria and the Punta del Cobre district, Chile: Early Cretaceous iron oxide Cu-Au(-ZnAg) mineralization, in Porter, T.M., ed., Hydrothermal iron-oxide coppergold & related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 163175. Martin, W., Daz, R., Nuez, R., Olivares, R., Caldern, C., and Caldern, P., 1997, The updated Candelaria geologic mine model: Congreso Geolgico Chileno, 8th, Antofagasta, Actas, p. 10631067. Mathur, R.D., 2000, Re-Os isotopes of base metal porphyry deposits: Unpublished Ph.D. thesis, Tucson, Arizona, University of Arizona, 153 p. Mathur, R.D., Marschik, R., Ruiz, J., Munizaga, F., Leveille, R.A., and Martin, W., 2002, Age of mineralization of the Candelaria iron oxide Cu-Au deposit, and the origin of the Chilean iron belt based on Re-Os isotopes: ECONOMIC GEOLOGY, v. 97, in press. Mnard, J.J., 1995, Relationship between altered pyroxene diorite and the magnetite mineralization in the Chilean iron belt, with emphasis on the El Algarrobo iron deposits (Atacama region, Chile): Mineralium Deposita, v. 30, p. 268274. Mpodozis, C., and Allmendinger, R., 1993, Extentional tectonics, Cretaceous Andes, northern Chile (27S): Geological Society of America Bulletin, v. 105, p. 14621477. Nakamura, N., 1974, Determination of REE, Ba, Fe, Mg, Na, and K in carbonaceous and ordinary chondrites: Geochimica et Cosmochimica Acta, v. 38, p. 757775. Ohmoto, H, and Lasaga, A.C., 1982, Kinetics of reactions between aqueous sulphates and sulphides in hydrothermal systems: Geochimica et Cosmochimica Acta, v. 46, p. 17271745. Ohmoto, H., and Rye, R.O., 1979, Isotopes of sulfur and carbon, in Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits: New York, John Wiley and Sons, p. 509567. Orrego, M., and Zamora, R., 1991, Manto Ruso: Un Yacimiento de Cobre ligado a la Falla de Atacama, Norte de Chile: Congreso Geolgico Chileno, 6th, Via del Mar, Resumenes Expandidos, 1991, p. 174178. Prez, E., Cooper, M.R., and Covacevich, V., 1990, Aptian ammonite-based age for the Pabelln Formation, Atacama region, Chile: Revista Geolgica de Chile, v. 17, p. 181185. Pollard, P.J., 2000, Evidence of a magmatic fluid and metal source for Feoxide Cu-Au mineralization, in Porter, T.M., ed., Hydrothermal iron-oxide copper-gold & related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 2741. Rabbia, O.M., Frutos, J., Pop, N., Isache, C., Sanhueza, V., and Edelstein, O., 1996, Caractersticas isotpicas de la mineralizacin de Cu (-Fe) de Mina Carola, distrito minero Punta del Cobre, norte de Chile: Congreso Geolgic Argentino, 8th, Buenos Aires, Actas, p. 241254. Rahmdohr, P., 1980, The ore minerals and their intergrowths. Oxford, Pergamon Press, 1207 p. Requia, K., 2002, The Archean Salobo iron oxide Cu-Au deposit, Carajs mineral province, Brazil: Terre et Environnement, Universit de Genve, in press. Ripley, E.M., 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. 10171042. Rotherham, J.F., 1997, Origin and fluid chemistry of the Starra ironstones and high-grade Au-Cu mineralization, Cloncurry district, Mount Isa inlier, Australia: Unpublished Ph.D. thesis, Townsville, James Cook University, 251 p.

Dilles, J.H., and Einaudi, M.T., 1992, Wall-rock alteration and hydrothermal flow paths about the Ann Mason porphyry copper deposit, Nevada: A 6-km vertical reconstruction: ECONOMIC GEOLOGY, v. 87, p. 10632001. Einaudi, M.T., 1982, General features and origin of skarns associated with porphyry copper plutons, southwestern North America, in Titley, S.R., ed., Advances in geology of the porphyry copper deposits, southwestern North America: Tucson, Arizona, Univerity of Arizona Press, p. 185210. Einaudi, M.T., Meinert, L.D., and Newberry, R.J., 1981, Skarn deposits: ECONOMIC GEOLOGY 75TH ANNIVERSARY VOLUME, p. 317391. Espinoza, S., 1990, The Atacama-Coquimbo ferrifereous belt, northern Chile: Society for Geology Applied to Mineral Deposits Special Publication 8, p. 395412. Espinoza, S., Vivallo, W., and Henriquez, F., 1994, Geologia y genesis de mineralizacin metalica en el distrito ferrfero de Cerro Imn, Copiap, Chile: Congreso Geolgico Chileno, 7th, Concepcin, Actas, p. 799802. Fleck, R.J., Sutter, F.J., and Elliot, D.H., 1977, Interpretation of discordant 40 Ar/39Ar age-spectra of Mesozoic tholeiites from Antarctica: Geochimica et Cosmochimica Acta, v. 41, p. 1532. Flores, P., 1997, Geologa y mineralizacin del yacimiento de cobre Santos y el enterno geollico. III. Regin Chile: Congreso Geolgico Chileno, 8th, Antofagasta, Actas, p. 956960. Friedman, I., and ONeil, J.R., 1977, Data of geochemistry: U.S. Geological Survey Professional Paper 440-KK, 12 p. Gow, P.A., Wall, V.J., Oliver, N.H.S., and Valenta, R.K., 1994, Proterozoic iron oxide (Cu-U-Au-REE) deposits: Further evidence of hydrothermal origins: Geology, v. 22, p. 633636. Gradstein, F.M., Agterberg, F.P., Ogg, J.G., Hardenbol, J., Van Veen, P., Thierry, J., and Huang, 1995, A Triassic, Jurassic and Creataceous time scale: Society for Sedimentary Geology Special Publication 54, p. 95126. Grocott, J., Taylor, G.K., and Treloar, P.J., 1993, Mesozoic extensional and strike-slip fault systems in magmatic arc rocks of the Andean plate boundary zone, northern Chile [ext. abs.]: International Symposium on Andean Geodynamics, Oxford, U.K., Extended Abstracts, p. 187190. Haller, A., de, 2000, The Ral-Condestable iron oxides-Cu-Au deposit, Lima department, Per: Preliminary results: Annual GEODE-Fennoscandian Shield Field Workshop on Paleoproterozoic and Archean Greenstone Belts and VMS Districts in the Fennoscandian Shield, 2nd, Gllivare-Kiruna, Sweden, August 28-September 1, 2000, Research Report, p. 811. 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. Hitzman, M.W., 2000, Iron oxide-Cu-Au deposits: What, where, when, and why, in Porter, T.M., ed., Hydrothermal iron-oxide copper-gold & related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 925. Hitzman, M.W., Oreskes, N., and Einaudi, M.T., 1992, Geological characteristics and tectonic setting of Proterozoic iron oxide (Cu-U-Au-REE) deposits: Precambrian Research, v. 58, p. 241287. Hopf, S., 1987, Petrographische, Mineralogische und Geochemische Beobachtungen an der Cu-Lagersttte Agustina/Distrikt Punta del Cobre/Chile: Unpublished Diplomarbeit, Heidelberg, Universitt Heidelberg, 144 p. 1990, The Agustina mine, a volcanic-hosted copper deposit in northern Chile: Society for Geology Applied to Mineral Deposits Special Publication 8, p. 421434. Hopper, D., and Correa, A., 2000, The Panulcillo and Teresa de Colmo copper deposits: Contrasting examples of Fe-ox Cu-Au mineralization from the Coastal Cordillera of Chile, in Porter, T.M., ed., Hydrothermal iron-oxide copper-gold & related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 177189. Jurgan, H., 1977, Strukturelle und lithofazielle Entwicklung des andinen Unterkreide-Beckens im Norden Chiles (Provinz Atacama): Geotektonische Forschung, v. 52, 138 p. Lang, J.R., Stanley, C.R., Thompson, J.F.H., and Dunne, K.P.E., 1995, NaK-Ca magmatic-hydrothermal alteration in alkalic porphyry Cu-Au deposits, British Columbia: Mineralogical Association of Canada Short Course Series, v. 23, p. 339366. Lara, L., Gelcich, S., Carrasco, J., and Daz, A., 1996, Arc and forearc brittle deformation in transpressive regime of the Lower Cretaceous, Coastal Range 26-27S, Chile: Microtectonic antecedents [ext. abs.]: International Symposium on Andean Geodynamics (ISAG), 3rd, St. Malo, France, Extended Abstracts, p. 415418. Lino, S., 1984, Geologa de la mina Agustina, Comuna de Tierra Amarilla, Provincia de Copiap, Region de Atacama: Unpublished Memoria de Titulo, Antofagasta, Universidad Catolica del Norte, 161 p.

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

1824

CANDELARIA-PUNTA DEL COBRE IRON OXIDE Cu-Au(-Zn-Ag) DEPOSITS, CHILE Rotherham, J.F., Blake, K.L., Cartwright, I., and Williams, P.J., 1998, Stable isotope evidence for the origin of the Mesoproterozoic Starra Au-Cu deposit, Cloncurry district, Northwest Queensland: ECONOMIC GEOLOGY, v. 93, p. 14351449. Ryan, P.J., Lawrence, A.L., Jenkins, R.A., Matthews, J.P., Zamora, J.C., Marino, E., and Urqueta, I., 1995, The Candelaria copper-gold deposit, Chile: Arizona Geological Society Digest, v. 20, p. 625645. Samson, S.D., and, Alexander, E.C., Jr., 1987, Calibration of the interlaboratory 40Ar/39Ar dating standard, Mmhb-1: Chemical Geology, v. 66, p. 2734. Scheuber, E., and Andriessen, P.A.M., 1990, The kinematic and geodynamic significance of the Atacama fault zone, northern Chile: Journal of Structural Geology, v. 12, p. 243257. Scheuber, E., Hammerschmidt, K., and Friedrichsen, H., 1995, 40Ar/39Ar and Rb-Sr analyses from ductile shear zones from the Atacama fault zone, northern Chile: The age of deformation: Tectonophysics, v. 250, p. 6187. Segerstrom, K., 1962, Structural effects related to hydration of anhydrite, Copiap area, Chile: U.S. Geological Survey Professional Paper 450C, p. C28C30. Segerstrom, K. and Parker, R.L., 1959, Cuadrngulo Cerrillos, Provincia de Atacama: Santiago, Instituto de Investigaciones Geolgicas, Carta Geolgica de Chile, v. 1, no. 2, 33 p. Segerstrom, K. and Ruiz, C., 1962, Geologa del Cuadrngulo Copiap, Provincia de Atacama: Santiago, Instituto de Investigaciones Geolgicas, Carta Geolgica de Chile, v. 3, no. 1, 115 p. Segerstrom, K., Thomas, H., and Tilling, R.I., 1963, Cuadrngulo Pintadas, Provincia de Atacama: Santiago, Instituto de Investigaciones Geolgicas, Carta Geolgica de Chile, no. 52, 12 p. Sillitoe, R.H., and Clark, A.H., 1969, Copper and copper iron sulphides as the initial products of supergene oxidation, Copiap mining district, northern Chile: American Mineralogist, v. 54, p. 16841710. Sylvester, H., and Linke, M., 1993, Structural control of intrusions and hydrothermal alteration zones by intersecting fault systems in the Creataceous magmatic arc of the southern Central Andes 27 S, III. Region, Chile: Zentralblatt fr Geologie und Palontologie, Teil I, p. 361376. Sylvester, H., and Palacios, C., 1992, Transpressional structures in the Andes between the Atacama fault zone and the West Fissure system at 27S, III Region, Chile: Zentralblatt fr Geologie und Palontologie, Teil I, p. 16451658. Taylor, G.K., Grocott, J., Pope, A., and Randall, D.E., 1998, Mesozoic fault systems, deformation and fault block rotation in the Andean forearc; a crustal scale strike-slip duplex in the Coastal Cordillera of northern Chile: Tectonophysics, v. 13, p. 93109. 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.

1825

Tilling, R., 1962, Batholith emplacement and contact metamorphism in the Paipote-Tierra Amarilla area, Atacama province, Chile: Unpublished Ph.D. thesis, New Haven, Connecticut, Yale University, 195 p. 1963, Disequilibrium skarns of the Tierra Amarilla aureole, Atacama province, Chile [abs.]: Geological Society of America Special Paper 76, p. 167. 1976, El Batolito Andino cerca de Copiap, Provinca de Atacama. Geologa y Petrologa: Revista Geolgica de Chile, v. 3, p. 124. Ullrich, T.D., and Clark, A.H., 1997, Paragenetic sequence of mineralization in the main orebody, Candelaria copper-gold deposit, Chile: Phoenix, Arizona, Phelps Dodge Exploration Corporation, unpublished internal report 3, 36 p. 1998, Evolution of the Candelaria Cu-Au deposit, III Region, Chile [abs.]: Geological Society of America Abstracts with Programs, p. A-75. 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. Vidal, C., Injoque-Espinoza, J., Sidder, G.B., and Mukasa, S.B., 1990, Amphibolitic Cu-Fe skarn deposits in the central coast of Peru: ECONOMIC GEOLOGY, v. 85, p. 14471461. Vila, T., Lindsay, N., and Zamora, R., 1996, Geology of the Manto Verde copper deposit, northern Chile: A specularite-rich, hydrothermal-tectonic breccia related to the Atacama fault zone: Society of Economic Geologists Special Publication 5, p. 157170. Vivallo, W., Espinoza, S., and Henrquez, F., 1995, Metasomatismo y alteracin hidrotermal en el Distrito Ferrfero Cerro Negro Norte, Copiap, Chile: Revista Geolgica de Chile, v. 22, p. 7588. Williams, P.J., 1998, Metalliferous economic geology of the Mt. Isa Eastern Succession, northwest Queensland: Australian Journal of Earth Sciences, v. 45, p. 329341. Williams, P.J., Guoyi, D., Pollard, P.J., and Perring, C.S., 1999, Fluid inclusion geochemistry of Cloncurry (Fe)-Cu-Au deposits, in Stanley et al., eds., Mineral deposits: Processes to processing: Rotterdam, Balkema, v. 1, p. 111114. Wilson, J., Dallmeyer, R.D., and Grocott, J., 2000, New 40Ar/39Ar dates from the Las Tazas Complex, northern Chile: Tectonic significance: Journal of South American Earth Sciences, v. 13, p. 115122. Zentilli, M., 1974, Geological evolution and metallogenetic relationships in the Andes of northern Chile between 26 and 29 S: Unpublished Ph.D. thesis, Kingston, Ontario, Queens University, 394 p.

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

1825

1826

MARSCHIK AND FONTBOT

APPENDIX The 40Ar/39Ar analysis was carried out by the Geochronological Research Laboratory of the New Mexico Bureau of Mines and Mineral Resources, Socorro, New Mexico, using the furnace incremental-heating technique. Sample preparation and irradiation: The biotite and amphibole concentrates were prepared using standard handpicking techniques. The samples were loaded into a machined Al disk and irradiated for 24 h in L67 position in the Ford Memorial Reactor, University of Michigan. Neutron flux monitor was Fish Canyon Tuff sanidine (FC-1) with an assigned age of 27.84 Ma (Deino and Potts, 1990) relative to Mmhb-1 at 520.4 Ma (Samson and Alexander, 1987). Instrumentation: A Mass Analyzer Products 215-50 mass spectrometer online with automated all-metal extraction system was used. Mineral separates were step-heated in an Mo double-vacuum resistance furnace. Heating duration was 8 min. Reactive gases were removed by reaction with three SAES GP-50 getters, two operated at ~450C and one at 20C. Gas was also exposed to a tungsten filament operated at ~2,000C. Analytical parameters: Electron multiplier sensitivity averaged 2.81 1016 moles/pA. Total system blank and background for the furnace averaged 324, 1.2, 0.4, 1.1, 1.4 1018 moles at masses 40, 39, 38, 37, and 36, respectively. J factors were determined to a precision of 0.1 percent by CO2 laser fusion of four single crystals from each of four radial positions around the irradiation tray or each monitor packet in quartz tube. Correction factors for interfering nuclear reactions were determined using K glass and CaF2 and are as follows: (40Ar/39Ar)K = 0.0237 0.00096; (36Ar/37Ar)Ca = 0.000289 0.000008; and (39Ar/37Ar)Ca = 0.000747 0.0003.

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

1826

You might also like