Sei sulla pagina 1di 24

Economic Geology, v. 112, pp.

295318

Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with


Cretaceous Porphyry Cu Mo Au and Intrusion-Related Iron Oxide Cu-Au Deposits
in Northern Chile*
Jeremy P. Richards,1, Gloria P. Lpez,1 Jing-Jing Zhu,1,2 Robert A. Creaser,1 Andrew J. Locock,1 and A. Hamid Mumin3
1 Department of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta T6G 2E3, Canada
2 State Key Laboratory of Ore Deposit Geochemistry, Institute of Geochemistry, Chinese Academy of Sciences, Guiyang 550081, PR China
3 Department of Geology, Brandon University, 270-18th Street, Brandon, Manitoba R7A 6A9, Canada

Abstract
Porphyry Cu Mo Au and iron oxide copper-gold (IOCG) deposits share many similarities (e.g., Fe, Cu,
and Au contents), but also have important differences (e.g., the predominance of sulfide minerals in porphyry
deposits and iron oxides in IOCG deposits). Genetic comparisons are complicated by the broad definition of
IOCG deposits; here we restrict our study to IOCG deposits that are related to igneous intrusive systems. In
the Mesozoic Coastal Cordillera of northern Chile, both porphyry and IOCG deposits occur in close spatial and
temporal proximity, offering the chance to examine what controls their different modes of formation. From
detailed examination of the timing, geochemistry, and tectonic setting of associated igneous rocks, based on
new and published data, we find that rocks associated with mid-Cretaceous IOCG deposits (~125110 Ma) are
largely indistinguishable from those associated with slightly earlier (>125 Ma) and later (<110 Ma) porphyry
Cu Mo Au deposits. Magmas related to IOCG deposits were formed during a brief period of back-arc
transtension in the mid-Cretaceous and are, on average, somewhat more mafic (dioritic), locally alkaline, and
isotopically primitive compared to granodioritic magmas associated with porphyry deposits formed during nor-
mal contractional arc tectonics in the later Cretaceous. However, these compositional ranges overlap, and the
differences are not clear enough to be diagnostic.
We measured the SO3 content of igneous apatite from selected samples of these rocks to test the hypothesis
that the difference in sulfur content of the ore deposits was due to differences in sulfur content of the associated
magmas. Early igneous apatite crystals occurring as inclusions in silicate phenocrysts from the Carmen de And-
acollo porphyry Cu-Au deposit (Re-Os molybdenite ages of 103.9 0.5 Ma, 103.6 0.5 Ma) are significantly
richer in S (0.25 0.17 wt % SO3, n = 69) than similar apatite crystals from two IOCG deposits (Candelaria,
Casualidad) and a sample of regional mid-Cretaceous igneous rock from near Productora (0.04 0.02 wt %
SO3, n = 76). Using published partition coefficients for S between apatite and oxidized silicate melt, we semi-
quantitatively estimate corresponding magmatic sulfur contents of ~0.02 wt % S in the Carmen de Andacollo
magmas versus ~0.001 to 0.005 wt % S in the IOCG-associated magmas. This is an order of magnitude dif-
ference, and the opposite of what would be expected if the difference were due to bulk magma composition
(sulfur solubility is generally higher in mafic magmas, whereas here the S content is higher in the more felsic
porphyries). We conclude that the porphyry-forming magmas indeed had higher S contents than the IOCG-
related magmas and suggest that these differences reflect different petrogenetic processes. During normal sub-
duction, magmas derived from the metasomatized mantle wedge are hydrous, moderately oxidized, and S rich,
and have the potential to generate S-rich porphyry-type deposits. In contrast, in back-arc extensional settings,
upwelling asthenospheric melts carry a weaker subduction signature, including lower S contents. Interaction of
these S-poor magmas with previously subduction modified upper plate lithosphere is more likely to give rise to
S-poor IOCG deposits.

Introduction oxides (as opposed to Fe sulfides), with or without Cu and Au


Sulfur-rich porphyry Cu Mo Au deposits are formed by mineralization (Hitzman, 2000; Williams et al., 2005, 2010;
the precipitation of sulfide minerals (pyrite, chalcopyrite, Hunt et al., 2007; Groves et al., 2010). The broadness of this
molybdenite) from hydrothermal fluids exsolved from shal- definition, as well as the capacity of oxidized saline fluids to
lowly emplaced calc-alkaline magmas in volcanic arcs, typi- transport Fe Cu Au in a variety of geologic settings, has
cally generated in response to oceanic lithosphere subduction led to controversy over the origin(s) of this group of deposits
(Burnham, 1979; Richards, 2003; Cooke et al., 2005; Sil- (e.g., Barton and Johnson, 1996, 2000; Pollard, 2000; Williams
litoe, 2010). In contrast, sulfur-poor iron oxide copper-gold et al., 2005; Williams, 2010; Barton, 2014). However, within
(IOCG) deposits include a wide range of different deposit this group there is a subset of deposits that is more clearly
types, broadly linked by the prevalence of hydrothermal Fe associated with igneous rocks and fluids of possible magmatic-
hydrothermal origin (Pollard, 2000; Sillitoe, 2003). Richards
Corresponding author: e-mail, Jeremy.Richards@ualberta.ca and Mumin (2013a, b) have referred to these as magmatic-
*A digital supplement to this paper is available at http://economicgeology.org/ hydrothermal IOCG deposits, but, for simplicity, we use the
and at http://econgeol.geoscienceworld.org/. general term IOCG below. They share many features with

2017 Gold Open Access: this paper is published under the terms of the CC-BY license.

Submitted: October 2, 2015


0361-0128/17/4470/295-24 295 Accepted: August 8, 2016
296 RICHARDS ET AL.

porphyry systems (e.g., Cu-Au-Mo-Fe metal association and 1,000km (Fig. 1a; Sillitoe, 2003). The largest IOCG depos-
broad tectonic setting and magmatic affinity) but also have its within this part of the belt are Candelaria (116110 Ma;
important differences (such as more extensive development 501Mt at 0.54% Cu, 0.13 g/t Au, and 2.06 g/t Ag) and Man-
of high-temperature sodic, sodic-calcic, and potassic iron toverde (121117 Ma; 440 Mt at 0.56 % Cu and 0.12 g/t Au).
alteration envelopes, and more restricted development of The largest porphyry Cu-Au deposit is Carmen de Andacollo
lower-temperature phyllic and argillic alteration in IOCG (104 Ma; proven and probable reserves of 417 Mt at 0.34%
systems compared to porphyry systems; Hitzman et al., 1992; Cu and 0.12 g/t Au; Table 1).
Mumin et al., 2010). Small Early Cretaceous (>125 Ma; Berriasian-Barremian)
Richards and Mumin (2013b) explained some of these dif- porphyry Cu-Au deposits occur at 22S (e.g., Antucoya,
ferences, in particular the smaller acidic alteration zones in 142Ma; Tovaku, 132 Ma; Maksaev et al., 2006) and at 33S
IOCG deposits, as reflecting lower abundances of S (SO2) in (e.g., Colliguay, ~129 Ma; Maksaev et al., 2010) and appear to
the ore-forming fluids compared to porphyry fluids. As S-rich have formed during a short period of synchronous transpres-
porphyry fluids cool, the SO2 disproportionates to form H2S sion at 22S (Maksaev et al., 2006) or continental arc extension
and H2SO4 (sulfuric acid), leading to widespread develop- at 33S (Creixell, 2007). In contrast, IOCG (and magnetite-
ment of acidic alteration at shallow levels (Burnham, 1979; apatite) deposits formed predominantly in the mid-Creta-
Candela, 1992; Field et al., 2005; Richards, 2011b, 2015); this ceous (~125~110 Ma; Aptian-Albian; Gelcich et al., 2005;
happens to a lesser extent in S-poor IOCG fluids. Another Arvalo et al., 2006; Rieger et al., 2010; Tornos et al., 2010)
important difference is the broader range of metals found in and are located in a tectonically distinct but spatially super-
IOCG versus porphyry deposits, including, in some deposits, imposed belt from 25 to 34S. These deposits have been
the presence of abundant U, rare earth elements (REEs), P, described as spatially and temporally related to the culmina-
Co, Ni, and Bi. The increased variety of metals is attributed tion of a period of back-arc extension that developed along the
to the much greater influence of fluid reactions with crustal continental margin from the Late Jurassic to mid-Cretaceous
rocks and the resultant metal fluxing that occurs in giant (Oyarzun et al., 1999; Grocott and Taylor, 2002; Sillitoe, 2003)
IOCG hydrothermal systems (in addition to magmatic contri- or, alternatively, as having formed in response to the initiation
butions), and the more common occurrence of mafic country of basin inversion in the mid-Cretaceous (Chen et al., 2013).
rocks around many IOCGs (Somarin and Mumin, 2012; Rich- Porphyry Cu-Au-(Mo) deposits, including the large Carmen
ards and Mumin, 2013b; Barton, 2014). de Andacollo porphyry Cu-Au deposit (104Ma), again began
The Mesozoic Coastal Cordillera of northern Chile is to form in the later Cretaceous (<110 Ma; Cenomanian-
uniquely suited to compare the characteristics and controls Turonian) between 25 and 32S, following the resumption
on ore formation in porphyry and IOCG deposits because of arc magmatism during or after basin inversion (Maksaev
both deposit types occur in a broadly coeval (Cretaceous), et al., 2010). We group igneous rocks and associated ore
50- to 80-km-wide belt that runs parallel to the coast for over deposits into three broad temporal groupings that relate to
24S
Extension Sinistral transtension Contraction
N Legend
Chile (1) (i)
(1,2) (3) (1)
Tropezon (110 Ma) Porphyry CuMoAu
Casualidad (10094 Ma) (> or < 1 Mt Cu)
26S Todos los Santos (118 Ma)
Santo Domingo (124 Ma) IOCG (> or < 1 Mt Cu)
Mantoverde (121117 Ma) Inca de Oro
Cerro Negro Norte (116 Ma) (9088 Ma) (ii) Transitional
(iii)
Candelaria (116110 Ma)
Punta del Cobre (116 Ma) Los Toros (98 Ma)
28S Structural setting based on
tectonostratigraphic evidence
Cortadera (87 Ma) (4)
Productora (129 Ma) Dos Amigos (108104 Ma) Extension
Los Loros (91 Ma)
La Verde (88 Ma) (5)
Trapiche (122120 Ma) Cachiyuyo (111 Ma) (7) Contraction
Las Campanas (90 Ma)
Totora (120 Ma)
30S Pajonales (117 Ma) ?
Frontera (112 Ma) Structural setting based on
Panulcillo (115 Ma) Andacollo (104 Ma) (i) (7) (v) (7)
(6) (iv) structural evidence
Extension
Espino (9386 Ma) Llahuin (92 Ma)
Transtension
32S (9)
(vii)
Argentina (8)
Contraction
(vi)
Pacific
Ocean Colliguay (129 Ma) Transpression
(10) (i) (viii) (viii) (11)
34S
72W 70W 68W 145 135 125 115 105 95 85
Age (Ma)
Fig. 1. Distribution, timing, and tectonic setting of Cretaceous porphyry Cu-Au and IOCG deposits in the Mesozoic Coastal
Cordillera of northern Chile. (a) Geographic distribution of deposits and spatial relationship to the Atacama fault system; ages
of deposits in Ma are shown in parentheses. (b) Temporal and tectonic separation of porphyry and IOCG deposits. Sources
of data are provided in Table 1 and Appendix 1.
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 297

Table 1. Cretaceous Porphyry and IOCG Deposits of the Coastal Cordillera of Northern Chile

Metal
Deposit/prospect association Resource Age (Ma) References

Porphyry
Inca de Oro Cu-Au-Ag 389 Mt at 0.39% Cu, 0.1 g/t Au 9088 (U-Pb Zr) Maksaev et al. (2010)
Los Toros Cu 98 Maksaev et al. (2007)
Dos Amigos Cu-Au 36 Mt at 0.36% Cu 108106 (U-Pb Zr) Almonacid (2007)
Totora Cu-Au 120121 (U-Pb Zr) Maksaev and Llaumet (2015)
La Union (Frontera) Cu-Au 50.5 Mt at 0.4% Cu, 0.2 g/t Au 112 (U-Pb Zr) Creixell et al. (2015);
Maksaev and Llaumet (2015)
Pajonales Cu-Au 117 (U-Pb Zr) Morelli (2008)
Punta Colorada Cu-Au 109 (U-Pb Zr) Creixell et al. (2015)
Cachiyuyo Cu-Au 111 (U-Pb Zr) Creixell et al. (2015)
La Verde Cu-Au 88 (U-Pb Zr) Creixell et al. (2015)
Elisa Cu-Au 92 (U-Pb Zr) Creixell et al. (2015)
Cortadera Cu-Au 87 (U-Pb Zr) Creixell et al. (2015)
Las Campanas Cu-Au 90 (U-Pb Zr) Creixell et al. (2015)
Los Loros Cu-Au 91 Maksaev et al. (2010)
Carmen de Andacollo Cu-Au-Mo Proven and probable reserves 104 (U-Pb Zr; Re-Os Mo) Maksaev et al. (2010); Teck
of 417 Mt at 0.34% Cu and Resources Limited, 2016;
0.12 g/t Au this study
Llahuin Cu-Au-Mo 145 Mt at 0.4% Cu equiv 92 (Ar-Ar Bt) Maksaev et al. (2010)
Colliguay Cu-Au 129 (K-Ar WR) Maksaev et al. (2010)

IOCG
Tropezon Cu-(Mo-Au) ~1.5 Mt Cu 110 (U-Pb Zr) Tornos et al. (2010)
Casualidad Cu-Au 400 Mt at 0.55% Cu 100 (U-Pb Zr), 10094 (Ar-Ar Bt, Kovacic (2014)
Kspar), 9484 (Ar-Ar Act)
Todos Los Santos Cu-Au 118 (Ar-Ar Act) Gelcich et al. (2005)
Santo Domingo Cu-Fe-Au 417 Mt at 0.25% Cu, 27% Fe, 124 (U-Pb Zr, Tn) Daroch et al. (2015)
0.032 Au g/t
Mantoverde Cu-Au 440 Mt at 0.56% Cu, 0.12 g/t Au 121117 (K-Ar Ser) Rieger et al. (2010)
Cerro Negro Norte Fe-(Cu-Au) 100 Mt at 65% Fe 116 (U-Pb Tn) Raab (2002)
Candelaria-Punta del Cobre Cu-Au-Ag 501 Mt at 0.54% Cu, 0.13 g/t Au 116110 (Ar-Ar Bt; Re-Os Mo) Arvalo et al. (2006);
Marschik and Fontbot (2001)
Productora1 Cu-Au-(Mo) 214.3 Mt at 0.48% Cu, 0.1 g/t Au, 128.9 (Re-Os Mo), 130 (U-Pb Zr) Marquardt et al. (2015)
138 ppm Mo
Trapiche Cu-Au 122120 (Ar-Ar Act) Creixell et al. (2015)
Panulcillo Cu-Au 0.5 Mt at 2.753% Cu, 0.51 g/t Au 115 (K-Ar Phl) Diaz and Corvaln (2015)
El Espino Cu-Au 145 Mt at 0.55% Cu, 0.22 g/t Au 9389 (U-Pb Zr); 8886 Del Real and Arriagada (2015);
(Ar-Ar Act, Ser, Kspar) Lpez et al. (2014)

Abbreviations: Act = actinolite, Bt = secondary biotite, Kspar = secondary potassium feldspar, Mo = molybdenite, Phl = phlogopite, Ser = sericite, Tn =
titanite, WR = whole rock, Zr = zircon
1Deposit type is transitional between porphyry and IOCG

tectonic setting as follows: early Cretaceous (extension), mid- origin, and that their contrasting styles of ore formation relate
Cretaceous (transtension), and late Cretaceous (contraction). to differences in the tectonic setting and chemistry of the
Because the tectonic setting changed diachronously from associated magmas. We combine new and published geochro-
north to south over intervals of 10 to 15 m.y., these terms do nological and geochemical data for igneous rocks associated
not correspond exactly to formal stratigraphic period subdivi- with Cretaceous porphyry and IOCG deposits in the Coastal
sions (epochs), and there is some temporal overlap between Cordillera with published structural information to show that
groups. porphyry and IOCG deposits formed in distinct tectono-
Evidence for a magmatic hydrothermal origin for several of magmatic settings. The bulk compositions of the associated
these Chilean IOCG deposits has been reported, including at igneous rocks are almost indistinguishable, but we present
Candelaria and Mantoverde (Marschik and Kendrick, 2015) data from analysis of igneous apatite to suggest that the por-
and Tropezon (Tornos et al., 2010, 2012), and several smaller phyry-forming magmas were S rich compared to their IOCG
IOCG deposits occur in close proximity to coeval intrusions counterparts, consistent with the difference in S content of
with alteration patterns apparently centered on those plutons porphyry (S rich) and IOCG (S poor) ore deposits.
(e.g., El Trapiche veins, Creixell et al., 2009; El Espino, Lpez
et al., 2014). In contrast, Barton and Johnson (1996, 2000) Mesozoic Tectonomagmatic Setting and
argue that the ore-forming fluids were basinal brines, albeit Metallogeny of Northern Chile
with convection driven by the heat from coeval magmatism. Arc magmatism related to subduction has taken place along the
In this paper, we adopt the hypothesis that the Chilean IOCG Chilean segment of the Gondwana supercontinental margin
and porphyry deposits are both of magmatic-hydrothermal since the late Paleozoic (Parada et al., 2007). Mid-Jurassic to
298 RICHARDS ET AL.

Cretaceous magmatism mostly developed ~100 km to the west of magnetite-apatite deposits (Gelcich et al., 2005; Creix-
of the Paleozoic arc (Parada et al., 2007). Extensional tecton- ell et al., 2009) and then IOCG deposits during the main
ics affected the arc from the mid-Jurassic to early Cretaceous, stage of rifting and back-arc basin development in the mid-
with intra-arc and back-arc basin formation occurring in the Cretaceous (~125~110 Ma, Aptian-Albian; Oyarzun et al.,
mid-Cretaceous, followed by tectonic inversion in the mid- to 1999). This tectonic relationship appears to be consistent with
late Cretaceous (Charrier et al., 2007). These tectonic changes IOCG deposits globally, which are commonly found to be
are interpreted to reflect alternating periods of subduction associated with extensional events, including postcollisional,
coupling and decoupling between the down-going and overrid- intracontinental, back-arc, and intra-arc rift settings (Williams
ing plates, with the generation of periods dominated by con- et al., 2005; Corriveau and Mumin, 2010; Skirrow, 2010).
tractional and extensional tectonics, respectively, in the upper Porphyry Cu-Au-(Mo) deposit formation returned in the late
plate (Scheuber and Gonzalez, 1999). Of particular relevance Cretaceous with the resumption of subduction-related mag-
to this study, early Cretaceous extension and mid-Cretaceous matism during and after tectonic inversion. Apparent excep-
transtension occurred during a period of low convergence rate tions to this three-part deposit distribution (e.g., the small,
and weak plate coupling that may reflect slab rollback. This was mid-Cretaceous Totora, Pajonales, and Cachiyuyo porphy-
followed by contraction in the mid- to late Cretaceous, caused ries, and the late Cretaceous Casualidad and El Espino IOCG
by an increase in convergence rate in response to global-scale deposits; Fig. 1) may reflect uncertainties in age determina-
plate reorganization (Matthews et al., 2012). tion, lithospheric heterogeneities, or delayed response to tec-
Early Cretaceous magmatism in the Coastal Cordillera tonic changes.
between 25 and 34S was generated in a broadly extensional
tectonic regime that caused rifting and crustal thinning (Parada Sampling
et al., 2007). It was characterized by relatively primitive calc- Fifty-four samples of volcanic and intrusive igneous rocks associ-
alkaline to shoshonitic lavas that built thick (510 km) subaerial ated with the Carmen de Andacollo, Frontera, and Dos Amigos
volcanic sequences (Vergara et al., 1995; Morata and Aguirre, porphyry Cu Au Mo deposits, the transitional Productora
2003; Parada et al., 2005; Charrier et al., 2007; Girardi, 2014). deposit, and the Candelaria, Casualidad, Espino, and Man-
By the mid-Cretaceous, the tectonic setting had become pre- toverde IOCG deposits from the Coastal Cordillera of north-
dominantly transtensional (Brown et al., 1993; Arvalo et al., ern Chile between 26 and 32S (Fig. 1) were collected from
2003) and was characterized by episodic volcanism and sedi- drill core and outcrops in July 2015. Where possible, intrusive
ment deposition in intra-arc and back-arc shallow marine and rocks thought to be directly related to mineralization (coeval,
continental basins (Fig. 1b; Morata and Aguirre, 2003). The cospatial) were collected; other samples were from broadly
transition from extensional to transtensional tectonics was coeval intrusions or volcanic rock outcrops in the vicinity of
slightly diachronous from north to south, and represents the the deposits (based on regional geologic maps). Least-altered
progression of continental arc rifting southward with time, samples were targeted for collection, but most of the rocks have
beginning in the north in the Barremian (~130 Ma) and reach- undergone either low-grade regional metamorphism (result-
ing its maximum in the Aptian-Albian (~120 Ma). Plutonic com- ing in minor chloritization of ferromagnesian silicate miner-
plexes were emplaced during both tectonic stages, controlled by als and partial saussuritization of plagioclase) or hydrothermal
crustal-scale fault zones that evolved into the Atacama fault sys- alteration due to proximity to the ore deposits. Two samples of
tem in the Valanginian-Barremian (144126 Ma; Brown et al., quartz-molybdenite veins were collected from the Carmen de
1993). The Atacama fault is an early dip-slip and later sinistral Andacollo porphyry Cu-Au deposit for Re-Os dating.
transtensional N-trending fault system that extends for more The rocks were studied petrographically to determine the
than ~1,000 km parallel to the continental margin (Fig. 1a) and degree of alteration and to seek unaltered accessory minerals
is interpreted to have formed in response to oblique subduction such as zircon and apatite for electron microprobe analysis.
(Scheuber and Andriessen, 1990; Brown et al., 1993; Palacios
et al., 1993); it is the primary structural control on the location Analytical Methods
of IOCG deposits in Chile (Sillitoe, 2003; Creixell et al., 2009).
Crustal extension ended in the Albian-Cenomanian (11095 Whole-rock geochemistry
Ma) with a return to contractional tectonics that produced Thirty-nine least-altered samples from the collected suite
crustal shortening and thickening, basin closure, rapid uplift, were submitted to Activation Laboratories Ltd. (Ancaster,
and a marked eastward shift of magmatism (Parada et al., 2002, Ontario) for analysis using the 4E-Research analytical pack-
2005; Arancibia, 2004; Maksaev et al., 2010). age, which combines instrumental neutron activation analysis
The spatiotemporal distribution of Cretaceous IOCG and and lithium metaborate/tetraborate fusion inductively coupled
porphyry Cu-Au deposits in northern Chile between 25 and plasma-mass spectrometry for determination of 62 elements.
34S is shown in relation to these tectonomagmatic periods Analyses of standards and duplicates indicate that accuracy
in Figure 1b. Here it can be seen that, although the deposits for major elements is typically within 5 relative %, and 10 rela-
roughly overlap spatially, IOCG deposits are broadly sepa- tive % for minor and trace elements. The ferrous iron (FeO)
rated in time from younger porphyry Cu-Au deposits at ~110 content of the rocks was also determined by titration.
to 100Ma, which also marks the time of the major tectonic
change from transtension to contraction. During the initial Electron microprobe analyses
stages of arc rifting in the Late Jurassic-early Cretaceous, Polished thin sections of all samples were examined for the
only a few small porphyries were developed (e.g., Antu- presence of igneous accessory minerals such as zircon and
coya, Colliguay). This was followed by the early formation apatite. Because of the relatively mafic nature of many of
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 299

the samples, few contained zircon in thin section, so zircon crystals, and standards were as follows: F K, PC0, topaz;
chemistry was not attempted. A larger number of samples Na K, TAP, tugtupite; Si K, TAP, topaz; P K, PET,
contained apatite, although care was needed to distinguish fluorapatite; S K, PET, anhydrite (intensity data aggregated
between igneous and hydrothermal (or late-stage) apatite. from two spectrometers; Donovan et al., 2011); Cl K, PET,
The latter was common in the groundmass of altered samples, tugtupite; and Ca K, PET, fluorapatite. The calculated limits
and clearly hydrothermal apatite (intergrown with hydrother- of detection (as element, rounded to the nearest 10 ppm) at
mal minerals such as quartz and sulfides) typically had thin, 99% confidence were as follows: F, 570 ppm; Na, 130 ppm;
elongated prismatic habits (200-m length; Fig. 2a). Apatite Si, 120ppm; P, 170 ppm; S, 80 ppm; Cl, 290 ppm; and Ca,
crystals with unequivocally igneous origin were distinguished 280 ppm. To check for interference from third-order P K
by their inclusion within phenocrysts (mainly plagioclase and on F K, the signal from synthetic GaP was examined with
biotite). The igneous apatite crystals had stubby prismatic the PC0 crystal: no significant interference from P on F was
habits (2050-m length; Fig. 2b-d). detected under the conditions of analysis.
Compositional data were acquired with a Cameca SX100 For sessions in which nine elements (F, Si, P, S, Cl, Ca,
electron microprobe using wavelength-dispersive spectros- Mn, Fe, and As) were measured, the following conditions
copy and Probe for EPMA software (Donovan et al., 2015). were used: 15-kV accelerating voltage, 20-nA beam cur-
For sessions in which seven elements (F, Na, Si, P, S, Cl, rent, and 5-m beam diameter. The X-ray lines, analyzing
and Ca) were measured, the following conditions were used: crystals, standards, and count times (seconds) on both peaks
10-kV accelerating voltage, 20-nA beam current, and 5- to and backgrounds were as follows: F K, PC0, topaz, 60 s;
10-m beam diameter. Total count times of 30 s were used Si K, TAP, topaz, 40 s; P K, PET, fluorapatite, 30 s; S K,
for both peaks and backgrounds. The X-ray lines, analyzing PET, anhydrite, 30 and 40 s (intensity data aggregated from

(a) CAN-2 Ap (b) CAS-2


0.01% 0.01%
sulfide
Plag Plag
Qz

Plag Qz
Amph
0.08%
0.07%
Qz Ap
sulfide Plag
Qz
0.02%
Qz Ap Amph
sulfide Plag

(c) CDA-8 (d) CDA-2


Qz
Biotite
Ap 0.45%
0.06% Biotite
Ap
0.41%
0.27%
Ap
Chl

0.15%
Chl

Qz

Fig. 2. Photomicrographs of apatite crystals in samples from (a) Candelaria (CAN-2), (b) Casualidad (CAS-2), and (c, d)
Carmen de Andacollo (CDA-8, CDA-2). Concentrations of SO3 in apatite crystals are shown in red (wt %); higher concentra-
tions are observed in apatite from porphyry-related samples (Carmen de Andacollo) compared to the IOCG-related samples
(Candelaria, Casualidad). Some apatite microphenocrysts from Carmen de Andacollo show zonation from SO3-rich cores to
SO3-poorer rims (d). Abbreviations: Amph = amphibole, Ap = apatite, Chl = chlorite, Plag = plagioclase, Qz = quartz.
300 RICHARDS ET AL.

measurements on two spectrometers; Donovan et al., 2011); agreement with the U-Pb date for magmatism of Maksaev et
Cl K, PET, tugtupite, 40 s; Ca K, PET, fluorapatite, 30 s; al. (2010).
Mn K, LIF, spessartine, 40 s; Fe K, LIF, spessartine, 40 s;
and As L, TAP, synthetic GaAs, 40 s. The calculated limits Cretaceous Igneous Geochemistry
of detection (as element, rounded to the nearest 10 ppm) at Whole-rock major and trace element geochemical data for 125
99% confidence were as follows: F, 460 ppm; Si, 110 ppm; igneous rocks coeval with either IOCG or porphyry deposits
P, 260 ppm; S, 70 ppm; Cl, 150 ppm; Ca, 170 ppm; Mn, from the Coastal Cordillera between 25 and 34S were com-
170ppm; Fe, 160ppm; and As, 180 ppm. piled from the literature and combined with our 40 new anal-
In all sessions, time-dependent intensity corrections for F yses (new data are listed in Table 3 and Supplementary Table
and Cl were carried out (peak count times divided into six S1, and data from the literature are listed in Supplementary
intervals) with Probe for EPMA software (Donovan et al., Table S2). These rocks are divided into three groups, based
2015), following Nielsen and Sigurdsson (1981), Stormer et on the tectonomagmatic periods described above: early Cre-
al. (1993), and Henderson (2011). Intensity data for all ele- taceous, related to a few small porphyry Cu-Au deposits and
ments were reduced following the methods of Armstrong early stages of arc rifting; mid-Cretaceous, related to IOCG
(1995). Oxygen was calculated by stoichiometry and included deposits and back-arc transtension; and late Cretaceous,
in the data reduction, as was the correction for oxygen equiva- related to porphyry Cu-Au-(Mo) deposits and a return to con-
lence of the halogens (F and Cl). tractional tectonics and arc magmatism. Note that, because
of diachronous changes in tectonic style from north to south
Re-Os dating over periods of 10 to 15 m.y., the terms early, mid-, and
A molybdenite mineral separate was made for each sample late Cretaceous do not correspond exactly to formal strati-
by metal-free crushing followed by gravity and magnetic con- graphic divisions, but are approximately separated at ~125
centration methods described in detail by Selby and Creaser and ~110 Ma, respectively.
(2004). The 187Re and 187Os concentrations in molybdenite Volcanic and plutonic rocks from the three tectonomag-
were determined by isotope dilution mass spectrometry using matic groups are mostly metaluminous, calc-alkaline to high-
Carius tube, solvent extraction, anion chromatography, and K calc-alkaline, and range in composition from andesite
negative thermal ionization mass spectrometry techniques. (diorite) to dacite (granodiorite). Samples of igneous rocks
A mixed double spike containing known amounts of isotopi- directly associated with porphyry and IOCG deposits show
cally enriched 185Re, 190Os, and 188Os was used (Markey et al., a similar compositional range on a total alkali-silica diagram
2007). A ThermoScientific Triton mass spectrometer with a (Fig. 3), although it is evident that the porphyry-related rocks
Faraday collector was used for isotopic analysis. Total blanks are mostly more felsic (diorite to granodiorite) compared to
for Re and Os are less than 3 and 2 pg, respectively, which are those associated with IOCG deposits (mostly gabbroic diorite
insignificant for the Re and Os concentrations in molybde- to diorite). The apparently alkaline composition of several of
nite. The molybdenite HLP-5 (Markey et al., 1998) was ana- these samples (especially from IOCG deposits) might be due
lyzed as a standard, and over a period of two years an average to variable degrees of hydrothermal alteration (sodic-calcic or
Re-Os date of 221.56 0.40 Ma (1 s.d. uncertainty, n = 10) potassic), which was unavoidable in these suites of ore-asso-
was obtained. This value is within the uncertainty of the 221.0 ciated rocks. However, on a plot of immobile element ratios
1.0 Ma age reported by Markey et al. (1998). (Zr/Ti versus Nb/Y; Fig. 4; Winchester and Floyd, 1977), a
small number of mid-Cretaceous and IOCG-related samples
Re-Os Age of the Carmen de Andacollo do plot in more alkaline fields (alkali gabbro and syenite), sug-
Porphyry Cu-Au Deposit gesting that some of these samples are genuinely alkaline in
The Carmen de Andacollo porphyry Cu-Au deposit has previ- composition. Nevertheless, the majority of the samples over-
ously been dated at 104 3.3 Ma by U-Pb analysis of zircons lap in the gabbro-diorite-granodiorite fields in Figure 4, with
in the ore-forming porphyry intrusions (Maksaev et al., 2010), no clear distinction between age groups or association with
and at 104 3 and 98 2 Ma by K-Ar analysis of phyllic- and porphyry and IOCG deposits.
potassic-altered rocks, respectively (Reyes, 1991). We have On primitive mantle-normalized extended trace element
dated two samples of molybdenite from B-type quartz veins in (Fig. 5) and chondrite-normalized REE (Fig. 6) diagrams,
the deposit, which yielded statistically indistinguishable ages the porphyry and IOCG suites display almost indistinguish-
of 103.9 0.5 and 103.6 0.5 Ma (2s errors; Table 2), in good able patterns: enrichments in large-ion lithophile elements

Table 2. Re-Os Molybdenite Age Data from the Carmen de Andacollo Porphyry Cu-Au Deposit

Total Model
Re 187Re 187Os common age 2 with
Sample no. Location Description (ppm) 2 (ppb) 2 (ppb) 2 Os (pg) (Ma)1 (Ma)2

CDA-4 DDH 11-41 Quartz-molybdenite vein 405.2 1.3 254,683 805 441.4 0.3 2.6 103.9 0.5
at 396.4 m in Porphyry D
CDA-12 DDH 13-09 Quartz-molybdenite vein 556.9 1.8 350,059 1,106 604.8 0.4 1.0 103.6 0.5
at 383.2 m in andesitic country rock

1Model age calculated from the equation t = ln (187Os/187Re + 1)/, where t = model age and = 187Re decay constant, and assuming no initial radiogenic Os
2 = 187Re decay constant, 1.666 1011 yr1 (Smoliar et al., 2006)
Table 3. Whole-Rock and Trace Element Geochemical Data for Igneous Rock Samples from the Coastal Cordillera of Northern Chile Between 25 and 34S
Sample no. ESP-1 ESP-2 ESP-3 CDA-1 CDA-2 CDA-3 CDA-7 CDA-9 CDA-10 FR-1
Frontera-Mine
Locality Espino Espino Espino Andacollo Andacollo Andacollo Andacollo Andacollo Andacollo La Union
Biotite, Less altered, gypsum- Quartz-
Chlorite, disseminated chalcopyrite- chalcopyrite
Alteration/ calcite veins Chlorite, chalcopyrite- Chlorite- iron oxide veinlets, veinlets, gypsum- Epidote, clays,
mineralization (skarn in wall rock) K-feldspar pyrite (sericite) 2 biotite iron oxide weak potassic
TAS Gabbroic Gabbroic Gabbroic Basaltic trachyandesite
classification1 diorite diorite diorite Monzonite Granodiorite Granodiorite (mugearite) Granodiorite Monzonite Granodiorite

SiO2 (wt %) 51.45 51.85 49.78 57.38 62.92 62.79 49.92 62.05 54.67 63.54
Al2O3 (wt %) 15.55 16.81 16.78 16.23 15.73 15.85 18.19 15.86 14.24 14.63
Fe2O3 (wt %) 0.98 0.69 1.35 2.33 0.97 0 5.83 0.56 1.07 3.79
FeO (wt %) 6.3 6.4 6.8 2.8 2.7 4.1 4 3.7 4.9 2.5
MnO (wt %) 0.119 0.168 0.167 0.13 0.074 0.048 0.152 0.095 0.118 0.173
MgO (wt %) 7.87 8 7.43 3.22 1.06 2.63 4.28 3.39 6.77 1.72
CaO (wt %) 8.06 9.43 7.15 5.03 4.75 4.49 3.78 5.27 5.74 3.77
Na2O (wt %) 3.44 3.11 2.98 4.05 2.68 3.46 4.86 4.34 2.83 4.11
K2O (wt %) 1.36 0.35 1.3 2.64 2.89 2.31 2.91 1.48 2.9 2.36
TiO2 (wt %) 0.647 0.693 0.711 0.558 0.375 0.489 0.929 0.517 0.651 0.368
P2O5 (wt %) 0.12 0.12 0.12 0.2 0.13 0.13 0.25 0.2 0.33 0.12
LOI (wt %) 2.33 2.25 5.59 5.37 5.19 3.31 4.54 2.76 4.29 1.31
LOI 2 (wt %) 1.63 1.53 4.82 5.06 4.89 2.85 4.1 2.35 3.74 1.03
Total (wt %) 98.94 100.6 100.9 100.3 99.78 100.1 100.1 100.6 99.06 98.66
Total 2 (wt %) 98.23 99.87 100.2 99.94 99.48 99.62 99.64 100.2 98.51 98.38
Fe2O3(T) (wt %) 7.99 7.8 8.91 5.45 3.97 4.56 10.28 4.67 6.52 6.57
Cs (ppm) 9.4 1.8 3.5 4.4 4 2.2 4.1 2 3.8 0.7
Tl (ppm) 0.27 <0.05 <0.05 0.23 0.24 0.29 0.08 0.2 0.66 0.1
Rb (ppm) 71 9 35 68 78 63 68 43 105 37
Ba (ppm) 153 209 293 508 611 683 638 217 270 626
Th (ppm) 0.52 0.49 0.48 2.9 5.76 5.21 2.44 6.07 4.96 3.13
U (ppm) 0.24 0.2 0.28 1.1 1.48 2.07 0.71 1.75 1.99 0.74
Nb (ppm) 1.8 1.7 1.7 2.7 2.9 2.6 2 2.8 3.3 2.5
Ta (ppm) 0.14 0.16 0.14 0.22 0.33 0.31 0.16 0.28 0.25 0.28
La (ppm) 7.46 6.11 7.7 15.5 18.8 16.3 16 15.9 23 11.8
Ce (ppm) 15.9 14.2 16.9 32.7 36 31.9 32.7 35.3 49.6 23.1
Pb (ppm) <5 <5 <5 6 <5 <5 18 <5 <5 10
Pr (ppm) 2.26 2.02 2.26 4.15 4.26 3.86 4.33 4.75 6.3 2.87
Sr (ppm) 489 499 365 621 234 584 738 507 358 348
Nd (ppm) 9.61 8.96 10.4 16.8 16.1 15.4 18 19.6 24.9 11.4
Zr (ppm) 58 60 60 92 91 86 66 113 115 69
Hf (ppm) 1.5 1.7 1.6 2.4 2.5 2.4 2 2.9 3 1.9
Sm (ppm) 2.74 2.31 2.58 3.55 3.16 3.09 4.25 4.06 5.26 2.54
Eu (ppm) 0.863 0.748 0.845 1.17 0.84 1.03 1.27 1.19 1.42 0.836
Sb (ppm) 1.7 0.8 1.6 1.7 1 <0.1 1.6 0.2 <0.1 2.6
Gd (ppm) 2.58 2.55 2.57 2.9 2.43 2.55 3.42 3.28 4 2.38
Tb (ppm) 0.46 0.44 0.43 0.41 0.35 0.38 0.54 0.44 0.56 0.39
Dy (ppm) 2.91 2.71 2.77 2.27 1.91 2.06 3.13 2.45 2.9 2.4
Y (ppm) 15 15 14 13 12 13 15 14 15 15
Ho (ppm) 0.58 0.56 0.55 0.44 0.36 0.39 0.59 0.46 0.54 0.5
Er (ppm) 1.65 1.73 1.67 1.26 1.08 1.09 1.73 1.29 1.49 1.5
Tm (ppm) 0.256 0.258 0.237 0.19 0.161 0.163 0.262 0.201 0.214 0.237
Yb (ppm) 1.74 1.68 1.66 1.2 1.02 1.07 1.74 1.33 1.4 1.7
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE

Lu (ppm) 0.29 0.277 0.263 0.183 0.163 0.17 0.285 0.21 0.223 0.28
Cr (ppm) 369 389 375 26.5 32.5 43.1 3.4 102 372 8.4
Ni (ppm) 130 140 123 17 11 16 13 38 128 3
Sc (ppm) 28.1 28.6 28.7 13.8 9.34 13.9 22.3 13.9 22.9 8.22
V (ppm) 197 208 207 145 100 141 335 133 181 83
Co (ppm) 31.5 28.6 32.1 15.4 8.1 12.3 26.3 13 23.5 10.7
Cu (ppm) 84 250 621 5 813 1,140 12 376 1,430 655
301
Table 3. (Cont.)
302
Sample no. PR-1 PR-2 PR-3 PR-4 PR-5 PR-6 PR-7 PR-8 CAN-1
Productora- Productora Productora Productora
Rancho Hill, Productora- Productora- Productora (iron district (iron district (iron district
Locality Cachiyuyito stock Alice porphyry Alice porphyry Productora (Ruta 5 batholith) regional) regional) regional) Candelaria
Weak disseminated
pyrite-chalcopyrite, Sodic
Alteration/ Weak albite- but otherwise (albite-chlorite) Moderately Weak albite-
mineralization actinolite fairly fresh replacing hornblende altered actinolite Minor epidote
TAS classification1 Granodiorite Granodiorite Granodiorite Basaltic andesite Granodiorite Diorite Granodiorite Diorite Monzonite

SiO2 (wt %) 62.33 66.72 62.29 51.69 64.1 60.49 65.25 57.03 58.93
Al2O3 (wt %) 14.57 15.72 14.26 14.93 16.38 16.41 15.17 17.23 17.87
Fe2O3 (wt %) 2.1 0.43 6.87 2.02 1.26 0.69 2.19 0.68 2.59
FeO (wt %) 2.5 2 4.3 4.3 3.3 2.2 3 3 2.7
MnO (wt %) 0.102 0.042 0.105 0.131 0.116 0.054 0.09 0.153 0.055
MgO (wt %) 2.44 1.8 1.53 9.41 2.02 3.87 1.73 4.31 1.89
CaO (wt %) 6.17 5.77 3.66 6.12 5.64 8.66 4.39 10.1 3.81
Na2O (wt %) 3.58 4.02 3.46 4.19 3.69 4.26 3.79 4.61 6.08
K2O (wt %) 2.97 0.75 1.12 0.5 1.63 0.41 2.29 0.58 2.62
TiO2 (wt %) 0.953 0.501 0.413 0.717 0.42 1.083 0.705 1.032 0.437
P2O5 (wt %) 0.19 0.09 0.1 0.1 0.13 0.22 0.14 0.19 0.23
LOI (wt %) 0.89 1.94 0.86 4.41 1.14 1.07 1.1 1.41 1.15
LOI 2 (wt %) 0.61 1.72 0.38 3.93 0.77 0.82 0.76 1.08 0.85
Total (wt %) 99.07 100 99.44 99.01 100.2 99.67 100.2 100.7 98.67
Total 2 (wt %) 98.79 99.8 98.96 98.53 99.82 99.42 99.83 100.3 98.37
Fe2O3(T) (wt %) 4.88 2.65 11.65 6.8 4.94 3.14 5.53 4.02 5.6
Cs (ppm) 0.6 0.8 1.1 0.4 1.3 0.2 0.3 0.5 0.3
Tl (ppm) <0.05 <0.05 <0.05 <0.05 0.09 <0.05 <0.05 <0.05 <0.05
Rb (ppm) 56 20 38 17 39 6 33 14 34
Ba (ppm) 643 118 130 41 432 128 590 150 843
Th (ppm) 12 3.19 3.94 1.63 3.61 7.13 4.44 2.8 3.33
U (ppm) 1.44 0.61 1.16 2.4 1.09 1.58 0.53 0.73 1.05
RICHARDS ET AL.

Nb (ppm) 6.8 2.3 5.5 1.2 3.8 4.4 4 3.3 3.1


Ta (ppm) 0.63 0.24 0.3 0.1 0.45 0.41 0.35 0.27 0.24
La (ppm) 10.9 6.9 12.5 10.8 14.6 12.2 11.8 4.22 18.3
Ce (ppm) 27.2 15.7 25.5 21.4 30.8 33.4 29.8 12.4 36.8
Pb (ppm) <5 <5 <5 <5 <5 <5 <5 <5 <5
Pr (ppm) 3.88 2.24 3.14 2.37 3.72 4.6 4.37 2.06 4.59
Sr (ppm) 285 313 184 144 384 364 228 330 426
Nd (ppm) 18.2 9.13 12.5 8.97 14.3 19.5 20.4 10.5 18.9
Zr (ppm) 189 89 95 61 113 171 181 133 127
Hf (ppm) 4.8 2.2 2.4 1.7 2.8 4.2 4.6 3.5 3
Sm (ppm) 4.96 2.06 2.68 2.04 2.99 4.83 5.24 3.39 3.7
Eu (ppm) 1.2 0.794 0.896 0.628 0.918 1.41 1.24 1.12 1.06
Sb (ppm) <0.1 0.3 0.5 0.5 0.5 0.5 <0.1 0.2 0.7
Gd (ppm) 5.29 2.17 2.6 2.42 2.58 4.64 5.54 4.28 3.28
Tb (ppm) 0.88 0.38 0.43 0.44 0.41 0.76 0.97 0.81 0.49
Dy (ppm) 5.5 2.37 2.67 2.82 2.48 4.67 6.09 5.27 2.91
Y (ppm) 29 18 12 17 16 28 36 31 19
Ho (ppm) 1.12 0.49 0.56 0.61 0.5 0.95 1.26 1.08 0.58
Er (ppm) 3.2 1.51 1.75 1.9 1.5 2.84 3.62 3.16 1.76
Tm (ppm) 0.473 0.236 0.27 0.302 0.242 0.433 0.534 0.46 0.279
Yb (ppm) 3.12 1.67 1.86 2.19 1.68 3.05 3.59 3.11 1.91
Lu (ppm) 0.506 0.284 0.307 0.371 0.285 0.502 0.576 0.502 0.316
Cr (ppm) 47.1 25.7 230 434 34.8 86.8 26.5 31.3 9.1
Ni (ppm) 13 5 17 96 9 15 5 7 6
Sc (ppm) 20 13.8 11.7 34.5 9.7 27.5 19.9 31.2 9.25
V (ppm) 126 94 82 251 77 196 110 190 99
Co (ppm) 11.4 6.9 18.4 8.9 8.2 11.3 12.9 6.5 7.7
Cu (ppm) 25 194 1,300 7 11 5 19 15 42
Table 3. (Cont.)
Sample no. CAN-2 CAN-3 CAN-4 CAS-1 CAS-2 CAS-3 CAS-4 CAS-5 CAS-8 CAS-9
Candelaria-
Coquimbana
Locality Candelaria Candelaria mine Casualidad Casualidad Casualidad Casualidad Casualidad Casualidad Casualidad
Albitic alteration Potassic + Propylitic
Alteration/ + hematite-Cu Weak Weak propylitic chalcopyrite (chlorite-epidote)
mineralization veins nearby propylitic (chlorite-epidote) veinlets on potassic
TAS Gabbroic Gabbroic Foid
classification1 Monzonite diorite diorite Monzodiorite Diorite Monzodiorite Granodiorite monzosyenite Monzodiorite Diorite

SiO2 (wt %) 58.34 52.88 52.8 52.87 57.82 51.26 65.03 51.48 54.19 56.66
Al2O3 (wt %) 18.38 17.53 16.58 15.12 17.99 16.96 16.7 15.99 18.14 16.93
Fe2O3 (wt %) 3.01 5.86 3.55 5.83 2.03 6.5 0.19 4.52 0.96 2.62
FeO (wt %) 3.4 3.9 6 5.9 3.3 4.9 3.5 7.3 2.1 3.3
MnO (wt %) 0.106 0.051 0.101 0.146 0.153 0.063 0.044 0.085 0.092 0.126
MgO (wt %) 1.68 4.13 4.45 5.55 3.16 5.83 1.54 5.38 3.5 4.01
CaO (wt %) 5.67 8.04 8.46 4.12 6 1.84 4.77 1.22 10.46 8.02
Na2O (wt %) 4.27 4.05 3.85 3.94 4.26 6.13 4.03 2.08 4.25 4.12
K2O (wt %) 3.66 1.1 1.17 2.29 1.15 0.71 1.25 8.12 1.49 1.19
TiO2 (wt %) 0.84 1.104 1.123 0.822 0.478 0.783 0.351 0.768 0.854 0.746
P2O5 (wt %) 0.45 0.45 0.42 0.18 0.12 0.15 0.13 0.17 0.32 0.14
LOI (wt %) 0.76 1.07 0.62 2.29 1.98 3.35 1.2 1.5 2.43 2.09
LOI 2 (wt %) 0.38 0.63 -0.06 1.63 1.61 2.8 0.81 0.68 2.19 1.72
Total (wt %) 100.9 100.6 99.8 99.72 98.8 99.02 99.13 99.42 99.03 100.3
Total 2 (wt %) 100.6 100.2 99.12 99.06 98.43 98.47 98.74 98.61 98.8 99.95
Fe2O3(T) (wt %) 6.79 10.19 10.22 12.39 5.7 11.95 4.08 12.64 3.29 6.29
Cs (ppm) 1.7 0.6 0.7 1.5 1.4 0.6 0.7 2.8 0.3 0.4
Tl (ppm) 0.05 <0.05 <0.05 0.08 <0.05 <0.05 0.06 0.33 <0.05 0.05
Rb (ppm) 90 28 31 48 26 16 26 178 44 31
Ba (ppm) 776 263 346 269 352 67 341 3,137 134 147
Th (ppm) 7.47 4.95 4.76 4.53 0.59 0.83 2.31 2.47 4.06 2.39
U (ppm) 2.2 1.41 1.65 1.23 0.23 0.28 0.36 1.31 0.59 0.78
Nb (ppm) 7.4 4 4.1 3.2 1.7 1.6 3.1 2.8 5.5 2.5
Ta (ppm) 0.57 0.29 0.3 0.27 0.16 0.13 0.41 0.2 0.31 0.23
La (ppm) 33 25.5 22.4 15.5 7.71 7.89 9.72 9.88 6.37 12.6
Ce (ppm) 72.6 57.8 51 36.8 16.9 18 19.2 23.3 28 30.2
Pb (ppm) 6 <5 <5 <5 <5 <5 <5 <5 <5 <5
Pr (ppm) 9.47 7.71 6.99 4.94 2.17 2.6 2.16 3.15 5.24 4.09
Sr (ppm) 500 561 503 236 713 77 460 209 409 409
Nd (ppm) 39.5 33 30.2 20 9.29 12.3 7.83 13.8 26.9 16.6
Zr (ppm) 100 124 130 135 75 66 77 125 96 116
Hf (ppm) 3 3.3 3.5 3.7 1.9 2 2.2 3.3 2.8 3.1
Sm (ppm) 8.81 7.4 7.06 4.6 2.27 3.17 1.28 3.78 8.07 3.92
Eu (ppm) 2.05 1.69 1.86 1.18 0.852 0.923 0.628 0.953 1.66 1.16
Sb (ppm) 0.6 0.7 0.8 0.7 0.6 0.5 <0.1 0.2 0.6 <0.1
Gd (ppm) 6.78 6.38 6.2 4.25 2.37 3.06 1.07 3.72 7.93 3.39
Tb (ppm) 1.03 0.98 0.94 0.63 0.37 0.48 0.15 0.67 1.28 0.51
Dy (ppm) 5.63 5.52 5.32 3.88 2.33 2.98 0.89 4.14 7.79 3.07
Y (ppm) 30 27 26 19 14 14 7 20 40 17
Ho (ppm) 1.08 1.06 1.06 0.79 0.46 0.6 0.17 0.84 1.46 0.59
Er (ppm) 3.06 3.08 2.89 2.24 1.38 1.69 0.46 2.39 4.14 1.86
Tm (ppm) 0.441 0.452 0.411 0.326 0.224 0.249 0.068 0.352 0.645 0.281
Yb (ppm) 2.96 2.91 2.77 2.26 1.59 1.67 0.45 2.44 4.08 1.92
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE

Lu (ppm) 0.48 0.449 0.467 0.362 0.259 0.262 0.073 0.39 0.569 0.308
Cr (ppm) 16.1 32.7 26 183 26.2 76.8 35.7 57 26.6 39.2
Ni (ppm) 6 14 12 42 14 32 5 33 12 15
Sc (ppm) 15.2 29.4 33 25.4 13 29.1 6.84 22.3 23.7 24.4
V (ppm) 127 316 315 201 117 233 59 83 171 213
Co (ppm) 14.6 21.3 22.5 28.2 15.3 32.7 20.2 37.9 8.9 17.8
Cu (ppm) 154 21 125 886 90 16 1,130 452 76 18
303
Table 3. (Cont.)
304
Sample no. CAS-10 CAS-11 MV-1 MV-5 MV-10 MV-11 MV-12 DA-1 DA-2 DA-3
Mantoverde- Mantoverde- Dos Amigos- Dos Amigos-
Locality Casualidad Casualidad Laura Laura Mantoverde Mantoverde Mantoverde Dos Amigos Tricolor Tricolor
2 K-feldspar Potassic Potassic (K-feldspar, Potassic (some Potassic + minor Potassic +
Alteration/ Weak Weak propylitic with chlorite (K-feldspar, biotite, chlorite) + Potassic (some weathering to chalcopyrite- chalcopyrite-
mineralization chlorite (chlorite-epidote) overprint biotite, epidote) minor chalcopyrite Potassic weathering) clay on fractures) pyrite pyrite
TAS Basaltic trachy- Basaltic trachy-
classification1 Granodiorite Diorite Monzodiorite Monzodiorite Syenite andesite (shoshonite) andesite (shoshonite) Granodiorite Granodiorite Granodiorite
SiO2 (wt %) 61.64 59.77 46.54 50.79 57.39 51.21 49.41 64.7 65.92 66.5
Al2O3 (wt %) 14.94 15.46 15.14 16.8 16.64 14.75 16.08 13.91 14.88 14.69
Fe2O3 (wt %) 2.22 2.12 1.29 2.92 2.37 8.54 6.3 4.74 2.18 2.15
FeO (wt %) 3.7 4.4 4.7 5.9 4.1 2.9 2.9 4.3 3.2 4
MnO (wt %) 0.096 0.146 0.647 0.279 0.061 0.164 0.411 0.065 0.109 0.124
MgO (wt %) 2.44 2.94 4.65 5.25 3.06 3.23 3.27 1.76 1.55 1.5
CaO (wt %) 4.68 5.54 9.42 6.67 0.6 4.65 5.2 1.7 3.02 3.04
Na2O (wt %) 3.14 2.96 3.2 3.74 0.17 1.18 3.7 3.88 3.9 3.57
K2O (wt %) 2.97 2.76 2.17 1.85 9.81 4.05 4.13 1.83 1.48 1.94
TiO2 (wt %) 0.819 0.801 0.678 0.787 1.196 1.001 1.023 0.401 0.368 0.351
P2O5 (wt %) 0.19 0.19 0.1 0.13 0.04 0.18 0.24 0.09 0.12 0.16
LOI (wt %) 1.42 1.33 9.96 2.99 2.91 6.76 5.86 1.19 1.41 1.39
LOI 2 (wt %) 1 0.84 9.43 2.33 2.45 6.43 5.54 0.71 1.05 0.94
Total (wt %) 98.67 98.9 99.03 98.76 98.81 98.94 98.86 99.05 98.51 99.86
Total 2 (wt %) 98.26 98.41 98.51 98.1 98.36 98.62 98.53 98.57 98.15 99.41
Fe2O3(T) (wt %) 6.34 7.01 6.52 9.48 6.93 11.77 9.52 9.52 5.74 6.6
Cs (ppm) 1.8 1.7 1.8 0.5 0.7 1 1.2 2.1 0.6 0.5
Tl (ppm) 0.11 0.2 <0.05 <0.05 0.12 <0.05 0.13 0.33 0.1 0.09
Rb (ppm) 97 101 37 36 162 93 64 40 24 30
Ba (ppm) 427 368 2,230 801 1,440 377 2,753 303 654 615
Th (ppm) 13.3 11.6 2.33 3.25 3.14 2.58 4.16 2.66 3.97 3.94
U (ppm) 3.62 3.38 0.65 0.93 1.08 0.92 1.22 0.3 0.98 0.85
Nb (ppm) 5.9 5.2 2 2.9 7.1 4 5.5 3.6 3.3 3.3
Ta (ppm) 0.56 0.44 0.16 0.22 0.54 0.3 0.38 0.3 0.36 0.35
RICHARDS ET AL.

La (ppm) 22.1 20 6.32 12.9 3.43 6.87 14.7 5.81 13.4 10.8
Ce (ppm) 50.4 45.4 16.1 26.5 6.76 16.4 34 12.3 26.9 21.6
Pb (ppm) <5 8 <5 7 <5 <5 <5 7 <5 <5
Pr (ppm) 6.53 5.92 2.34 3.57 0.81 2.16 4.53 1.68 3.13 2.64
Sr (ppm) 297 323 417 490 63 50 224 269 329 321
Nd (ppm) 26.9 24.2 10.7 14.5 3.55 9.89 19.8 7.18 12.7 10.3
Zr (ppm) 259 203 62 81 108 132 124 62 69 64
Hf (ppm) 6.8 5.6 1.8 2.2 3.1 3.4 3.2 1.8 2.1 1.9
Sm (ppm) 6.45 5.64 2.69 3.44 0.87 2.4 4.7 1.8 2.75 2.31
Eu (ppm) 1.09 1.16 0.889 1.07 0.229 0.636 1.46 0.567 0.868 0.753
Sb (ppm) 1.4 1.8 5.3 1.3 0.8 1.9 1.3 0.5 0.7 1
Gd (ppm) 5.57 5.26 2.57 3.36 0.95 2.82 4.51 1.83 2.21 2.03
Tb (ppm) 0.92 0.87 0.43 0.52 0.19 0.51 0.71 0.31 0.38 0.34
Dy (ppm) 5.57 5.2 2.57 3.03 1.25 3.52 4.38 1.92 2.46 2.14
Y (ppm) 29 27 13 14 8 18 21 11 16 14
Ho (ppm) 1.12 1.05 0.5 0.6 0.27 0.8 0.86 0.4 0.52 0.45
Er (ppm) 3.24 3.04 1.34 1.68 0.88 2.51 2.43 1.2 1.61 1.37
Tm (ppm) 0.477 0.455 0.192 0.247 0.144 0.385 0.365 0.192 0.261 0.218
Yb (ppm) 3.23 3 1.27 1.64 1 2.61 2.44 1.35 1.85 1.55
Lu (ppm) 0.516 0.473 0.201 0.256 0.168 0.415 0.375 0.229 0.31 0.268
Cr (ppm) 32.1 43.6 63.3 54.3 36.3 94.7 59.7 34.4 13.4 23
Ni (ppm) 8 11 22 26 24 36 15 9 3 3
Sc (ppm) 20 23.4 27.8 27.7 22.7 24.6 27.9 8.78 7.32 7.01
V (ppm) 147 169 198 228 143 209 261 92 74 68
Co (ppm) 14.8 20.5 25.7 28.4 29 15.4 22.3 5 7.5 10.3
Cu (ppm) 111 122 12 122 214 14 230 3,190 580 1,580
LOI = loss on ignition
1Classification
schemes of Le Maitre et al. (2002) for extrusive rocks and Middlemost (1994) for plutonic rocks
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 305

Fig. 3. Total alkali-silica diagram (Le Maitre et al., 2002) showing the compositions of Cretaceous igneous rocks associated
with porphyry (blue symbols) and IOCG deposits (red symbols) in the Coastal Cordillera of northern Chile; Productora
(purple symbol) shows characteristics of both porphyry and IOCG deposits. Data are from Table 3, excluding the following
samples for reasons of extreme alteration or lack of correlation with the main suites: CAS-4 (clasts in igneous breccia); CAS-8,
9, 10, 11 (regional plutons, relationship to Casualidad deposit unclear); CDA-7 (Quebrada Marquesa Formation, andesitic
country rock); MV-10 (strong potassic alteration); MV-11, MV-12 (altered volcanic rock outcrops, relationship to Mantoverde
deposit unclear); PR-1 (Cachiyuyito stock, possibly unrelated). All of the remaining samples shown are nevertheless altered
to varying degrees, and we attribute the apparently alkaline compositions of some rocks to this effect.

1
early Cretaceous
mid-Cretaceous
late Cretaceous
IOCG
Porphyry
Rhyolite/
Granite
Trachyte
0.1
Rhyodacite, dacite/
Granodiorite
Zr/Ti

Trachyandesite/
Andesite/ Syenite
Diorite

0.01
Andesite, basalt/ Alkali basalt/
Diorite, gabbro Alkali gabbro

Sub-alkaline basalt/
Sub-alkaline gabbro

0.001
0.01 0.1 1 10
Nb/Y
Fig. 4. Zr/Ti vs. Nb/Y discrimination diagram (Winchester and Floyd, 1977) for Cretaceous igneous rocks from the Coastal
Cordillera of Chile between 25 and 34S. Regional igneous geochemical data from the literature are grouped according to
the three temporal-tectonomagmatic groups, as defined in the text. Igneous rocks related to mid-Cretaceous IOCG and late
Cretaceous porphyry deposits collected during this study are plotted for comparison. Regional Cretaceous data from Irwin
et al. (1988), Vergara et al. (1995), Cisternas et al. (1999), Parada et al. (1999, 2002), Grocott and Taylor (2002), Morata and
Aguirre (2003), Creixell (2007), Hasler (2007), and Lpez et al. (2014).
306 RICHARDS ET AL.

Fig. 5. Primitive mantle-normalized extended trace element diagrams (normalization values of Sun and McDonough, 1989)
for selected igneous rocks associated with (a) porphyry and (b) IOCG deposits in the Coastal Cordillera of northern Chile.

(LILEs: Rb, Ba, Th, U, K) and light rare earth elements of subduction-related igneous rocks and reflect enrichments
(LREEs); negative anomalies for Nb, Ta, and Ti; relative in fluid-mobile LILEs, retention of Nb, Ta, and Ti in insoluble
depletions in compatible elements and middle to heavy rare Fe-Ti oxides, and fractionation of amphibole (which preferen-
earth elements (MREEs, HREEs); and flat to listric-shaped tially partitions MREEs; Gill, 1981; Green and Pearson, 1985;
patterns from MREEs to HREEs. Such patterns are typical Klein et al., 1997). The only noticeable difference between
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 307

Fig. 6. C1 chondrite-normalized extended trace element diagrams (normalization values of Sun and McDonough, 1989)
for selected igneous rocks associated with (a) porphyry and (b) IOCG deposits in the Coastal Cordillera of northern Chile.

these suites is the slightly greater depletion of MREEs- In an attempt to assess the relative oxidation states of the
HREEs in the porphyry-related suite, which can be attributed various suites of rocks, we have assessed whole-rock Fe2O3/
to the more felsic (fractionated) nature of these rocks. This FeO ratios where reported. While many of these samples are
characteristic is also observed in the ratios of Sr/Y and La/Yb, altered to varying degrees, which will clearly affect the Fe2O3/
which are generally more elevated in the felsic porphyry suite FeO ratio, most of the samples have values ranging from 0.18
compared to the more mafic IOCG suite (Fig. 7). to 5.10, corresponding to moderately or strongly oxidized
308 RICHARDS ET AL.

Fig. 7. (a) Sr/Y versus Y and (b) La/Yb versus Yb plots for selected igneous rocks associated with porphyry and IOCG deposits
from the Coastal Cordillera of northern Chile. The more adakite like compositions of many of the porphyry rocks compared
to many of the IOCG-related rocks likely reflect the more felsic compositions and greater degrees of fractionation (especially
of amphibole) of the former. Note that, with only one exception, all of these rocks have lower La/Yb ratios than adakites.
Fields for adakite-like rocks from Richards and Kerrich (2007); regional Cretaceous data from Irwin et al. (1988), Vergara et
al. (1995), Cisternas et al. (1999), Parada et al. (1999, 2002), Grocott and Taylor (2002), Morata and Aguirre (2003), Creixell
(2007), Hasler (2007), and Lpez et al. (2014).

rocks (using the criteria of Blevin, 2004). Importantly, there Published Sr-Nd isotope compositions for 53 Cretaceous
is no clear difference in oxidation state between the three igneous rocks from the Coastal Cordillera between 25 and
temporal-tectonomagmatic groups and porphyry- and IOCG- 34S (Supplementary Table S3) are plotted in Figure 8.
related rocks, which all appear to be similarly oxidized. Early Cretaceous rocks have evolved isotopic compositions
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 309

10
Depleted MORB mantle
early Cretaceous
mid-Cretaceous
late Cretaceous

Nd 5

Jurassic igneous rocks

0
Crustal contamination trend

Late Paleozoic
intrusive rocks

-5
0.702 0.703 0.704 0.705 0.706 0.707 0.708 0.709 0.710

(87Sr/86Sr)i
Fig. 8. Nd vs. initial 87Sr/86Sr ratios for Cretaceous igneous rocks from the Coastal Cordillera of Chile between 25 and
34S. Mid-Cretaceous rocks associated with IOCG deposits mostly fall at the primitive end of a range of data reflecting
various degrees of crustal contamination of mantle-derived magmas, whereas early and late Cretaceous rocks associated
with porphyry deposits range to significantly more isotopically evolved compositions. Data from Parada et al. (2005), Hasler
(2007), Morata et al. (2008), and Girardi (2014). Depleted MORB mantle field at ~110 Ma from Pilet et al. (2011). Jurassic
igneous rock field from Parada et al. (1999), Creixell (2007), and Girardi (2014). Paleozoic intrusive rock field from Parada
et al. (1999).

suggesting extensive crustal contamination. In contrast, the However, the apatite-melt partition coefficient formula of
mid- and later Cretaceous rocks cluster at relatively primitive Peng et al. (1997), which is derived for relatively oxidized
compositions, albeit still with some crustal contamination. arc magmas, can be used to obtain a semiquantitative esti-
mate of magmatic S content. We have estimated the apatite
Apatite Compositions saturation temperature of four samples from the Carmen de
Electron microprobe analyses of magmatic and hydrothermal Andacollo porphyry, the Candelaria and Casualidad IOCG
apatite from three samples from the Carmen de Andacollo deposits, and a sample of regional mid-Cretaceous igneous
porphyry and seven samples from or near IOCG deposits rock near Productora (using the equation of Piccoli and Can-
(Productora, Candelaria, and Casualidad) are listed in Supple- dela, 1994, 2002, which is derived from the data of Harrison
mentary Table S4, and SO3 analyses of magmatic apatite are and Watson, 1984) and used this temperature in the equa-
summarized in Table 4 (along with data from the literature). tions of Peng et al. (1997) and Parat et al. (2011) to derive
The results show a clear difference between igneous apatite estimates of magmatic S content. The data reported in Table
from rocks associated with the Carmen de Andacollo porphyry 5 suggest that the average S content of the magma associ-
deposit (0.25 0.17 wt % SO3, n = 69) compared with those ated with the Carmen de Andacollo porphyry deposit was
from IOCG deposits (0.04 0.02 wt % SO3, n = 76). A simi- ~0.02 wt %S (up to 0.06 wt % S)significantly higher than
lar relationship is observed for apatite from porphyry deposits the average values for magma associated with IOCG depos-
(0.120.60 wt % SO3) and IOCG deposits (0.070.13 wt % its in the region (0.0010.005 wt % S; Table 5). An alterna-
SO3) reported in the literature (Table 4). Hydrothermal or tive method for calculating magmatic sulfur content from
late-stage igneous apatite (e.g., edges of apatite micropheno- apatite SO3 compositions is provided by Parat et al. (2011),
crysts; Fig. 2d) showed lower and more variable SO3 contents and these values are also listed in Table 5. The results differ
(Supplementary Table S4), consistent with observations else- somewhat in absolute values compared to the results using
where in the literature (Streck and Dilles, 1998; Van Hoose the Peng et al. (1997) formula, but not in the relative enrich-
et al., 2013). ment in S of the porphyry-related magmas compared to the
The SO3 content of apatite varies as a complex function IOCG-related magmas.
of magmatic temperature, oxidation state, and sulfur fugac- In order to eliminate the possibility that contrasting mag-
ity (Peng et al., 1997; Parat and Holtz, 2005; Parat et al., matic oxidation state was responsible for this difference in
2011), and an accurate calculation of magmatic sulfur con- apatite SO3 content, we have estimated magmatic fO2 values
tent from apatite SO3 compositions is not currently possible. for these samples from the average MnO contents of apatite
310 RICHARDS ET AL.

Table 4. Summary of SO3 Compositions of Igneous Apatite from Chilean and Other Porphyry and IOCG Deposits

SO3 (wt %)

Locality Deposit type Minimum Maximum Average s.d. n Source

Carmen de Andacollo Porphyry Cu-Au 0.05 0.81 0.25 0.17 69 This study
Productora Transitional IOCG 0.02 0.07 0.04 0.02 14 This study
Candelaria IOCG 0.02 0.09 0.03 0.02 32 This study
Casualidad IOCG 0.02 0.08 0.03 0.02 30 This study
Yerington Porphyry Cu 0.26 0.99 0.60 0.36 4 Streck and Dilles (1998)
Santo Tomas II (Philex) Porphyry Cu n.a. 0.45 0.30 0.07 n.a. Imai (2002)
Santo Tomas II (Philex) Porphyry Cu n.a. 0.62 0.25 0.07 n.a. Imai (2002)
Clifton Porphyry Cu n.a. 0.76 0.29 0.14 n.a. Imai (2002)
Clifton Porphyry Cu n.a. 0.48 0.27 0.08 n.a. Imai (2002)
Bumolo (Waterhole) Porphyry Cu n.a. 0.59 0.21 0.15 n.a. Imai (2002)
Camp 6 (Black Mountain) Porphyry Cu n.a. 0.48 0.12 0.09 n.a. Imai (2002)
Dizon Porphyry Cu n.a. 0.57 0.17 0.08 n.a. Imai (2002)
Taysan Porphyry Cu n.a. 0.55 0.20 0.11 n.a. Imai (2002)
Duolong Porphyry Cu-Au 0.44 0.82 n.a. n.a. n.a. Li et al. (2012)
Mt. Isa IOCG n.a. n.a. 0.13 0.09 n.a. Piccoli and Candela (2002)
Mt. Isa IOCG n.a. n.a. 0.07 n.a. n.a. Belousova et al. (2002)
Tjrrojkka IOCG 0.02 0.13 0.08 0.03 n.a. Edfelt (2007)
Se-Chahun IOCG (IOA) <0.03 0.11 n.a. n.a. n.a. Bonyadi et al. (2011)

n.a. = data not available, s.d. = standard deviation

using the equation of Miles et al. (2014, Table 5). These data We interpret these data to indicate that there was little fun-
confirm that all of the rocks are moderately oxidized (DFMQ damental change in the source of magmatism throughout the
0.42.7), although we note that this fO2 calculation relies on Cretaceous, except for a hint of more mafic alkaline magma-
electron microprobe analyses of Mn in apatite at concentra- tism during the mid-Cretaceous transtensional rifting phase.
tions close to the detection limit, and so cannot be considered Published Sr and Nd isotope compositions of Cretaceous rocks
highly accurate. There is also debate about the applicability of from the Coastal Cordillera indicate mixing between depleted,
this method to rocks of different magmatic temperature and mantle-derived magmas and isotopically evolved continen-
composition (Marks et al., 2016; Miles et al., 2016), although tal crust, with early Cretaceous magmas showing the highest
the samples compared here are of broadly similar composi- degrees of contamination, likely by late Paleozoic lower and
tion and likely temperature. Consequently, we consider that mid-crustal rocks (Fig. 8, Damm et al., 1990; Lucassen et al.,
the formula of Peng et al. (1997), derived for oxidized arc 1999, 2001, 2002). Although the isotopic ranges of mid- and
rocks, should be approximately valid, and the calculated abun- late Cretaceous rocks overlap, there is a significant clustering
dances of magmatic S should be relatively correct, even if the of mid-Cretaceous data at the most primitive end of the range
absolute values are questionable (cf. Streck and Dilles, 1998). (eNd 6, 87Sr/86Sri 0.7033). We suggest that back-arc exten-
These data support the hypothesis that the porphyry-forming sional tectonics in the mid-Cretaceous may have facilitated
magmas were richer in S than the IOCG-related magmas. the greater involvement of primitive, mildly alkaline magmas
derived from partial melting of upwelling mantle astheno-
Discussion sphere during crustal extension and thinning (Fig. 9). Never-
theless, a significant amount of crustal processing is evident
Igneous geochemistry for all of these Cretaceous magmas, as is commonly observed
Whole-rock geochemical compositions of Cretaceous igneous in continental arc rocks (Hildreth and Moorbath, 1988). The
rocks from the Coastal Cordillera of Chile between 25 and more felsic nature and higher Sr/Y and La/Yb ratios of the later
34S show minimal differences beyond those expected from Cretaceous porphyry-associated magmas (Fig. 7) may reflect
normal fractionation processes (Figs. 35). Rocks associated longer residence times in deep crustal magma chambers under
with early Cretaceous arc rifting (with minor porphyry Cu-Au contractional tectonic conditions (Hildreth and Moorbath,
deposits), mid-Cretaceous back-arc transtension (with IOCG 1988; Haschke et al., 2002; Richards and Kerrich, 2007), with
deposits), and later Cretaceous contractional arc magmatism more extensive fractionation of amphibole (Lang and Titley,
(with porphyry Cu-Au-(Mo) deposits) have compositions typi- 1998; Richards and Kerrich, 2007; Richards, 2011a).
cal of subduction-related magmatism. On average, however, We co nclude that it is not possible uniquely to distinguish
later Cretaceous arc rocks are more felsic (Fig. 3), with pro- between igneous rocks associated with porphyry and IOCG
nounced listric-shaped REE patterns (Fig. 6) and higher Sr/Y deposits in the Coastal Cordillera of Chile using their whole-
and La/Yb ratios compared to mid-Cretaceous rocks associ- rock geochemical or isotopic compositions alone, although the
ated with IOCG deposits (Fig. 7). There is also a suggestion IOCG-associated rocks are, on average, more mafic (locally
that some of the mid-Cretaceous magmas and those associ- alkaline) and isotopically slightly more primitive than the por-
ated with IOCG deposits may have had slightly more mafic, phyry-related rocks, consistent with their back-arc as opposed
alkaline compositions (alkali gabbros to syenites). to main arc tectonomagmatic setting.

Table 5. Estimates of Magmatic Temperature and Oxidation State from Igneous Apatite and Whole-Rock Compositions

Apatite SO3
Whole- Whole- Apatite (wt %) (n)3 Average Average Maximum Maximum
Deposit Sample rock SiO2 rock P2O5 MnO Estimated Estimated [maximum, magmatic S magmatic S magmatic S magmatic S
(type) no. (wt %)1 (wt %)1 AST2 (wt %) (n)3 log fO24 DFMQ5 minimum] content (wt %)6 content (wt %)7 content (wt %)6 content (wt %)7

Carmen de CDA-2 66.74 0.14 905C n.a. n.a. n.a. 0.23 0.18 (34) 0.0160 0.0088 0.0571 0.1308
Andacollo [0.05, 0.81] 0.0120 0.0230
(porphyry) CDA-88
63.67 0.21 917C 0.12 0.04 (18) 11.8 1.5 0.4 1.7 0.27 0.16 (35) 0.0230 0.0079 0.0513 0.0364
[0.08, 0.61] 0.0137 0.0102

Productora PR-5 64.95 0.13 878C n.a. n.a. n.a. 0.04 (1) 0.0017 0.0010 0.0017 0.0010
(transitional PR-8 57.66 0.19 830C 0.08 0.01 (8) 11.1 0.9 2.7 1.1 0.04 0.02 (13) 0.0009 0.0010 0.0015 0.0012
IOCG) [0.02, 0.07] 0.0003 0.0001

Candelaria CAN-2 58.45 0.45 951C n.a. n.a. n.a. 0.04 0.02 (3) 0.0051 0.0010 0.0077 0.0011
(IOCG) [0.03, 0.06] 0.0022 0.0001
CAN-3 53.36 0.45 880C 0.12 0.06 (6) 11.8 1.9 1.1 2.1 0.03 0.02 (29) 0.0017 0.0010 0.0043 0.0013
[0.02, 0.09] 0.0007 0.0001

Casualidad CAS-4 66.67 0.13 897C n.a. n.a. n.a. 0.06 0.03(3) 0.0030 0.0011 0.0050 0.0013
(IOCG) [0.02, 0.08 ] 0.0015 0.0002
CAS-10 63.65 0.20 911C n.a. n.a. n.a. 0.04 0.02(4) 0.0035 0.0010 0.0046 0.0011
[0.02, 0.06] 0.0020 0.0001
CAS-11 61.56 0.20 886C 0.08 0.03 (10) 11.0 1.2 1.7 1.4 0.03 0.01(23) 0.0016 0.0009 0.0034 0.0012
[0.02, 0.07] 0.0006 0.00007

n.a. = no data available


1Normalized to 100 wt % (Supplementary Table S1)

2 2 5
2Apatite saturation temperature (AST) calculated from whole-rock SiO and P O concentrations using the equation of Piccoli and Candela (1994), recast from the data of Harrison and Watson (1984)
3Average of all igneous apatite analyses (Supplementary Table S4)
4Average log f
O2 calculated from MnO contents of igneous apatite crystals (Supplementary Table S4) using the equation of Miles et al. (2014):
log fO2 = 0.0022 ( 0.0003) Mn (ppm) 9.75 ( 0.46)
5Average log f
O2 value relative to the fayalite-magnetite-quartz (FMQ), where the value of FMQ at different temperatures (provided by the AST estimate) is calculated using the equation of Myers
and Eugster (1983): log fO2 = 24,441.9/T (K) + 8.290 ( 0.167); the FMQ error is the sum of the 0.167 error on FMQ and the uncertainty in the sample log fO2estimate from the previous column
3
6Estimated from apatite SO contents (Supplementary Table S4) using the temperature-dependent apatite-melt partition coefficient formula of Peng et al. (1997):

lnKD = 21,130/T 16.2 (where T is in Kelvin)


3 3 3
7Estimated from apatite SO contents (Supplementary Table S4) using the apatite-melt partition coefficient formula of Parat et al. (2011): SO apatite (wt %) = 0.157 ln[SO ] (melt, wt %) + 0.9834 (r2

= 0.62)
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE

2 2 5
8SiO and P O compositions for this sample taken from sample CDA-9 (same lithology as CDA-8 but less altered)
311
312 RICHARDS ET AL.

Mild extension
difficult to constrain these parameters due to the widespread
(a) Early Cretaceous
effects of oxidation during hydrothermal alteration in ore-
Small porphyry Volcanic arc
associated rocks and the loss of S (as SO2) from igneous rocks
deposits
Sea level
Minor upper crustal
as they crystallize. Nevertheless, Fe2O3/FeO ratios of igneous
0 km
Oceanic crust Continental
crust
plutonism rocks from these different associations show no clear differ-
Lower crustal
Oceanic mantle
lithosphere
MASH zone ence, and all are moderately to strongly oxidized (acknowl-
Basaltic
underplating SCLM
edging that many of these rocks are hydrothermally altered,
1000C which likely leads to some secondary oxidation). We would
Partial melting of
hydrated mantle 100 km expect the oxidation state of the magma to make a difference
De

wedge
to the behavior of dissolved sulfur only if the rocks were sig-
h yd

nificantly reduced (resulting in S being dominantly present


ra

Asthenosphere
t
io

Asthenosphere
n

as sulfide) compared to oxidized (with S dissolved as sulfate;


10
10

00
00

Carroll and Rutherford, 1985; Jugo et al., 2005). All of the


C
C

200 km
(b) Mid-Cretaceous Sinistral rocks in these suites appear to be at least moderately oxidized,
transtension as expected in arc rocks, so S should have behaved in a similar
Minor volcanism
Upper crustal
batholith and IOCG
way in the magma.
formation
0 km
In order to test the hypothesis that the magmas that formed
S-rich porphyry deposits were more S rich than those form-
ing S-poor IOCG deposits, we analyzed the compositions of
1000C igneous apatite crystals trapped as inclusions in silicate phe-
De

Back-arc
nocryst phases. Our results show that apatite from intrusive
hy

asthenospheric
dra

upwelling
100 km rocks associated with the Carmen de Andacollo porphyry
tio

Cu-Au deposit contains significantly more S (0.25 0.17 wt%


n

Partial-melting in lower crust,


amphibolite (black), or
hydrated lithosphere SO3; Table 4) than that from rocks associated with or near
1000C

Partial-melt flow lines IOCG deposits (Productora, Candelaria, and Casualidad;


Fluid flow lines
0.04 0.02wt % SO3; Table 4). No reliable formula currently
200 km
(c) Late Cretaceous Contraction exists to convert apatite SO3 compositions to magmatic S con-
to Cenozoic tents, but the partition coefficients of Peng et al. (1997) can
Large porphyry
deposits Volcanic arc be used to provide a semiquantitative estimate. The results
Upper crustal 0 km
of these calculations suggest that Carmen de Andacollo por-
Continental
crust
batholith phyry magmas contained ~0.02 wt % S, compared to 0.001 to
Lower crustal
MASH zone 0.005wt% S in the IOCG-associated magmasan order of
De
hy Basaltic
underplating 1000C
magnitude difference (Table 5). This is the opposite to what
dr
at
io
n SCLM would be expected if the difference were simply due to the
of Partial melting of
oc
ea hydrated mantle 100 km felsic versus more mafic nature of the porphyry- versus IOCG-
ni wedge
c
lith
os
related rocks, because S solubility decreases with decreasing
ph
10

Asthenosphere er FeO content of magmas (Wallace and Carmichael, 1992). We


00

e
C

10
00
C
conclude that the IOCG-related magmas were significantly
200 km
S undersaturated compared to the porphyry-forming mag-
mas and, consequently, could only generate S-poor magmatic
Fig. 9. Schematic model illustrating the evolution of the tectonomagmatic hydrothermal fluids and ore deposits.
setting along the margin of northern Chile between 25 and 34S during the
Cretaceous, and its relationship to porphyry and IOCG deposit formation. Metallogenic implications
SCLM = subcontinental lithospheric mantle.
The differences in tectonic setting and S content of magmas
associated with porphyry and IOCG deposits in the Meso-
zoic Coastal Cordillera of Chile suggest contrasting models
Magmatic sulfur content for their formation (Fig. 9). Tectonomagmatic processes that
Porphyry and IOCG deposits are most fundamentally dif- give rise to magmas with the potential to form porphyry Cu
ferentiated by the much greater abundance of sulfur in the Mo Au deposits are well established (Richards, 2003; Cooke
former (predominantly fixed as pyrite in phyllic alteration et al., 2005) and involve the evolution of hydrous, oxidized,
zones) compared to the latter (where Fe oxides, magnetite S-rich, calc-alkaline magmas derived from partial melting
and hematite, predominate). We have hypothesized that this in the metasomatized suprasubduction zone asthenospheric
might reflect a difference in the chemistry of magmas asso- mantle wedge (Fig. 9a, c). A continuous flux of oxidized sulfur
ciated with these contrasting deposit types, but the analysis and water (and other fluid-mobile components such as Cl and
presented above shows that no clear differences can be found LILEs) from the subducting slab supplies the large amount
in their major and trace element compositions, although the of S that is typically present in Phanerozoic arc magmas and
IOCG-related rocks are somewhat more mafic than the por- gives them the potential to form S-rich porphyry deposits.
phyry intrusions. An input of oxidized sulfur is specifically required because
Two other possible explanations are differences in the oxi- reduced sulfur will separate early from the magma as sulfide
dation state or sulfur content of the original magma, but it is liquid and will deplete the magma of chalcophile elements
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 313

prior to their potential segregation into late-exsolving mag- porphyry Cu-Au deposits. Some overlap in space and time
matic hydrothermal fluids (Hamlyn et al., 1985; Spooner, between these deposit styles may result from local mantle and
1993; Richards, 1995, 2003). crustal heterogeneities, as well as delayed magmatic responses
If this supply of oxidized sulfur ceases (e.g., subduction to tectonic changes. We suggest that other deposit types in
stops due to arc migration, reversal, or collision) or moves the wider IOCG family diverge from this point of intersection
away from the region of magma production (e.g., subduction with porphyry systems by the greater involvement of external
rollback), then derived magmas will be relatively S poor (Wal- fluids and crustal metal fluxing.
lace and Edmonds, 2011). This may occur in back-arc settings,
where slab rollback and back-arc extension result in the gen- Acknowledgments
eration of magmas from upwelling asthenosphere distal to the This paper is based in part on the published findings of
subduction zone. These magmas have compositions reflect- numerous researchers, whose work we hope we have appro-
ing more primitive intra-plate characteristics, although they priately cited. This research was funded by a Discovery
retain some arc geochemical features (Kimura and Yoshida, Grant (RGPIN203099) and an Engage Grant (EGP 485693-
2006; Caulfield et al., 2012; Timm et al., 2012). 15) from the Natural Sciences and Engineering Research
We hypothesize that this may have been the case during the Council of Canada to JPR, and a postdoctoral fellowship
mid-Cretaceous back-arc extensional phase in northern Chile, from the University of Alberta to GPL. Claire Chamberlain
and that magmas formed at this time had lower magmatic S and Andrew Davies from Teck Resources Ltd. are particu-
contents than normal arc magmas due to their distance from larly thanked for their support of the Engage Grant and for
the subduction zone. Interaction of these relatively primi- access to and accommodation at the Carmen de Andacollo
tive, S-undersaturated asthenospheric magmas with previ- mine. We also thank all the geologists who helped us at the
ously subduction metasomatized upper plate lithosphere mine or project sites, particularly Andres Castillo from Teck
may have resulted in remobilization of metals and volatiles Resources Ltd., Cristian Vasquez from Hot Chili Limited,
by dissolution of residual Cu-Aurich sulfides and melting Osvaldo Beltran from Pucobre, Cristian Leal from the Dos
of hydrous silicates from lower crustal arc cumulates (Keays, Amigos Mine, Bernardino Garay and Orlando Rivera from
1987; Ackerman et al., 2009; Richards, 2009). Derivative mag- Codelco Chile, and Jose Armando Rodriguez from Anglo
mas would be compositionally similar to normal arc magmas American. Special thanks go to Brian Townley for lending
(being largely derived from subduction-modified lithosphere) us regional maps. Finally, we are also grateful to managers
but would only have the potential to form S-poor magmatic- Jorge Skarmeta (Codelco Chile), Vicente Irarrazabal (Anglo
hydrothermal ore deposits upon upper crustal emplacement American), Rodrigo Arce (Anglo American), Wilfredo Tabilo
(Fig. 9b; Core et al., 2006; Richards, 2009; Pettke et al., 2010; and Marcelo Bruna (Pucobre), Rodrigo Diaz (Hot Chili), and
Richards and Mumin, 2013a, b). Deposits formed under such Melanie Leighton (Hot Chili) for giving permission to access
conditions would share many traits with normal arc porphy- their sites and to sample drill core. Jared Geiger is thanked for
ries but would be deficient in sulfur and sulfide minerals, and his help with sample preparation at the University of Alberta.
rich instead in Fe oxides. We refer to these as IOCG deposits Garry Davidson is thanked for his careful review of an ear-
and note that they share many similarities with magnetite- lier version of the manuscript, and Murray Hitzman and John
rich, alkalic-type porphyry Cu-Au deposits formed in back- Dilles are thanked for their editorial contributions, which
arc settings (Harris et al., 2013), suggesting a continuum of have helped to improve its quality.
broadly subduction related magmatic-hydrothermal deposit
types, differentiated largely by their sulfur contents and the REFERENCES
sulfur contents of the generative magmas (Richards and Ackerman, L., Walker, R.J., Puchtel, I.S., Pitcher, L., Jelnek, E., and Strnad,
Mumin, 2013b). L., 2009, Effects of melt percolation on highly siderophile elements and Os
isotopes in subcontinental lithospheric mantle: A study of the upper mantle
Conclusions profile beneath Central Europe: Geochimica et Cosmochimica Acta, v. 73,
p. 24002414.
By studying an area where porphyry and IOCG deposits occur Almonacid, T., 2007, Geologa de la zona de alteracin hidrotermal de
in relatively close spatial and temporal proximity, we have Domeyko y del yacimiento de cobre Dos Amigos, Regin de Atacama,
shown that small changes in tectonic conditions can gener- Chile: Unpublished M.Sc. thesis, Santiago, Chile, Universidad de Chile,
163 p.
ate magmas with broadly similar bulk compositions but sub- Arancibia, G., 2004, Mid-Cretaceous crustal shortening: Evidence from
stantially different sulfur contents, such that distinctive ore a regional-scale ductile shear zone in the Coastal Range of central Chile
deposit types are formed. Specifically, porphyry Cu Mo (32S): Journal of South American Earth Sciences, v. 17, p. 209226.
Au deposits form from relatively S rich arc magmas under Arvalo, C., Grocott, J., and Welkner, D., 2003, The Atacama fault system in
conditions of upper plate contraction or transpression (e.g., the Huasco province: Universidad de Concepcin, Sociedad Geolgica de
Chile, Congreso Geolgico Chileno, X, Concepcin, October 610, 2003,
the late Cretaceous period in Chile) and suboptimally under Abstract 10955, CD-ROM.
extension (early Cretaceous; Tosdal and Richards, 2001). In Arvalo, C., Grocott, J., Martin, W., Pringle, M., and Taylor, G., 2006, Struc-
contrast, magmas generated in back-arc (extensional) settings tural setting of the Candelaria Fe oxide Cu-Au deposit, Chilean Andes
distal to the subduction zone are likely to be relatively S poor (2730'S): Economic Geology, v. 101, p. 819841.
and somewhat more mafic (locally alkaline, and isotopically Armstrong, J.T., 1995, CITZAF: A package of correction programs for the
quantitative electron microbeam X-ray-analysis of thick polished materials,
primitive) in composition. Where such magmas interact with thin-films, and particles: Microbeam Analysis, v. 4, p. 177200.
previously subduction modified lithosphere during ascent, Barton, M.D., 2014, Iron oxide(-Cu-Au-REE-P-Ag-U-Co) systems, in
they may remobilize metals, leading to the potential forma- Turekian, K.K., and Holland, H.D., eds., Treatise on Geochemistry, 2nd
tion of S-poor ore deposits, such as IOCG and alkalic-type edition: Elsevier, p. 515541.
314 RICHARDS ET AL.

Barton, M.D., and Johnson, D.A., 1996, Evaporitic-source model for igneous- Damm, K.-W., Pichowiak, S., Harmon, R.S., Todt, W., Kelley, S., Omarini,
related Fe oxide-(REE-Cu-Au-U) mineralization: Geology, v. 24, p.259262. R., and Niemeyer, H., 1990, Pre-Mesozoic evolution of the central Andes;
2000, Alternative brine sources for Fe-oxide(-Cu-Au) systems: Implica- the basement revisited: Geological Society of America, Special Paper 241,
tions for hydrothermal alteration and metals, in Porter, T.M., ed., Hydro- p. 101126.
thermal iron oxide copper-gold and related deposits: A global perspective: Daroch, G., Anguita, N., and Cortes, M., 2015, Geologa y zonacin hidroter-
Adelaide, Australian Mineral Foundation, p. 4360. mal del depsito Fe(-Cu-Au) Santo Domingo, Regin de Atacama - Chile:
Belousova, E.A., Griffin, W.L., OReilly, S.Y., and Fisher, N.I., 2002, Apa- Colegio de Gelogos de Chile, Sociedad Geolgica de Chile, Congreso
tite as an indicator mineral for mineral exploration: Trace-element com- Geologico Chileno, XIV, La Serena, Chile, October 48, 2015, Extended
positions and their relationship to host rock type: Journal of Geochemical abstracts, 4 p.
Exploration, v. 76, p. 4569. Del Real, I., and Arriagada, C., 2015, Inversin tectnica positiva en el dis-
Blevin, P.L., 2004, Redox and compositional parameters for interpreting the trito El Espino: Relaciones entre deformacin, magmatismo y mineral-
granitoid metallogeny of eastern Australia: Implications for gold-rich ore izacion IOCG, Provincia de Choapa, Chile: Colegio de Gelogos de Chile,
systems: Resource Geology, v. 54, p. 241252. Sociedad Geolgica de Chile, Congreso Geologico Chileno, XIV, La Ser-
Bonyadi, Z., Davidson, G.J., Mehrabi, B., Meffre, S., and Ghazban, F., 2011, ena, Chile, October 48, 2015, Extended abstracts, 4 p.
Significance of apatite REE depletion and monazite inclusions in the brec- Diaz, A., and Corvaln, M., 2015, Modelo gentico preliminar de mina Panul-
ciated Se-Chahun iron oxide-apatite deposit, Bafq district, Iran: Insights cillo y su relacin con los depsitos de Fe-P y del tipo IOCG presentes
from paragenesis and geochemistry: Chemical Geology, v. 281, p. 253269. en la Provincia Metalognica de la Cordillera de la Costa (PMCC), regin
Brown, M., Diz, F., and Grocott, J., 1993, Displacement history of the Ata- de Coquimbo, norte de Chile: Colegio de Gelogos de Chile, Sociedad
cama fault system 2500'S2700'S, northern Chile: Geological Society of Geolgica de Chile, Congreso Geolgico Chileno, XIV, La Serena, Chile,
America Bulletin, v. 105, p. 11651174. October 48, 2015, Extended abstracts, 4 p.
Burnham, C.W., 1979, Magmas and hydrothermal fluids, in Barnes, H.L., ed., Donovan, J.J., Lowers, H.A., and Rusk, B.G., 2011, Improved electron probe
Geochemistry of hydrothermal ore deposits, 2nd edition: New York, John microanalysis of trace elements in quartz: American Mineralogist, v. 96,
Wiley and Sons, p. 71136. p.274282.
Candela, P.A., 1992, Controls on ore metal ratios in granite-related ore sys- Donovan, J.J., Kremser, D., Fournelle, J.H., and Goemann, K., 2015, Probe
tems: An experimental and computational approach: Transactions of the for EPMA: Acquisition, automation and analysis, version 11: Eugene, Ore-
Royal Society of Edinburgh, Earth Sciences, v. 83, p. 317326. gon, Probe Software, Inc., http://www.probesoftware.com.
Carroll, M.R., and Rutherford, M.J., 1985, Sulfide and sulfate saturation in Edfelt, ., 2007, The Tjrrojkka apatite-iron and Cu (-Au) deposits, north-
hydrous silicate melts: Journal of Geophysical Research, v. 90, Supplement, ern Sweden: Unpublished Ph.D. thesis, Lule, Sweden, Lule University
p. C601C612. of Technology, 164 p.
Caulfield, J., Turner, S., Arculus, R., Dale, C., Jenner, F., Pearce, J., Macpher- Emparan, C., and Pineda, G., 2006, Geologa del area Andacollo-Puerto
son, C., and Handley, H., 2012, Mantle flow, volatiles, slab-surface tem- Aldea: Carta Geologica de Chile N96, escala 1:100,000: Chile, Sernageo-
peratures and melting dynamics in the north Tonga arc-Lau back-arc basin: min, 87 p.
Journal of Geophysical Research, v. 117, 17 p., doi:10.1029/2012JB009526.
Ferrando, R., Roperch, P., Morata, D.., Arriagada, C., Ruffet, G., and Cr-
Charrier, R., Pinto, L., and Rodriguez, M.P., 2007, Tectonostratigraphic evo-
dova, M.L., 2014, A paleomagnetic and magnetic fabric study of the Illapel
lution of the Andean orogen in Chile, in Moreno, T., and Gibbons, W., eds.,
plutonic complex, Coastal Range, central Chile: Implications for emplace-
The geology of Chile: London, Geological Society, p. 21114.
ment mechanism and regional tectonic evolution during the mid-Creta-
Chen, H., Cooke, D.R., and Baker, M.J., 2013, Mesozoic iron oxide copper-
ceous: Journal of South American Earth Sciences, v. 50, p. 1226.
gold mineralization in the Central Andes and the Gondwana Superconti-
Field, C.W., Zhanga, L., Dilles, J.H., Rye, R.O., and Reed, M.H., 2005, Sulfur
nent breakup: Economic Geology, v. 108, p. 3744.
and oxygen isotopic record in sulfate and sulfide minerals of early, deep,
Cisternas, M.E., Frutos, J., Galindo, E., and Spiro, B., 1999, Lavas con bitu-
pre-Main stage porphyry Cu-Mo and late Main stage base-metal mineral
men en el Cretcico Inferior de Copiap, Regin de Atacama, Chile: Petro-
qumica e importancia metalognica: Revista Geolgica de Chile, v. 26, deposits, Butte district, Montana: Chemical Geology, v. 215, p. 6193.
p.205226. Gana, P., and Zentilli, M., 2000, Historia termal y exhumacin de intrusivos
Cooke, D.R., Hollings, P., and Walshe, J.L., 2005, Giant porphyry deposits: de la Cordillera de la Costa de Chile central: Sociedad Geolgica de Chile,
Characteristics, distribution, and tectonic controls: Economic Geology, Congreso Geologico Chileno, IX, Puerto Varas, July 31August 4, 2000,
v.100, p. 801818. Proceedings, p. 664668.
Core, D.P., Kesler, S.E., and Essene, E.J., 2006, Unusually Cu-rich magmas Gelcich, S., Davis, D.W., and Spooner, E.T.C., 2005, Testing the apatite-
associated with giant porphyry copper deposits: Evidence from Bingham, magnetite geochronometer: U-Pb and 40Ar/39Ar geochronology of plutonic
Utah: Geology, v. 34, p. 4144. rocks, massive magnetite-apatite tabular bodies, and IOCG mineralization
Corriveau, L., and Mumin, A.H., 2010, Exploring for iron oxide copper-gold in northern Chile: Geochimica et Cosmochimica Acta, v. 69, p. 33673384.
(Ag-Bi-Co-U) deposits: Case examples, classifications and exploration vec- Gill, J.B., 1981, Orogenic andesites and plate tectonics: New York, Springer-
tors: Geological Association of Canada, Short Course Notes 20, p. 112. Verlag, 390 p.
Creixell, C., 2007, Petrognesis y emplazamiento de enjambre de diques mf- Girardi, J.D., 2014, Comparison of Mesozoic magmatic evolution and iron
icos Mesozoicos de Chile Central: Implicancias tectnicas en el desarrollo oxide-copper-gold (IOCG) mineralization, central Andes and western
del arco Jursico-Cretcico temprano: Unpublished Ph.D. thesis, Santiago, North America: Unpublished Ph.D. thesis, Tucson, University of Arizona,
Chile, Universidad de Chile, 275 p. 363 p.
Creixell, C., Arvalo, C., and Fanning, M., 2009, Geochronology of the Cre- Green, T.H., and Pearson, N.J., 1985, Experimental determination of REE
taceous magmatism from the Coastal Cordillera of north-central Chile partition coefficients between amphibole and basaltic to andesitic liquids
(2915' to 2930' S): Metallogenic implications: Universidad de Chile, at high pressure: Geochimica et Cosmochimica Acta, v. 49, p. 14651468.
Sociedad Geolgica de Chile, Congreso Geologico Chileno, XII, Santiago, Grocott, J., and Taylor, G.K, 2002, Magmatic arc fault systems, deformation
Chile, November 2226, 2009, Abstract no. S11_009, 3 p. partitioning and emplacement of granitic complexes in the Coastal Cordil-
Creixell, C., Parada, M.A., Morata, D.,Vsquez, P., Prez de Arce, C., and lera, north Chilean Andes (2530'S to 2700'S): Journal of the Geological
Arriagada, C., 2011, Middle-Late Jurassic to Early Cretaceous transtension Society, London, v. 159, p. 425442.
and transpression during arc building in central Chile: Evidence from mafic Groves, D.I., Bierlein, F.P., Meinert, L.D., and Hitzman, M.W., 2010, Iron
dike swarms: Andean Geology, v. 38, p. 3763. oxide copper-gold (IOCG) deposits through Earth history: Implications for
Creixell, C., Fuentes, J., Bierma, H., and Salazar, E., 2015, Tectnica regional origin, lithospheric setting, and distinction from other epigenetic iron oxide
y metalognesis asociada al emplazamiento de la franja de prfidos cuprfe- deposits: Economic Geology, v. 105, p. 641654.
ros cretcicos del norte de Chile (2830 S): Universidad de Chile, Socie- Hamlyn, P.R., Keays, R.R., Cameron, W.E., Crawford, A.J., and Waldron,
dad Geolgica de Chile, Congreso Geolgico Chileno, XIV, La Serena, H.M., 1985, Precious metals in magnesian low-Ti lavas: Implications for
Chile, October 48, 2015, Extended abstracts, 5 p. metallogenesis and sulfur saturation in primary magmas: Geochimica et
Dallmeyer, D., Brown, M., Grocott, J., Taylor, G.K., and Treloar, P., 1996, Cosmochimica Acta, v. 49, p. 17971811.
Mesozoic magmatic and tectonic events within the Andean plate boundary Harris, A.C., Cooke, D.R., Blackwell, J.L., Fox, N., and Orovan, E.A., 2013,
zone, north Chile: Journal of Geology, v. 104, p. 1940. Volcanotectonic setting of world-class alkalic porphyry and epithermal Au
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 315

Cu deposits of the Southwest Pacific: Society of Economic Geologists, Chile (~22). Implications for the composition of the Andean crust: Journal
Special Publication, no. 17, p. 337359. of Petrology, v. 40, p. 15271551.
Harrison, T.M., and Watson, E.B., 1984, The behavior of apatite during Lucassen, F., Becchio, R., Harmon, R., Kasemann, S., Franz, G., Trumbull,
crustal anatexis: Equilibrium and kinetic considerations: Geochimica et R., Wilke, H.G., Romer, R.L., and Dulski, P., 2001, Composition and den-
Cosmochimica Acta, v. 48, p. 14671477. sity model of the continental crust at an active continental marginthe
Haschke, M., Siebel, W., Gnther, A., and Scheuber, E., 2002, Repeated Central Andes between 21 and 27S: Tectonophysics, v. 341, p. 195223.
crustal thickening and recycling during the Andean orogeny in north Lucassen, F., Harmon, R., Franz, G., Romer, R.L., Becchio, R., and Siebel,
Chile (2126S): Journal of Geophysical Research, v. 107, p. ECV 6-118, W., 2002, Lead evolution of the pre-Mesozoic crust in the Central Andes
doi:10.1029/2001JB000328. (1827): Progressive homogenisation of Pb: Chemical Geology, v. 186,
Hasler, K., 2007, Petrognesis del magmatismo bimodal y metamorfismo de p.183197.
muy bajo grado del Cretcico inferior de la Cordillera de la Costa, Chile Maksaev, V., and Llaumet, C., 2015, Gua de Excursin El Pleito Mina
Central: Unpublished M.Sc. thesis, Santiago, Chile, Universidad de Chile, Dos Amigos: Colegio de Gelogos de Chile, Sociedad Geolgica de Chile,
281 p. Congreso Geologico Chileno, XIV, La Serena, Chile, October 48, 2015,
Henderson, C.E., 2011, Protocols and pitfalls of electron microprobe analysis Proceedings, 13 p.
of apatite: Unpublished M.Sc. thesis, Ann Arbor, Michigan, University of Maksaev, V., Munizaga, F., Fanning, M., Palacios, C., and Tapia, J., 2006,
Michigan, 68 p. SHRIMP U-Pb dating of the Antucoya porphyry copper deposit: New
Hildreth, W., and Moorbath, S., 1988, Crustal contributions to arc magma- evidence for an Early Cretaceous porphyry-related metallogenic epoch
tism in the Andes of central Chile: Contributions to Mineralogy and Petrol- in the Coastal Cordillera of northern Chile: Mineralium Deposita, v. 41,
ogy, v. 98, p. 455489. p.637644.
Hitzman, M.W., 2000, Iron oxide-Cu-Au deposits: What, where, when, and Maksaev, V., Munizaga, F., Valencia, V., and Barra, F., 2009, LA-ICP-MS zir-
why?, in Porter, T.M., ed., Hydrothermal iron oxide copper-gold and related con U-Pb geochronology to constrain the age of post-Neocomian continen-
deposits: A global perspective, v. 1: Adelaide, PGC Publishing, p. 925. tal deposits of the Cerrillos Formation, Atacama Region, northern Chile:
Hitzman, M.W., Oreskes, N., and Einaudi, M.T., 1992, Geological character- Andean Geology, v. 36, p. 264287.
istics and tectonic setting of Proterozoic iron oxide (Cu-U-Au-REE) depos- Maksaev, V., Almonacid, T.A., Munizaga, F., Valencia, V., McWilliams, M.,
its: Precambrian Research, v. 58, p. 241287. and Barra, F., 2010, Geochronological and thermochronological constraints
Hunt, J.A., Baker, T., and Thorkelson, D.J., 2007, A review of iron oxide on porphyry copper mineralization in the Domeyko alteration zone, north-
copper-gold deposits, with focus on the Wernecke Breccias, Yukon, Can- ern Chile: Andean Geology, v. 37, p. 144176.
ada, as an example of a non-magmatic end member and implications for Markey, R.J., Stein, H.J., and Morgan, J.W., 1998, Highly precise Re-Os
IOCG genesis and classification: Exploration and Mining Geology, v. 16, dating for molybdenite using alkaline fusion and NTIMS: Talanta, v. 45,
p. 209232. p.935946.
Imai, A., 2002, Metallogenesis of porphyry Cu deposits of the western Luzon Markey, R.J., Stein, H.J., Hannah, J.L., Selby, D., and Creaser, R.A., 2007,
arc, Philippines: K-Ar ages, SO3 contents of microphenocrystic apatite and Standardizing Re-Os geochronology: A new molybdenite reference mate-
significance of intrusive rocks: Resource Geology, v. 52, p. 147161. rial (Henderson, USA) and the stoichiometry of Os salts: Chemical Geol-
Irwin, J.J., Garca, C., Herv, F., and Brook, M., 1988, Geology of part of
ogy, v. 244, p. 7487.
a long-lived dynamic plate margin: The coastal cordillera of north-central
Marks, M.A.W., Scharrer, M., Ladenburgerand, S., and Markl, G., 2016,
Chile, latitude 3051'31S: Canadian Journal of Earth Sciences, v. 25,
Comment on Apatite: A new redox proxy for silicic magmas? [Geochimica
p.603624.
et Cosmochimica Acta 132 (2014) 101119]: Geochimica et Cosmochimica
Jugo, P.J., Luth, R.W., and Richards, J.P., 2005, Experimental data on the
Acta, v. 183, p. 267270.
speciation of sulfur as a function of oxygen fugacity in basaltic melts: Geo-
Marquardt, M., Cembrano, J., Bissig, T., and Vzquez, C., 2015, Mid Creta-
chimica et Cosmochimica Acta, v. 69, p. 497503.
ceous Cu-Au (Mo) mineralization in the Vallenar district: New Re-Os age
Keays, R.R., 1987, Principles of mobilization (dissolution) of metals in mafic
constraints from Productora deposit, northern Chile: Colegio de Gelogos
and ultramafic rocksthe role of immiscible magmatic sulphides in the
generation of hydrothermal gold and volcanogenic massive sulphide depos- de Chile, Sociedad Geolgica de Chile, Congreso Geologico Chileno, XIV,
its: Ore Geology Reviews, v. 2, p. 4763. La Serena, Chile, October 48, 2015, Extended abstracts, 4 p.
Kimura, J.-I., and Yoshida, T., 2006, Contributions of slab fluid, mantle wedge Marschik, R., and Fontbot, L., 2001, The Punta del Cobre Formation, Punta
and crust to the origin of Quaternary lavas in the NE Japan arc: Journal of del Cobre-Candelaria area, northern Chile: Journal of South American
Petrology, v. 47, p. 21852232. Earth Science, v. 14, p. 401433.
Klein, M., Stosch, H.-G., and Seck, H.A., 1997, Partitioning of high field- Marschik, R., and Kendrick, M.A., 2015, Noble gas and halogen constraints
strength and rare-earth elements between amphibole and quartz-dioritic on fluid sources in iron oxide-copper-gold mineralization: Mantoverde and
to tonalitic melts: An experimental study: Chemical Geology, v. 138, La Candelaria, northern Chile: Mineralium Deposita, v. 50, p. 357371.
p.257271. Matthews, K.J., Seton, M., and Mller, R.D., 2012, A global-scale plate
Kovacic, P., 2014, Geologa y metalognesis del yacimiento tipo IOCG reorganization event at 105100 Ma: Earth and Planetary Science Letters,
Casualidad, distrito Sierra Overa, Segunda Regin de Antofagasta, Chile: v.355356, p. 283298.
Unpublished M.Sc. thesis, Antofagasta, Chile, Universidad Catolica del Middlemost, E.A.K., 1994, Naming materials in the magma/igneous rock sys-
Norte, 118 p. tem: Earth-Science Reviews, v. 37, p. 215224.
Lang, J.R., and Titley, S.R., 1998, Isotopic and geochemical characteristics of Miles, A.J., Graham, C.M., Hawkesworth, C.J., Gillespie, M.R., Hinton,
Laramide magmatic systems in Arizona and implications for the genesis of R.W., Bromiley, G.D., and EMMAC, 2014, Apatite: A new redox proxy for
porphyry copper deposits: Economic Geology, v. 93, p. 138170. silicic magmas?: Geochimica et Cosmochimica Acta, v. 132, p. 101119.
Le Maitre, R.W., Streckeisen, A., Zanettin, B., Le Bas, M.J., Bonin, B., and 2016, Reply to comment by Marks et al. (2016) on Apatite: A new redox
Bateman, P., 2002, Igneous rocks: A classification and glossary of terms, 2nd proxy for silicic magmas? [Geochimica et Cosmochimica Acta 132 (2014)
edition: Cambridge University Press, 256 p. 101119]: Geochimica et Cosmochimica Acta, v. 183, p. 271273.
Li, J., Li, G., Qin, K., Xiao, B., Chen, L., and Zhao, J., 2012, Mineralogy Morata, D., and Aguirre, L., 2003, Extensional Lower Cretaceous volcanism
and mineral chemistry of the Cretaceous Duolong gold-rich porphyry cop- in the Coastal Range (2920'30S), Chile: Geochemistry and petrogenesis:
per deposit in the Bangongco arc, northern Tibet: Resource Geology, v. 62, Journal of South American Earth Sciences, v. 16, p. 459476.
p.1941. Morata, D., Fraud, G., Aguirre, L., Arancibia, G., Belmar, M., Morales, S.,
Lpez, G., 2012, The El Espino iron-oxide copper gold district, Coastal Cor- and Carrillo, J., 2008, Geochronology of the Lower Cretaceous volcanism
dillera of north-central Chile: Unpublished Ph.D. thesis, Golden, Colorado, from the Coastal Range (2920'30S): Chile: Revista Geolgica de Chile,
Colorado School of Mines, 163 p. v. 35, p. 123145.
Lpez, G.P., Hitzman, M.W., and Nelson, E.P., 2014, Alteration patterns and Morelli, P., 2008, Estudio geolgico del sistema de alteracin hidrotermal de
structural controls of the El Espino IOCG mining district, Chile: Minera- Pajonales, Provincia de Vallenar, Regin de Atacama: Unpublished under-
lium Deposita, v. 49, p. 235259. graduate thesis, Santiago, Chile, Universidad de Chile, 44 p.
Lucassen, F., Franz, G., Thirlwall, M.F., and Mezger, K., 1999, Crustal recy- Mumin, A.H., Somarin, A.K., Jones, B., Corriveau, L., Ootes, L., and Camier,
cling of metamorphic basement: Late Paleozoic granitoids of northern J., 2010, The IOCG-porphyry-epithermal continuum of deposit types in
316 RICHARDS ET AL.

the Great Bear magmatic zone, Northwest Territories, Canada: Geological 2011b, Magmatic to hydrothermal metal fluxes in convergent and col-
Association of Canada, Short Course Notes 20, p. 5978. lided margins: Ore Geology Reviews, v. 40, p. 126.
Myers, J., and Eugster, H., 1983, The system Fe-Si-O: Oxygen buffer calibra- 2015, The oxidation state, and sulfur and Cu contents of arc magmas:
tions to 1,500K: Contributions to Mineralogy and Petrology, v. 82, p. 7590. Implications for metallogeny: Lithos, v. 233, p. 2745.
Nielsen, C.H., and Sigurdsson, H., 1981, Quantitative methods for electron Richards, J.P., and Kerrich, R., 2007, Special paper: Adakite-like rocks: Their
microprobe analysis of sodium in natural and synthetic glasses: American diverse origins and questionable role in metallogenesis: Economic Geology,
Mineralogist, v. 66, p. 547552. v. 102, p. 537576.
Oyarzun, R., Rodriguez, M., Pinchera, M., Doblas, M., and Helle, S., 1999, Richards, J.P., and Mumin, A.H., 2013a, Magmatic-hydrothermal processes
The Candelaria (Cu-Fe-Au) and Punta del Cobre (Cu-Fe) deposits (Copi- within an evolving Earth: Iron oxide-copper-gold and porphyry Cu Mo
ap, Chile): A case for extension-related granitoid emplacement and miner- Au deposits: Geology, v. 41, p. 767770.
alization processes?: Mineralium Deposita, v. 34, p. 799801. 2013b, Lithospheric fertilization and mineralization by arc magmas:
Palacios, C.M., Townley, B.C., Lahsen, A.A., and Egaa, A.M., 1993, Geo- Genetic links and secular differences between porphyry copper molybde-
logical development and mineralization in the Atacama segment of the num gold and magmatic-hydrothermal iron oxide copper-gold deposits:
South American Andes, northern Chile (2615'2725'S): Geologische Society of Economic Geologists, Special Publication, no. 17, p. 277299.
Rundschau, v. 82, p. 652662. Rieger, A., Marschik, R., and Diaz, M., 2010, The Mantoverde district, north-
Parada, M.A., Nystrm, J.O., and Levi, B., 1999, Multiple sources for the ern Chile: An example of distal portions of zoned IOCG systems, in Porter,
Coastal batholith of central Chile (3134S): Geochemical and Sr-Nd isoto- T.M., ed., Hydrothermal iron oxide copper-gold and related deposits: A
pic evidence and tectonic implications: Lithos, v. 46, p. 505521. global perspective, v. 3: Adelaide, PGC Publishing, p. 273284.
Parada, M.A., Larrondo, P., Guiresse, C., and Roperch, P., 2002, Mag- Ring, U., Willner, A.P., Layer, P.W., and Richter, P.P., 2012, Jurassic to Early
matic gradients in the Cretaceous Caleu pluton (central Chile): Injections Cretaceous postaccretional sinistral transpression in north-central Chile
of pulses from a stratified magma reservoir: Gondwana Research, v. 5, (latitudes 3132S): Geological Magazine, v. 149, p. 208220.
p.307324. Scheuber, E., and Andriessen, P.A.M., 1990, The kinematic and geodynamic
Parada, M.A., Fraud, G., Fuentes, F., Aguirre, L., Morata, D., and Lar- significance of the Atacama fault zone, northern Chile: Journal of Structural
rondo, P., 2005, Ages and cooling history of the Early Cretaceous Caleu Geology, v. 12, p. 243257.
pluton: Testimony of a switch from a rifted to a compressional continental Scheuber, E., and Gonzalez, G., 1999, Tectonics of the Jurassic-Early Creta-
margin in central Chile: Journal of the Geological Society, London, v. 162, ceous magmatic arc of the north Chilean Coastal Cordillera (2226S): A
p. 273287. story of crustal deformation along a convergent plate boundary: Tectonics,
Parada, M.A., Lpez-Escobar, L., Oliveros, V., Fuentes, F., Morata, D., v. 18, p. 895910.
Caldern, M., Aguirre, L., Fraud, G., Espinoza, F., Moreno, H., Figueroa, Selby, D., and Creaser, R.A., 2004, Macroscale NTIMS and microscale
O., Muoz Bravo, J., Vsquez, R.T., and Stern, C.R., 2007, Andean magma- LA-MC-ICP-MS Re-Os isotopic analysis of molybdenite: Testing spatial
tism, in Moreno, T., and Gibbons, W., eds., The geology of Chile: London, restrictions for reliable Re-Os age determinations, and implications for the
Geological Society, p. 115146. decoupling of Re and Os within molybdenite: Geochimica et Cosmochi-
Parat, F., and Holtz, F., 2005, Sulfur partition coefficient between apatite mica Acta, v. 68, p. 38973908.
and rhyolite: The role of bulk S content: Contributions to Mineralogy and Sillitoe, R.H., 2003, Iron oxide-copper-gold deposits: An Andean view: Min-
Petrology, v. 150, p. 643651. eralium Deposita, v. 38, p. 787812.
Parat, F., Holtz, F., and Klgel, A., 2011, S-rich apatite-hosted glass inclu- 2010, Porphyry copper systems: Economic Geology, v. 105, p. 341.
sions in xenoliths from La Palma: Constraints on the volatile partitioning Skirrow, R.G., 2010, Hematite-group IOCG + U ore systems: Tectonic
in evolved alkaline magmas: Contributions to Mineralogy and Petrology, settings, hydrothermal characteristics, and Cu-Au and U mineralizing pro-
v.162, p. 463478. cesses: Geological Association of Canada, Short Course Notes 20, p. 3959.
Peng, G., Luhr, J., and McGee, J., 1997, Factors controlling sulfur concentra- Smoliar, M.I., Alexander, C.M.OD., Walker, R.J., and Jacobesen, S.B., 2006,
tions in volcanic apatite: American Mineralogist, v. 82, p. 12101224. Re-Os isochron for Allegan (H5): Reconciling Re-Os and U-Pb chronolo-
Pettke, T., Oberli, F., and Heinrich, C.A., 2010, The magma and metal source gies: European Association of Geochemistry, Geochemical Society, Annual
of giant porphyry-type ore deposits, based on lead isotope microanalysis Lunar and Planetary Science Conference, 37th, League City, Texas, March
of individual fluid inclusions: Earth and Planetary Science Letters, v. 296, 1317, 2006, Abstract no. 1468, 2 p.
p.267277. Somarin, A.K., and Mumin, A.H., 2012, The Paleo-Proterozoic high heat pro-
Piccoli, P., and Candela, P., 1994, Apatite in felsic rocks: A model for the duction Richardson granite, Great Bear magmatic zone, Northwest Terri-
estimation of initial halogen concentrations in the Bishop Tuff (Long Valley) tories, Canada: Source of U for Port Radium?: Resource Geology, v. 6263,
and Tuolumne intrusive suite (Sierra Nevada batholith) magmas: American p. 227242.
Journal of Science, v. 294, p. 92135. Spooner, E.T.C., 1993, Magmatic sulphide/volatile interaction as a mecha-
2002, Apatite in igneous systems: Reviews in Mineralogy and Geochem- nism for producing chalcophile element enriched, Archean Au-quartz,
istry, v. 48, p. 255292. epithermal Au-Ag and Au skarn hydrothermal ore fluids: Ore Geology
Pilet, S., Baker, M.B., Mntener, O., and Stolper, E.M., 2011, Monte Carlo Reviews, v. 7, p. 359379.
simulations of metasomatic enrichment in the lithosphere and implications Stormer, Jr., J.C., Pierson, M.L., and Tacker, R.C., 1993, Variation of F and
for the source of alkaline basalts: Journal of Petrology, v. 52, p. 14151442. Cl X-ray intensity due to anisotropic diffusion in apatite during electron
Pollard, P.J., 2000, Evidence of a magmatic fluid and metal source for Fe- microprobe analysis: American Mineralogist, v. 78, p. 641648.
oxide Cu-Au mineralization, in Porter, T.M., ed., Hydrothermal iron oxide Streck, M.J., and Dilles, J.H., 1998, Sulfur content of oxidized arc magmas
copper-gold and related deposits: A global perspective, v. 1: Adelaide, PGC as recorded in apatite from a porphyry copper batholith: Geology, v. 26,
Publishing, p. 2741. p.523526.
Raab, A., 2002, Geology of the Cerro Negro Norte Fe-oxide (Cu-Au) district, Sun, S.-s., and McDonough, W.F., 1989, Chemical and isotopic systematics of
Coastal Cordillera, northern Chile: Unpublished M.Sc. thesis, Corvallis, oceanic basalts: Implications for mantle composition and processes: Geo-
Oregon, Oregon State University, 296 p. logical Society of London, Special Publication, no. 42, p. 313345.
Reyes, M., 1991, The Andacollo strata-bound gold deposit, Chile, and Taylor, G.K., Grocott, J., Pope, A., and Randall, D.E., 1998, Mesozoic fault
its position in a porphyry copper-gold system: Economic Geology, v. 86, systems, deformation and fault block rotation in the Andean forearc: A
p.13011316. crustal scale strike-slip duplex in the Coastal Cordillera of northern Chile:
Richards, J.P., 1995, Alkalic-type epithermal gold depositsa review: Min- Tectonophysics, v. 299, p. 93109.
eralogical Association of Canada, Short Course Series, v. 23, p. 367400. Teck Resources Limited, 2016, 2015 Annual Information Form, 117 p.,
2003, Tectono-magmatic precursors for porphyry Cu-(Mo-Au) deposit http://www.teck.com/media/Investors-aif_march_2016_T5.1.2.pdf.
formation: Economic Geology, v. 98, p. 15151533. Timm, C., de Ronde, C.E.J., Leybourne, M.I., Layton-Matthews, D., and
2009, Postsubduction porphyry Cu-Au and epithermal Au deposits: Graham, I.J., 2012, Sources of chalcophile and siderophile elements in Ker-
Products of remelting of subduction-modified lithosphere: Geology, v. 37, madec arc lavas: Economic Geology, v. 107, p. 15271538.
p. 247250. Tornos, F., Velasco, F., Morata, D., Barra, F., and Rojo, M., 2010, The Trope-
2011a, High Sr/Y arc magmas and porphyry Cu Mo Au deposits: Just zn Cu-Mo-(Au) deposit, northern Chile: The missing link between IOCG
add water: Economic Geology, v. 106, p. 10751081. and porphyry copper systems?: Mineralium Deposita, v. 45, p. 313321.
TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 317

Tornos, F., Wiedenbeck, M., and Velasco, F., 2012, The boron isotope geo- Williams, P.J., 2010, Magnetite-group IOCGs with special reference to
chemistry of tourmaline-rich alteration in the IOCG systems of north- Cloncurry (NW Queensland) and northern Sweden: Settings, alteration,
ern Chile: Implications for a magmatic-hydrothermal origin: Mineralium deposit characteristics, fluid sources, and their relationship to apatite-
Deposita, v. 47, p. 483499. rich iron ores: Geological Association of Canada, Short Course Notes 20,
Tosdal, R.M., and Richards, J.P., 2001, Magmatic and structural controls on p.2338.
the development of porphyry CuMoAu deposits: Reviews in Economic Williams, P.J., Barton, M.D., Fontbot, L., de Haller, A., Johnson, D.A.,
Geology, v. 14, p. 157181. Mark, G., Marschik, R., and Oliver, N.H.S., 2005, Iron oxide copper-gold
Van Hoose, A.E., Streck, M.J., Pallister, J.S., and Wlle, M., 2013, Sulfur evo- deposits: Geology, space-time distribution, and possible modes of origin:
lution of the 1991 Pinatubo magmas based on apatite: Journal of Volcanol- Economic Geology 100th Anniversary Volume, p. 371406.
ogy and Geothermal Research, v. 257, p. 7289. Williams, P.J., Mark, A.K., and Xavier, R.P., 2010, Sources of ore fluid com-
Vergara, M., Levi, B., Nystrom, J.O., and Cancino, A., 1995, Jurassic and ponents in IOCG deposits, in Porter, T.M., ed., Hydrothermal iron oxide
Early Cretaceous island arc volcanism, extension, and subsidence in the copper-gold and related deposits: A global perspective, v. 3, Advances in the
Coast Range of central Chile: Geological Society of America Bulletin, understanding of IOCG deposits: Adelaide, PGC Publishing, p. 107116.
v.107, p. 14271440. Winchester, J.A., and Floyd, P.A., 1977, Geochemical discrimination of dif-
Wall, R., Selles, D., and Gana, P., 1999, Area Tiltil-Santiago, regin Metro- ferent magma series and their differentiation products using immobile ele-
politana: Chile, Sernageomin, scale 1,000,000, 1 sheet. ments: Chemical Geology, v. 20, p. 325343.
Wallace, P.J., and Carmichael, I.S.E., 1992, Sulfur in basaltic magmas: Geo-
chimica et Cosmochimica Acta, v. 56, p. 18631874.
Wallace, P.J., and Edmonds, M., 2011, The sulfur budget in magmas: Evi-
dence from melt inclusions, submarine glasses, and volcanic gas emissions:
Reviews in Mineralogy and Geochemistry, v. 73, p. 215246.
318 RICHARDS ET AL.

APPENDIX
Sources of Data for Figure 1

Tectonostratigraphic time boundaries in Figure 1b (gray lines) Black lines represent the time boundaries of different tec-
were defined based on correlation of similar formations and tonic settings based on a compilation of kinematics and tim-
their interpreted tectonic settings, as follows: ing of movement on structures along the Coastal Cordillera,
as follows:
i. Punta del Cobre, Arqueros, and Veta Negra formations
(Charrier et al., 2007); 1. Atacama fault system at 25.5 to 27S (Grocott and Tay-
ii. Contact between Punta del Cobre and Bandurrias forma- lor, 2002);
tions (Marschik and Fontbot, 2001); 2. Atacama fault system at 25 to 27S (Brown et al., 1993;
iii. Lower member of Cerrillos Formation (Maksaev et al., Taylor et al., 1998);
2009); 3. Atacama fault system at 26 to 27.5S (Dallmeyer et al.,
iv. Ka1/Ka2 boundary, Arqueros Formation (Morata et al., 1996);
2008); 4. Atacama fault system at 29S (Arvalo et al., 2003);
v. Quebrada Marquesa/Viita formations (Emparan and 5. Dos Amigos fault system (Almonacid, 2007);
Pineda, 2006); 6. Bahia Agua Dulce shear zone at 32S (Ring et al., 2012);
vi. Arqueros Formation (Ferrando et al., 2014); 7. El Romeral fault at 30S (Emparan and Pineda, 2006);
vii. Quebrada Marquesa Formation (Lpez, 2012); 8. Silla del Gobernador fault at 32S (Arancibia, 2004);
viii. Las Chilcas Formation (Wall et al., 1999). 9. Magnetic fabric in Illapel plutonic complex (Ferrando et
al., 2014);
10. Dike swarms at 33S (Creixell et al., 2011);
11. Rapid exhumation of intrusions at 33S (Gana and
Zentilli, 2000).

Potrebbero piacerti anche