24
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. López, 1 Jing-Jing Zhu, 1,2 Robert A. Creaser, 1 Andrew J. Locock, 1 and A. Hamid Mumin 3 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 (~125–110 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 Sulfur-rich porphyry Cu ± Mo ± Au deposits are formed by the precipitation of sulfide minerals (pyrite, chalcopyrite, molybdenite) from hydrothermal fluids exsolved from shal- lowly emplaced calc-alkaline magmas in volcanic arcs, typi- cally generated in response to oceanic lithosphere subduction (Burnham, 1979; Richards, 2003; Cooke et al., 2005; Sil- litoe, 2010). In contrast, sulfur-poor iron oxide copper-gold (IOCG) deposits include a wide range of different deposit types, broadly linked by the prevalence of hydrothermal Fe oxides (as opposed to Fe sulfides), with or without Cu and Au mineralization (Hitzman, 2000; Williams et al., 2005, 2010; Hunt et al., 2007; Groves et al., 2010). The broadness of this definition, as well as the capacity of oxidized saline fluids to transport Fe ± Cu ± Au in a variety of geologic settings, has led to controversy over the origin(s) of this group of deposits (e.g., Barton and Johnson, 1996, 2000; Pollard, 2000; Williams et al., 2005; Williams, 2010; Barton, 2014). However, within this group there is a subset of deposits that is more clearly associated with igneous rocks and fluids of possible magmatic- hydrothermal origin (Pollard, 2000; Sillitoe, 2003). Richards and Mumin (2013a, b) have referred to these as magmatic- hydrothermal IOCG deposits, but, for simplicity, we use the general term IOCG below. They share many features with 0361-0128/17/4470/295-24 295 Submitted: October 2, 2015 Accepted: August 8, 2016 Economic Geology, v. 112, pp. 295–318 Corresponding author: e-mail, [email protected] *A digital supplement to this paper is available at http://economicgeology.org/ and at http://econgeol.geoscienceworld.org/. © 2017 Gold Open Access: this paper is published under the terms of the CC-BY license.

Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

  • Upload
    others

  • View
    7

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

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. López,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

AbstractPorphyry 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 (~125–110 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.

IntroductionSulfur-rich porphyry Cu ± Mo ± Au deposits are formed by the precipitation of sulfide minerals (pyrite, chalcopyrite, molybdenite) from hydrothermal fluids exsolved from shal-lowly emplaced calc-alkaline magmas in volcanic arcs, typi-cally generated in response to oceanic lithosphere subduction (Burnham, 1979; Richards, 2003; Cooke et al., 2005; Sil-litoe, 2010). In contrast, sulfur-poor iron oxide copper-gold (IOCG) deposits include a wide range of different deposit types, broadly linked by the prevalence of hydrothermal Fe

oxides (as opposed to Fe sulfides), with or without Cu and Au mineralization (Hitzman, 2000; Williams et al., 2005, 2010; Hunt et al., 2007; Groves et al., 2010). The broadness of this definition, as well as the capacity of oxidized saline fluids to transport Fe ± Cu ± Au in a variety of geologic settings, has led to controversy over the origin(s) of this group of deposits (e.g., Barton and Johnson, 1996, 2000; Pollard, 2000; Williams et al., 2005; Williams, 2010; Barton, 2014). However, within this group there is a subset of deposits that is more clearly associated with igneous rocks and fluids of possible magmatic-hydrothermal origin (Pollard, 2000; Sillitoe, 2003). Richards and Mumin (2013a, b) have referred to these as magmatic-hydrothermal IOCG deposits, but, for simplicity, we use the general term IOCG below. They share many features with

0361-0128/17/4470/295-24 295Submitted: October 2, 2015

Accepted: August 8, 2016

Economic Geology, v. 112, pp. 295–318

† Corresponding author: e-mail, [email protected]*A digital supplement to this paper is available at http://economicgeology.org/

and at http://econgeol.geoscienceworld.org/.

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

Page 2: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

296 RICHARDS ET AL.

porphyry systems (e.g., Cu-Au-Mo-Fe metal association and broad tectonic setting and magmatic affinity) but also have important differences (such as more extensive development of high-temperature sodic, sodic-calcic, and potassic iron alteration envelopes, and more restricted development of lower-temperature phyllic and argillic alteration in IOCG systems compared to porphyry systems; Hitzman et al., 1992; Mumin et al., 2010).

Richards and Mumin (2013b) explained some of these dif-ferences, in particular the smaller acidic alteration zones in IOCG deposits, as reflecting lower abundances of S (SO2) in the ore-forming fluids compared to porphyry fluids. As S-rich porphyry fluids cool, the SO2 disproportionates to form H2S and H2SO4 (sulfuric acid), leading to widespread develop-ment of acidic alteration at shallow levels (Burnham, 1979; Candela, 1992; Field et al., 2005; Richards, 2011b, 2015); this happens to a lesser extent in S-poor IOCG fluids. Another important difference is the broader range of metals found in IOCG versus porphyry deposits, including, in some deposits, the presence of abundant U, rare earth elements (REEs), P, Co, Ni, and Bi. The increased variety of metals is attributed to the much greater influence of fluid reactions with crustal rocks and the resultant metal fluxing that occurs in giant IOCG hydrothermal systems (in addition to magmatic contri-butions), and the more common occurrence of mafic country rocks around many IOCGs (Somarin and Mumin, 2012; Rich-ards and Mumin, 2013b; Barton, 2014).

The Mesozoic Coastal Cordillera of northern Chile is uniquely suited to compare the characteristics and controls on ore formation in porphyry and IOCG deposits because both deposit types occur in a broadly coeval (Cretaceous), 50- to 80-km-wide belt that runs parallel to the coast for over

1,000 km (Fig. 1a; Sillitoe, 2003). The largest IOCG depos-its within this part of the belt are Candelaria (116–110 Ma; 501 Mt at 0.54% Cu, 0.13 g/t Au, and 2.06 g/t Ag) and Man-toverde (121–117 Ma; 440 Mt at 0.56 % Cu and 0.12 g/t Au). The largest porphyry Cu-Au deposit is Carmen de Andacollo (104 Ma; proven and probable reserves of 417 Mt at 0.34% Cu and 0.12 g/t Au; Table 1).

Small Early Cretaceous (>125 Ma; Berriasian-Barremian) porphyry Cu-Au deposits occur at 22°S (e.g., Antucoya, 142 Ma; Tovaku, 132 Ma; Maksaev et al., 2006) and at 33°S (e.g., Colliguay, ~129 Ma; Maksaev et al., 2010) and appear to have formed during a short period of synchronous transpres-sion at 22°S (Maksaev et al., 2006) or continental arc extension at 33°S (Creixell, 2007). In contrast, IOCG (and magnetite-apatite) deposits formed predominantly in the mid-Creta-ceous (~125–~110 Ma; Aptian-Albian; Gelcich et al., 2005; Arévalo et al., 2006; Rieger et al., 2010; Tornos et al., 2010) and are located in a tectonically distinct but spatially super-imposed belt from 25° to 34°S. These deposits have been described as spatially and temporally related to the culmina-tion of a period of back-arc extension that developed along the continental margin from the Late Jurassic to mid-Cretaceous (Oyarzun et al., 1999; Grocott and Taylor, 2002; Sillitoe, 2003) or, alternatively, as having formed in response to the initiation of basin inversion in the mid-Cretaceous (Chen et al., 2013).Porphyry Cu-Au-(Mo) deposits, including the large Carmen de Andacollo porphyry Cu-Au deposit (104 Ma), again began to form in the later Cretaceous (<110 Ma; Cenomanian-Turonian) between 25° and 32°S, following the resumption of arc magmatism during or after basin inversion (Maksaev et al., 2010). We group igneous rocks and associated ore deposits into three broad temporal groupings that relate to

Tropezon (110 Ma)Casualidad (100–94 Ma)

Santo Domingo (124 Ma)Todos los Santos (118 Ma)

Mantoverde (121–117 Ma)Cerro Negro Norte (116 Ma)

Candelaria (116–110 Ma)Punta del Cobre (116 Ma)

Productora (129 Ma)

Trapiche (122–120 Ma)

Panulcillo (115 Ma)

Espino (93–86 Ma)

Colliguay (129 Ma)

Llahuin (92 Ma)

Frontera (112 Ma)Andacollo (104 Ma)

Cachiyuyo (111 Ma)

Los Loros (91 Ma)La Verde (88 Ma)

Los Toros (98 Ma)

N

Porphyry Cu±Mo±Au (> or < 1 Mt Cu)

IOCG (> or < 1 Mt Cu)

Transitional

Extension

Contraction

Extension

Structural setting based onstructural evidence

Structural setting based ontectonostratigraphic evidence

Transtension

Contraction

Transpression

Age (Ma)

Inca de Oro(90–88 Ma)

Dos Amigos (108–104 Ma)Cortadera (87 Ma)

Pajonales (117 Ma)Totora (120 Ma)Las Campanas (90 Ma)

Chile

ArgentinaPaci�cOcean

(1) (i)

(ii)

(iv)

(vi)

(viii)

(vii)

(v)

(viii)

(iii)

(i)

(i)

(6)

(4)

(5)

(7)(7)

(7)

?

(9)

(11)

(8)

(10)

(3) (1)(1,2)

72°W34°S

32°S

30°S

28°S

26°S

24°S

70°W 68°W 145 135 125 115

Extension Sinistral transtension ContractionLegend

105 95 85

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.

Page 3: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 297

tectonic setting as follows: early Cretaceous (extension), mid-Cretaceous (transtension), and late Cretaceous (contraction). Because the tectonic setting changed diachronously from north to south over intervals of 10 to 15 m.y., these terms do not correspond exactly to formal stratigraphic period subdivi-sions (epochs), and there is some temporal overlap between groups.

Evidence for a magmatic hydrothermal origin for several of these Chilean IOCG deposits has been reported, including at Candelaria and Mantoverde (Marschik and Kendrick, 2015) and Tropezon (Tornos et al., 2010, 2012), and several smaller IOCG deposits occur in close proximity to coeval intrusions with alteration patterns apparently centered on those plutons (e.g., El Trapiche veins, Creixell et al., 2009; El Espino, López et al., 2014). In contrast, Barton and Johnson (1996, 2000) argue that the ore-forming fluids were basinal brines, albeit with convection driven by the heat from coeval magmatism.

In this paper, we adopt the hypothesis that the Chilean IOCG and porphyry deposits are both of magmatic-hydrothermal

origin, and that their contrasting styles of ore formation relate to differences in the tectonic setting and chemistry of the associated magmas. We combine new and published geochro-nological and geochemical data for igneous rocks associated with Cretaceous porphyry and IOCG deposits in the Coastal Cordillera with published structural information to show that porphyry and IOCG deposits formed in distinct tectono-magmatic settings. The bulk compositions of the associated igneous rocks are almost indistinguishable, but we present data from analysis of igneous apatite to suggest that the por-phyry-forming magmas were S rich compared to their IOCG counterparts, consistent with the difference in S content of porphyry (S rich) and IOCG (S poor) ore deposits.

Mesozoic Tectonomagmatic Setting and Metallogeny of Northern Chile

Arc magmatism related to subduction has taken place along the Chilean segment of the Gondwana supercontinental margin since the late Paleozoic (Parada et al., 2007). Mid-Jurassic to

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

MetalDeposit/prospect association Resource Age (Ma) References

PorphyryInca de Oro Cu-Au-Ag 389 Mt at 0.39% Cu, 0.1 g/t Au 90–88 (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 108–106 (U-Pb Zr) Almonacid (2007)Totora Cu-Au 120–121 (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 studyLlahuin 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)

IOCGTropezon 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), 100–94 (Ar-Ar Bt, Kovacic (2014) Kspar), 94–84 (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/tMantoverde Cu-Au 440 Mt at 0.56% Cu, 0.12 g/t Au 121–117 (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 116–110 (Ar-Ar Bt; Re-Os Mo) Arévalo et al. (2006); Marschik and Fontboté (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 MoTrapiche Cu-Au 122–120 (Ar-Ar Act) Creixell et al. (2015)Panulcillo Cu-Au 0.5 Mt at 2.75–3% Cu, 0.5–1 g/t Au 115 (K-Ar Phl) Diaz and Corvalán (2015)El Espino Cu-Au 145 Mt at 0.55% Cu, 0.22 g/t Au 93–89 (U-Pb Zr); 88–86 Del Real and Arriagada (2015); (Ar-Ar Act, Ser, Kspar) López 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

Page 4: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

298 RICHARDS ET AL.

Cretaceous magmatism mostly developed ~100 km to the west of the Paleozoic arc (Parada et al., 2007). Extensional tecton-ics affected the arc from the mid-Jurassic to early Cretaceous, with intra-arc and back-arc basin formation occurring in the mid-Cretaceous, followed by tectonic inversion in the mid- to late Cretaceous (Charrier et al., 2007). These tectonic changes are interpreted to reflect alternating periods of subduction coupling and decoupling between the down-going and overrid-ing plates, with the generation of periods dominated by con-tractional and extensional tectonics, respectively, in the upper plate (Scheuber and Gonzalez, 1999). Of particular relevance to this study, early Cretaceous extension and mid-Cretaceous transtension occurred during a period of low convergence rate and weak plate coupling that may reflect slab rollback. This was followed by contraction in the mid- to late Cretaceous, caused by an increase in convergence rate in response to global-scale plate reorganization (Matthews et al., 2012).

Early Cretaceous magmatism in the Coastal Cordillera between 25° and 34°S was generated in a broadly extensional tectonic regime that caused rifting and crustal thinning (Parada et al., 2007). It was characterized by relatively primitive calc-alkaline to shoshonitic lavas that built thick (5–10 km) subaerial volcanic sequences (Vergara et al., 1995; Morata and Aguirre, 2003; Parada et al., 2005; Charrier et al., 2007; Girardi, 2014). By the mid-Cretaceous, the tectonic setting had become pre-dominantly transtensional (Brown et al., 1993; Arévalo et al., 2003) and was characterized by episodic volcanism and sedi-ment deposition in intra-arc and back-arc shallow marine and continental basins (Fig. 1b; Morata and Aguirre, 2003). The transition from extensional to transtensional tectonics was slightly diachronous from north to south, and represents the progression of continental arc rifting southward with time, beginning in the north in the Barremian (~130 Ma) and reach-ing its maximum in the Aptian-Albian (~120 Ma). Plutonic com-plexes were emplaced during both tectonic stages, controlled by crustal-scale fault zones that evolved into the Atacama fault sys-tem in the Valanginian-Barremian (144–126 Ma; Brown et al., 1993). The Atacama fault is an early dip-slip and later sinistral transtensional N-trending fault system that extends for more than ~1,000 km parallel to the continental margin (Fig. 1a) and is interpreted to have formed in response to oblique subduction (Scheuber and Andriessen, 1990; Brown et al., 1993; Palacios et al., 1993); it is the primary structural control on the location of IOCG deposits in Chile (Sillitoe, 2003; Creixell et al., 2009). Crustal extension ended in the Albian-Cenomanian (110–95 Ma) with a return to contractional tectonics that produced crustal shortening and thickening, basin closure, rapid uplift, and a marked eastward shift of magmatism (Parada et al., 2002, 2005; Arancibia, 2004; Maksaev et al., 2010).

The spatiotemporal distribution of Cretaceous IOCG and porphyry Cu-Au deposits in northern Chile between 25° and 34°S is shown in relation to these tectonomagmatic periods in Figure 1b. Here it can be seen that, although the deposits roughly overlap spatially, IOCG deposits are broadly sepa-rated in time from younger porphyry Cu-Au deposits at ~110 to 100 Ma, which also marks the time of the major tectonic change from transtension to contraction. During the initial stages of arc rifting in the Late Jurassic-early Cretaceous, only a few small porphyries were developed (e.g., Antu-coya, Colliguay). This was followed by the early formation

of magnetite-apatite deposits (Gelcich et al., 2005; Creix-ell et al., 2009) and then IOCG deposits during the main stage of rifting and back-arc basin development in the mid-Cretaceous (~125–~110  Ma, Aptian-Albian; Oyarzun et al., 1999). This tectonic relationship appears to be consistent with IOCG deposits globally, which are commonly found to be associated with extensional events, including postcollisional, intracontinental, back-arc, and intra-arc rift settings (Williams et al., 2005; Corriveau and Mumin, 2010; Skirrow, 2010). Porphyry Cu-Au-(Mo) deposit formation returned in the late Cretaceous with the resumption of subduction-related mag-matism during and after tectonic inversion. Apparent excep-tions to this three-part deposit distribution (e.g., the small, mid-Cretaceous Totora, Pajonales, and Cachiyuyo porphy-ries, and the late Cretaceous Casualidad and El Espino IOCG deposits; Fig. 1) may reflect uncertainties in age determina-tion, lithospheric heterogeneities, or delayed response to tec-tonic changes.

SamplingFifty-four samples of volcanic and intrusive igneous rocks associ-ated with the Carmen de Andacollo, Frontera, and Dos Amigos porphyry Cu ± Au ± Mo deposits, the transitional Productora deposit, and the Candelaria, Casualidad, Espino, and Man-toverde IOCG deposits from the Coastal Cordillera of north-ern Chile between 26° and 32°S (Fig. 1) were collected from drill core and outcrops in July 2015. Where possible, intrusive rocks thought to be directly related to mineralization (coeval, cospatial) were collected; other samples were from broadly coeval intrusions or volcanic rock outcrops in the vicinity of the deposits (based on regional geologic maps). Least-altered samples were targeted for collection, but most of the rocks have undergone either low-grade regional metamorphism (result-ing in minor chloritization of ferromagnesian silicate miner-als and partial saussuritization of plagioclase) or hydrothermal alteration due to proximity to the ore deposits. Two samples of quartz-molybdenite veins were collected from the Carmen de Andacollo porphyry Cu-Au deposit for Re-Os dating.

The rocks were studied petrographically to determine the degree of alteration and to seek unaltered accessory minerals such as zircon and apatite for electron microprobe analysis.

Analytical Methods

Whole-rock geochemistry

Thirty-nine least-altered samples from the collected suite were submitted to Activation Laboratories Ltd. (Ancaster, Ontario) for analysis using the 4E-Research analytical pack-age, which combines instrumental neutron activation analysis and lithium metaborate/tetraborate fusion inductively coupled plasma-mass spectrometry for determination of 62 elements. Analyses of standards and duplicates indicate that accuracy for major elements is typically within 5 relative %, and 10 rela-tive % for minor and trace elements. The ferrous iron (FeO) content of the rocks was also determined by titration.

Electron microprobe analyses

Polished thin sections of all samples were examined for the presence of igneous accessory minerals such as zircon and apatite. Because of the relatively mafic nature of many of

Page 5: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 299

the samples, few contained zircon in thin section, so zircon chemistry was not attempted. A larger number of samples contained apatite, although care was needed to distinguish between igneous and hydrothermal (or late-stage) apatite. The latter was common in the groundmass of altered samples, and clearly hydrothermal apatite (intergrown with hydrother-mal minerals such as quartz and sulfides) typically had thin, elongated prismatic habits (≤200-µm length; Fig. 2a). Apatite crystals with unequivocally igneous origin were distinguished by their inclusion within phenocrysts (mainly plagioclase and biotite). The igneous apatite crystals had stubby prismatic habits (20–50-µm length; Fig. 2b-d).

Compositional data were acquired with a Cameca SX100 electron microprobe using wavelength-dispersive spectros-copy and Probe for EPMA software (Donovan et al., 2015). For sessions in which seven elements (F, Na, Si, P, S, Cl, and Ca) were measured, the following conditions were used: 10-kV accelerating voltage, 20-nA beam current, and 5- to 10-µm beam diameter. Total count times of 30 s were used for both peaks and backgrounds. The X-ray lines, analyzing

crystals, and standards were as follows: F Kα, PC0, topaz; Na Kα, TAP, tugtupite; Si Kα, TAP, topaz; P Kα, PET, fluorapatite; S Kα, PET, anhydrite (intensity data aggregated from two spectrometers; Donovan et al., 2011); Cl Kα, PET, tugtupite; and Ca Kα, PET, fluorapatite. The calculated limits of detection (as element, rounded to the nearest 10 ppm) at 99% confidence were as follows: F, 570 ppm; Na, 130 ppm; Si, 120 ppm; P, 170 ppm; S, 80 ppm; Cl, 290 ppm; and Ca, 280  ppm. To check for interference from third-order P Kα on F Kα, the signal from synthetic GaP was examined with the PC0 crystal: no significant interference from P on F was detected under the conditions of analysis.

For sessions in which nine elements (F, Si, P, S, Cl, Ca, Mn, Fe, and As) were measured, the following conditions were used: 15-kV accelerating voltage, 20-nA beam cur-rent, and 5-µm beam diameter. The X-ray lines, analyzing crystals, standards, and count times (seconds) on both peaks and backgrounds were as follows: F Kα, PC0, topaz, 60 s; Si Kα, TAP, topaz, 40 s; P Kα, PET, fluorapatite, 30 s; S Kα, PET, anhydrite, 30 and 40 s (intensity data aggregated from

0.01%

0.02%

0.07%

0.06%

0.27%

0.15%

0.08%

0.45%

0.41%

0.01%

Amph

Biotite

Biotite

Amph

Ap

Ap

Ap

Ap

Ap

Plag

Chl

Chl

Plag

Plag

Plag

PlagQz

Qz

Qz

Qz

Qz

Qz

Qz

sulfide

sulfide

sulfide

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

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

Ap

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.

Page 6: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

300 RICHARDS ET AL.

measurements on two spectrometers; Donovan et al., 2011); Cl Kα, PET, tugtupite, 40 s; Ca Kα, PET, fluorapatite, 30 s; Mn Kα, LIF, spessartine, 40 s; Fe Kα, LIF, spessartine, 40 s; and As Lα, TAP, synthetic GaAs, 40 s. The calculated limits of detection (as element, rounded to the nearest 10 ppm) at 99% confidence were as follows: F, 460 ppm; Si, 110 ppm; P, 260  ppm; S, 70 ppm; Cl, 150 ppm; Ca, 170 ppm; Mn, 170 ppm; Fe, 160 ppm; and As, 180 ppm.

In all sessions, time-dependent intensity corrections for F and Cl were carried out (peak count times divided into six intervals) with Probe for EPMA software (Donovan et al., 2015), following Nielsen and Sigurdsson (1981), Stormer et al. (1993), and Henderson (2011). Intensity data for all ele-ments were reduced following the methods of Armstrong (1995). Oxygen was calculated by stoichiometry and included in the data reduction, as was the correction for oxygen equiva-lence of the halogens (F and Cl).

Re-Os dating

A molybdenite mineral separate was made for each sample by metal-free crushing followed by gravity and magnetic con-centration methods described in detail by Selby and Creaser (2004). The 187Re and 187Os concentrations in molybdenite were determined by isotope dilution mass spectrometry using Carius tube, solvent extraction, anion chromatography, and negative thermal ionization mass spectrometry techniques. A mixed double spike containing known amounts of isotopi-cally enriched 185Re, 190Os, and 188Os was used (Markey et al., 2007). A ThermoScientific Triton mass spectrometer with a Faraday collector was used for isotopic analysis. Total blanks for Re and Os are less than 3 and 2 pg, respectively, which are insignificant for the Re and Os concentrations in molybde-nite. The molybdenite HLP-5 (Markey et al., 1998) was ana-lyzed as a standard, and over a period of two years an average Re-Os date of 221.56 ± 0.40 Ma (1 s.d. uncertainty, n = 10) was obtained. This value is within the uncertainty of the 221.0 ± 1.0 Ma age reported by Markey et al. (1998).

Re-Os Age of the Carmen de Andacollo Porphyry Cu-Au Deposit

The Carmen de Andacollo porphyry Cu-Au deposit has previ-ously been dated at 104 ± 3.3 Ma by U-Pb analysis of zircons in the ore-forming porphyry intrusions (Maksaev et al., 2010), and at 104 ± 3 and 98 ± 2 Ma by K-Ar analysis of phyllic- and potassic-altered rocks, respectively (Reyes, 1991). We have dated two samples of molybdenite from B-type quartz veins in the deposit, which yielded statistically indistinguishable ages of 103.9 ± 0.5 and 103.6 ± 0.5 Ma (2s errors; Table 2), in good

agreement with the U-Pb date for magmatism of Maksaev et al. (2010).

Cretaceous Igneous GeochemistryWhole-rock major and trace element geochemical data for 125 igneous rocks coeval with either IOCG or porphyry deposits from the Coastal Cordillera between 25° and 34°S were com-piled from the literature and combined with our 40 new anal-yses (new data are listed in Table 3 and Supplementary Table S1, and data from the literature are listed in Supplementary Table S2). These rocks are divided into three groups, based on the tectonomagmatic periods described above: early Cre-taceous, related to a few small porphyry Cu-Au deposits and early stages of arc rifting; mid-Cretaceous, related to IOCG deposits and back-arc transtension; and late Cretaceous, related to porphyry Cu-Au-(Mo) deposits and a return to con-tractional tectonics and arc magmatism. Note that, because of diachronous changes in tectonic style from north to south over periods of 10 to 15 m.y., the terms “early,” “mid-,” and “late Cretaceous” do not correspond exactly to formal strati-graphic divisions, but are approximately separated at ~125 and ~110 Ma, respectively.

Volcanic and plutonic rocks from the three tectonomag-matic groups are mostly metaluminous, calc-alkaline to high-K calc-alkaline, and range in composition from andesite (diorite) to dacite (granodiorite). Samples of igneous rocks directly associated with porphyry and IOCG deposits show a similar compositional range on a total alkali-silica diagram (Fig. 3), although it is evident that the porphyry-related rocks are mostly more felsic (diorite to granodiorite) compared to those associated with IOCG deposits (mostly gabbroic diorite to diorite). The apparently alkaline composition of several of these samples (especially from IOCG deposits) might be due to variable degrees of hydrothermal alteration (sodic-calcic or potassic), which was unavoidable in these suites of ore-asso-ciated rocks. However, on a plot of immobile element ratios (Zr/Ti versus Nb/Y; Fig. 4; Winchester and Floyd, 1977), a small number of mid-Cretaceous and IOCG-related samples do plot in more alkaline fields (alkali gabbro and syenite), sug-gesting that some of these samples are genuinely alkaline in composition. Nevertheless, the majority of the samples over-lap in the gabbro-diorite-granodiorite fields in Figure 4, with no clear distinction between age groups or association with porphyry and IOCG deposits.

On primitive mantle-normalized extended trace element (Fig. 5) and chondrite-normalized REE (Fig. 6) diagrams, the porphyry and IOCG suites display almost indistinguish-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 ±2s withSample no. Location Description (ppm) ±2s (ppb) ±2s (ppb) ±2s Os (pg) (Ma)1 λ (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 DCDA-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)/λ, where t = model age and λ = 187Re decay constant, and assuming no initial radiogenic Os2λ = 187Re decay constant, 1.666 × 10–11 yr–1 (Smoliar et al., 2006)

Page 7: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 301Ta

ble

3. W

hole

-Roc

k an

d Tr

ace

Ele

men

t Geo

chem

ical

Dat

a fo

r Ig

neou

s R

ock

Sam

ples

from

the

Coa

stal

Cor

dille

ra o

f Nor

ther

n C

hile

Bet

wee

n 25

° an

d 34

°S

Sam

ple

no.

ESP

-1

ESP

-2

ESP

-3

CD

A-1

C

DA

-2

CD

A-3

C

DA

-7

CD

A-9

C

DA

-10

FR

-1

Fro

nter

a-M

ine

Loc

ality

E

spin

o E

spin

o E

spin

o A

ndac

ollo

A

ndac

ollo

A

ndac

ollo

A

ndac

ollo

A

ndac

ollo

A

ndac

ollo

L

a U

nion

Bio

tite,

Les

s al

tere

d, g

ypsu

m-

Qua

rtz-

C

hlor

ite,

diss

emin

ated

chal

copy

rite

- ch

alco

pyri

teA

ltera

tion/

ca

lcite

vei

ns

C

hlor

ite,

chal

copy

rite

- C

hlor

ite-

iron

oxi

de v

einl

ets,

ve

inle

ts, g

ypsu

m-

Epi

dote

, cla

ys,

min

eral

izat

ion

(ska

rn in

wal

l roc

k)

K

-fel

dspa

r py

rite

(s

eric

ite)

2° b

iotit

e ir

on o

xide

w

eak

pota

ssic

TAS

Gab

broi

c G

abbr

oic

Gab

broi

c

Bas

altic

trac

hyan

desi

tecl

assi

ficat

ion1

di

orite

di

orite

di

orite

M

onzo

nite

G

rano

dior

ite

Gra

nodi

orite

(m

ugea

rite

) G

rano

dior

ite

Mon

zoni

te

Gra

nodi

orite

SiO

2 (w

t %)

51.4

5 51

.85

49.7

8 57

.38

62.9

2 62

.79

49.9

2 62

.05

54.6

7 63

.54

Al 2O

3 (w

t %)

15.5

5 16

.81

16.7

8 16

.23

15.7

3 15

.85

18.1

9 15

.86

14.2

4 14

.63

Fe 2

O3 (

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

0.16

8 0.

167

0.13

0.

074

0.04

8 0.

152

0.09

5 0.

118

0.17

3M

gO (w

t %)

7.87

8

7.43

3.

22

1.06

2.

63

4.28

3.

39

6.77

1.

72C

aO (w

t %)

8.06

9.

43

7.15

5.

03

4.75

4.

49

3.78

5.

27

5.74

3.

77N

a 2O

(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

TiO

2 (w

t %)

0.64

7 0.

693

0.71

1 0.

558

0.37

5 0.

489

0.92

9 0.

517

0.65

1 0.

368

P 2O

5 (w

t %)

0.12

0.

12

0.12

0.

2 0.

13

0.13

0.

25

0.2

0.33

0.

12L

OI

(wt %

) 2.

33

2.25

5.

59

5.37

5.

19

3.31

4.

54

2.76

4.

29

1.31

LO

I 2

(wt %

) 1.

63

1.53

4.

82

5.06

4.

89

2.85

4.

1 2.

35

3.74

1.

03To

tal (

wt %

) 98

.94

100.

6 10

0.9

100.

3 99

.78

100.

1 10

0.1

100.

6 99

.06

98.6

6To

tal 2

(wt %

) 98

.23

99.8

7 10

0.2

99.9

4 99

.48

99.6

2 99

.64

100.

2 98

.51

98.3

8F

e 2O

3(T

) (w

t %)

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.

7T

l (pp

m)

0.27

<0

.05

<0.0

5 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

37B

a (p

pm)

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.

13U

(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 (p

pm)

0.14

0.

16

0.14

0.

22

0.33

0.

31

0.16

0.

28

0.25

0.

28L

a (p

pm)

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

.1Pb

(ppm

) <5

<5

<5

6

<5

<5

18

<5

<5

10Pr

(ppm

) 2.

26

2.02

2.

26

4.15

4.

26

3.86

4.

33

4.75

6.

3 2.

87Sr

(ppm

) 48

9 49

9 36

5 62

1 23

4 58

4 73

8 50

7 35

8 34

8N

d (p

pm)

9.61

8.

96

10.4

16

.8

16.1

15

.4

18

19.6

24

.9

11.4

Zr (p

pm)

58

60

60

92

91

86

66

113

115

69H

f (pp

m)

1.5

1.7

1.6

2.4

2.5

2.4

2 2.

9 3

1.9

Sm (p

pm)

2.74

2.

31

2.58

3.

55

3.16

3.

09

4.25

4.

06

5.26

2.

54E

u (p

pm)

0.86

3 0.

748

0.84

5 1.

17

0.84

1.

03

1.27

1.

19

1.42

0.

836

Sb (p

pm)

1.7

0.8

1.6

1.7

1 <0

.1

1.6

0.2

<0.1

2.

6G

d (p

pm)

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.

4Y

(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

0.25

8 0.

237

0.19

0.

161

0.16

3 0.

262

0.20

1 0.

214

0.23

7Yb

(ppm

) 1.

74

1.68

1.

66

1.2

1.02

1.

07

1.74

1.

33

1.4

1.7

Lu

(ppm

) 0.

29

0.27

7 0.

263

0.18

3 0.

163

0.17

0.

285

0.21

0.

223

0.28

Cr

(ppm

) 36

9 38

9 37

5 26

.5

32.5

43

.1

3.4

102

372

8.4

Ni (

ppm

) 13

0 14

0 12

3 17

11

16

13

38

12

8 3

Sc (p

pm)

28.1

28

.6

28.7

13

.8

9.34

13

.9

22.3

13

.9

22.9

8.

22V

(ppm

) 19

7 20

8 20

7 14

5 10

0 14

1 33

5 13

3 18

1 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

25

0 62

1 5

813

1,14

0 12

37

6 1,

430

655

Page 8: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

302 RICHARDS ET AL.Ta

ble

3. (

Con

t.)

Sam

ple

no.

PR-1

PR

-2

PR-3

PR

-4

PR-5

PR

-6

PR-7

PR

-8

CA

N-1

Pr

oduc

tora

-

Pr

oduc

tora

Pr

oduc

tora

Pr

oduc

tora

R

anch

o H

ill,

Prod

ucto

ra-

Prod

ucto

ra-

Pr

oduc

tora

(ir

on d

istr

ict

(iron

dis

tric

t (ir

on d

istr

ict

Loc

ality

C

achi

yuyi

to s

tock

A

lice

porp

hyry

A

lice

porp

hyry

Pr

oduc

tora

(R

uta

5 ba

thol

ith)

regi

onal

) re

gion

al)

regi

onal

) C

ande

lari

a

Wea

k di

ssem

inat

ed

py

rite

-cha

lcop

yrite

, So

dic

Alte

ratio

n/

Wea

k al

bite

- bu

t oth

erw

ise

(alb

ite-c

hlor

ite)

Mod

erat

ely

W

eak

albi

te-

min

eral

izat

ion

actin

olite

fa

irly

fres

h re

plac

ing

horn

blen

de

alte

red

ac

tinol

ite

Min

or e

pido

te

TAS

clas

sific

atio

n1

Gra

nodi

orite

G

rano

dior

ite

Gra

nodi

orite

B

asal

tic a

ndes

ite

Gra

nodi

orite

D

iori

te

Gra

nodi

orite

D

iori

te

Mon

zoni

te

SiO

2 (w

t %)

62.3

3 66

.72

62.2

9 51

.69

64.1

60

.49

65.2

5 57

.03

58.9

3A

l 2O3 (

wt %

) 14

.57

15.7

2 14

.26

14.9

3 16

.38

16.4

1 15

.17

17.2

3 17

.87

Fe 2

O3 (

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

0.04

2 0.

105

0.13

1 0.

116

0.05

4 0.

09

0.15

3 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.

81N

a 2O

(wt %

) 3.

58

4.02

3.

46

4.19

3.

69

4.26

3.

79

4.61

6.

08K

2O (w

t %)

2.97

0.

75

1.12

0.

5 1.

63

0.41

2.

29

0.58

2.

62Ti

O2 (

wt %

) 0.

953

0.50

1 0.

413

0.71

7 0.

42

1.08

3 0.

705

1.03

2 0.

437

P 2O

5 (w

t %)

0.19

0.

09

0.1

0.1

0.13

0.

22

0.14

0.

19

0.23

LO

I (w

t %)

0.89

1.

94

0.86

4.

41

1.14

1.

07

1.1

1.41

1.

15L

OI

2 (w

t %)

0.61

1.

72

0.38

3.

93

0.77

0.

82

0.76

1.

08

0.85

Tota

l (w

t %)

99.0

7 10

0 99

.44

99.0

1 10

0.2

99.6

7 10

0.2

100.

7 98

.67

Tota

l 2 (w

t %)

98.7

9 99

.8

98.9

6 98

.53

99.8

2 99

.42

99.8

3 10

0.3

98.3

7F

e 2O

3(T

) (w

t %)

4.88

2.

65

11.6

5 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.

3T

l (pp

m)

<0.0

5 <0

.05

<0.0

5 <0

.05

0.09

<0

.05

<0.0

5 <0

.05

<0.0

5R

b (p

pm)

56

20

38

17

39

6 33

14

34

Ba

(ppm

) 64

3 11

8 13

0 41

43

2 12

8 59

0 15

0 84

3T

h (p

pm)

12

3.19

3.

94

1.63

3.

61

7.13

4.

44

2.8

3.33

U (p

pm)

1.44

0.

61

1.16

2.

4 1.

09

1.58

0.

53

0.73

1.

05N

b (p

pm)

6.8

2.3

5.5

1.2

3.8

4.4

4 3.

3 3.

1Ta

(ppm

) 0.

63

0.24

0.

3 0.

1 0.

45

0.41

0.

35

0.27

0.

24L

a (p

pm)

10.9

6.

9 12

.5

10.8

14

.6

12.2

11

.8

4.22

18

.3C

e (p

pm)

27.2

15

.7

25.5

21

.4

30.8

33

.4

29.8

12

.4

36.8

Pb (p

pm)

<5

<5

<5

<5

<5

<5

<5

<5

<5Pr

(ppm

) 3.

88

2.24

3.

14

2.37

3.

72

4.6

4.37

2.

06

4.59

Sr (p

pm)

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

.9Zr

(ppm

) 18

9 89

95

61

11

3 17

1 18

1 13

3 12

7H

f (pp

m)

4.8

2.2

2.4

1.7

2.8

4.2

4.6

3.5

3Sm

(ppm

) 4.

96

2.06

2.

68

2.04

2.

99

4.83

5.

24

3.39

3.

7E

u (p

pm)

1.2

0.79

4 0.

896

0.62

8 0.

918

1.41

1.

24

1.12

1.

06Sb

(ppm

) <0

.1

0.3

0.5

0.5

0.5

0.5

<0.1

0.

2 0.

7G

d (p

pm)

5.29

2.

17

2.6

2.42

2.

58

4.64

5.

54

4.28

3.

28T

b (p

pm)

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 (p

pm)

29

18

12

17

16

28

36

31

19H

o (p

pm)

1.12

0.

49

0.56

0.

61

0.5

0.95

1.

26

1.08

0.

58E

r (p

pm)

3.2

1.51

1.

75

1.9

1.5

2.84

3.

62

3.16

1.

76T

m (p

pm)

0.47

3 0.

236

0.27

0.

302

0.24

2 0.

433

0.53

4 0.

46

0.27

9Yb

(ppm

) 3.

12

1.67

1.

86

2.19

1.

68

3.05

3.

59

3.11

1.

91L

u (p

pm)

0.50

6 0.

284

0.30

7 0.

371

0.28

5 0.

502

0.57

6 0.

502

0.31

6C

r (p

pm)

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 (p

pm)

20

13.8

11

.7

34.5

9.

7 27

.5

19.9

31

.2

9.25

V (p

pm)

126

94

82

251

77

196

110

190

99C

o (p

pm)

11.4

6.

9 18

.4

8.9

8.2

11.3

12

.9

6.5

7.7

Cu

(ppm

) 25

19

4 1,

300

7 11

5

19

15

42

Page 9: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 303Ta

ble

3. (

Con

t.)

Sam

ple

no.

CA

N-2

C

AN

-3

CA

N-4

C

AS-

1 C

AS-

2 C

AS-

3 C

AS-

4 C

AS-

5 C

AS-

8 C

AS-

9

C

ande

lari

a-

Coq

uim

bana

Loc

ality

C

ande

lari

a C

ande

lari

a m

ine

Cas

ualid

ad

Cas

ualid

ad

Cas

ualid

ad

Cas

ualid

ad

Cas

ualid

ad

Cas

ualid

ad

Cas

ualid

ad

A

lbiti

c al

tera

tion

Pota

ssic

+

Prop

yliti

cA

ltera

tion/

+

hem

atite

-Cu

Wea

k W

eak

prop

yliti

c

chal

copy

rite

(c

hlor

ite-e

pido

te)

min

eral

izat

ion

vein

s ne

arby

prop

yliti

c (c

hlor

ite-e

pido

te)

ve

inle

ts

on

pota

ssic

TAS

G

abbr

oic

Gab

broi

c

F

oid

clas

sific

atio

n1

Mon

zoni

te

dior

ite

dior

ite

Mon

zodi

orite

D

iori

te

Mon

zodi

orite

G

rano

dior

ite

mon

zosy

enite

M

onzo

dior

ite

Dio

rite

SiO

2 (w

t %)

58.3

4 52

.88

52.8

52

.87

57.8

2 51

.26

65.0

3 51

.48

54.1

9 56

.66

Al 2O

3 (w

t %)

18.3

8 17

.53

16.5

8 15

.12

17.9

9 16

.96

16.7

15

.99

18.1

4 16

.93

Fe 2

O3 (

wt %

) 3.

01

5.86

3.

55

5.83

2.

03

6.5

0.19

4.

52

0.96

2.

62F

eO (w

t %)

3.4

3.9

6 5.

9 3.

3 4.

9 3.

5 7.

3 2.

1 3.

3M

nO (w

t %)

0.10

6 0.

051

0.10

1 0.

146

0.15

3 0.

063

0.04

4 0.

085

0.09

2 0.

126

MgO

(wt %

) 1.

68

4.13

4.

45

5.55

3.

16

5.83

1.

54

5.38

3.

5 4.

01C

aO (w

t %)

5.67

8.

04

8.46

4.

12

6 1.

84

4.77

1.

22

10.4

6 8.

02N

a 2O

(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.

19Ti

O2 (

wt %

) 0.

84

1.10

4 1.

123

0.82

2 0.

478

0.78

3 0.

351

0.76

8 0.

854

0.74

6P 2

O5 (

wt %

) 0.

45

0.45

0.

42

0.18

0.

12

0.15

0.

13

0.17

0.

32

0.14

LO

I (w

t %)

0.76

1.

07

0.62

2.

29

1.98

3.

35

1.2

1.5

2.43

2.

09L

OI

2 (w

t %)

0.38

0.

63

-0.0

6 1.

63

1.61

2.

8 0.

81

0.68

2.

19

1.72

Tota

l (w

t %)

100.

9 10

0.6

99.8

99

.72

98.8

99

.02

99.1

3 99

.42

99.0

3 10

0.3

Tota

l 2 (w

t %)

100.

6 10

0.2

99.1

2 99

.06

98.4

3 98

.47

98.7

4 98

.61

98.8

99

.95

Fe 2

O3(

T) (

wt %

) 6.

79

10.1

9 10

.22

12.3

9 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.

4T

l (pp

m)

0.05

<0

.05

<0.0

5 0.

08

<0.0

5 <0

.05

0.06

0.

33

<0.0

5 0.

05R

b (p

pm)

90

28

31

48

26

16

26

178

44

31B

a (p

pm)

776

263

346

269

352

67

341

3,13

7 13

4 14

7T

h (p

pm)

7.47

4.

95

4.76

4.

53

0.59

0.

83

2.31

2.

47

4.06

2.

39U

(ppm

) 2.

2 1.

41

1.65

1.

23

0.23

0.

28

0.36

1.

31

0.59

0.

78N

b (p

pm)

7.4

4 4.

1 3.

2 1.

7 1.

6 3.

1 2.

8 5.

5 2.

5Ta

(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

.6C

e (p

pm)

72.6

57

.8

51

36.8

16

.9

18

19.2

23

.3

28

30.2

Pb (p

pm)

6 <5

<5

<5

<5

<5

<5

<5

<5

<5

Pr (p

pm)

9.47

7.

71

6.99

4.

94

2.17

2.

6 2.

16

3.15

5.

24

4.09

Sr (p

pm)

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 (p

pm)

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.

1Sm

(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

0.92

3 0.

628

0.95

3 1.

66

1.16

Sb (p

pm)

0.6

0.7

0.8

0.7

0.6

0.5

<0.1

0.

2 0.

6 <0

.1G

d (p

pm)

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 (p

pm)

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.

59E

r (p

pm)

3.06

3.

08

2.89

2.

24

1.38

1.

69

0.46

2.

39

4.14

1.

86T

m (p

pm)

0.44

1 0.

452

0.41

1 0.

326

0.22

4 0.

249

0.06

8 0.

352

0.64

5 0.

281

Yb (p

pm)

2.96

2.

91

2.77

2.

26

1.59

1.

67

0.45

2.

44

4.08

1.

92L

u (p

pm)

0.48

0.

449

0.46

7 0.

362

0.25

9 0.

262

0.07

3 0.

39

0.56

9 0.

308

Cr

(ppm

) 16

.1

32.7

26

18

3 26

.2

76.8

35

.7

57

26.6

39

.2N

i (pp

m)

6 14

12

42

14

32

5

33

12

15Sc

(ppm

) 15

.2

29.4

33

25

.4

13

29.1

6.

84

22.3

23

.7

24.4

V (p

pm)

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

.8C

u (p

pm)

154

21

125

886

90

16

1,13

0 45

2 76

18

Page 10: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

304 RICHARDS ET AL.Ta

ble

3. (

Con

t.)

Sam

ple

no.

CA

S-10

C

AS-

11

MV-

1 M

V-5

MV-

10

MV-

11

MV-

12

DA

-1

DA

-2

DA

-3

Man

tove

rde-

M

anto

verd

e-

Dos

Am

igos

- D

os A

mig

os-

Loc

ality

C

asua

lidad

C

asua

lidad

L

aura

L

aura

M

anto

verd

e M

anto

verd

e M

anto

verd

e D

os A

mig

os

Tric

olor

Tr

icol

or

2° K

-fel

dspa

r Po

tass

ic

Pota

ssic

(K-f

elds

par,

Pota

ssic

(som

e Po

tass

ic +

min

or

Pota

ssic

+

Alte

ratio

n/

Wea

k W

eak

prop

yliti

c w

ith c

hlor

ite

(K-f

elds

par,

biot

ite, c

hlor

ite) +

Pota

ssic

(som

e w

eath

erin

g to

ch

alco

pyri

te-

chal

copy

rite

-m

iner

aliz

atio

n ch

lori

te

(chl

orite

-epi

dote

) ov

erpr

int

biot

ite, e

pido

te)

min

or c

halc

opyr

ite

Pota

ssic

w

eath

erin

g)

clay

on

frac

ture

s)

pyri

te

pyri

teTA

S

Bas

altic

trac

hy-

Bas

altic

trac

hy-

clas

sific

atio

n1

Gra

nodi

orite

D

iori

te

Mon

zodi

orite

M

onzo

dior

ite

Syen

ite

ande

site

(sho

shon

ite)

ande

site

(sho

shon

ite)

Gra

nodi

orite

G

rano

dior

ite

Gra

nodi

orite

SiO

2 (w

t %)

61.6

4 59

.77

46.5

4 50

.79

57.3

9 51

.21

49.4

1 64

.7

65.9

2 66

.5A

l 2O3 (

wt %

) 14

.94

15.4

6 15

.14

16.8

16

.64

14.7

5 16

.08

13.9

1 14

.88

14.6

9F

e 2O

3 (w

t %)

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

0.14

6 0.

647

0.27

9 0.

061

0.16

4 0.

411

0.06

5 0.

109

0.12

4M

gO (w

t %)

2.44

2.

94

4.65

5.

25

3.06

3.

23

3.27

1.

76

1.55

1.

5C

aO (w

t %)

4.68

5.

54

9.42

6.

67

0.6

4.65

5.

2 1.

7 3.

02

3.04

Na 2

O (w

t %)

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

TiO

2 (w

t %)

0.81

9 0.

801

0.67

8 0.

787

1.19

6 1.

001

1.02

3 0.

401

0.36

8 0.

351

P 2O

5 (w

t %)

0.19

0.

19

0.1

0.13

0.

04

0.18

0.

24

0.09

0.

12

0.16

LO

I (w

t %)

1.42

1.

33

9.96

2.

99

2.91

6.

76

5.86

1.

19

1.41

1.

39L

OI

2 (w

t %)

1 0.

84

9.43

2.

33

2.45

6.

43

5.54

0.

71

1.05

0.

94To

tal (

wt %

) 98

.67

98.9

99

.03

98.7

6 98

.81

98.9

4 98

.86

99.0

5 98

.51

99.8

6To

tal 2

(wt %

) 98

.26

98.4

1 98

.51

98.1

98

.36

98.6

2 98

.53

98.5

7 98

.15

99.4

1F

e 2O

3(T

) (w

t %)

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

<0.0

5 <0

.05

0.12

<0

.05

0.13

0.

33

0.1

0.09

Rb

(ppm

) 97

10

1 37

36

16

2 93

64

40

24

30

Ba

(ppm

) 42

7 36

8 2,

230

801

1,44

0 37

7 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 (p

pm)

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.

3Ta

(ppm

) 0.

56

0.44

0.

16

0.22

0.

54

0.3

0.38

0.

3 0.

36

0.35

La

(ppm

) 22

.1

20

6.32

12

.9

3.43

6.

87

14.7

5.

81

13.4

10

.8C

e (p

pm)

50.4

45

.4

16.1

26

.5

6.76

16

.4

34

12.3

26

.9

21.6

Pb (p

pm)

<5

8 <5

7

<5

<5

<5

7 <5

<5

Pr (p

pm)

6.53

5.

92

2.34

3.

57

0.81

2.

16

4.53

1.

68

3.13

2.

64Sr

(ppm

) 29

7 32

3 41

7 49

0 63

50

22

4 26

9 32

9 32

1N

d (p

pm)

26.9

24

.2

10.7

14

.5

3.55

9.

89

19.8

7.

18

12.7

10

.3Zr

(ppm

) 25

9 20

3 62

81

10

8 13

2 12

4 62

69

64

Hf (

ppm

) 6.

8 5.

6 1.

8 2.

2 3.

1 3.

4 3.

2 1.

8 2.

1 1.

9Sm

(ppm

) 6.

45

5.64

2.

69

3.44

0.

87

2.4

4.7

1.8

2.75

2.

31E

u (p

pm)

1.09

1.

16

0.88

9 1.

07

0.22

9 0.

636

1.46

0.

567

0.86

8 0.

753

Sb (p

pm)

1.4

1.8

5.3

1.3

0.8

1.9

1.3

0.5

0.7

1G

d (p

pm)

5.57

5.

26

2.57

3.

36

0.95

2.

82

4.51

1.

83

2.21

2.

03T

b (p

pm)

0.92

0.

87

0.43

0.

52

0.19

0.

51

0.71

0.

31

0.38

0.

34D

y (p

pm)

5.57

5.

2 2.

57

3.03

1.

25

3.52

4.

38

1.92

2.

46

2.14

Y (p

pm)

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.

37T

m (p

pm)

0.47

7 0.

455

0.19

2 0.

247

0.14

4 0.

385

0.36

5 0.

192

0.26

1 0.

218

Yb (p

pm)

3.23

3

1.27

1.

64

1 2.

61

2.44

1.

35

1.85

1.

55L

u (p

pm)

0.51

6 0.

473

0.20

1 0.

256

0.16

8 0.

415

0.37

5 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

23N

i (pp

m)

8 11

22

26

24

36

15

9

3 3

Sc (p

pm)

20

23.4

27

.8

27.7

22

.7

24.6

27

.9

8.78

7.

32

7.01

V (p

pm)

147

169

198

228

143

209

261

92

74

68C

o (p

pm)

14.8

20

.5

25.7

28

.4

29

15.4

22

.3

5 7.

5 10

.3C

u (p

pm)

111

122

12

122

214

14

230

3,19

0 58

0 1,

580

LO

I =

loss

on

igni

tion

1 Cla

ssifi

catio

n sc

hem

es o

f Le

Mai

tre

et a

l. (2

002)

for

extr

usiv

e ro

cks

and

Mid

dlem

ost (

1994

) for

plu

toni

c ro

cks

Page 11: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

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.

Rhyolite/Granite

Trachyandesite/Syenite

Rhyodacite, dacite/Granodiorite

Andesite/Diorite

Andesite, basalt/Diorite, gabbro

Sub-alkaline basalt/Sub-alkaline gabbro

Alkali basalt/Alkali gabbro

Trachyte

Zr/T

i

Nb/Y0.001

0.01

0.1

1

0.01 0.1 1 10

early Cretaceousmid-Cretaceouslate CretaceousIOCGPorphyry

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 34°S. 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 López et al. (2014).

Page 12: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

306 RICHARDS ET AL.

(LILEs: Rb, Ba, Th, U, K) and light rare earth elements (LREEs); negative anomalies for Nb, Ta, and Ti; relative depletions in compatible elements and middle to heavy rare earth elements (MREEs, HREEs); and flat to listric-shaped patterns from MREEs to HREEs. Such patterns are typical

of subduction-related igneous rocks and reflect enrichments in fluid-mobile LILEs, retention of Nb, Ta, and Ti in insoluble Fe-Ti oxides, and fractionation of amphibole (which preferen-tially partitions MREEs; Gill, 1981; Green and Pearson, 1985; Klein et al., 1997). The only noticeable difference between

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.

Page 13: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 307

these suites is the slightly greater depletion of MREEs-HREEs in the porphyry-related suite, which can be attributed to the more felsic (fractionated) nature of these rocks. This characteristic is also observed in the ratios of Sr/Y and La/Yb, which are generally more elevated in the felsic porphyry suite compared to the more mafic IOCG suite (Fig. 7).

In an attempt to assess the relative oxidation states of the various suites of rocks, we have assessed whole-rock Fe2O3/FeO ratios where reported. While many of these samples are altered to varying degrees, which will clearly affect the Fe2O3/FeO ratio, most of the samples have values ranging from 0.18 to 5.10, corresponding to moderately or strongly oxidized

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.

Page 14: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

308 RICHARDS ET AL.

rocks (using the criteria of Blevin, 2004). Importantly, there is no clear difference in oxidation state between the three temporal-tectonomagmatic groups and porphyry- and IOCG-related rocks, which all appear to be similarly oxidized.

Published Sr-Nd isotope compositions for 53 Cretaceous igneous rocks from the Coastal Cordillera between 25° and 34°S (Supplementary Table S3) are plotted in Figure 8. Early Cretaceous rocks have evolved isotopic compositions

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 López et al. (2014).

Page 15: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 309

suggesting extensive crustal contamination. In contrast, the mid- and later Cretaceous rocks cluster at relatively primitive compositions, albeit still with some crustal contamination.

Apatite CompositionsElectron microprobe analyses of magmatic and hydrothermal apatite from three samples from the Carmen de Andacollo porphyry and seven samples from or near IOCG deposits (Productora, Candelaria, and Casualidad) are listed in Supple-mentary Table S4, and SO3 analyses of magmatic apatite are summarized in Table 4 (along with data from the literature). The results show a clear difference between igneous apatite from rocks associated with the Carmen de Andacollo porphyry deposit (0.25 ± 0.17 wt % SO3, n = 69) compared with those from IOCG deposits (0.04 ± 0.02 wt % SO3, n = 76). A simi-lar relationship is observed for apatite from porphyry deposits (0.12–0.60 wt % SO3) and IOCG deposits (0.07–0.13  wt % SO3) reported in the literature (Table 4). Hydrothermal or late-stage igneous apatite (e.g., edges of apatite micropheno-crysts; Fig. 2d) showed lower and more variable SO3 contents (Supplementary Table S4), consistent with observations else-where in the literature (Streck and Dilles, 1998; Van Hoose et al., 2013).

The SO3 content of apatite varies as a complex function of magmatic temperature, oxidation state, and sulfur fugac-ity (Peng et al., 1997; Parat and Holtz, 2005; Parat et al., 2011), and an accurate calculation of magmatic sulfur con-tent from apatite SO3 compositions is not currently possible.

However, the apatite-melt partition coefficient formula of Peng et al. (1997), which is derived for relatively oxidized arc magmas, can be used to obtain a semiquantitative esti-mate of magmatic S content. We have estimated the apatite saturation temperature of four samples from the Carmen de Andacollo porphyry, the Candelaria and Casualidad IOCG deposits, and a sample of regional mid-Cretaceous igneous rock near Productora (using the equation of Piccoli and Can-dela, 1994, 2002, which is derived from the data of Harrison and Watson, 1984) and used this temperature in the equa-tions of Peng et al. (1997) and Parat et al. (2011) to derive estimates of magmatic S content. The data reported in Table 5 suggest that the average S content of the magma associ-ated with the Carmen de Andacollo porphyry deposit was ~0.02 wt % S (up to 0.06 wt % S)—significantly higher than the average values for magma associated with IOCG depos-its in the region (0.001–0.005 wt % S; Table 5). An alterna-tive method for calculating magmatic sulfur content from apatite SO3 compositions is provided by Parat et al. (2011), and these values are also listed in Table 5. The results differ somewhat in absolute values compared to the results using the Peng et al. (1997) formula, but not in the relative enrich-ment in S of the porphyry-related magmas compared to the IOCG-related magmas.

In order to eliminate the possibility that contrasting mag-matic oxidation state was responsible for this difference in apatite SO3 content, we have estimated magmatic fO2 values for these samples from the average MnO contents of apatite

ε Nd

(87Sr/86Sr)i

Jurassic igneous rocks

Crustal contamination trend

Depleted MORB mantle

Late Paleozoicintrusive rocks

-5

0

5

10

0.702 0.703 0.704 0.705 0.706 0.707 0.708 0.709 0.710

early Cretaceous

mid-Cretaceous

late Cretaceous

Fig. 8. εNd vs. initial 87Sr/86Sr ratios for Cretaceous igneous rocks from the Coastal Cordillera of Chile between 25° and 34°S. 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).

Page 16: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

310 RICHARDS ET AL.

using the equation of Miles et al. (2014, Table 5). These data confirm that all of the rocks are moderately oxidized (DFMQ ≈ 0.4–2.7), although we note that this fO2 calculation relies on electron microprobe analyses of Mn in apatite at concentra-tions close to the detection limit, and so cannot be considered highly accurate. There is also debate about the applicability of this method to rocks of different magmatic temperature and composition (Marks et al., 2016; Miles et al., 2016), although the samples compared here are of broadly similar composi-tion and likely temperature. Consequently, we consider that the formula of Peng et al. (1997), derived for oxidized arc rocks, should be approximately valid, and the calculated abun-dances of magmatic S should be relatively correct, even if the absolute values are questionable (cf. Streck and Dilles, 1998). These data support the hypothesis that the porphyry-forming magmas were richer in S than the IOCG-related magmas.

Discussion

Igneous geochemistry

Whole-rock geochemical compositions of Cretaceous igneous rocks from the Coastal Cordillera of Chile between 25° and 34°S show minimal differences beyond those expected from normal fractionation processes (Figs. 3–5). Rocks associated with early Cretaceous arc rifting (with minor porphyry Cu-Au deposits), mid-Cretaceous back-arc transtension (with IOCG deposits), and later Cretaceous contractional arc magmatism (with porphyry Cu-Au-(Mo) deposits) have compositions typi-cal of subduction-related magmatism. On average, however, later Cretaceous arc rocks are more felsic (Fig. 3), with pro-nounced listric-shaped REE patterns (Fig. 6) and higher Sr/Y and La/Yb ratios compared to mid-Cretaceous rocks associ-ated with IOCG deposits (Fig. 7). There is also a suggestion that some of the mid-Cretaceous magmas and those associ-ated with IOCG deposits may have had slightly more mafic, alkaline compositions (alkali gabbros to syenites).

We interpret these data to indicate that there was little fun-damental change in the source of magmatism throughout the Cretaceous, except for a hint of more mafic alkaline magma-tism during the mid-Cretaceous transtensional rifting phase. Published Sr and Nd isotope compositions of Cretaceous rocks from the Coastal Cordillera indicate mixing between depleted, mantle-derived magmas and isotopically evolved continen-tal crust, with early Cretaceous magmas showing the highest degrees of contamination, likely by late Paleozoic lower and mid-crustal rocks (Fig. 8, Damm et al., 1990; Lucassen et al., 1999, 2001, 2002). Although the isotopic ranges of mid- and late Cretaceous rocks overlap, there is a significant clustering of mid-Cretaceous data at the most primitive end of the range (εNd ≈ 6, 87Sr/86Sri ≈ 0.7033). We suggest that back-arc exten-sional tectonics in the mid-Cretaceous may have facilitated the greater involvement of primitive, mildly alkaline magmas derived from partial melting of upwelling mantle astheno-sphere during crustal extension and thinning (Fig. 9). Never-theless, a significant amount of crustal processing is evident for all of these Cretaceous magmas, as is commonly observed in continental arc rocks (Hildreth and Moorbath, 1988). The more felsic nature and higher Sr/Y and La/Yb ratios of the later Cretaceous porphyry-associated magmas (Fig. 7) may reflect longer residence times in deep crustal magma chambers under contractional tectonic conditions (Hildreth and Moorbath, 1988; Haschke et al., 2002; Richards and Kerrich, 2007), with more extensive fractionation of amphibole (Lang and Titley, 1998; Richards and Kerrich, 2007; Richards, 2011a).

We co nclude that it is not possible uniquely to distinguish between igneous rocks associated with porphyry and IOCG deposits in the Coastal Cordillera of Chile using their whole-rock geochemical or isotopic compositions alone, although the IOCG-associated rocks are, on average, more mafic (locally alkaline) and isotopically slightly more primitive than the por-phyry-related rocks, consistent with their back-arc as opposed to main arc tectonomagmatic setting.

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 studyProductora Transitional IOCG 0.02 0.07 0.04 0.02 14 This studyCandelaria IOCG 0.02 0.09 0.03 0.02 32 This studyCasualidad IOCG 0.02 0.08 0.03 0.02 30 This studyYerington 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)Tjårrojåkka 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

Page 17: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 311

Tabl

e 5.

Est

imat

es o

f Mag

mat

ic T

empe

ratu

re a

nd O

xida

tion

Stat

e fr

om I

gneo

us A

patit

e an

d W

hole

-Roc

k C

ompo

sitio

ns

Apa

tite

SO3

Who

le-

Who

le-

A

patit

e

(w

t %) (

n)3

Ave

rage

A

vera

ge

Max

imum

M

axim

umD

epos

it Sa

mpl

e ro

ck S

iO2

rock

P2O

5

MnO

E

stim

ated

E

stim

ated

[m

axim

um,

mag

mat

ic S

m

agm

atic

S

mag

mat

ic S

m

agm

atic

S (t

ype)

no

. (w

t %)1

(w

t %)1

A

ST2

(wt %

) (n)

3 lo

g f O

24 D

FM

Q5

min

imum

] co

nten

t (w

t %)6

co

nten

t (w

t %)7

co

nten

t (w

t %)6

co

nten

t (w

t %)7

Car

men

de

C

DA

-2

66.7

4 0.

14

905°

C

n.a.

n.

a.

n.a.

0.

23 ±

0.1

8 (3

4)

0.01

60 ±

0.

0088

±

0.05

71

0.13

08A

ndac

ollo

[0.0

5, 0

.81]

0.

0120

0.

0230

(por

phyr

y)

CD

A-8

8 63

.67

0.21

91

7°C

0.

12 ±

0.0

4 (1

8)

–11.

8 ±

1.5

0.4

± 1.

7 0.

27 ±

0.1

6 (3

5)

0.02

30 ±

0.

0079

±

0.05

13

0.03

64

[0

.08,

0.6

1]

0.01

37

0.01

02

Prod

ucto

ra

PR-5

64

.95

0.13

87

8°C

n.

a.

n.a.

n.

a.

0.04

(1)

0.00

17

0.00

10

0.00

17

0.00

10(t

rans

ition

al

PR-8

57

.66

0.19

83

0°C

0.

08 ±

0.0

1 (8

) –1

1.1

± 0.

9 2.

7 ±

1.1

0.04

± 0

.02

(13)

0.

0009

±

0.00

10 ±

0.

0015

0.

0012

IOC

G)

[0

.02,

0.0

7]

0.00

03

0.00

01

Can

dela

ria

C

AN

-2

58.4

5 0.

45

951°

C

n.a.

n.

a.

n.a.

0.

04 ±

0.0

2 (3

) 0.

0051

±

0.00

10 ±

0.

0077

0.

0011

(IO

CG

)

[0.0

3, 0

.06]

0.

0022

0.

0001

C

AN

-3

53.3

6 0.

45

880°

C

0.12

± 0

.06

(6)

–11.

8 ±

1.9

1.1

± 2.

1 0.

03 ±

0.0

2 (2

9)

0.00

17 ±

0.

0010

±

0.00

43

0.00

13

[0

.02,

0.0

9]

0.00

07

0.00

01

Cas

ualid

ad

CA

S-4

66.6

7 0.

13

897°

C

n.a.

n.

a.

n.a.

0.

06 ±

0.0

3(3)

0.

0030

±

0.00

11 ±

0.

0050

0.

0013

(IO

CG

)

[0.0

2, 0

.08

] 0.

0015

0.

0002

C

AS-

10

63.6

5 0.

20

911°

C

n.a.

n.

a.

n.a.

0.

04 ±

0.0

2(4)

0.

0035

±

0.00

10 ±

0.

0046

0.

0011

[0.0

2, 0

.06]

0.

0020

0.

0001

C

AS-

11

61.5

6 0.

20

886°

C

0.08

± 0

.03

(10)

–1

1.0

± 1.

2 1.

7 ±

1.4

0.03

± 0

.01(

23)

0.00

16 ±

0.

0009

±

0.00

34

0.00

12

[0

.02,

0.0

7]

0.00

06

0.00

007

n.a.

= n

o da

ta a

vaila

ble

1 Nor

mal

ized

to 1

00 w

t % (S

uppl

emen

tary

Tab

le S

1)2 A

patit

e sa

tura

tion

tem

pera

ture

(AST

) cal

cula

ted

from

who

le-r

ock

SiO

2 and

P2O

5 con

cent

ratio

ns u

sing

the

equa

tion

of P

icco

li an

d C

ande

la (1

994)

, rec

ast f

rom

the

data

of H

arri

son

and

Wat

son

(198

4)3 A

vera

ge o

f all

igne

ous

apat

ite a

naly

ses

(Sup

plem

enta

ry T

able

S4)

4 Ave

rage

log

f O2 c

alcu

late

d fr

om M

nO c

onte

nts

of ig

neou

s ap

atite

cry

stal

s (S

uppl

emen

tary

Tab

le S

4) u

sing

the

equa

tion

of M

iles

et a

l. (2

014)

:

log

f O2 =

–0.

0022

(± 0

.000

3) M

n (p

pm) –

9.7

5 (±

0.4

6)5 A

vera

ge lo

g f O

2 val

ue r

elat

ive

to th

e fa

yalit

e-m

agne

tite-

quar

tz (Δ

FM

Q),

whe

re th

e va

lue

of F

MQ

at d

iffer

ent t

empe

ratu

res

(pro

vide

d by

the

AST

est

imat

e) is

cal

cula

ted

usin

g th

e eq

uatio

n of

Mye

rs

and

Eug

ster

(198

3): l

og f O

2 = –

24,4

41.9

/T (K

) + 8

.290

(± 0

.167

); th

e ΔF

MQ

err

or is

the

sum

of t

he ±

0.16

7 er

ror

on F

MQ

and

the

unce

rtai

nty

in th

e sa

mpl

e lo

g f O

2est

imat

e fr

om th

e pr

evio

us c

olum

n6 E

stim

ated

from

apa

tite

SO3 c

onte

nts

(Sup

plem

enta

ry T

able

S4)

usi

ng th

e te

mpe

ratu

re-d

epen

dent

apa

tite-

mel

t par

titio

n co

effic

ient

form

ula

of P

eng

et a

l. (1

997)

:

lnK

D =

21,

130/

T –

16.

2 (w

here

T is

in K

elvi

n)7 E

stim

ated

from

apa

tite

SO3 c

onte

nts

(Sup

plem

enta

ry T

able

S4)

usi

ng th

e ap

atite

-mel

t par

titio

n co

effic

ient

form

ula

of P

arat

et a

l. (2

011)

: SO

3 apa

tite

(wt %

) = 0

.157

× ln

[SO

3] (m

elt,

wt %

) + 0

.983

4 (r

2 =

0.62

)8 S

iO2 a

nd P

2O5 c

ompo

sitio

ns fo

r th

is s

ampl

e ta

ken

from

sam

ple

CD

A-9

(sam

e lit

holo

gy a

s C

DA

-8 b

ut le

ss a

ltere

d)

Page 18: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

312 RICHARDS ET AL.

Magmatic sulfur content

Porphyry and IOCG deposits are most fundamentally dif-ferentiated by the much greater abundance of sulfur in the former (predominantly fixed as pyrite in phyllic alteration zones) compared to the latter (where Fe oxides, magnetite and hematite, predominate). We have hypothesized that this might reflect a difference in the chemistry of magmas asso-ciated with these contrasting deposit types, but the analysis presented above shows that no clear differences can be found in their major and trace element compositions, although the IOCG-related rocks are somewhat more mafic than the por-phyry intrusions.

Two other possible explanations are differences in the oxi-dation state or sulfur content of the original magma, but it is

difficult to constrain these parameters due to the widespread effects of oxidation during hydrothermal alteration in ore-associated rocks and the loss of S (as SO2) from igneous rocks as they crystallize. Nevertheless, Fe2O3/FeO ratios of igneous rocks from these different associations show no clear differ-ence, and all are moderately to strongly oxidized (acknowl-edging that many of these rocks are hydrothermally altered, which likely leads to some secondary oxidation). We would expect the oxidation state of the magma to make a difference to the behavior of dissolved sulfur only if the rocks were sig-nificantly reduced (resulting in S being dominantly present as sulfide) compared to oxidized (with S dissolved as sulfate; Carroll and Rutherford, 1985; Jugo et al., 2005). All of the rocks in these suites appear to be at least moderately oxidized, as expected in arc rocks, so S should have behaved in a similar way in the magma.

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 igneous apatite crystals trapped as inclusions in silicate phe-nocryst phases. Our results show that apatite from intrusive rocks associated with the Carmen de Andacollo porphyry Cu-Au deposit contains significantly more S (0.25 ± 0.17 wt % SO3; Table 4) than that from rocks associated with or near IOCG deposits (Productora, Candelaria, and Casualidad; 0.04 ± 0.02 wt % SO3; Table 4). No reliable formula currently exists to convert apatite SO3 compositions to magmatic S con-tents, but the partition coefficients of Peng et al. (1997) can be used to provide a semiquantitative estimate. The results of these calculations suggest that Carmen de Andacollo por-phyry magmas contained ~0.02 wt % S, compared to 0.001 to 0.005 wt % S in the IOCG-associated magmas—an order of magnitude difference (Table 5). This is the opposite to what would be expected if the difference were simply due to the felsic versus more mafic nature of the porphyry- versus IOCG-related rocks, because S solubility decreases with decreasing FeO content of magmas (Wallace and Carmichael, 1992). We conclude that the IOCG-related magmas were significantly S undersaturated compared to the porphyry-forming mag-mas and, consequently, could only generate S-poor magmatic hydrothermal fluids and ore deposits.

Metallogenic implications

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 for their formation (Fig. 9). Tectonomagmatic processes that give rise to magmas with the potential to form porphyry Cu ± Mo ± Au deposits are well established (Richards, 2003; Cooke et al., 2005) and involve the evolution of hydrous, oxidized, S-rich, calc-alkaline magmas derived from partial melting in the metasomatized suprasubduction zone asthenospheric mantle wedge (Fig. 9a, c). A continuous flux of oxidized sulfur and water (and other fluid-mobile components such as Cl and LILEs) from the subducting slab supplies the large amount of S that is typically present in Phanerozoic arc magmas and gives them the potential to form S-rich porphyry deposits. An input of oxidized sulfur is specifically required because reduced sulfur will separate early from the magma as sulfide liquid and will deplete the magma of chalcophile elements

(b) Mid-Cretaceous

Back-arcasthenospheric

upwelling

Sinistral transtension

Minor volcanism

Volcanic arc

Minor upper crustalplutonismLower crustalMASH zoneOceanic mantle

lithosphere

Oceanic crustSea level

(a) Early Cretaceous

SCLM

Continental crust

Mild extension

Basaltic underplating

Partial melting ofhydrated mantle

wedge

AsthenosphereAsthenosphere

Volcanic arc

Upper crustalbatholith

Lower crustalMASH zone

Dehydration of oceanic lithosphere

(c) Late Cretaceousto Cenozoic

SCLM

Continental crust

Contraction

Basaltic underplating

Partial melting ofhydrated mantle

wedge

Asthenosphere

1000°C

1000°C

1000°C

1000°C

1000°C

1000°C

Upper crustalbatholith and IOCG

formation

Small porphyry deposits

Dehydration

Dehydration1000°C

1000°C

Large porphyry deposits

Partial-melting in lower crust, amphibolite (black), orhydrated lithospherePartial-melt flow lines

Fluid flow lines

0 km

100 km

200 km

0 km

100 km

200 km

0 km

100 km

200 km

0 km

100 km

200 km

0 km

100 km

200 km

Fig. 9. Schematic model illustrating the evolution of the tectonomagmatic setting along the margin of northern Chile between 25° and 34°S during the Cretaceous, and its relationship to porphyry and IOCG deposit formation. SCLM = subcontinental lithospheric mantle.

Page 19: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 313

prior to their potential segregation into late-exsolving mag-matic hydrothermal fluids (Hamlyn et al., 1985; Spooner, 1993; Richards, 1995, 2003).

If this supply of oxidized sulfur ceases (e.g., subduction stops due to arc migration, reversal, or collision) or moves away from the region of magma production (e.g., subduction rollback), then derived magmas will be relatively S poor (Wal-lace and Edmonds, 2011). This may occur in back-arc settings, where slab rollback and back-arc extension result in the gen-eration of magmas from upwelling asthenosphere distal to the subduction zone. These magmas have compositions reflect-ing more primitive intra-plate characteristics, although they retain some arc geochemical features (Kimura and Yoshida, 2006; Caulfield et al., 2012; Timm et al., 2012).

We hypothesize that this may have been the case during the mid-Cretaceous back-arc extensional phase in northern Chile, and that magmas formed at this time had lower magmatic S contents than normal arc magmas due to their distance from the subduction zone. Interaction of these relatively primi-tive, S-undersaturated asthenospheric magmas with previ-ously subduction metasomatized upper plate lithosphere may have resulted in remobilization of metals and volatiles by dissolution of residual Cu-Au–rich sulfides and melting of hydrous silicates from lower crustal arc cumulates (Keays, 1987; Ackerman et al., 2009; Richards, 2009). Derivative mag-mas would be compositionally similar to normal arc magmas (being largely derived from subduction-modified lithosphere) but would only have the potential to form S-poor magmatic-hydrothermal ore deposits upon upper crustal emplacement (Fig. 9b; Core et al., 2006; Richards, 2009; Pettke et al., 2010; Richards and Mumin, 2013a, b). Deposits formed under such conditions would share many traits with normal arc porphy-ries but would be deficient in sulfur and sulfide minerals, and rich instead in Fe oxides. We refer to these as IOCG deposits and note that they share many similarities with magnetite-rich, alkalic-type porphyry Cu-Au deposits formed in back-arc settings (Harris et al., 2013), suggesting a continuum of broadly subduction related magmatic-hydrothermal deposit types, differentiated largely by their sulfur contents and the sulfur contents of the generative magmas (Richards and Mumin, 2013b).

ConclusionsBy studying an area where porphyry and IOCG deposits occur in relatively close spatial and temporal proximity, we have shown that small changes in tectonic conditions can gener-ate magmas with broadly similar bulk compositions but sub-stantially different sulfur contents, such that distinctive ore deposit types are formed. Specifically, porphyry Cu ± Mo ± Au deposits form from relatively S rich arc magmas under conditions of upper plate contraction or transpression (e.g., the late Cretaceous period in Chile) and suboptimally under extension (early Cretaceous; Tosdal and Richards, 2001). In contrast, magmas generated in back-arc (extensional) settings distal to the subduction zone are likely to be relatively S poor and somewhat more mafic (locally alkaline, and isotopically primitive) in composition. Where such magmas interact with previously subduction modified lithosphere during ascent, they may remobilize metals, leading to the potential forma-tion of S-poor ore deposits, such as IOCG and alkalic-type

porphyry Cu-Au deposits. Some overlap in space and time between these deposit styles may result from local mantle and crustal heterogeneities, as well as delayed magmatic responses to tectonic changes. We suggest that other deposit types in the wider IOCG family diverge from this point of intersection with porphyry systems by the greater involvement of external fluids and crustal metal fluxing.

AcknowledgmentsThis paper is based in part on the published findings of numerous researchers, whose work we hope we have appro-priately cited. This research was funded by a Discovery Grant (RGPIN203099) and an Engage Grant (EGP 485693-15) from the Natural Sciences and Engineering Research Council of Canada to JPR, and a postdoctoral fellowship from the University of Alberta to GPL. Claire Chamberlain and Andrew Davies from Teck Resources Ltd. are particu-larly thanked for their support of the Engage Grant and for access to and accommodation at the Carmen de Andacollo mine. We also thank all the geologists who helped us at the mine or project sites, particularly Andres Castillo from Teck Resources Ltd., Cristian Vasquez from Hot Chili Limited, Osvaldo Beltran from Pucobre, Cristian Leal from the Dos Amigos Mine, Bernardino Garay and Orlando Rivera from Codelco Chile, and Jose Armando Rodriguez from Anglo American. Special thanks go to Brian Townley for lending us regional maps. Finally, we are also grateful to managers Jorge Skarmeta (Codelco Chile), Vicente Irarrazabal (Anglo American), Rodrigo Arce (Anglo American), Wilfredo Tabilo and Marcelo Bruna (Pucobre), Rodrigo Diaz (Hot Chili), and Melanie Leighton (Hot Chili) for giving permission to access their sites and to sample drill core. Jared Geiger is thanked for his help with sample preparation at the University of Alberta. Garry Davidson is thanked for his careful review of an ear-lier version of the manuscript, and Murray Hitzman and John Dilles are thanked for their editorial contributions, which have helped to improve its quality.

REFERENCES

Ackerman, L., Walker, R.J., Puchtel, I.S., Pitcher, L., Jelínek, E., and Strnad, L., 2009, Effects of melt percolation on highly siderophile elements and Os isotopes in subcontinental lithospheric mantle: A study of the upper mantle profile beneath Central Europe: Geochimica et Cosmochimica Acta, v. 73, p. 2400–2414.

Almonacid, T., 2007, Geología de la zona de alteración hidrotermal de Domeyko y del yacimiento de cobre Dos Amigos, Región de Atacama, Chile: Unpublished M.Sc. thesis, Santiago, Chile, Universidad de Chile, 163 p.

Arancibia, G., 2004, Mid-Cretaceous crustal shortening: Evidence from a regional-scale ductile shear zone in the Coastal Range of central Chile (32°S): Journal of South American Earth Sciences, v. 17, p. 209–226.

Arévalo, C., Grocott, J., and Welkner, D., 2003, The Atacama fault system in the Huasco province: Universidad de Concepción, Sociedad Geológica de Chile, Congreso Geológico Chileno, X, Concepción, October 6–10, 2003, Abstract 10955, CD-ROM.

Arévalo, C., Grocott, J., Martin, W., Pringle, M., and Taylor, G., 2006, Struc-tural setting of the Candelaria Fe oxide Cu-Au deposit, Chilean Andes (27°30'S): Economic Geology, v. 101, p. 819–841.

Armstrong, J.T., 1995, CITZAF: A package of correction programs for the quantitative electron microbeam X-ray-analysis of thick polished materials, thin-films, and particles: Microbeam Analysis, v. 4, p. 177–200.

Barton, M.D., 2014, Iron oxide(-Cu-Au-REE-P-Ag-U-Co) systems, in Turekian, K.K., and Holland, H.D., eds., Treatise on Geochemistry, 2nd edition: Elsevier, p. 515–541.

Page 20: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

314 RICHARDS ET AL.

Barton, M.D., and Johnson, D.A., 1996, Evaporitic-source model for igneous-related Fe oxide-(REE-Cu-Au-U) mineralization: Geology, v. 24, p. 259–262.

——2000, Alternative brine sources for Fe-oxide(-Cu-Au) systems: Implica-tions for hydrothermal alteration and metals, in Porter, T.M., ed., Hydro-thermal iron oxide copper-gold and related deposits: A global perspective: Adelaide, Australian Mineral Foundation, p. 43–60.

Belousova, E.A., Griffin, W.L., O’Reilly, S.Y., and Fisher, N.I., 2002, Apa-tite as an indicator mineral for mineral exploration: Trace-element com-positions and their relationship to host rock type: Journal of Geochemical Exploration, v. 76, p. 45–69.

Blevin, P.L., 2004, Redox and compositional parameters for interpreting the granitoid metallogeny of eastern Australia: Implications for gold-rich ore systems: Resource Geology, v. 54, p. 241–252.

Bonyadi, Z., Davidson, G.J., Mehrabi, B., Meffre, S., and Ghazban, F., 2011, Significance of apatite REE depletion and monazite inclusions in the brec-ciated Se-Chahun iron oxide-apatite deposit, Bafq district, Iran: Insights from paragenesis and geochemistry: Chemical Geology, v. 281, p. 253–269.

Brown, M., Diáz, F., and Grocott, J., 1993, Displacement history of the Ata-cama fault system 25°00'S–27°00'S, northern Chile: Geological Society of America Bulletin, v. 105, p. 1165–1174.

Burnham, C.W., 1979, Magmas and hydrothermal fluids, in Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 2nd edition: New York, John Wiley and Sons, p. 71–136.

Candela, P.A., 1992, Controls on ore metal ratios in granite-related ore sys-tems: An experimental and computational approach: Transactions of the Royal Society of Edinburgh, Earth Sciences, v. 83, p. 317–326.

Carroll, M.R., and Rutherford, M.J., 1985, Sulfide and sulfate saturation in hydrous silicate melts: Journal of Geophysical Research, v. 90, Supplement, p. C601–C612.

Caulfield, J., Turner, S., Arculus, R., Dale, C., Jenner, F., Pearce, J., Macpher-son, C., and Handley, H., 2012, Mantle flow, volatiles, slab-surface tem-peratures and melting dynamics in the north Tonga arc-Lau back-arc basin: Journal of Geophysical Research, v. 117, 17 p., doi:10.1029/2012JB009526.

Charrier, R., Pinto, L., and Rodriguez, M.P., 2007, Tectonostratigraphic evo-lution of the Andean orogen in Chile, in Moreno, T., and Gibbons, W., eds., The geology of Chile: London, Geological Society, p. 21–114.

Chen, H., Cooke, D.R., and Baker, M.J., 2013, Mesozoic iron oxide copper-gold mineralization in the Central Andes and the Gondwana Superconti-nent breakup: Economic Geology, v. 108, p. 37–44.

Cisternas, M.E., Frutos, J., Galindo, E., and Spiro, B., 1999, Lavas con bitu-men en el Cretácico Inferior de Copiapó, Región de Atacama, Chile: Petro-química e importancia metalogénica: Revista Geológica de Chile, v.  26, p. 205–226.

Cooke, D.R., Hollings, P., and Walshe, J.L., 2005, Giant porphyry deposits: Characteristics, distribution, and tectonic controls: Economic Geology, v. 100, p. 801–818.

Core, D.P., Kesler, S.E., and Essene, E.J., 2006, Unusually Cu-rich magmas associated with giant porphyry copper deposits: Evidence from Bingham, Utah: Geology, v. 34, p. 41–44.

Corriveau, L., and Mumin, A.H., 2010, Exploring for iron oxide copper-gold (Ag-Bi-Co-U) deposits: Case examples, classifications and exploration vec-tors: Geological Association of Canada, Short Course Notes 20, p. 1–12.

Creixell, C., 2007, Petrogénesis y emplazamiento de enjambre de diques máf-icos Mesozoicos de Chile Central: Implicancias tectónicas en el desarrollo del arco Jurásico-Cretácico temprano: Unpublished Ph.D. thesis, Santiago, Chile, Universidad de Chile, 275 p.

Creixell, C., Arévalo, C., and Fanning, M., 2009, Geochronology of the Cre-taceous magmatism from the Coastal Cordillera of north-central Chile (29°15' to 29°30' S): Metallogenic implications: Universidad de Chile, Sociedad Geológica de Chile, Congreso Geologico Chileno, XII, Santiago, Chile, November 22–26, 2009, Abstract no. S11_009, 3 p.

Creixell, C., Parada, M.A., Morata, D.,Vásquez, P., Pérez de Arce, C., and Arriagada, C., 2011, Middle-Late Jurassic to Early Cretaceous transtension and transpression during arc building in central Chile: Evidence from mafic dike swarms: Andean Geology, v. 38, p. 37–63.

Creixell, C., Fuentes, J., Bierma, H., and Salazar, E., 2015, Tectónica regional y metalogénesis asociada al emplazamiento de la franja de pórfidos cuprífe-ros cretácicos del norte de Chile (28°–30° S): Universidad de Chile, Socie-dad Geológica de Chile, Congreso Geológico Chileno, XIV, La Serena, Chile, October 4–8, 2015, Extended abstracts, 5 p.

Dallmeyer, D., Brown, M., Grocott, J., Taylor, G.K., and Treloar, P., 1996, Mesozoic magmatic and tectonic events within the Andean plate boundary zone, north Chile: Journal of Geology, v. 104, p. 19–40.

Damm, K.-W., Pichowiak, S., Harmon, R.S., Todt, W., Kelley, S., Omarini, R., and Niemeyer, H., 1990, Pre-Mesozoic evolution of the central Andes; the basement revisited: Geological Society of America, Special Paper 241, p. 101–126.

Daroch, G., Anguita, N., and Cortes, M., 2015, Geología y zonación hidroter-mal del depósito Fe(-Cu-Au) Santo Domingo, Región de Atacama - Chile: Colegio de Geólogos de Chile, Sociedad Geológica de Chile, Congreso Geologico Chileno, XIV, La Serena, Chile, October 4–8, 2015, Extended abstracts, 4 p.

Del Real, I., and Arriagada, C., 2015, Inversión tectónica positiva en el dis-trito El Espino: Relaciones entre deformación, magmatismo y mineral-izacion IOCG, Provincia de Choapa, Chile: Colegio de Geólogos de Chile, Sociedad Geológica de Chile, Congreso Geologico Chileno, XIV, La Ser-ena, Chile, October 4–8, 2015, Extended abstracts, 4 p.

Diaz, A., and Corvalán, M., 2015, Modelo genético preliminar de mina Panul-cillo y su relación con los depósitos de Fe-P y del tipo IOCG presentes en la Provincia Metalogénica de la Cordillera de la Costa (PMCC), región de Coquimbo, norte de Chile: Colegio de Geólogos de Chile, Sociedad Geológica de Chile, Congreso Geológico Chileno, XIV, La Serena, Chile, October 4–8, 2015, Extended abstracts, 4 p.

Donovan, J.J., Lowers, H.A., and Rusk, B.G., 2011, Improved electron probe microanalysis of trace elements in quartz: American Mineralogist, v. 96, p. 274–282.

Donovan, J.J., Kremser, D., Fournelle, J.H., and Goemann, K., 2015, Probe for EPMA: Acquisition, automation and analysis, version 11: Eugene, Ore-gon, Probe Software, Inc., http://www.probesoftware.com.

Edfelt, Å., 2007, The Tjårrojåkka apatite-iron and Cu (-Au) deposits, north-ern Sweden: Unpublished Ph.D. thesis, Luleå, Sweden, Luleå University of Technology, 164 p.

Emparan, C., and Pineda, G., 2006, Geología del area Andacollo-Puerto Aldea: Carta Geologica de Chile N°96, escala 1:100,000: Chile, Sernageo-min, 87 p.

Ferrando, R., Roperch, P., Morata, D.., Arriagada, C., Ruffet, G., and Cór-dova, M.L., 2014, A paleomagnetic and magnetic fabric study of the Illapel plutonic complex, Coastal Range, central Chile: Implications for emplace-ment mechanism and regional tectonic evolution during the mid-Creta-ceous: Journal of South American Earth Sciences, v. 50, p. 12–26.

Field, C.W., Zhanga, L., Dilles, J.H., Rye, R.O., and Reed, M.H., 2005, Sulfur and oxygen isotopic record in sulfate and sulfide minerals of early, deep, pre-Main stage porphyry Cu-Mo and late Main stage base-metal mineral deposits, Butte district, Montana: Chemical Geology, v. 215, p. 61–93.

Gana, P., and Zentilli, M., 2000, Historia termal y exhumación de intrusivos de la Cordillera de la Costa de Chile central: Sociedad Geológica de Chile, Congreso Geologico Chileno, IX, Puerto Varas, July 31–August 4, 2000, Proceedings, p. 664–668.

Gelcich, S., Davis, D.W., and Spooner, E.T.C., 2005, Testing the apatite-magnetite geochronometer: U-Pb and 40Ar/39Ar geochronology of plutonic rocks, massive magnetite-apatite tabular bodies, and IOCG mineralization in northern Chile: Geochimica et Cosmochimica Acta, v. 69, p. 3367–3384.

Gill, J.B., 1981, Orogenic andesites and plate tectonics: New York, Springer-Verlag, 390 p.

Girardi, J.D., 2014, Comparison of Mesozoic magmatic evolution and iron oxide-copper-gold (“IOCG”) mineralization, central Andes and western North America: Unpublished Ph.D. thesis, Tucson, University of Arizona, 363 p.

Green, T.H., and Pearson, N.J., 1985, Experimental determination of REE partition coefficients between amphibole and basaltic to andesitic liquids at high pressure: Geochimica et Cosmochimica Acta, v. 49, p. 1465–1468.

Grocott, J., and Taylor, G.K, 2002, Magmatic arc fault systems, deformation partitioning and emplacement of granitic complexes in the Coastal Cordil-lera, north Chilean Andes (25°30'S to 27°00'S): Journal of the Geological Society, London, v. 159, p. 425–442.

Groves, D.I., Bierlein, F.P., Meinert, L.D., and Hitzman, M.W., 2010, Iron oxide copper-gold (IOCG) deposits through Earth history: Implications for origin, lithospheric setting, and distinction from other epigenetic iron oxide deposits: Economic Geology, v. 105, p. 641–654.

Hamlyn, P.R., Keays, R.R., Cameron, W.E., Crawford, A.J., and Waldron, H.M., 1985, Precious metals in magnesian low-Ti lavas: Implications for metallogenesis and sulfur saturation in primary magmas: Geochimica et Cosmochimica Acta, v. 49, p. 1797–1811.

Harris, A.C., Cooke, D.R., Blackwell, J.L., Fox, N., and Orovan, E.A., 2013, Volcanotectonic setting of world-class alkalic porphyry and epithermal Au

Page 21: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

TECTONIC SETTINGS AND S CONTENTS OF MAGMAS ASSOCIATED WITH CRETACEOUS DEPOSITS, CHILE 315

± Cu deposits of the Southwest Pacific: Society of Economic Geologists, Special Publication, no. 17, p. 337–359.

Harrison, T.M., and Watson, E.B., 1984, The behavior of apatite during crustal anatexis: Equilibrium and kinetic considerations: Geochimica et Cosmochimica Acta, v. 48, p. 1467–1477.

Haschke, M., Siebel, W., Günther, A., and Scheuber, E., 2002, Repeated crustal thickening and recycling during the Andean orogeny in north Chile (21°–26°S): Journal of Geophysical Research, v. 107, p. ECV 6-1–18, doi:10.1029/2001JB000328.

Hasler, K., 2007, Petrogénesis del magmatismo bimodal y metamorfismo de muy bajo grado del Cretácico inferior de la Cordillera de la Costa, Chile Central: Unpublished M.Sc. thesis, Santiago, Chile, Universidad de Chile, 281 p.

Henderson, C.E., 2011, Protocols and pitfalls of electron microprobe analysis of apatite: Unpublished M.Sc. thesis, Ann Arbor, Michigan, University of Michigan, 68 p.

Hildreth, W., and Moorbath, S., 1988, Crustal contributions to arc magma-tism in the Andes of central Chile: Contributions to Mineralogy and Petrol-ogy, v. 98, p. 455–489.

Hitzman, M.W., 2000, Iron oxide-Cu-Au deposits: What, where, when, and why?, in Porter, T.M., ed., Hydrothermal iron oxide copper-gold and related deposits: A global perspective, v. 1: Adelaide, PGC Publishing, p. 9–25.

Hitzman, M.W., Oreskes, N., and Einaudi, M.T., 1992, Geological character-istics and tectonic setting of Proterozoic iron oxide (Cu-U-Au-REE) depos-its: Precambrian Research, v. 58, p. 241–287.

Hunt, J.A., Baker, T., and Thorkelson, D.J., 2007, A review of iron oxide copper-gold deposits, with focus on the Wernecke Breccias, Yukon, Can-ada, as an example of a non-magmatic end member and implications for IOCG genesis and classification: Exploration and Mining Geology, v. 16, p. 209–232.

Imai, A., 2002, Metallogenesis of porphyry Cu deposits of the western Luzon arc, Philippines: K-Ar ages, SO3 contents of microphenocrystic apatite and significance of intrusive rocks: Resource Geology, v. 52, p. 147–161.

Irwin, J.J., García, C., Hervé, F., and Brook, M., 1988, Geology of part of a long-lived dynamic plate margin: The coastal cordillera of north-central Chile, latitude 30°51'–31°S: Canadian Journal of Earth Sciences, v. 25, p. 603–624.

Jugo, P.J., Luth, R.W., and Richards, J.P., 2005, Experimental data on the speciation of sulfur as a function of oxygen fugacity in basaltic melts: Geo-chimica et Cosmochimica Acta, v. 69, p. 497–503.

Keays, R.R., 1987, Principles of mobilization (dissolution) of metals in mafic and ultramafic rocks—the role of immiscible magmatic sulphides in the generation of hydrothermal gold and volcanogenic massive sulphide depos-its: Ore Geology Reviews, v. 2, p. 47–63.

Kimura, J.-I., and Yoshida, T., 2006, Contributions of slab fluid, mantle wedge and crust to the origin of Quaternary lavas in the NE Japan arc: Journal of Petrology, v. 47, p. 2185–2232.

Klein, M., Stosch, H.-G., and Seck, H.A., 1997, Partitioning of high field-strength and rare-earth elements between amphibole and quartz-dioritic to tonalitic melts: An experimental study: Chemical Geology, v. 138, p. 257–271.

Kovacic, P., 2014, Geología y metalogénesis del yacimiento tipo IOCG Casualidad, distrito Sierra Overa, Segunda Región de Antofagasta, Chile: Unpublished M.Sc. thesis, Antofagasta, Chile, Universidad Catolica del Norte, 118 p.

Lang, J.R., and Titley, S.R., 1998, Isotopic and geochemical characteristics of Laramide magmatic systems in Arizona and implications for the genesis of porphyry copper deposits: Economic Geology, v. 93, p. 138–170.

Le Maitre, R.W., Streckeisen, A., Zanettin, B., Le Bas, M.J., Bonin, B., and Bateman, P., 2002, Igneous rocks: A classification and glossary of terms, 2nd edition: Cambridge University Press, 256 p.

Li, J., Li, G., Qin, K., Xiao, B., Chen, L., and Zhao, J., 2012, Mineralogy and mineral chemistry of the Cretaceous Duolong gold-rich porphyry cop-per deposit in the Bangongco arc, northern Tibet: Resource Geology, v. 62, p. 19–41.

López, G., 2012, The El Espino iron-oxide copper gold district, Coastal Cor-dillera of north-central Chile: Unpublished Ph.D. thesis, Golden, Colorado, Colorado School of Mines, 163 p.

López, G.P., Hitzman, M.W., and Nelson, E.P., 2014, Alteration patterns and structural controls of the El Espino IOCG mining district, Chile: Minera-lium Deposita, v. 49, p. 235–259.

Lucassen, F., Franz, G., Thirlwall, M.F., and Mezger, K., 1999, Crustal recy-cling of metamorphic basement: Late Paleozoic granitoids of northern

Chile (~22°). Implications for the composition of the Andean crust: Journal of Petrology, v. 40, p. 1527–1551.

Lucassen, F., Becchio, R., Harmon, R., Kasemann, S., Franz, G., Trumbull, R., Wilke, H.G., Romer, R.L., and Dulski, P., 2001, Composition and den-sity model of the continental crust at an active continental margin—the Central Andes between 21° and 27°S: Tectonophysics, v. 341, p. 195–223.

Lucassen, F., Harmon, R., Franz, G., Romer, R.L., Becchio, R., and Siebel, W., 2002, Lead evolution of the pre-Mesozoic crust in the Central Andes (18–27°): Progressive homogenisation of Pb: Chemical Geology, v. 186, p. 183–197.

Maksaev, V., and Llaumet, C., 2015, Guía de Excursión El Pleito – Mina Dos Amigos: Colegio de Geólogos de Chile, Sociedad Geológica de Chile, Congreso Geologico Chileno, XIV, La Serena, Chile, October 4–8, 2015, Proceedings, 13 p.

Maksaev, V., Munizaga, F., Fanning, M., Palacios, C., and Tapia, J., 2006, SHRIMP U-Pb dating of the Antucoya porphyry copper deposit: New evidence for an Early Cretaceous porphyry-related metallogenic epoch in the Coastal Cordillera of northern Chile: Mineralium Deposita, v. 41, p. 637–644.

Maksaev, V., Munizaga, F., Valencia, V., and Barra, F., 2009, LA-ICP-MS zir-con U-Pb geochronology to constrain the age of post-Neocomian continen-tal deposits of the Cerrillos Formation, Atacama Region, northern Chile: Andean Geology, v. 36, p. 264–287.

Maksaev, V., Almonacid, T.A., Munizaga, F., Valencia, V., McWilliams, M., and Barra, F., 2010, Geochronological and thermochronological constraints on porphyry copper mineralization in the Domeyko alteration zone, north-ern Chile: Andean Geology, v. 37, p. 144–176.

Markey, R.J., Stein, H.J., and Morgan, J.W., 1998, Highly precise Re-Os dating for molybdenite using alkaline fusion and NTIMS: Talanta, v. 45, p. 935–946.

Markey, R.J., Stein, H.J., Hannah, J.L., Selby, D., and Creaser, R.A., 2007, Standardizing Re-Os geochronology: A new molybdenite reference mate-rial (Henderson, USA) and the stoichiometry of Os salts: Chemical Geol-ogy, v. 244, p. 74–87.

Marks, M.A.W., Scharrer, M., Ladenburgerand, S., and Markl, G., 2016, Comment on “Apatite: A new redox proxy for silicic magmas?” [Geochimica et Cosmochimica Acta 132 (2014) 101–119]: Geochimica et Cosmochimica Acta, v. 183, p. 267–270.

Marquardt, M., Cembrano, J., Bissig, T., and Vázquez, C., 2015, Mid Creta-ceous Cu-Au (Mo) mineralization in the Vallenar district: New Re-Os age constraints from Productora deposit, northern Chile: Colegio de Geólogos de Chile, Sociedad Geológica de Chile, Congreso Geologico Chileno, XIV, La Serena, Chile, October 4–8, 2015, Extended abstracts, 4 p.

Marschik, R., and Fontboté, L., 2001, The Punta del Cobre Formation, Punta del Cobre-Candelaria area, northern Chile: Journal of South American Earth Science, v. 14, p. 401–433.

Marschik, R., and Kendrick, M.A., 2015, Noble gas and halogen constraints on fluid sources in iron oxide-copper-gold mineralization: Mantoverde and La Candelaria, northern Chile: Mineralium Deposita, v. 50, p. 357–371.

Matthews, K.J., Seton, M., and Müller, R.D., 2012, A global-scale plate reorganization event at 105–100 Ma: Earth and Planetary Science Letters, v. 355–356, p. 283–298.

Middlemost, E.A.K., 1994, Naming materials in the magma/igneous rock sys-tem: Earth-Science Reviews, v. 37, p. 215–224.

Miles, A.J., Graham, C.M., Hawkesworth, C.J., Gillespie, M.R., Hinton, R.W., Bromiley, G.D., and EMMAC, 2014, Apatite: A new redox proxy for silicic magmas?: Geochimica et Cosmochimica Acta, v. 132, p. 101–119.

——2016, Reply to comment by Marks et al. (2016) on “Apatite: A new redox proxy for silicic magmas?” [Geochimica et Cosmochimica Acta 132 (2014) 101–119]: Geochimica et Cosmochimica Acta, v. 183, p. 271–273.

Morata, D., and Aguirre, L., 2003, Extensional Lower Cretaceous volcanism in the Coastal Range (29°20'–30°S), Chile: Geochemistry and petrogenesis: Journal of South American Earth Sciences, v. 16, p. 459–476.

Morata, D., Féraud, G., Aguirre, L., Arancibia, G., Belmar, M., Morales, S., and Carrillo, J., 2008, Geochronology of the Lower Cretaceous volcanism from the Coastal Range (29°20'–30°S): Chile: Revista Geológica de Chile, v. 35, p. 123–145.

Morelli, P., 2008, Estudio geológico del sistema de alteración hidrotermal de Pajonales, Provincia de Vallenar, Región de Atacama: Unpublished under-graduate thesis, Santiago, Chile, Universidad de Chile, 44 p.

Mumin, A.H., Somarin, A.K., Jones, B., Corriveau, L., Ootes, L., and Camier, J., 2010, The IOCG-porphyry-epithermal continuum of deposit types in

Page 22: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

316 RICHARDS ET AL.

the Great Bear magmatic zone, Northwest Territories, Canada: Geological Association of Canada, Short Course Notes 20, p. 59–78.

Myers, J., and Eugster, H., 1983, The system Fe-Si-O: Oxygen buffer calibra-tions to 1,500K: Contributions to Mineralogy and Petrology, v. 82, p. 75–90.

Nielsen, C.H., and Sigurdsson, H., 1981, Quantitative methods for electron microprobe analysis of sodium in natural and synthetic glasses: American Mineralogist, v. 66, p. 547–552.

Oyarzun, R., Rodriguez, M., Pinchera, M., Doblas, M., and Helle, S., 1999, The Candelaria (Cu-Fe-Au) and Punta del Cobre (Cu-Fe) deposits (Copi-apó, Chile): A case for extension-related granitoid emplacement and miner-alization processes?: Mineralium Deposita, v. 34, p. 799–801.

Palacios, C.M., Townley, B.C., Lahsen, A.A., and Egaña, A.M., 1993, Geo-logical development and mineralization in the Atacama segment of the South American Andes, northern Chile (26°15'–27°25'S): Geologische Rundschau, v. 82, p. 652–662.

Parada, M.A., Nyström, J.O., and Levi, B., 1999, Multiple sources for the Coastal batholith of central Chile (31–34°S): Geochemical and Sr-Nd isoto-pic evidence and tectonic implications: Lithos, v. 46, p. 505–521.

Parada, M.A., Larrondo, P., Guiresse, C., and Roperch, P., 2002, Mag-matic gradients in the Cretaceous Caleu pluton (central Chile): Injections of pulses from a stratified magma reservoir: Gondwana Research, v. 5, p. 307–324.

Parada, M.A., Féraud, G., Fuentes, F., Aguirre, L., Morata, D., and Lar-rondo, P., 2005, Ages and cooling history of the Early Cretaceous Caleu pluton: Testimony of a switch from a rifted to a compressional continental margin in central Chile: Journal of the Geological Society, London, v. 162, p. 273–287.

Parada, M.A., López-Escobar, L., Oliveros, V., Fuentes, F., Morata, D., Calderón, M., Aguirre, L., Féraud, G., Espinoza, F., Moreno, H., Figueroa, O., Muñoz Bravo, J., Vásquez, R.T., and Stern, C.R., 2007, Andean magma-tism, in Moreno, T., and Gibbons, W., eds., The geology of Chile: London, Geological Society, p. 115–146.

Parat, F., and Holtz, F., 2005, Sulfur partition coefficient between apatite and rhyolite: The role of bulk S content: Contributions to Mineralogy and Petrology, v. 150, p. 643–651.

Parat, F., Holtz, F., and Klügel, A., 2011, S-rich apatite-hosted glass inclu-sions in xenoliths from La Palma: Constraints on the volatile partitioning in evolved alkaline magmas: Contributions to Mineralogy and Petrology, v. 162, p. 463–478.

Peng, G., Luhr, J., and McGee, J., 1997, Factors controlling sulfur concentra-tions in volcanic apatite: American Mineralogist, v. 82, p. 1210–1224.

Pettke, T., Oberli, F., and Heinrich, C.A., 2010, The magma and metal source of giant porphyry-type ore deposits, based on lead isotope microanalysis of individual fluid inclusions: Earth and Planetary Science Letters, v. 296, p. 267–277.

Piccoli, P., and Candela, P., 1994, Apatite in felsic rocks: A model for the estimation of initial halogen concentrations in the Bishop Tuff (Long Valley) and Tuolumne intrusive suite (Sierra Nevada batholith) magmas: American Journal of Science, v. 294, p. 92–135.

——2002, Apatite in igneous systems: Reviews in Mineralogy and Geochem-istry, v. 48, p. 255–292.

Pilet, S., Baker, M.B., Müntener, O., and Stolper, E.M., 2011, Monte Carlo simulations of metasomatic enrichment in the lithosphere and implications for the source of alkaline basalts: Journal of Petrology, v. 52, p. 1415–1442.

Pollard, P.J., 2000, Evidence of a magmatic fluid and metal source for Fe-oxide Cu-Au mineralization, in Porter, T.M., ed., Hydrothermal iron oxide copper-gold and related deposits: A global perspective, v. 1: Adelaide, PGC Publishing, p. 27–41.

Raab, A., 2002, Geology of the Cerro Negro Norte Fe-oxide (Cu-Au) district, Coastal Cordillera, northern Chile: Unpublished M.Sc. thesis, Corvallis, Oregon, Oregon State University, 296 p.

Reyes, M., 1991, The Andacollo strata-bound gold deposit, Chile, and its position in a porphyry copper-gold system: Economic Geology, v. 86, p. 1301–1316.

Richards, J.P., 1995, Alkalic-type epithermal gold deposits—a review: Min-eralogical Association of Canada, Short Course Series, v. 23, p. 367–400.

——2003, Tectono-magmatic precursors for porphyry Cu-(Mo-Au) deposit formation: Economic Geology, v. 98, p. 1515–1533.

——2009, Postsubduction porphyry Cu-Au and epithermal Au deposits: Products of remelting of subduction-modified lithosphere: Geology, v. 37, p. 247–250.

——2011a, High Sr/Y arc magmas and porphyry Cu ± Mo ± Au deposits: Just add water: Economic Geology, v. 106, p. 1075–1081.

——2011b, Magmatic to hydrothermal metal fluxes in convergent and col-lided margins: Ore Geology Reviews, v. 40, p. 1–26.

——2015, The oxidation state, and sulfur and Cu contents of arc magmas: Implications for metallogeny: Lithos, v. 233, p. 27–45.

Richards, J.P., and Kerrich, R., 2007, Special paper: Adakite-like rocks: Their diverse origins and questionable role in metallogenesis: Economic Geology, v. 102, p. 537–576.

Richards, J.P., and Mumin, A.H., 2013a, Magmatic-hydrothermal processes within an evolving Earth: Iron oxide-copper-gold and porphyry Cu ± Mo ± Au deposits: Geology, v. 41, p. 767–770.

——2013b, Lithospheric fertilization and mineralization by arc magmas: Genetic links and secular differences between porphyry copper ± molybde-num ± gold and magmatic-hydrothermal iron oxide copper-gold deposits: Society of Economic Geologists, Special Publication, no. 17, p. 277–299.

Rieger, A., Marschik, R., and Diaz, M., 2010, The Mantoverde district, north-ern Chile: An example of distal portions of zoned IOCG systems, in Porter, T.M., ed., Hydrothermal iron oxide copper-gold and related deposits: A global perspective, v. 3: Adelaide, PGC Publishing, p. 273–284.

Ring, U., Willner, A.P., Layer, P.W., and Richter, P.P., 2012, Jurassic to Early Cretaceous postaccretional sinistral transpression in north-central Chile (latitudes 31–32°S): Geological Magazine, v. 149, p. 208–220.

Scheuber, E., and Andriessen, P.A.M., 1990, The kinematic and geodynamic significance of the Atacama fault zone, northern Chile: Journal of Structural Geology, v. 12, p. 243–257.

Scheuber, E., and Gonzalez, G., 1999, Tectonics of the Jurassic-Early Creta-ceous magmatic arc of the north Chilean Coastal Cordillera (22°–26°S): A story of crustal deformation along a convergent plate boundary: Tectonics, v. 18, p. 895–910.

Selby, D., and Creaser, R.A., 2004, Macroscale NTIMS and microscale LA-MC-ICP-MS Re-Os isotopic analysis of molybdenite: Testing spatial restrictions for reliable Re-Os age determinations, and implications for the decoupling of Re and Os within molybdenite: Geochimica et Cosmochi-mica Acta, v. 68, p. 3897–3908.

Sillitoe, R.H., 2003, Iron oxide-copper-gold deposits: An Andean view: Min-eralium Deposita, v. 38, p. 787–812.

——2010, Porphyry copper systems: Economic Geology, v. 105, p. 3–41.Skirrow, R.G., 2010, “Hematite-group” IOCG + U ore systems: Tectonic

settings, hydrothermal characteristics, and Cu-Au and U mineralizing pro-cesses: Geological Association of Canada, Short Course Notes 20, p. 39–59.

Smoliar, M.I., Alexander, C.M.O’D., Walker, R.J., and Jacobesen, S.B., 2006, Re-Os isochron for Allegan (H5): Reconciling Re-Os and U-Pb chronolo-gies: European Association of Geochemistry, Geochemical Society, Annual Lunar and Planetary Science Conference, 37th, League City, Texas, March 13–17, 2006, Abstract no. 1468, 2 p.

Somarin, A.K., and Mumin, A.H., 2012, The Paleo-Proterozoic high heat pro-duction Richardson granite, Great Bear magmatic zone, Northwest Terri-tories, Canada: Source of U for Port Radium?: Resource Geology, v. 62–63, p. 227–242.

Spooner, E.T.C., 1993, Magmatic sulphide/volatile interaction as a mecha-nism for producing chalcophile element enriched, Archean Au-quartz, epithermal Au-Ag and Au skarn hydrothermal ore fluids: Ore Geology Reviews, v. 7, p. 359–379.

Stormer, Jr., J.C., Pierson, M.L., and Tacker, R.C., 1993, Variation of F and Cl X-ray intensity due to anisotropic diffusion in apatite during electron microprobe analysis: American Mineralogist, v. 78, p. 641–648.

Streck, M.J., and Dilles, J.H., 1998, Sulfur content of oxidized arc magmas as recorded in apatite from a porphyry copper batholith: Geology, v. 26, p. 523–526.

Sun, S.-s., and McDonough, W.F., 1989, Chemical and isotopic systematics of oceanic basalts: Implications for mantle composition and processes: Geo-logical Society of London, Special Publication, no. 42, p. 313–345.

Taylor, G.K., Grocott, J., Pope, A., and Randall, D.E., 1998, Mesozoic fault systems, deformation and fault block rotation in the Andean forearc: A crustal scale strike-slip duplex in the Coastal Cordillera of northern Chile: Tectonophysics, v. 299, p. 93–109.

Teck Resources Limited, 2016, 2015 Annual Information Form, 117 p., http://www.teck.com/media/Investors-aif_march_2016_T5.1.2.pdf.

Timm, C., de Ronde, C.E.J., Leybourne, M.I., Layton-Matthews, D., and Graham, I.J., 2012, Sources of chalcophile and siderophile elements in Ker-madec arc lavas: Economic Geology, v. 107, p. 1527–1538.

Tornos, F., Velasco, F., Morata, D., Barra, F., and Rojo, M., 2010, The Trope-zón Cu-Mo-(Au) deposit, northern Chile: The missing link between IOCG and porphyry copper systems?: Mineralium Deposita, v. 45, p. 313–321.

Page 23: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

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-chemistry of tourmaline-rich alteration in the IOCG systems of north-ern Chile: Implications for a magmatic-hydrothermal origin: Mineralium Deposita, v. 47, p. 483–499.

Tosdal, R.M., and Richards, J.P., 2001, Magmatic and structural controls on the development of porphyry Cu±Mo±Au deposits: Reviews in Economic Geology, v. 14, p. 157–181.

Van Hoose, A.E., Streck, M.J., Pallister, J.S., and Wälle, M., 2013, Sulfur evo-lution of the 1991 Pinatubo magmas based on apatite: Journal of Volcanol-ogy and Geothermal Research, v. 257, p. 72–89.

Vergara, M., Levi, B., Nystrom, J.O., and Cancino, A., 1995, Jurassic and Early Cretaceous island arc volcanism, extension, and subsidence in the Coast Range of central Chile: Geological Society of America Bulletin, v. 107, p. 1427–1440.

Wall, R., Selles, D., and Gana, P., 1999, Area Tiltil-Santiago, región Metro-politana: Chile, Sernageomin, scale 1,000,000, 1 sheet.

Wallace, P.J., and Carmichael, I.S.E., 1992, Sulfur in basaltic magmas: Geo-chimica et Cosmochimica Acta, v. 56, p. 1863–1874.

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. 215–246.

Williams, P.J., 2010, “Magnetite-group” IOCGs with special reference to Cloncurry (NW Queensland) and northern Sweden: Settings, alteration, deposit characteristics, fluid sources, and their relationship to apatite-rich iron ores: Geological Association of Canada, Short Course Notes 20, p. 23–38.

Williams, P.J., Barton, M.D., Fontboté, L., de Haller, A., Johnson, D.A., Mark, G., Marschik, R., and Oliver, N.H.S., 2005, Iron oxide copper-gold deposits: Geology, space-time distribution, and possible modes of origin: Economic Geology 100th Anniversary Volume, p. 371–406.

Williams, P.J., Mark, A.K., and Xavier, R.P., 2010, Sources of ore fluid com-ponents in IOCG deposits, in Porter, T.M., ed., Hydrothermal iron oxide copper-gold and related deposits: A global perspective, v. 3, Advances in the understanding of IOCG deposits: Adelaide, PGC Publishing, p. 107–116.

Winchester, J.A., and Floyd, P.A., 1977, Geochemical discrimination of dif-ferent magma series and their differentiation products using immobile ele-ments: Chemical Geology, v. 20, p. 325–343.

Page 24: Contrasting Tectonic Settings and Sulfur Contents of ...€¦ · Contrasting Tectonic Settings and Sulfur Contents of Magmas Associated with Cretaceous Porphyry Cu ± Mo ± Au and

318 RICHARDS ET AL.

Tectonostratigraphic time boundaries in Figure 1b (gray lines) were defined based on correlation of similar formations and their interpreted tectonic settings, as follows:

i. Punta del Cobre, Arqueros, and Veta Negra formations (Charrier et al., 2007);

ii. Contact between Punta del Cobre and Bandurrias forma-tions (Marschik and Fontboté, 2001);

iii. Lower member of Cerrillos Formation (Maksaev et al., 2009);

iv. Ka1/Ka2 boundary, Arqueros Formation (Morata et al., 2008);

v. Quebrada Marquesa/Viñita formations (Emparan and Pineda, 2006);

vi. Arqueros Formation (Ferrando et al., 2014);vii. Quebrada Marquesa Formation (López, 2012);viii. Las Chilcas Formation (Wall et al., 1999).

Black lines represent the time boundaries of different tec-tonic settings based on a compilation of kinematics and tim-ing of movement on structures along the Coastal Cordillera, as follows:

1. Atacama fault system at 25.5° to 27°S (Grocott and Tay-lor, 2002);

2. Atacama fault system at 25° to 27°S (Brown et al., 1993; Taylor et al., 1998);

3. Atacama fault system at 26° to 27.5°S (Dallmeyer et al., 1996);

4. Atacama fault system at 29°S (Arévalo et al., 2003); 5. Dos Amigos fault system (Almonacid, 2007); 6. Bahia Agua Dulce shear zone at 32°S (Ring et al., 2012); 7. El Romeral fault at 30°S (Emparan and Pineda, 2006); 8. Silla del Gobernador fault at 32°S (Arancibia, 2004); 9. Magnetic fabric in Illapel plutonic complex (Ferrando et

al., 2014);10. Dike swarms at 33°S (Creixell et al., 2011);11. Rapid exhumation of intrusions at 33°S (Gana and

Zentilli, 2000).

APPENDIXSources of Data for Figure 1