JOURNAL OF PETROLOGY VOLUME 53 NUMBER 6 PAGES 1091^1122 2012 doi:10.1093/petrology/egs010 ZirconTrace Element and O^Hf Isotope Analyses of Mineralized Intrusions from El Teniente Ore Deposit, Chilean Andes: Constraints on the Source and Magmatic Evolution of Porphyry Cu^Mo Related Magmas M. MUN‹OZ1*, R. CHARRIER1, C. M. FANNING2, V. MAKSAEV1 AND K. DECKART1 UNIVERSIDAD DE CHILE, DEPARTAMENTO DE GEOLOGI¤A, PLAZA ERCILLA 803, CASILLA 13518, CORREO 21, 1 SANTIAGO, CHILE 2 RESEARCH SCHOOL OF EARTH SCIENCES, AUSTRALIAN NATIONAL UNIVERSITY, CANBERRA, ACT 0200, AUSTRALIA RECEIVED DECEMBER 17, 2010; ACCEPTED JANUARY 30, 2012 ADVANCE ACCESS PUBLICATION MARCH 5, 2012 Intrusive rocks related to porphyry copper mineralization are part of the wide diversity of subduction-related, mantle-derived, igneous rocks generated in convergent margin settings. What differentiates them from barren igneous rocks results ultimately from the multi-component and multi-stage processes that condition magma composition in these settings. Unfortunately, the petrogenetic history is largely obscured by the pervasive alteration that affects rocks in these deposits. We address this issue through the study of zircon grains from El Teniente, one of the largest known porphyry Cu^Mo deposits in the world. El Teniente belongs to the Miocene^Pliocene Cu^Mo belt of the Central Chilean Andes, which formed in a short timespan during the Cenozoic constructive period of the orogen. Previously U^Pb dated zircon grains were selected for re-examination of their morphological characteristics and in situ analysis of chemical (rare earth element, Hf, Y and Ti contents) and isotopic (Hf, O) composition. They are from six intermediate to felsic syn- to late-mineralization, intrusive units covering a timespan of 1·6 Myr. The El Teniente zircons have compositional and morphological characteristics indicating crystallization from a series of cogenetic melts. However, a minor hydrothermal imprint is documented in the presence of crystals with mottled surfaces that correspond to thin high U^Th overgrowth rims (low-luminescent features in cathodoluminescence images). In terms of any other chemical and isotopic characteristic, these are indistinguishable from the main mineral populations. Zircons define morphological and chemical trends reflecting an evolution towards more differentiated magma compositions, lower crystallization temperatures and increased cooling rates with decreasing age of intrusion. Hf and O isotopic compositions are remarkably uniform at grain, sample and deposit scale. This, together with the general absence of older inherited zircon components, the lack of correlations between isotopic signature and whole-rock composition and high initial eHf values (total average 7·4 1·2; 2s), rules out involvement of any significant crustal contamination in the genesis of the El Teniente magmas. The Hf isotopic composition indicates a relatively juvenile source, but with some crustal residence time. The d18OZrc weighted mean of 4·76 0·12ø (2s; 61 analyses) is at the lower limit of the normal mantle zircon range of 5·3 0·6ø (2s), and might reflect crystallization from low-18O magmas. Hf isotopic compositions have a restricted range in initial eHf values between þ6 and þ10, identical to preceding Cenozoic barren magmatic activity in Central Chile. These igneous rocks are the product of nearly 25 Myr of subduction-related magmatic activity, developed under contrasting tectonic regimes and margin configurations.This suggests a primary control of the isotopic signature by a stable long-lived MASH-type (melting, assimilation, storage and homogenization) reservoir in the deep lithosphere. In the context of the Cenozoic evolution of Central Chile we argue that dehydration melting in the enriched * Corresponding author. Telephone: 56 2 9784533. Fax: 56 2 6963050. E-mail: marmunoz@ing.uchile.cl The Author 2012. Published by Oxford University Press. All rights reserved. For Permissions, please e-mail: journals.permissions@ oup.com JOURNAL OF PETROLOGY VOLUME 53 MASH reservoir occurred as a consequence of increasing crustal thickness, and was prompted by a high-temperature thermal regime resulting from long-lasting preceding magmatism. This process can also fractionate O to generate low-18O magmas. At the time of El Teniente formation, dehydration melting occurred coevally with arc migration, which probably influenced the fertility of the magmas by increasing the melt component derived from this process relative to the component derived from primary basalt differentiation. At a regional scale, such reactions are expected to occur as a consequence of progressive crustal thickening during the constructive period of the Andes, and can explain the simultaneous generation of porphyry deposits in the Miocene^Pliocene Cu^Mo belt of Central Chile. El Teniente; porphyry Cu^Mo deposit; zircon; O^Hf isotopes; dehydration melting KEY WORDS: I N T RO D U C T I O N The relation between porphyry Cu^Mo deposit formation and active continental margin magmatism has been well documented (e.g. Lindgren, 1933; Sillitoe, 1972; Burnham, 1979; Cline & Bodnar, 1991; Hedenquist & Lowenstern, 1994). The magmas are the main source of H2O, S, and Cu, among other elements and compounds, of the hydrothermal systems whose evolution results in the formation of these ore deposits. However, whereas magmatic and hydrothermal processes occur widely in subduction-related arc settings, the formation of large economic porphyry copper deposits is restricted. These deposits constitute localized chemical and mineralogical anomalies, formed during a relatively short timespan and at specific moments during the lifetime of the host magmatic arc (Maksaev & Zentilli, 1988; McKee & Noble, 1989; Cornejo et al., 1997; Richards et al., 2001). The many studies performed on these deposits have well characterized the numerous tectonic, structural, magmatic and chemical conditions optimal for their formation. However, the genesis of magmas related to porphyry copper mineralization remains a highly debated issue. Different models have been proposed, ranging from those invoking primary enrichment owing to key processes during magma genesis (Kay & Mpodozis, 2001; Oyarzu¤n et al., 2001; Mungall, 2002; Core et al., 2006; Shafiei et al., 2009), to those considering them as the products of the convergence of normal processes operating in arcs (Richards, 2003, 2005; Stern & Skewes, 2005; Chiaradia et al., 2009; Stern et al., 2010). These models are not necessarily exclusive, but further research is needed to understand more accurately the processes and/or components involved in the genesis of porphyry copper related magmas at their source, as well as during their subsequent evolution in their passage through the upper lithosphere. The main problem in studying intrusive rocks in such deposits is the widespread pervasive hydrothermal NUMBER 6 JUNE 2012 alteration that has modified their primary textural and chemical characteristics. This has greatly restricted reliable characterization of geochemical and isotopic primary signatures by conventional methods. However, the study of single minerals, made possible by the development of microanalytical techniques over the past few decades, allows an insight into the primary characteristics of heavily altered rocks by selectively avoiding the effects of whole-rock alteration. Zircon is a common accessory mineral particularly well suited for this kind of study. The physical and chemical stability of zircon, its resistance to high-temperature diffusive re-equilibration (Watson & Cherniak, 1997; Cherniak & Watson, 2003), and its tendency to incorporate numerous trace and radiogenic elements make it an ideal mineral to see through the subsequent alteration commonly seen on the whole-rock scale. Additionally, zircon is abundant in intermediate to felsic igneous rocks such as those related to porphyry copper deposits, making it a valuable tool to track and characterize the petrogenetic evolution of the magmas from which it crystallized. The Chilean continental margin is ideal to examine different aspects of porphyry copper systems as it hosts numerous deposits of this type. Its geological evolution has been largely linked to abundant intrusive and volcanic activity as a consequence of plate convergence. The current volcanic arc is located along the axis of the main range and is represented by the Chilean Northern and Southern volcanic zones (Fig. 1). Trenchwards, igneous rocks cropping out in different north^south-trending belts are the remnants of several arcs developed along the continental margin since Paleozoic times (Mpodozis & Ramos, 1989; Charrier et al., 2007), some of which host numerous porphyry copper deposits. Their formation is restricted to a short time interval, during the last stages of the related arc lifespan, characterized by the waning of widespread intrusive and volcanic activity in a regime of crustal shortening, thickening, and uplifting (for reviews see Camus, 2003; Maksaev et al., 2007; Charrier et al., 2009). The El Teniente porphyry Cu^Mo deposit is the youngest known deposit of this type within the Chilean continental margin and one of the largest in the world, with 93·7 Mt (megatonnes) of copper, based on current resources plus past production (16 756 Mt at 0·558% Cu; CODELCO, 2010). It belongs to the north^south-trending Neogene metallogenic belt of Central Chile, which extends along the western slope of the Chilean Andes (32^34·58S, Fig. 1). This belt includes other giant ore deposits and constitutes one of the most richly endowed copper provinces in the world, with more than 220 Mt of contained Cu (Camus, 2003; Antofagasta plc, 2009; CODELCO, 2010). The objective of the present study is to track the petrogenetic evolution of El Teniente related magmas, from the source to emplacement levels, as recorded by zircon 1092 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Fig. 1. (a) Schematic map showing the main tectonic features of the southeastern Nazca Plate and Chilean continental margin. The map shows the distribution of volcanoes constituting the Northern and Southern Chilean volcanic zones (CVZ, SVZ), the volcanic gap between 288 and 338S, the location of porphyry Cu^Mo deposits of the Central Chile Neogene Metallogenic Belt (black stars), and the Wadati^Benioff zone contours of the convergent zone (dotted line, Isacks, 1988). LP, Los Pelambres; RB-LB, R|¤o Blanco^Los Bronces; ET, El Teniente; RR, Rosario de Rengo. (b) Geological map showing the distribution of the main lithological units and structural features of the central Chilean^ Argentinean Andes where El Teniente and other porphyry Cu^Mo deposits are emplaced. Upper inset shows the distribution of main morphostructural units forming the Andean orogen in this area. Circled numbers indicate the location of the areas discussed in the text: 1, Abanico and Farellones formations NE of the R|¤o Blanco^Los Bronces deposit; 2, San Francisco batholith and Yerba Loca pluton; 3, La Gloria pluton and Cerro Meso¤n Alto stock. PFZ, Pocuro Fault Zone; AFTB, Aconcagua Fold and Thrust Belt. 1093 JOURNAL OF PETROLOGY VOLUME 53 grains and to frame this evolution within the continental margin global geodynamic setting during ore deposit formation. For this purpose, isotopic (O, Hf) and trace element [Ti, Y, Hf, rare earth elements (REE)] compositions have been determined in zircons with previously known U^Pb ages and U^Th contents (Maksaev et al., 2004). These analyses are complemented by a study of the external morphology and internal structure of the zircons by standard optical methods and cathodoluminescence (CL) images. The main results reveal patterns of magmatic evolution and a common source for the different intrusive pulses of the El Teniente deposit. This source is indistinguishable from that of the preceding barren magmatic activity in the region. We argue that this is a consequence of long-lived MASH-type processes, as originally defined by Hildreth & Moorbath (1988), where ascending, subduction-related, mantle-derived magmas are hybridized in deep lithospheric zones of melting, assimilation, storage and homogenization. Such processes can control the isotopic characteristics of the magmatism during extended periods of time and throughout the contrasting tectonic regimes within the margin, as observed for this portion of the Andean range. We argue that El Teniente mineralization-related fertile magmas are a mixture of melt components derived from dehydration melting in the enriched MASH reservoir and from primary basalt differentiation. Additionally, we show that this process can fractionate O to generate low-d18O magmas, and finally we discuss how dehydration melting at the base of a thickened crust favors the formation of giant porphyry copper deposits simultaneously and in a regional context. G E N E R A L B AC KG RO U N D Regional geology and geodynamic setting The three major known porphyry Cu^Mo deposits of the Central Chile Neogene Metallogenic Belt are located between 328 and 348S: Los Pelambres^El Pacho¤n (328S), R|¤o Blanco^Los Bronces (33808’S), and El Teniente (34804’S; Fig. 1a). These deposits formed between late Miocene and Pliocene times, the northernmost one being older (10^12 Ma; Perello¤ et al., 2009) than the remaining two, which are considered coeval (6^4 Ma; Maksaev et al., 2004; Deckart et al., 2005). They are distributed along the western slope of the Andean range through two morphostructurally different segments of the continental margin separated at 338S. This latitude also coincides with the current locus of subduction of the Juan Ferna¤ndez Ridge, the southern limit of the flat-slab subduction segment (27^338S) and the beginning of the Chilean Southern Volcanic Zone (33^468S; Fig. 1a). The western slope of the Andean Principal Cordillera in Central Chile is dominated by Cenozoic igneous rocks distributed along an 60 km wide north^south-trending belt (Fig. 1b). Older units are exposed to the east, Mesozoic NUMBER 6 JUNE 2012 sedimentary sequences crop out near the Chilean^ Argentinean border and Triassic volcanic rocks and crystalline Paleozoic basement compose the Argentinean Frontal Cordillera (Fig. 1b). In marked contrast to other metallogenic belts in the Chilean continental margin, major trench-parallel structures spatially related to the Neogene Metallogenic Belt appear to be absent. The main structural systems developed in this part of the Andean range are the west-vergent Pocuro^San Ramo¤n Fault, bounding the Principal Cordillera to the west, and the east-vergent Aconcagua Fold and Thrust Belt affecting the Mesozoic deposits near the Chilean^Argentinean border (Fig. 1b). However, local structures have been described for each deposit and have been related to reactivation of pre-Cenozoic basement structures during the development of the Andean Cordillera (e.g. Rivera & Falco¤n, 2000). The Central Chilean Andes Cenozoic magmatic rocks are the product of a prolonged and intense period of arc-related igneous activity lasting from 536 to 6 Ma. They form a nearly 5500 m thick volcanic^volcaniclastic sequence that makes up the Farellones and Abanico (¼ Coya Machal|¤ ) formations (Charrier et al., 2002). The Abanico Formation was deposited during Oligocene^early Miocene times in a tholeiitic arc setting, overlying an 30^35 km thick continental crust, during basin development under crustal extension (Charrier et al., 2002; Kay et al., 2005). Continued volcanic activity after basin inversion, between 21 and 15 Ma (Charrier et al., 2002), led to deposition of the Farellones Formation during the Miocene. Igneous rocks formed during this time show a progressively more calc-alkaline affinity with respect to the preceding magmatism and were formed in an arc setting over a progressively thickening crust. This has been inferred to have reached no more than 45^50 km (Charrier et al., 2002; Kay et al., 2005), the current estimated crustal thickness under the Central Chilean Cenozoic magmatic belt (Tassara et al., 2006). Diminished magmatic activity followed these episodes and is represented in numerous isolated intrusive bodies and less abundant volcanic rocks throughout the region. Overall, these rocks young to the east until reaching the current active volcanic zone near the Chilean^Argentinean border, revealing the progressive arc migration that followed Farellones Formation deposition (Stern & Skewes, 1995; Kay et al., 2005). The Abanico basin inversion during early to middle Miocene times marks the onset of the constructive period of the Andean orogen in Central Chile, characterized by shortening, thickening, and uplifting processes (Mpodozis & Ramos, 1989; Giambiagi & Ramos, 2002; Kay et al., 2005; Far|¤ as et al., 2008, 2010). During this period, shortening was accommodated by different structural systems and migrated to the east in three stages: (1) early to middle Miocene inversion of basin-bounding normal 1094 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS faults currently bounding Abanico Formation outcrops; (2) middle to late Miocene development of the Aconcagua Fold and Thrust Belt; (3) late Miocene to Pliocene activity of high-angle reverse faults that uplifted the Frontal Cordillera in Argentina and a series of out-of-sequence thrusting in the eastern Principal Cordillera (Fig. 1b; Mpodozis & Ramos, 1989; Charrier et al., 2002; Giambiagi & Ramos, 2002; Fock et al., 2006). Concomitant progressive crustal thickening during this evolution has been inferred mostly from the geochemical signatures of coeval igneous rocks from the Abanico and Farellones formations (Kay & Mpodozis, 2002; Kay et al., 2005). Overall uplift of the Andean Principal Cordillera occurred during late Miocene^Pliocene times, resulting in an accumulated 2 km of uplift in 2 Myr during an event taking place sometime between 10 and 4 Ma (Far|¤ as et al., 2008). Additionally, high exhumation rates have been recognized locally and regionally in the area during this same time interval, particularly in the Western Principal Cordillera south of 338S where the R|¤o Blanco^Los Bronces and El Teniente deposits are located (Skewes & Holmgren, 1993; Maksaev et al., 2009). El Teniente porphyry Cu^Mo deposit El Teniente is genetically linked to late Miocene^early Pliocene magmatic^hydrothermal processes (Howell & Molloy, 1960; Cuadra, 1986; Skewes et al., 2002; Camus, 2003). The deposit is hosted by a mafic volcano-plutonic complex, known as the Teniente Mafic Complex, composed of pervasively altered dark grey to black basalt, andesite, diabase sills, and gabbro intrusions forming a 450 km3 laccolith emplaced in the Farellones Formation (Fig. 2; Lindgren & Bastin, 1922; Skewes et al., 2002). Intensive alteration has prevented an accurate age determination for this intrusive unit. However, an apatite fission-track age of 8·9 2·8 Ma has been obtained for corresponding rocks near the mine (all ages are indicated with 2s error level; Skewes et al., 2002; Maksaev et al., 2004), which agrees with K^Ar ages ranging between 12 and 6 Ma for the Farellones Formation in the region (Cuadra, 1986; Kay et al., 2005). A series of felsic to intermediate stocks and dikes intruded the Teniente Mafic Complex between 6·5 and 2·9 Ma (Fig. 2; Cuadra, 1986; Maksaev et al., 2004). They are quartz-diorites, tonalites and granodiorites, with subordinate diorites and hornblende-rich andesitic dykes. The main intrusive igneous events are represented by intrusions forming the Sewell Stock, A Porphyry, Central and Northern diorites, Teniente Porphyry, and Late Dacite and Hornblende dikes (these correspond to informal names widely adopted in the literature; Fig. 2). The youngest igneous activity within the deposit is represented by the postmineralization and alteration-free Late Hornblende Dikes (3·8^2·9 Ma: Cuadra, 1986; Maksaev et al., 2004). Multiple magmatic and hydrothermal breccia complexes complete Fig. 2. Geology of the El Teniente 4 LHD level (2354 m) from the El Teniente mine. Locations of the samples discussed from main intrusive units are indicated with white diamonds along with their corresponding labels and zircon U^Pb age (2s; Maksaev et al., 2004). Two ages are shown for samples with a bimodal distribution; the older ages correspond to the dominant age peak. the deposit geology (Fig. 2). The Braden Pipe, the largest breccia and main lithological feature, is an inverted cone-shaped, weakly mineralized, diatreme body composed of two facies (Fig. 2). These represent late to post-mineralization events dated between 4·4 and 4·8 Ma (Cuadra, 1986; Maksaev et al., 2004). Hypogene mineralization at El Teniente is mostly distributed within a dense vein stockwork and a variety of magmatic^hydrothermal breccias (Camus, 1975; Cuadra, 1986; Skewes et al., 2002). Alteration assemblages have been classically divided into three main hypogene stages; however, this is an oversimplified scheme as a variety of localized hydrothermal events have been identified, reflecting an evolution by multi-stage processes (Cuadra, 1986; Skewes et al., 2002, 2005; Maksaev et al., 2004; Cannell et al., 2005; Vry et al., 2010). The superimposition of multiple, discrete, magmatic and hydrothermal events has led 1095 JOURNAL OF PETROLOGY VOLUME 53 to complex alteration and mineralization patterns; however, this is also probably responsible for the high metal concentrations in El Teniente (Skewes et al., 2002, 2005; Maksaev et al., 2004; Cannell et al., 2005; Vry et al., 2010). Isotopic evidence has shown that the main budget of metals, water and sulfur is of magmatic origin (Kusakabe et al., 1984, 1990; Skewes et al., 2001). However, in the last decade, there has been an intensive debate over the agents responsible for their introduction into the deposit (Skewes et al., 2002, 2005; Maksaev et al., 2004; Cannell et al., 2005, 2007; Stern & Skewes, 2005; Skewes & Stern, 2007; Vry et al., 2010). Since the early investigations, El Teniente has been considered to be a typical porphyry deposit in terms of its alteration assemblages, vein and breccia style, and spatial and temporal relationships between Cu^Mo mineralization and felsic porphyries (e.g. Howell & Molloy, 1960; Camus, 1975; Cuadra, 1986). The felsic porphyries are considered to have been conduits for mineralizing fluids sourced from a deeper-level magma chamber (e.g. Cannell et al., 2005; Klemm et al., 2007; Vry et al., 2010). Skewes et al. (2002) argued that the felsic intrusions correspond to small, late, copper-poor stocks that merely redistributed earlier copper mineralization that had already been introduced by previously unmapped early formed breccia pipes. This mineralization would ultimately originate as a fluid discharge from an unexposed evolving magma chamber of batholithic dimensions (Skewes et al., 2005; Stern & Skewes, 2005; Stern et al., 2010). The differing interpretations arise mainly as a consequence of the complexity of the mineralization and alteration patterns, together with the variable interpretations of the time relationships for the intrusive events as determined by the numerous K^Ar, Ar^Ar, Re^Os, and U^Pb ages reported. However, these studies agree that ultimately the breccias, veining, and intrusions are linked to a deep magma chamber located below the mine level and are thus derived from the evolution of a common magmatic system. Recently published data on alteration and mineralization patterns and their relation to intrusive events have been used to argue strongly in favor of the porphyry-style model, indicating that El Teniente represents a nested, but otherwise typical, porphyry Cu^Mo deposit (Vry et al., 2010). Sampled units and previous zircon U^Pb age data We address the petrogenesis of the intrusive rocks related to the El Teniente deposit through new morphological, chemical and isotopic studies of selected single zircon grains that have previously been dated [the complete U^Pb dataset has been given by Maksaev et al. (2004); a selection of these data relevant for this study is presented in Table 1]. The zircon grains are from six intrusive units that cover a timespan of 1·6 Myr and are located inside the mine, within the limits of the ore body, that is the NUMBER 6 JUNE 2012 Sewell Stock, the A Porphyry, the Northern and Central diorites, the Teniente Porphyry, and a Late Dacite Dike (Fig. 2). The Sewell Stock, located in the southeastern part of the deposit (Fig. 2), is the oldest intrusion and largest (30 km3) compared with the younger intrusive bodies (51km3). It shows two textural varieties with transitional contacts, suggesting emplacement as a composite intrusion (Faunes, 1981). The A Porphyry and the Central and Northern diorites are thin, cylindrical to irregularly shaped intrusions located in the southeastern and eastern portion of the deposit (Fig. 2). The Teniente Porphyry is a north^south-trending tabular stock whose southern edge has been truncated by the Braden Breccia (Fig. 2). Similar to the Sewell Stock, two textural varieties have been identified in the Teniente Porphyry and have been attributed to at least two independent intrusive pulses from a common magmatic source (Rojas, 2003). The Late Dacite Dike belongs to a series of 2^15 m wide felsic dikes located mainly in the SW and NE part of the deposit. They occur as concentric dikes surrounding the Braden Breccia and also as NE^SW- to NW^SE-trending planar dikes (Fig. 2). Compositionally, all these intrusive units are felsic dacitic igneous rocks ranging between 60 and 69 wt % SiO2, except for the relatively more mafic A Porphyry, which has an andesitic composition with SiO2 contents between 56 and 62 wt % (Rojas, 2003; Cannell et al., 2005; Gonza¤lez, 2006; Hitschfeld, 2006; Stern et al., 2007, 2010). They show strongly fractionated REE patterns with La/ YbN 9^44 along with high Sr/Y 24^253. These ‘adakite-like’ characteristics are shared by a few coeval intrusive rocks exposed near the deposit (Rabbia et al., 2000; Reich, 2001), but are otherwise absent in any of the earlier or later igneous rocks in the region (e.g. Kay et al., 2005). As with Central Chile Cenozoic magmatism in general, intrusive rocks from El Teniente are considered primarily as being derived from subduction-related, mantle-derived magmas. ‘Adakite-like’ characteristics have been attributed by Kay et al. (2005) to a combination of source contamination by subduction erosion and incorporation of a component derived from melting at the base of the thickened lower crust. Alternatively, Stern & Skewes (2005) and Stern et al. (2010) argued that they result mainly from the fractionation of igneous phases and extensive fluid transfer to the top of a crystallizing batholithic-size magma body. Zircon U^Pb age determinations have constrained crystallization ages that were previously biased by the use of K^Ar and Ar^Ar chronometers in the pervasively altered rocks of El Teniente (Fig. 2; Maksaev et al., 2004). However, the final interpretation of these ages has been open to debate (Cannell et al., 2005, 2007; Skewes et al., 2005, 2007). Maksaev et al. (2004) showed that zircon U^ Pb ages for the Sewell Stock, A Porphyry and Northern and Central diorites have a bimodal distribution with peaks at 6·4^6·1Ma and 5·6^5·4 Ma for the dominant and 1096 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Table 1: Sample locations and U,Th,Th/U and U^Pb age data for the zircon spots from the El Teniente deposit analyzed in this study (data from Maksaev et al., 2004) Spot Crystal U Th sector (ppm) (ppm) Th/U Age 2s Spot (Ma) A Porphyry, Sample TT150 Crystal U Th sector (ppm) (ppm) Th/U Age 2s (Ma) Northern Diorite, Sample TT102 1.1 r 62 48 0·78 6·6 0·8 1.1 r 80 37 0·46 6·4 0·6 3.1 r 76 62 0·82 6·8 0·6 2.1 c 55 37 0·68 6·7 0·8 4.1 r 73 60 0·82 6·4 0·6 3.1 r 174 133 0·76 5·9 0·6 5.2 r 64 52 0·81 5·7 0·8 4.1 r 95 51 0·53 5·9 0·6 6.1 r 542 920 1·70 6·5 0·2 6.1 c 185 148 0·80 9·0 0·8 7.1 r 53 32 0·60 6·1 0·8 8.1 r 138 86 0·62 5·9 0·6 11.1 c 36 18 0·51 6·7 1·0 8.2 c 511 235 0·46 27·8 0·6 12.1 r 46 33 0·73 6·8 1·0 11.1 c 381 252 0·66 79·4 1·4 13.1* r 1192 459 0·39 6·4 0·2 11.2 r 1569 564 0·36 6·1 0·2 13.2 c 69 58 0·84 6·2 0·6 12.2 c 60 34 0·56 5·3 0·8 13.1 c 566 209 0·37 6·3 0·4 Sewell Stock, Sample TT101 Teniente Porphyry, Sample TT94 1.1 r 3616 2782 0·77 6·4 0·4 1.1 r 154 107 0·70 5·4 0·4 1.2 c 45 24 0·55 6·0 0·8 2.1 r 146 117 0·80 5·3 0·4 3.1 r 89 41 0·46 5·8 0·6 4.1 r 97 56 0·57 4·7 0·6 3.2 c 68 54 0·79 6·3 0·8 5.1 r 197 93 0·47 4·8 0·4 4.2 c 53 42 0·78 6·3 0·8 6.1 r 142 86 0·61 5·6 0·4 7.1* r 2587 1101 0·43 6·1 0·2 6.2 c 115 87 0·75 5·0 0·8 7.2 c 93 71 0·77 5·6 0·6 7.1 r 239 212 0·88 5·4 0·4 9.1* r 3848 2519 0·65 6·0 0·3 8.1 c 255 318 1·25 5·1 0·4 9.2 c 155 158 1·02 6·3 0·4 9.1 c 262 241 0·92 6·0 0·4 10.1 r 362 298 0·82 6·2 0·4 9.2 r 300 183 0·61 5·4 0·4 12.1 r 86 34 0·40 5·2 1·0 15.1 r 178 104 0·59 5·2 0·4 17.1 r 206 182 0·88 5·3 0·4 Central Diorite, Sample TT90 Late Dacite Dike, Sample TT91 3.1 c 85 45 0·53 6·4 1·0 2.1 r 362 275 0·76 4·5 0·2 3.2 r 57 34 0·61 4·8 1·0 3.1 r 485 434 0·89 5·2 0·4 5.1 c 85 78 0·91 5·1 0·8 4.1 r 227 160 0·70 4·7 0·4 5.2 r 71 45 0·63 5·6 0·6 4.2 c 77 42 0·54 6·2 0·6 7.1 r 88 30 0·34 7·0 0·6 5.1 r 391 264 0·68 5·0 0·2 8.1 r 62 50 0·82 5·7 0·6 7.1 r 274 172 0·63 4·4 0·2 11.1 c 54 33 0·61 6·8 0·8 8.1 r 295 180 0·61 4·7 0·2 11.2 r 45 22 0·50 6·0 0·8 10.1 r 356 280 0·79 4·7 0·2 12.1 r 98 62 0·63 5·7 0·6 12.1 r 272 203 0·75 4·9 0·4 14.1 r 52 42 0·81 6·5 0·8 15.1 r 263 177 0·67 4·8 0·4 17.1 r 43 26 0·61 6·5 1·4 17.1 r 215 132 0·61 4·8 0·4 Sample TT150: zircon U–Pb age with bimodal distribution, 6·46 0·11 Ma and 5·67 0·19 Ma for the dominant and subordinate age groups, respectively. DDH-1337, 384 m; sample level 2044 m; northing 110N/easting 1735E. Sample TT102: zircon U–Pb age with bimodal distribution, 6·11 0·13 Ma and 5·49 0·19 Ma for the dominant and subordinate age groups, respectively. Level Teniente 6 UCL, 2161 m; northing 1016N/easting 1110 E. Sample TT101: zircon U–Pb age with bimodal distribution, 6·15 0·08 Ma and 5·59 0·17 Ma for the dominant and subordinate age groups, respectively. Level Teniente 4, 2347 m; northing –265N/easting 1365E. Sample TT94: zircon U–Pb age with unimodal distribution, 5·28 0·10 Ma. Level Teniente 6, 2161 m; northing 1050N/easting 450E. Sample TT90: zircon U–Pb age with bimodal distribution, 6·28 0·16 Ma and 5·50 0·24 Ma for the dominant and subordinate age groups, respectively. Level UCL Esmeralda, 2192 m; northing 250N/easting 1325E. Sample TT91: zircon U–Pb age with unimodal distribution, 4·82 0·09 Ma. Level UCL Esmeralda, 2192 m; northing 310N/easting 1030E. c, core; r, rim. *Weakly luminescent overgrowth rims observed in CL images. 1097 JOURNAL OF PETROLOGY VOLUME 53 subordinate age populations, respectively. Older ages have been interpreted to correspond to the intrusion age and younger ones to be related to hydrothermal activity. The remaining units show unimodal age distributions with peaks at 5·28 0·19 Ma for the Teniente Porphyry and 4·82 0·09 Ma for the Late Dacite Dike. A N A LY T I C A L T E C H N I Q U E S Zircon separation was carried out at the University of Chile mineral separation facility, and further selection along with imaging studies and compositional analyses were performed at the Research School of Earth Sciences of the National Australian University (RSES-ANU). Zircon grains were separated from total rock samples using standard crushing, washing, heavy liquid (specific gravity 2·96 and 3·3), and paramagnetic procedures. Hand selected zircon grains were placed onto double-sided tape, mounted in epoxy together with chips of the reference zircons (Temora and SL13), sectioned approximately in half, and polished. Reflected and transmitted light photomicrographs were prepared for all zircons, as were CL scanning electron microscope (SEM) images. These CL images were used to decipher the internal structures of the sectioned grains and to ensure that the 20 mm sensitive high-resolution ion microprobe (SHRIMP) spot was wholly within a single age component within the sectioned grains. REE data were acquired using SHRIMP II in spots adjacent to those analyzed for U^Pb^Th geochronology and belonging to the same crystal sector according to the CL images. The energy filtering method was used to reduce isobaric interferences (Ireland & Wlotzka, 1992). Operating conditions and data reduction methods are similar to those described by Hoskin (1998). REE detection limits are in the vicinity of 0·01ppm for analysis spots that are 30 mm across and a few micrometers deep. The in situ Ti analyses were also made using SHRIMP II in a separate session using methods similar to those described by Hiess et al. (2008). Previous U^Pb^Th analytical spots were lightly polished, then the same area within the grains was analyzed. Oxygen isotope analyses were made using the SHRIMP II fitted with a Cs source and electron gun for charge compensation following methods described by Ickert et al. (2008). The SHRIMP U^Pb, REE and Ti analytical spots, craters approximately 20 mm in diameter by 1^2 mm deep, were polished from the mount surface. The oxygen isotope analyses were then made on exactly the same location as used for the U^Pb analyses. Oxygen isotope ratios were determined in multiple collector mode using an axial continuous electron multiplier (CEM) triplet collector, and two floating heads with interchangeable CEM^Faraday cups. The Temora II reference zircon was analyzed to monitor and correct for isotope fractionation. The NUMBER 6 JUNE 2012 measured 18O/16O ratios and calculated d18O values have been normalized relative to a Temora II weighted mean d18O value of þ8·2ø (Ickert et al., 2008). Reproducibility for the Temora II reference zircon d18O value was 0·551 and 0·715ø (2s uncertainty) for the analytical sessions. As a secondary reference, zircons from the Duluth Gabbro sample FC1 analyzed in the same analytical sessions gave a d18O value of 5·405 0·348ø and 5·415 0·615ø (2s uncertainty), in agreement with data reported by Ickert et al. (2008). Lu^Hf isotopic measurements were conducted by laser ablation multi-collector inductively coupled plasma mass spectroscopy (LA-MC-ICPMS) using the RSES Neptune MC-ICPMS coupled with a 193 nm ArF Excimer laser; similar to procedures described by Munizaga et al. (2008). Laser ablation analyses were performed on the same locations within single zircon grains as used for both the U^ Pb and oxygen isotope analyses. For all analyses of unknowns or secondary standards, the laser spot size was c. 47 mm in diameter. The mass spectrometer was first tuned to optimal sensitivity using a large grain of zircon from the Monastery kimberlite. Isotopic masses were measured simultaneously in static-collection mode. A gas blank was acquired at regular intervals throughout the analytical session (every 10 analyses). The laser was fired with typically 5^8 Hz repetition rate and 60 mJ energy. Data were acquired for 100 s, but in many cases only a selected interval from the total acquisition was used in data reduction. Throughout the analytical session mostly FC1 and other widely used reference zircons (91500, Temora-2, Monastery and Mud Tank) were analyzed to monitor data quality. FC1 gave a weighted mean 176Hf/177Hf fractionation corrected ratio of 0·282173 12 (2s uncertainty) for 11 analyses, which is within uncertainty of reported solution values (Woodhead & Hergt, 2005; Vervoort, 2010). Signal intensity was typically c. 5^6 V for total Hf at the beginning of ablation, and decreased over the acquisition time to 2 V or less. Isobaric interferences of 176Lu and 176Yb on the 176Hf signal were corrected by monitoring signal intensities of 175Lu and 173Yb, 172Yb and 171Yb. The calculation of signal intensity for 176Hf also involved independent mass bias corrections for Lu and Yb. R E S U LT S Zircon crystal morphology and internal structure Zircon grains from the six studied units have mainly euhedral external morphologies and zoned inner structures typical of igneous zircons (Fig. 3). Overall, the crystals show the development of prism ({100} and {110}) and pyramidal faces ({211} and {101}), and internal oscillatory and sector zoning. However, there are marked variations of these characteristics between units. Disruptions of the 1098 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Fig. 3. Cathodoluminescence and corresponding transmitted light images of zircon grains from the El Teniente deposit. Intrusive units correspond to: (a) A Porphyry; (b) Sewell Stock; (c) Central Diorite; (d) Northern Diorite; (e) Teniente Porphyry; (f) Late Dacite Dike. Illustrated zircons are representative of the main morphological variations within each unit. Circles correspond to the U^Pb age spot location and analysis labels are indicated in the transmitted light images. (1) zircons with high U^Th, weakly luminescent rims, (2) inherited cores. A clear mottled texture is developed on the surfaces of grains (a) 11; (b) 4, 9, 3, 7, 1, 29; (c) 8, 1, 5; (d) 5, 6, 8. inner structure, such as internal cracks or resorption textures, are absent and inherited cores are extremely scarce (Table 1). Only a few grains show features suggesting a hydrothermal imprint (Fig. 3a and d). All these characteristics are detailed in the following discussion, where the intrusive units have been grouped according to the common morphological and structural parameters of their zircon populations. A Porphyry Zircons of the A Porphyry are the largest among the El Teniente zircons. They are commonly c. 100^300 mm equidimensional to stubby euhedral crystals, with aspect ratios mostly between 1 and 1·5 (Figs 3a and 4). CL images show the development of weak euhedral oscillatory zoning overprinted by strong sector zoning (Fig. 3a). Their morphology is characterized by well-developed {100} and {110} 1099 JOURNAL OF PETROLOGY VOLUME 53 NUMBER 6 JUNE 2012 prism forms and the presence of two pyramids, with the {101} form being predominant over the {211} form (Figs 3a and 4). Subordinately, crystals show mottled surfaces, observed in transmitted light images, that correspond to irregular crystal rims of low luminescent contrast in CL images (Fig. 3a: grain 11). These features are associated with high U and Th contents, but not with any other particular chemical or isotopic composition, or U^Pb age, among those determined in the whole population of A Porphyry zircon grains. Sewell Stock, Central Diorite and Northern Diorite These units have similar zircon populations that are less homogeneous in morphological types and size distribution than zircons from the rest of the units. Crystals are euhedral to subhedral with short prismatic and subordinately equidimensional habits and show oscillatory and sector zoning (Fig. 3b^d). They are of 100^250 mm length and 100^150 mm width, with aspect ratios between 1 and 3·5 (Fig. 4). Morphological types are varied, although overall there is a similar development of both prismatic and pyramidal forms (Figs 3b^d and 4). Several grains show euhedral to rounded cores that preserve inner, zoned, igneous structures. Their rims are oscillatory zoned euhedral overgrowths, which are usually brighter under CL (Fig. 3b: grain 3; Fig. 3c: grains 3 and 5; Fig. 3d: grains 5, 6 and 8). Chemical, isotopic and age determinations in both of these crystal sectors show no discernible difference, except in the Northern Diorite where three inherited cores are identified by their significantly older ages (Fig. 3d: grain 8). As in the A Porphyry, some grains from the Sewell Stock show a mottled texture on the surfaces (Fig. 3b: all grains), observed as weakly luminescent rims in CL images, which are associated with high U^Th contents. Teniente Porphyry and Late Dacite Dike Zircons from these units are euhedral, elongated, prismatic crystals of 100^250 mm length and 80^100 mm width (Fig. 3e and f). Aspect ratios are mostly in the range of 1·5^2·5 and 2^3 for the Teniente Porphyry and the Late Dacite dike, respectively (Fig. 4). Zircon grains are relatively simple with a single continuous pattern of oscillatory and/or sector zoning throughout the crystal (Fig. 3e and f). Morphological types are characterized by the {110} prismatic form and the predominance of the pyramidal form {211} over the {101} (Fig. 4). Few crystals develop the {100} prism, which is more common in zircons from the Teniente Porphyry than those from the Late Dacite Dike (Figs 3e, f and 4). Zircon chemistry and Ti-in-zircon thermometry Trace element data for El Teniente zircons are reported in Table 2. They include the determinations of REE, Y and Fig. 4. (a) Main morphological types of El Teniente zircons schematically illustrated in the typological classification diagram of Pupin (1980). (b) Histograms of aspect ratios (length/width) comparing distributions among the El Teniente intrusive units. n, number of crystals measured. Hf from this study. All zircons define a single group in terms of their REE contents and normalized REE patterns. REE abundances are moderately low among the values reported for crustal zircons (250^5000 ppm; 1100 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Table 2: REE, Yand Hf concentrations for zircons from El Teniente deposit Spot La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu EuN/ CeN/ CeIV/ EuN* CeN* CeIII (a) (b) (c) Y Hf A Porphyry 1.1 0·074 7·9 0·347 3·47 3·63 1·44 12·9 3·1 34·4 12·9 53·5 13·1 111·8 26·9 0·64 12·0 69·7 361·0 3.1 0·074 14·0 0·288 3·50 5·06 1·41 22·6 6·2 68·0 24·6 98·2 22·3 166·2 36·7 0·40 23·2 82·5 652·1 2428·3 2596·6 5.2 0·066 8·8 0·413 3·74 3·79 1·43 14·0 3·5 38·3 14·3 58·1 14·2 119·8 28·0 0·60 13·0 74·0 389·6 2466·0 6.1 0·136 47·6 1·585 14·27 23·87 9·60 117·0 29·8 308·3 108·9 422·8 97·5 714·7 167·9 0·55 24·8 57·5 3002·3 2209·0 7.1 0·059 5·7 0·043 0·44 0·69 0·30 3·3 1·0 11·4 4·5 20·0 5·3 49·9 11·3 0·60 27·4 576·4 130·7 2364·6 13.2 0·072 8·6 0·408 4·02 4·18 1·38 14·4 3·6 39·2 14·6 62·4 15·3 125·8 30·7 0·54 12·2 65·5 411·1 2307·6 Sewell Stock 3.2 0·062 11·2 0·229 2·69 3·62 1·33 14·2 3·7 39·8 14·6 61·1 15·1 126·6 31·3 0·57 22·7 142·4 405·6 2754·2 4.2 0·060 7·9 0·166 1·91 2·28 0·86 8·8 2·3 24·6 9·5 41·6 10·6 92·4 21·8 0·59 19·3 111·6 261·8 2676·6 7.1* 0·097 25·3 0·202 1·92 3·10 1·18 16·0 4·7 55·0 22·3 97·9 23·7 203·3 49·7 0·51 43·6 26·8 709·0 3365·7 7.2 0·075 9·3 0·109 1·12 1·59 0·66 7·8 2·4 30·8 13·7 63·7 17·0 146·8 38·1 0·57 25·0 228·3 395·8 2351·3 9.1* 0·086 6·2 0·175 1·76 2·01 0·76 7·2 2·0 21·6 8·4 37·5 9·7 84·8 20·6 0·61 12·3 147·1 234·8 2525·5 9.2 0·118 14·4 0·794 6·26 6·20 2·44 25·8 7·2 81·5 31·1 131·5 32·1 256·5 63·5 0·59 11·4 65·7 899·9 2217·1 10.1 0·099 17·5 0·053 0·53 1·04 0·41 6·4 2·0 25·1 10·8 49·0 13·1 117·3 28·8 0·49 58·8 223·7 313·7 3102·1 2629·2 Central Diorite 3.1 0·080 8·4 0·058 0·67 1·05 0·41 4·8 1·5 17·3 7·2 32·8 8·7 79·9 19·9 0·55 30·0 366·6 209·0 3.2 0·050 13·1 0·057 0·48 0·87 0·26 4·8 1·5 18·2 7·1 31·4 7·7 66·5 14·9 0·39 59·4 596·3 195·2 2796·8 5.2 0·079 8·8 0·339 3·52 4·03 1·44 14·2 3·5 37·7 13·7 55·6 13·2 110·2 26·7 0·58 13·0 35·0 368·6 2482·4 11.2 0·064 10·3 0·052 0·55 0·91 0·35 5·0 1·5 18·4 7·3 33·1 8·2 72·9 16·2 0·50 43·0 466·6 202·8 2771·8 14.1 0·081 9·1 0·342 3·60 4·00 1·42 15·1 3·8 41·6 15·9 62·9 14·9 119·6 27·7 0·56 13·2 37·4 421·7 2349·1 17.1 0·096 8·4 0·129 1·11 1·44 0·52 5·6 1·7 18·1 7·2 32·8 8·3 76·2 19·0 0·55 18·2 188·7 199·3 2290·6 Northern Diorite 1.1 0·067 9·2 0·042 0·49 0·70 0·27 3·6 1·1 13·0 5·5 24·2 6·2 61·3 13·8 0·52 41·9 619·3 151·0 2903·3 2.1 0·075 7·8 0·140 1·63 2·33 0·97 9·5 2·5 26·5 9·9 41·7 10·3 91·5 21·5 0·63 18·5 75·1 283·1 2625·8 8.1 0·061 9·4 0·097 1·05 1·50 0·61 7·3 2·2 27·0 11·9 54·8 14·5 126·6 32·6 0·56 29·5 330·6 340·0 2449·7 8.2 0·071 8·9 0·108 1·20 1·89 0·27 11·7 4·0 52·7 22·5 104·0 26·0 207·3 48·7 0·17 24·5 275·8 622·7 2775·5 11.1 0·516 12·6 0·414 1·72 2·07 0·52 12·5 4·5 63·7 29·7 145·7 38·9 328·2 85·9 0·31 6·6 588·2 849·3 2408·6 12.2 0·072 11·0 0·040 0·41 0·70 0·32 4·4 1·4 17·2 7·4 34·9 9·3 87·4 21·5 0·55 49·6 1122·1 211·0 3097·6 Teniente Porphyry 1.1 0·062 9·2 0·062 0·59 1·00 0·38 5·1 1·5 17·4 7·5 33·0 8·5 77·5 18·6 0·52 36·0 412·4 209·9 2895·5 5.1 0·069 12·9 0·043 0·43 0·82 0·30 4·8 1·5 19·7 8·5 38·1 9·7 90·0 22·0 0·46 57·4 1002·5 252·0 3226·5 6.1 0·080 18·1 0·068 0·77 1·27 0·46 7·2 2·2 27·1 11·0 49·3 12·2 108·6 24·6 0·47 59·4 651·2 320·4 2857·5 7.1 0·075 21·9 0·108 1·10 2·04 0·72 11·3 3·6 45·1 19·1 88·5 22·2 186·7 46·3 0·46 59·0 574·1 547·5 2672·7 9.1 0·075 9·4 0·120 1·52 2·38 0·93 11·1 3·1 33·4 12·5 53·0 13·2 112·5 26·9 0·55 23·9 107·3 348·1 2860·7 12.1 0·073 9·2 0·027 0·23 0·40 0·17 2·5 0·8 10·7 5·0 25·6 7·5 78·7 20·0 0·52 50·3 2717·8 172·8 3174·0 3026·6 Late Dacite Dike 4.1 0·088 26·6 0·096 1·01 1·71 0·64 9·7 2·9 36·6 14·5 63·6 16·1 135·8 31·9 0·48 70·0 738·0 405·8 4.2 0·069 8·5 0·087 1·03 1·36 0·52 6·6 1·9 23·7 10·4 48·1 12·7 113·8 29·2 0·53 26·6 354·1 278·7 2786·0 5.1 0·066 8·8 0·413 3·74 3·79 1·43 14·0 3·5 38·3 14·3 58·1 14·2 119·8 28·0 0·60 13·0 45·3 389·6 2466·0 10.1 0·064 38·9 0·155 1·71 2·68 1·03 13·4 3·7 43·3 16·9 71·8 17·5 144·2 34·3 0·53 94·6 482·2 462·2 3110·4 12.1 0·087 51·0 0·140 1·48 3·07 1·12 19·7 6·3 77·9 32·7 143·6 35·0 283·4 68·1 0·44 111·6 924·7 922·9 3073·1 17.1 0·063 22·4 0·062 0·60 1·22 0·52 7·9 2·5 31·3 13·4 60·9 15·1 136·9 36·3 0·52 87·2 1390·7 375·3 3351·9 All concentrations are reported in ppm. (a) Eu-anomaly calculated as EuN/(SmNGdN)1/2; (b) Ce-anomaly calculated as CeN/(LaNPrN)1/2; (c) CeIV/CeIII calculated according to Ballard et al. (2002) with whole-rock data for El Teniente intrusive rocks taken from Rojas (2003), Cannell (2004), González (2006) and Hitschfeld (2006). 1101 JOURNAL OF PETROLOGY VOLUME 53 Hoskin & Schaltegger, 2003) ranging from 114 to 727 ppm (Table 2). Chondrite-normalized REE patterns are characterized by a steep increase from LaN to LuN with a strong positive Ce-anomaly [(Ce/Ce*)N: 6^111] and a slight negative Eu-anomaly [(Eu/Eu*)N: 0·1^0·6; Fig. 5]. Zircon Ce4þ/Ce3þ ratios, which primarily depend on the oxidation state of the magma (Ballard et al., 2002), show a wide range of values between 26 and 2717 (Fig. 5). A slight decrease in (Eu/Eu*)N ratios and an increase in (Ce/Ce*)N and Ce4þ/Ce3þ ratios is observed with decreasing age (Fig. 5). Additionally, Ce4þ/Ce3þ and (Eu/Eu*)N values are mostly within the range defined by zircons from mineralization-related intrusions associated with porphyry copper deposits in northern Chile (Ballard et al., 2002; Fig. 5). Overall, zircon REE patterns and concentrations are typical of those reported for crustal zircons in general. U and Th contents can be grouped by crystal sectors showing simple igneous structures (fine oscillatory or sector zoning) and those from weakly luminescent overgrowth rims, both characteristics observed in CL images (Table 1; Fig. 3). For the former, U and Th are in the range of 36^1570 ppm and 18^920 ppm, respectively, and single grains show inner compositional variations up to one order of magnitude in concentration. Weakly luminescent overgrowth rims, observed only in zircons from the A Porphyry and the Sewell Stock, have comparatively higher U and Th contents in the range of 1125^4160 ppm and 329^3481ppm, respectively. These concentrations are one to two orders of magnitude higher than those of their respective grain cores (Table 1). Zircon Th/U ratios are relatively uniform between 0·5 and 1, irrespective of the unit or crystal sector considered (Table 1; Fig. 6). Hf concentration varies between 2300 and 2400 ppm (Table 1; Fig. 6) and Y between 130 and 920 ppm (Table 2; Fig. 6), both within the range reported for crustal zircons (Hoskin & Schaltegger, 2003). Overall, the Hf, U, and Th contents of zircons from the Teniente Porphyry and the Late Dacite Dike extend to progressively higher values relative to the older units (Fig. 6). Temperatures were estimated using the Ti-in-zircon thermometer from the equations of Ferry & Watson (2007) whose calibration assumes crystallization under rutile- and quartz-saturated conditions (aTiO2 ¼1; aSiO2 ¼1) at 10 kbar. Though present, rutile has not been reported in the ore deposit as part of the magmatic mineral assemblage and is generally considered to have a hydrothermal origin (Rabbia, 2002). Thus, an aTiO2 ¼ 0·6 has been used in the calculations in agreement with the general presence of other Fe^Ti oxides (e.g. Ferry & Watson, 2007; Fu et al., 2008). Calculated temperatures vary between 6138 and 8138C for a total range inTi concentration between 1·2 and 11·8 ppm (n ¼ 62; n is the number of analyses; Fig. 7; Table 3). Temperature decreases systematically with decreasing age from an average of NUMBER 6 JUNE 2012 Fig. 5. (a) Chondrite-normalized REE patterns for zircons from the El Teniente deposit; normalization values after McDonough & Sun (1995). The field defined by all analyses is highlighted and single grain patterns from the Sewell Stock and the A Porphyry are shown as examples. (b) Single grain and average CeN =CeN and CeIV =CeIII Zrc ratios vs age and EuN =EuN ratios. For single grains, the corresponding U^Pb ages were used for plotting; for averages, in samples with bimodal U^Pb age distributions, the oldest and dominant age peak was used.CeIV =CeIII Zrc calculated according to Ballard et al. (2002) with whole-rock chemical data for El Teniente intrusions taken from Rojas (2003), Cannell (2004), Gonza¤lez (2006) and Hitschfeld (2006). 777 508C in zircons from the A Porphyry to 681 788C in zircons from the Late Dacite Dike (dispersion from averages given at 2s, Fig. 7). Pressure corrections to these temperatures would lower these estimates by up to 508C, but most of them would still record higher temperatures than the highest estimated for the hydrothermal activity within the orebody (500^6008C; Cannell, 2004), and are within the general range of felsic to intermediate igneous rock crystallization temperatures (Fu et al., 2008). Zircon Hf and O isotopes The range in Hf and O isotopic compositions of El Teniente zircons is remarkably uniform between the 1102 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Fig. 6. Plots of U, Y and Hf concentration vs age and U^Th covariation of zircons from the El Teniente deposit. Zircon U (and Th) contents show a trend of increasing values with decreasing age despite the scatter induced by the high U analyses. A similar trend is seen in Hf contents whereas Y remains essentially constant. Th/U ratios are mostly between 0·5 and 1·0. Symbols as in Fig. 5. In the U^age and U^Th plots dashed lines indicate minimum concentrations shown by the weakly luminescent high U^Th rims identified from CL images. These have been omitted from the averages of the Sewell Stock and A Porphyry samples. The complete dataset of U^Th concentrations has been given by Maksaev et al. (2004). various intrusive units. Importantly, isotopic values show no correlation with respect to spot location within a single zircon grain (core or rim), any chemical parameter, the U^Pb ages, or between each other (Tables 4 and 5). Zircon Hf ratios are characterized by a high initial 176 Hf/177Hf that varies between 0·283010 and 0·282945 (n ¼ 57; Table 4). Corresponding initial eHf values range from þ8·4 to þ6·1, with an average of þ7·4 1·2, and intra-grain variation is less than 1 eHf unit (Fig. 8). These characteristics, together with the general absence of inherited zircons or zircon cores, are consistent with a common magmatic system for the various intrusions within the deposit. Two-stage depleted mantle (DM) model ages range from 480 to 630 Ma (T2DM ; Table 4). O isotope compositions for all Teniente zircons define a range of d18OZrc values between 5·6 and 3·6ø, with an arithmetic average of 4·7 1·0ø (Fig. 9; Table 5). At a Fig. 7. Variation of temperature obtained by Ti-in-zircon thermometry vs age for zircons from the El Teniente deposit. Crystallization temperatures define a trend of decreasing values with decreasing age. Symbols as in Fig. 5. 1103 JOURNAL OF PETROLOGY VOLUME 53 Table 3: Ti concentration and T8estimates for zircons from El Teniente deposit Spot Ti 2s (ppm) T8Zrc Spot (8C) A Porphyry Ti 2s (ppm) T8Zrc (8C) Northern Diorite 1.1 6·3 0·3 750 1.1 5·0 0·1 729 3.1 7·4 0·2 766 2.1 5·6 0·4 739 4.1 5·9 0·4 745 3.1 4·9 0·2 727 5.2 7·3 0·3 765 4.1 6·9 0·3 759 6.1 11·8 0·4 813 6.1 7·9 0·1 773 7.1 9·3 0·4 789 8.1 7·1 0·3 762 11.1 6·1 0·1 747 8.2 8·9 0·4 784 12.1 8·2 0·2 775 11.1 4·7 0·3 723 13.1* 5·0 0·2 730 12.2 3·9 1·0 707 13.2 9·4 0·3 790 13.1 1·3 0·1 617 Sewell Stock 6·7 0·5 756 1.1 2·7 0·2 677 1.2 7·9 0·3 772 4.1 3·2 0·2 690 3.1 3·9 0·2 708 5.1 1·7 0·1 640 3.2 4·9 0·1 727 6.1 2·5 0·1 669 4.2 6·0 0·1 746 6.2 8·1 0·3 774 7.1* 1·2 0·1 613 7.1 3·2 0·2 691 7.2 7·2 0·2 763 8.1 4·9 0·2 728 9.1* 3·4 0·4 695 9.2 2·7 0·1 677 9.2 8·6 0·2 780 9.1 3·1 0·2 687 10.1 3·0 0·1 685 12.1 1·3 0·3 620 10.2 4·8 0·1 725 Central Diorite Late Dacite Dike 3.1 8·1 0·3 775 2.1 2·7 0·1 675 3.2 8·2 0·3 776 3.1 1·3 0·1 620 5.2 6·3 0·2 750 4.1 3·2 0·1 690 5.1 7·3 0·3 765 4.2 7·2 0·4 763 7.1 2·8 0·1 680 5.1 3·5 0·1 697 8.1 7·9 0·5 772 7.1 3·3 0·1 693 11.1 6·2 0·3 749 8.1 3·7 0·1 703 11.2 6·0 0·4 746 10.1 2·1 0·1 656 12.1 7·2 0·3 763 12.1 2·4 0·1 668 14.1 8·4 0·1 779 17.1 1·8 0·1 645 17.1 7·8 0·3 772 JUNE 2012 et al., 1998). Despite some scatter, the entire dataset is remarkably uniform in that the d18OZrc values closely describe a unimodal normal distribution (Fig. 9). As such, the weighted mean of 4·76 0·12ø (MSWD ¼ 2·0; 61 analyses) should be a good representation of the zircon population. Although an MSWD of 2·0 for the whole dataset of d18OZrc might be indicating a scatter slightly in excess of that expected from analytical error, the uniform distribution of the data does not reveal the existence of different populations. Moreover, the absence of substantial differences between intrusive units and of any correlations with the crystal sector analyzed or compositional and age parameters further precludes this possibility. DISCUSSION Magmatic vs hydrothermal origin of zircon Teniente Porphyry 1.1 NUMBER 6 TZrc estimates have been calculated with equations from Ferry & Watson (2007) assuming an aSiO2 ¼ 1 and aTiO2 ¼ 0·6. single grain scale the d18O values are mostly within analytical uncertainty, with an intragrain variation lower than 0·8ø, whereas variations within samples of the different units are up to 1·1^1·7ø. It is noteworthy that the El Teniente d18OZrc average is within the lower limit of normal mantle zircon values of 5·3 0·6ø (2s; Valley A key question to be investigated in this study relates to the nature of the zircon, whether it is of magmatic origin or if there is a significant zircon fraction that can be interpreted to be of hydrothermal origin. As such, zircon compositional and morphological features may be a consequence of fluid circulation in hydrothermal systems with new zircon growth and/or recrystallization. Diffusion of most elements in zircon is unlikely, unless assisted by defects in the crystal structure as in cracked or metamict zircons (Cherniak & Watson, 2003). However, such areas can be recognized as being altered through imaging techniques and thus avoided during analysis. The principal evidence for a hydrothermal imprint in the El Teniente zircons comes from particular textural and chemical features shown by several grains. Texturally, they show a mottled external texture in transmitted light images. Similar features have been recognized in other hydrothermal deposits for zircons that chemically and isotopically preserve primary magmatic characteristics (van Dongen et al., 2010). In El Teniente this surficial mottling feature also corresponds to weakly luminescent high U^ Th rims (Fig. 3). The bimodal U^Pb age distribution shown by zircon populations from the four older units have been previously interpreted as representing the crystallization and alteration age of the respective intrusions (Maksaev et al., 2004). This interpretation is also supported by the agreement of the latter age with alteration, mineralization, and intrusive event ages obtained by different geochronological systems (Ar^Ar and Re^Os) on different mineral phases. The younger units, the Teniente Porphyry and the Late Dacite Dike, show unimodal zircon U^Pb age distributions interpreted as representing the crystallization ages. Samples studied from all units correspond to massive intrusive rocks that are heavily altered, as is characteristic for rocks within the deposit. Because the main objective of this work was to study the genesis of magmas related to porphyry copper mineralization, most of the analyses were made on zircon spots inferred to be 1104 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Table 4: Lu and Hf isotopic data for zircons from El Teniente deposit Spot 176 Hf/177Hf 2s 176 Lu/177Hf 2s 206 Pb/238U 2s eHf(t) 2s age (Ma) 2 TDM (crustal) A Porphyry 1.1 0·283001 0·000028 0·000256 0·000004 6·6 0·8 8·2 1·0 504 3.1 0·282965 0·000025 0·000300 0·000016 6·8 0·7 7·0 0·9 585 4.1 0·283000 0·000028 0·000415 0·000010 6·4 0·7 8·2 1·0 505 5.2 0·283010 0·000031 0·000801 0·000105 5·7 0·7 8·5 1·1 484 6.1 0·282988 0·000042 0·001838 0·000036 6·5 0·2 7·8 1·5 535 12.1 0·282988 0·000028 0·000337 0·000014 6·8 1·0 7·8 1·0 533 13.1* 0·282958 0·000034 0·000311 0·000008 6·4 0·2 6·7 1·2 601 13.2 0·282976 0·000026 0·000381 0·000016 6·2 0·7 7·3 0·9 561 1.1 0·282998 0·000048 0·001010 0·000058 6·4 0·3 8·1 1·7 512 3.1 0·282992 0·000028 0·000354 0·000025 5·8 0·6 7·9 1·0 525 3.2 0·283005 0·000028 0·000341 0·000016 6·3 0·7 8·4 1·0 494 4.2 0·282996 0·000025 0·000304 0·000011 6·3 0·8 8·1 0·9 515 7.1* 0·282998 0·000024 0·000236 0·000005 6·1 0·1 8·1 0·9 511 7.2 0·282996 0·000030 0·000615 0·000011 5·6 0·6 8·0 1·1 517 9.1* 0·282980 0·000058 0·000853 0·000040 6·0 0·3 7·5 2·0 553 9.2 0·282987 0·000024 0·000333 0·000008 6·3 0·5 7·8 0·9 535 10.1 0·282977 0·000026 0·000211 0·000002 6·2 0·3 7·4 0·9 559 Sewell Stock Central Diorite 3.1 0·282974 0·000023 0·000300 0·000005 6·4 0·9 7·3 0·8 566 3.2 0·282966 0·000032 0·000210 0·000002 4·8 1·1 7·0 1·1 584 5.1 0·282978 0·000026 0·000314 0·000010 5·1 0·7 7·4 0·9 556 5.2 0·282984 0·000022 0·000168 0·000006 5·6 0·6 7·6 0·8 542 7.1 0·282995 0·000032 0·000299 0·000020 7·0 0·6 8·0 1·1 518 8.1 0·282947 0·000026 0·000319 0·000015 5·7 0·7 6·3 0·9 627 11.1 0·283001 0·000027 0·000323 0·000007 6·8 0·8 8·3 0·9 503 11.2 0·282995 0·000030 0·000465 0·000069 6·0 0·8 8·0 1·1 517 12.1 0·282980 0·000027 0·000726 0·000025 5·7 0·6 7·5 0·9 551 14.1 0·283001 0·000029 0·000393 0·000007 6·5 0·8 8·2 1·0 505 17.1 0·282995 0·000029 0·000560 0·000026 6·5 1·4 8·0 1·0 517 1.1 0·282980 0·000022 0·000369 0·000003 6·4 0·6 7·5 0·8 551 2.1 0·282986 0·000048 0·000699 0·000059 6·7 0·8 7·7 1·7 538 4.1 0·282982 0·000042 0·000483 0·000044 5·9 0·6 7·6 1·5 546 8.1 0·283000 0·000040 0·000480 0·000016 5·9 0·5 8·2 1·4 506 8.2 0·282991 0·000037 0·000621 0·000047 27·8 0·6 8·3 1·3 513 11.1 0·282985 0·000028 0·001204 0·000025 79·4 1·5 9·2 1·0 499 12.2 0·282980 0·000029 0·000927 0·000006 5·3 0·7 7·5 1·0 552 13.1 0·282987 0·000023 0·000238 0·000005 6·3 0·3 7·7 0·8 535 1.1 0·282947 0·000029 0·000504 0·000021 5·4 0·4 6·3 1·0 626 2.1 0·282975 0·000041 0·000577 0·000050 5·3 0·4 7·3 1·4 564 4.1 0·282968 0·000029 0·000459 0·000031 4·7 0·5 7·0 1·0 580 5.1 0·282973 0·000048 0·000625 0·000029 4·8 0·3 7·2 1·7 567 Northern Diorite Teniente Porphyry (continued) 1105 JOURNAL OF PETROLOGY VOLUME 53 NUMBER 6 JUNE 2012 Table 4: Continued Spot 176 Hf/177Hf 2s 176 Lu/177Hf 2s 206 Pb/238U 2s eHf(t) 2s age (Ma) 2 TDM (crustal) 6.1 0·282960 0·000024 0·000479 0·000026 5·6 0·4 6·8 0·9 597 6.2 0·282985 0·000022 0·000430 0·000023 5·0 0·8 7·6 0·8 542 8.1 0·282952 0·000028 0·000678 0·000009 5·1 0·3 6·5 1·0 615 9.2 0·282978 0·000051 0·000340 0·000006 5·4 0·3 7·4 1·8 557 15.1 0·282982 0·000037 0·000537 0·000011 5·2 0·4 7·6 1·3 547 17.1 0·282996 0·000045 0·000592 0·000041 5·3 0·4 8·0 1·6 515 2.1 0·282956 0·000031 0·000577 0·000024 4·5 0·3 6·6 1·1 607 3.1 0·282945 0·000032 0·000760 0·000025 5·2 0·4 6·2 1·1 631 4.1 0·282951 0·000029 0·000449 0·000020 4·7 0·4 6·4 1·0 618 4.2 0·282968 0·000034 0·000280 0·000023 6·2 0·6 7·0 1·2 580 5.1 0·282947 0·000034 0·000876 0·000018 5·0 0·3 6·3 1·2 627 7.1 0·282957 0·000029 0·000729 0·000014 4·4 0·3 6·6 1·0 606 10.1 0·282968 0·000027 0·000388 0·000023 4·7 0·3 7·0 0·9 580 12.1 0·283002 0·000039 0·000398 0·000025 4·9 0·5 8·2 1·4 504 15.1 0·282962 0·000029 0·000497 0·000012 4·8 0·3 6·8 1·0 594 17.1 0·282995 0·000029 0·000560 0·000026 4·8 0·3 8·0 1·0 518 Late Dacite Dike Söderlund et al. (2004) 176Lu decay constant of 1·867 1011 has been used in these calculations. For eHf(t) values the chondritic values of Blichert-Toft & Albarède (1997) have been used along with the corresponding zircon spot age. Two-stage depleted mantle model age TDM2 was calculated using the present-day depleted mantle values of Vervoort & Blichert-Toft (1999) assuming a crustal average of 176Lu/177Hf ¼ 0·015 (Goodge & Vervoort, 2006). magmatic in origin. However, it is still necessary to consider to what extent a hydrothermal imprint might have affected the chemical and isotopic characteristics of the magmatic zircon grains, both for crystal sectors inferred to be magmatic and for those that are suspected of being affected by the hydrothermal imprint. The CL images of the El Teniente zircons show preservation of a euhedral morphology and a well-developed zoned structure, both common to simple zoned igneous zircon (Fig. 3). Exceptions to this norm are the weakly luminescent, high U^Th rims developed around some grains from the Sewell Stock and the A Porphyry (Fig. 3). These rims are not associated with any other particular compositional signature or U^Pb age group. Overall the zircon populations show rather homogeneous morphological features that are distinctive for different intrusive units. These characteristics argue against hydrothermal recrystallization for most zircons and suggest that either this process or new zircon growth is restricted to the limited development of overgrowth rims. Therefore, in general the zircon grains from a single intrusion are compositionally homogeneous, and in terms of Y þ REE contents and O^Hf isotopic composition they are indistinguishable between units. Discernible compositional variations arise in the somewhat higher relative contents of U, Th and Hf in zircons from the younger intrusive rocks, the Teniente Porphyry and the Late Dacite Dike, and the general trend of decreasing zircon Ti content with decreasing age (Figs 6 and 7). Zircon-Ti thermometry records temperatures that are higher than the maximum estimated for hydrothermal fluids during ore deposit formation (500^ 6008C; Cannell, 2004). Overall, all these characteristics are consistent with the zircons being formed by crystallization from a series of cogenetic melts at magmatic temperatures. Particular morphological and chemical features are probably a first-order consequence of the particular melt composition and variations in the magmatic evolution of each intrusive unit. The origin of the low d18OZrc in the El Teniente intrusive rocks, whether magmatic or hydrothermal, has profound implications in terms of petrogenetic processes. If this feature were produced by hydrothermal alteration of the zircon grains, which are devoid of crystal defects, the process would have occurred through isotopic exchange by volume diffusion inwards from grain boundaries. Watson & Cherniak (1997) have shown experimentally that even under wet conditions oxygen diffusion in zircon is extremely sluggish. For grains between 100 and 300 mm, the range of El Teniente zircons, timescales to achieve complete isotopic exchange at 6008C are between 216 and 1106 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Table 5: d18O (VSMOW) of zircons from El Teniente deposit Spot d18O 2s Spot (ø) d18O 2s (ø) A Porphyry Northern Diorite 1.1 5·13 0·58 1.1 5·37 0·60 3.1 4·59 0·62 2.1 4·89 0·59 4.1 5·13 0·63 4.1 5·63 0·60 5.2 4·34 0·62 6.1 4·61 0·59 6.1 4·47 0·57 8.1 5·43 0·59 7.1 5·52 0·59 8.2 4·62 0·61 11.1 4·85 0·57 11.1 4·88 0·61 12.1 5·51 0·59 11.2 4·38 0·62 13.1* 4·22 0·60 12.2 3·96 0·72 13.2 4·26 0·63 13.1 4·87 0·58 Sewell Stock Fig. 8. Initial eHf isotope ratios in zircon grains relative to the respective spot age showing the restricted range of values defined by intrusive units from the El Teniente deposit. Chondritic (CHUR: Blichert-Toft & Albare'de, 1997) and depleted mantle reservoirs (Vervoort & Blichert-Toft, 1999) are shown for reference. Symbols as in Fig. 5. Average analytical error bar at 2s is indicated. Teniente Porphyry 1.1 3·64 0·60 1.1 4·52 0·74 1.2 5·25 0·61 4.1 5·53 0·77 3.1 4·57 0·61 5.1 5·39 0·75 3.2 4·83 0·65 6.1 4·83 0·74 4.2 4·72 0·61 6.2 5·57 0·76 7.1* 4·45 0·64 8.1 5·59 0·75 7.2 5·15 0·60 9.1 4·12 0·75 9.1* 4·76 0·56 9.2 4·68 0·76 9.2 5·01 0·57 15.1 4·73 0·73 10.1 4·47 0·60 17.1 4·96 0·73 Central Diorite Late Dacite Dike 3.1 5·28 0·74 2.1 5·04 0·74 3.2 4·92 0·84 3.1 4·19 0·75 5.1 4·78 0·77 4.1 4·52 0·73 5.2 4·95 0·80 4.2 4·64 0·76 7.1 5·13 0·73 5.1 4·25 0·73 8.1 4·14 0·80 7.1 4·73 0·74 11.1 4·30 0·77 10.1 4·20 0·74 11.2 4·73 0·76 12.1 4·66 0·73 12.1 5·06 0·74 15.1 3·71 0·77 14.1 4·32 0·77 17.1 3·96 0·75 17.1 4·39 0·75 1945 Myr, respectively. Shorter timescales of 33^299 Myr are obtained using the empirical determinations for wet diffusion of Zheng & Fu (1998), but they are still unreasonably high when considering the timescales of hydrothermal processes within the El Teniente deposit (Cuadra, 1986; Maksaev et al., 2004; Cannell et al., 2005). Additionally, partial exchange by diffusion is expected to produce zircons with isotopic zonation, and the process should proceed faster in smaller grains than in larger ones. Both Fig. 9. Variation of d 18OZrc for the El Teniente deposit. (a) d18OZrc relative to the respective spot age. Average analytical error bar at 2s is indicated. (b) Histogram of d 18OZrc values for all analyses shown in (a) overlaid by a cumulative probability curve (calculated with Isoplot version 3.00; Ludwig, 2003). Despite scatter, El Teniente zircons show a simple unimodal distribution that extends to values lower than those for normal mantle zircons (Valley et al., 1998). 1107 JOURNAL OF PETROLOGY VOLUME 53 characteristics are absent from El Teniente. This strongly suggests preservation of magmatic d18OZrc values regardless of hydrothermal alteration processes and, consequently, the data reflect a primary magmatic signature. In summary, the morphological, chemical and isotopic characteristics of the El Teniente zircons support the conclusion that the areas analyzed relate to the primary magmatic signature. A possible hydrothermal imprint is evidenced by the presence of overgrowth rims and a mottled texture, and in the bimodal U^Pb age distribution in the older units. However, from the available data, the interpretation that the younger U^Pb ages represent hydrothermal alteration events relies heavily on their agreement with ages obtained for gangue and ore minerals by other geochronological methods (Ar^Ar and Re^Os; Maksaev et al 2004). Zircon records of magmatic evolution As noted above, zircon morphological characteristics are a first-order result of the melt composition from which they crystallize, along with the temperature and cooling rate during crystallization. With decreasing age the El Teniente zircons show a progressive reduction of the {100} prismatic form and an increase in the {110} form (Figs 3 and 4). In the Teniente Porphyry the {100} form is poorly developed, and it is almost completely absent in crystals from the Late Dacite Dike (Figs 3 and 4). The size relations between the two prismatic morphologies of zircon are strongly dependent on the chemical parameters of the growth medium. Development of a {110}-dominated form is favored by a growth blocking effect produced by the adsorption of protons, provided in the growth environment by H2O molecules and hydrated complexes, and of elements such as U, Th, P and Y (Benisek & Finger, 1993; Vavra, 1994), all of which are expected to increase in concentration with the degree of melt fractionation. In the El Teniente zircons this mechanism controlling prism development is documented by the higher concentration of elements as U, Th and Y shown by zircon populations from the younger units (Fig. 6). This morphological and chemical evolution reflects the progressively more differentiated nature of the younger intrusions. Additionally, zircon crystals show a systematic increase in aspect ratio with decreasing age (Fig. 4), a parameter that has been empirically shown to depend on cooling rate during crystallization. Thus, besides the decreasing crystallization temperatures recorded by the Ti-in-zircon thermometry, conditions of increasing cooling rate can be inferred from the zircon aspect ratios. Evolution of these parameters proceeds in accordance with the progressive waning of igneous activity within the ore deposit. Additionally, strong regional uplift and unroofing in the Andean range during this period (Far|¤ as et al., 2008; Maksaev et al., 2009) are probably also key factors controlling this evolution. NUMBER 6 JUNE 2012 Despite the compositional differences observed between the A Porphyry (SiO2 57%) and the remaining felsic porphyries (SiO2 67%), all the zircons share similar Hf^O isotopic compositions (Figs 8 and 9). This, along with the general absence of older inherited zircon components, argues against significant crustal contamination in the magmatic processes responsible for generating the different members of the suite. However, the El Teniente range of initial eHf values between þ6·4 to þ8·4 records an enrichment that, unless inherited from the source, could result from crustal contamination processes during the early stages of magma evolution. Recently published data on the Hf isotopic composition of Cenozoic igneous rocks from Central Chile show that these units share a similar signature to that of El Teniente (Montecinos et al., 2008; Deckart et al., 2010; Fig. 10). With the exception of intrusive units that are about 15 Ma in age, the Cenozoic magmatism has remained remarkably uniform, with initial eHf values between þ6 and þ10 for just over 25 Myr. This holds true despite the fact that it includes compositionally different rocks formed under contrasting tectonic regimes and margin configurations (Fig. 10). The predominantly basic to intermediate igneous rocks from the Oligocene^early Miocene Abanico Formation, formed in an extensional setting over a thinned crust and with minimal crustal contamination, are indistinguishable from the fractionated middle Miocene^Pliocene ones formed under a contractional regime in a progressively thickening crust (Fig. 10). This indicates that the control of Hf isotopic enrichment observed in El Teniente, and in the Cenozoic magmatism in general, resides in the mantle source, ruling out significant crustal contamination in their genesis. As discussed above, the coherent morphological, chemical and Hf isotopic characteristics of the El Teniente zircons indicates crystallization from a series of cogenetic melts. These observations fully agree with the model of Stern et al. (2010), who argued that the Late Miocene and Pliocene plutonic rocks that host the deposit were derived from a large, long-lived, thermally and chemically stratified, open-system magma chamber, or magmatic plumbing system. Moreover, this hypothesis has also been favored in explaining the structural patterns and the chemical evolution of the hydrothermal systems within the deposit (Cannell et al., 2005; Klemm et al., 2007). The El Teniente d18OZrc weighted mean of 4·76 0·12ø is considered to be a primary igneous feature, and therefore might result from crystallization from low-18O magmas. In the context of El Teniente, sourced by a long-lived magmatic system with an extensive record of hydrothermal activity, a likely scenario to produce this composition is the assimilation of hydrothermally altered wall-rocks. Simple mixing calculations indicate that the average of d18OZrc 4·7ø could be produced from a mantle-derived magma by 8^11% bulk 1108 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Fig. 10. (a) Zircon initial eHf isotope ratios of Oligocene^Pliocene igneous rocks from Central Chile. Gray shaded field highlights the relatively homogeneous Hf-isotopic signature shown by Central Chilean magmatism during the last 25 Myr. Legend: 1, Abanico and Farellones formations NE of the R|¤o Blanco^Los Bronces deposit (Montecinos et al., 2008); 2, San Francisco batholith; 3, Yerba Loca pluton; 4, Cerro Meso¤n Alto stock; 5, La Gloria pluton (Deckart et al., 2010); 6, El Teniente intrusive rocks; 7, El Teniente inherited cores. (b) Schematic representation of the main tectonic events in the Andean range for the studied region during the time span considered in (a). Vertical axis shows the Andean range segmented, from west to east, according to the main morphostructural units that form it. The main tectonic events illustrated include extensional tectonics related to Abanico Basin development, compressional tectonics related to basin inversion and subsequent eastward migrating shortening episodes, and the main Andean uplift event. Data from Giambiagi & Ramos (2002) and Far|¤ as et al. (2010). All locations and the distribution of Andean morphostructural units are indicated in Fig. 1. assimilation of crustal material with d18OWR of between ^2 and 0ø. Although limited, such an amount of crustal contamination restricts possible contaminants to the nearly 6 km thick sequence of Oligocene^Miocene igneous rocks into which the deposit is emplaced. If contaminated by Mesozoic or Paleozoic basement rocks, such as those that crop out near the Chilean^Argentinean border and that probably underlie the whole region (Fig. 2), then the El Teniente Hf isotopic composition would be shifted towards radiogenic values inconsistent with those observed. Wall-rock assimilation under upper crustal conditions can proceed either by assimilation^fractional crystallization (AFC)-type processes (DePaolo, 1981) or by the incorporation of a low-degree partial melt (Campbell & Turner, 1987; Huppert & Sparks, 1988). However, in El Teniente, characteristics such as the lack of zircon intragrain d18O variation and the homogeneous d18OZrc values shown by the intrusive rocks irrespective of the age of intrusion or composition strongly argue against AFC-type processes. A second possibility involves the formation of a magma layer by melting of the roof rocks of an underlying basaltic magma chamber that provides the heat for melting (Campbell & Turner, 1987; Huppert & Sparks, 1988). Models show that the melted material will remain at the 1109 JOURNAL OF PETROLOGY VOLUME 53 top of the chamber and be chemically isolated from the basaltic magma at the bottom, thus separating assimilation from crystallization in space and time. This process predicts the formation of granodiorite or tonalite melts and can occur at upper levels of the crust only where the rocks have been pre-heated by earlier intrusions. Although this scenario seems likely in the case of the El Teniente magmatism, this process is not consistent with two main observations: (1) the general absence of inherited zircons or zircon cores; (2) the simultaneous generation of melts of dacitic to andesitic composition which also have the same O isotopic signature. Although the occurrence of this process cannot be ruled out, it seems unlikely to be fully responsible for generating the diverse magmas with low-O isotopic compositions at El Teniente. Another process capable of lowering the magma d18O signature is through fractionation of O isotopes during degassing. However, high-temperature magma volatile loss is expected to produce a relatively minor shift towards lower d18O values in the resulting melt in the case of a pure H2O volatile phase, and even lower shifts are expected in the case of SO2 (Eiler, 2001). For example, high-temperature (46008C) melt-volatile fractionation predicts that 10% degassing of pure H2O from a gabbroic melt would produce a degassed melt with a d18O value of 0·11ø lower than the original one, and no fractionation is predicted for more felsic compositions such as granodioritic or granitic melts (Zhao & Zheng, 2003). Thus, this process seems unable to reproduce the values observed at El Teniente, which are on average 0·6ø lower than would be expected for zircon from a mantle-derived magma. Thus, as for Hf, the El Teniente O isotopic composition is inferred to be a characteristic inherited from the mantle source. Magma generation: constraints on the source and melting processes A major reconfiguration of the Chilean continental margin took place during the Cenozoic. The Oligocene^ early Miocene extensional Abanico Basin was inverted at around 21 Ma, and this was followed by a contractional regime in which shortening, thickening and uplift characterized the constructive period of the Andean orogen (Charrier et al., 2002, 2009; Giambiagi & Ramos, 2002; Kay et al., 2005; Far|¤ as et al., 2008, 2010). Whether directly related or not, Nazca and South American plate convergence parameters also varied during this period. Obliqueness abruptly diminished from 45^558 to 15^ 208 around 26 Ma, and the convergence rate increased from 6 cm a1 at 27 Ma to reach 12^20 cm a1 around 15 Ma and then decreased steadily to its present value of 8 cm a1 (Pardo-Casas & Molnar, 1987). The limited variation in Hf isotopic composition shown by Central Chilean igneous rocks throughout this period strongly suggests buffering by a stable isotopic reservoir. Hf depleted NUMBER 6 JUNE 2012 mantle model ages of 500^600 Ma highlight the enriched nature of this reservoir. The subcontinental lithospheric mantle and the lower crust constitute reservoirs able to constantly imprint such a signature in magmas irrespective of the margin configuration and thus are likely to be responsible for the observed isotopic composition of Chilean Cenozoic igneous rocks. The Hf isotopic data for El Teniente and other intrusive rocks of the region (Deckart et al., 2010) agree with and expand the results of Montecinos et al. (2008). Based on Pb, Sr and Hf isotope data, those workers showed that there are no significant variations in the Oligocene^middle Miocene magmatism and attributed the enriched isotopic signatures to a characteristic inherited from the subcontinental lithospheric mantle. Overall the Hf isotopic composition of the Cenozoic magmatism in Central Chile can be considered to have been derived from extensive and long-lived MASH-type processes. Such MASH processes were originally proposed by Hildreth & Moorbath (1988) to explain the chemical and isotopic variations in volcanic rocks along the Chilean Southern Volcanic Zone. MASH domains are deep lithospheric zones where ascending, subduction-related, mantle-derived magmas are hybridized generating relatively homogeneous magmas with chemical and isotopic characteristics specific of the MASH zone from which they evolved. An exception to the observed homogeneous Hf isotopic compositions in Central Chile is the higher initial Hf signatures (more depleted) recorded by intrusive units at around 15 Ma (Fig. 10). These correspond to the Yerba Loca stock and the oldest portions of the San Francisco Batholith (locations indicated in Fig. 2; Deckart et al., 2010). This period coincides with an important change in the structural evolution of the Andean orogen. Following Abanico Basin inversion, contractional deformation that was mostly concentrated in the western slope of the Andes migrated to the east, forming the Aconcagua Fold and Thrust Belt in the Eastern Principal Cordillera (Fig. 2; Giambiagi & Ramos, 2002). We suggest that during this reorganization a transient period of stress in the Andean range permitted rapid ascent of subduction-related, mantle-derived magmas that had little interaction with the upper lithosphere. Such a process is supported not only by the high initial eHf values recorded in the Yerba Loca stock, but also in the change towards less depleted compositions recorded in the San Francisco Batholith, with eHf values of þ8·5 to þ10·9 at 14 Ma and þ7·0 to þ8·6 at 11Ma (Fig. 10). As an alternative to the MASH model of Hildreth & Moorbath (1988), Stern (1991) proposed source contamination to explain the variable enrichment of the Chilean Southern Volcanic Zone magmas. In this model, incorporation into the mantle wedge of different amounts of subducted sediment and Paleozoic upper crust derived from forearc subduction erosion could be responsible for the 1110 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS observed isotopic signatures of the arc magmas. This process was also proposed to in part explain the adakite-like chemistry of the El Teniente intrusive rocks within the context of the ‘normal’ Cenozoic magmatism in Central Chile (Kay et al., 2005). According to this model, crustal blocks incorporated into the mantle wedge at peak times of subduction erosion during continental margin evolution are subjected to high-pressure metamorphism and partial melting, and generate the adakite-like magmas later emplaced into the upper plate. However, if such a process were responsible for the chemical differences between the El Teniente magmas and the preceding magmatism it would undoubtedly result in a distinct isotopic signature between them, which is not the case as recorded by the Hf isotopic data (Fig. 10). Several lines of evidence, other than just geochemical, support subduction erosion as a long-term process occurring along the Chilean continental margin (Stern, 1991; Laursen et al., 2002; Kay et al., 2005). However, the extent to which this process controls the chemical and isotopic signatures of Andean arc magmas is certainly variable. Subduction erosion products may enter the asthenospheric source of the Cenozoic magmas, but their impact on magma geochemistry is probably overwhelmed by the later imprint that they are subjected to in deep lithospheric MASH zones. In recent years, several workers have highlighted the possibility of melting subduction-modified lithosphere as a process for generating porphyry Cu Mo Au related magmas (Kay & Mpodozis, 2001; Richards, 2009; Shafiei et al., 2009). Arc magmatism is ultimately a means of material transfer from the oceanic slab and the asthenospheric mantle to the overriding plate. The interaction with, and differentiation of, these magmas deep in the lithosphere results in the formation of hydrous cumulate zones, of mafic to ultramafic composition, where significant accumulation of amphibole and the presence of residual sulfide phases can account for water and metal storage in the upper lithosphere (e.g. Davidson et al., 2007; Jagoutz et al., 2007; Richards, 2009). Thus, after long periods of arc magmatism, the lithospheric roots are expected to be a fertile reservoir from which potentially ore-forming magmas may be extracted. During the Cenozoic evolution of the Central Chilean Andes, progressive lithospheric thickening has been inferred to involve a change in mineralogy in the deep lithosphere from low-pressure amphibole-dominated to higher pressure garnet-dominated assemblages. This is evidenced in the chemical trends defined by the arc-related igneous rocks formed during this period (Kay & Mpodozis, 2001, 2002; Kay et al., 2005). Such a process involves the generation of an H2O-undersaturated melt, the H2O being provided by the dissociation of hydrous mineral phases. This dehydration melting reaction overall involves Hornblende þ Plagioclase Quartz ! Hornblende þ Pyroxene þ Garnet þ Melt, and is basically the metamorphic transition from amphibolite to eclogite facies (Sen & Dunn, 1994; Wolf & Wyllie, 1994). Experimental work has shown the ability of this process to generate hydrous silicate melts of felsic to intermediate composition, which share many chemical similarities with igneous rocks of the tonalite^trondhjemite series and adakites. Adakite-like characteristics such as intermediate to high SiO2 content, highly fractionated REE patterns, trondhjemitic affinities, and high concentrations of Sr and high Sr/ Y ratios, have all been recognized as characteristic of the El Teniente intrusive rocks (Rabbia et al., 2000; Kay et al., 2005; Stern & Skewes, 2005; Stern et al., 2010), as well as of many other ore deposit-related igneous rocks worldwide. Moreover, the Os isotope compositions of sulfides in Chilean porphyry copper deposits, including El Teniente, suggest significant crustal residence of the Os and thus a probable addition from the lower crust (Mathur et al., 2000). A key question is the ability of the crust at the time of formation of the El Teniente magmas to stabilize garnet through dehydration melting of basic precursors. Experimental studies have reproduced this reaction over a wide range of pressures and temperatures from 10 to 16 kbar and from 700 to 10008C (Wolf & Wyllie, 1993, 1994; Wyllie & Wolf, 1993; Sen & Dunn, 1994; Rapp & Watson, 1995; Lo¤pez & Castro, 2001), in agreement with observations from natural examples (e.g. Garrido et al., 2006; Berger et al., 2009). In terms of pressure, this indicates a minimum crustal thickness of 30^35 km. However, regarding temperature there is a thermal barrier to overcome to drive this process, as, assuming a conservative thermal gradient of 208C km1, the base of such a crust would reach no more than 600^7008C. In Central Chile, the increasing La/Yb ratios of the Cenozoic igneous rocks, particularly in the El Teniente region, suggest the increasing involvement of garnet as a high-pressure residual assemblage (Kay et al., 2005). Moreover, the current 45^50 km crustal thickness under this area must have been reached no later than 4 Ma, the time at which uplift of the Andean orogen to its current altitude was completed (Far|¤ as et al., 2008, 2010). Uplift has been inferred to be primarily the result of an isostatic crustal response to tectonic shortening and thickening processes. Thus, the chemical trends and tectonic evolution of the margin during the Cenozoic both indicate the presence of garnet in lower crustal assemblages. Less straightforward is how to elucidate the coeval thermal regime governing this zone, which would ultimately condition the occurrence of dehydration melting reactions. Since the original MASH model proposed by Hildreth & Moorbath (1988), several studies have focused on modeling the dynamics of magmatism-driven processes occurring at or close to the Moho in convergent margin settings 1111 JOURNAL OF PETROLOGY VOLUME 53 (Petford & Gallagher, 2001; Annen & Sparks, 2002; Annen et al., 2006). Annen et al. (2006) developed a comprehensive model built upon the concepts of underplating and high-pressure basalt differentiation, incorporating aspects of AFC- and MASH-type processes, to evaluate the effects of repeated basalt intrusion into the crust. Their work proposed the existence of ‘deep crustal hot zones’, which result from repeated deep intrusion of mantle-derived, hydrous, basalt sills, either by injection at a fixed depth at the Moho or randomly throughout the lower crust. This leads to a scenario in which evolved melts reaching middle to upper crustal levels are generated from H2O-rich parental basalts, both by partial crystallization of the basalts themselves and by partial melting of the surrounding crustal rocks through heat and H2O transfer from the cooling basalts. Further mixing between these two end-members can create a large range of intermediate and silicic melts with variable composition. Moreover, models indicate that ‘deep crustal hot zones’ are the place where much of the geochemical diversity of magmas originates, owing to substantial variations in melt proportions and temperature at such depths. Either at a fixed or random depth, repeated basalt intrusion induces a significant thermal perturbation into the lower crust and below the Moho, as a result of heat transfer into the country rocks, which will take several millions of years to decay (Annen & Sparks, 2002). Thermal gradients in both situations would allow the lower sections of a 30 km thick crust to reach temperatures much higher than 7508C (Fig. 11). These are high enough to drive dehydration melting of amphibolite in the stability field of garnet at pressures 410 kbar (Fig. 11). Additionally, the solidus curve of this reaction has an abrupt backbend towards lower temperatures that coincides with the garnet-in curve towards higher pressures (Fig. 11). According to different models, the melting temperature is lowered by 100^2008C once entering the stability field of garnet (Wyllie & Wolf, 1993; Wolf & Wyllie, 1994; Rapp & Watson, 1995; Lo¤pez & Castro, 2001). An amphibolitic lower crust can reach temperatures up to 750^8708C at pressures 510 kbar without melting, but undergo dehydration melting in response to an increase in crustal thickness (Fig. 11). These observations can certainly vary in terms of detail when considering the uncertainties related to the Annen et al. (2006) model, as well as those related to the phase relationships. However, the overall scenario is unlikely to change: repeated basalt intrusion at or close to the Moho will induce a thermal perturbation that will significantly widen the field in which dehydration melting can occur at the base of a 433 km thick crust (Fig. 11). Regarding El Teniente, continuous and extensive magmatic activity in Central Chile preceded the formation of the ore deposit for at least 30 Myr; thus a perturbed thermal gradient may have been established in the crust well before ore deposit formation. Moreover, NUMBER 6 JUNE 2012 Fig. 11. Pressure^temperature diagram showing the solidus for dehydration melting of amphibolite and the volume per cent of melt involved in this reaction (gray dashed lines; Lo¤pez & Castro, 2001). A characteristic of the amphibolite solidus is the abrupt back bend towards lower temperatures at pressures greater than 10 kbar (Wyllie & Wolf, 1993; Wolf & Wyllie, 1994; Rapp & Watson, 1995; Lo¤pez & Castro, 2001). Different geothermal gradients are indicated with dashed and dotted black lines. Linear gradients of 20 and 308C km1 were calculated assuming a geobarometric gradient of 0·33 bar km1; perturbed gradients from long-term basalt injection, at fixed and random depth, at the base of an initially 30 km thick crust are from Annen et al. (2006). Grt, garnet. if high enough temperatures (750^8708C) were reached in the lower crust before ore deposit formation, it could have undergone dehydration melting following an increase in crustal thickness and once the residual assemblages began to stabilize garnet. Two main processes can be invoked to explain the primary low-d18O composition of the El Teniente magmas: (1) differentiation from or involvement of low-d18O material in the source; (2) isotopic fractionation during magma generation. Regarding the former, low-d18O material in the MASH source would be expected to similarly influence the isotopic composition of the Cenozoic magmatism in the region. Currently there are no oxygen isotope data for these rocks; however, it is unlikely that they all share a low-d18O signature. The second possibility implies partial melting in a deep lithospheric MASH zone in which all Cenozoic magmatism has evolved. During partial melting fractionation of oxygen isotopes can be considerable and would have measurable effects on the generated melts (Eiler, 2001). To evaluate this possibility within a dehydration melting process a simple model has been constructed following the calculation approach of Eiler (2001) using the experimental results of Wolf & Wyllie (1994) and Getsinger et al. (2009). Wolf & Wyllie (1994) studied the progressive change in mineral assemblage and mineral and melt chemistry during dehydration melting of a natural low-K amphibolite (67·4% hornblende þ 32·5% plagioclase; wt % SiO2 ¼ 48·4%), at 10 kbar between 850 and 10008C. Getsinger et al. (2009) studied the impact on melt composition and residual mineral assemblage produced by melt segregation and accumulation during 1112 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS dehydration melting of an initial composition of a natural slightly alkaline metabasalt (45% hornblende þ 25% plagioclase þ10% biotite þ10% epidote þ 2% clinopyroxene, plus minor rutile, titanite and magnetite; wt % SiO2 ¼ 46·4%), at 14 kbar between 925 and 10008C. An initial d18OWR ¼ 5·8ø was assumed for the system, which corresponds to that of a mantle-derived basaltic magma (5·5ø) contaminated with 3% subducted sediment (15ø; Eiler, 2001), considering that a deep MASH zone would ultimately have the O isotope signature of subduction-related mantle-derived magmas. Results are displayed in Fig 12, where the calculated d18O for the whole system initial composition, residue, generated melt, and zircon in equilibrium with the latter are plotted against melt wt % SiO2 (further details of this calculation are given in the figure caption and the Appendix). Results from the two experiments are not directly comparable, but are complementary. The Wolf & Wyllie (1994) experiments reproduce an incremental reaction; incremental heating of a fixed initial composition overall produces increasing amounts of melt with a progressively more basic composition in each temperature step (Fig. 12). For the Getsinger et al. (2009) experiments the initial composition varies; different proportions of the original metabasalt are mixed with a 15% partial melt of this same composition forming composites that are then melted at different temperatures (Fig. 12). Models for both experiments show that dehydration melting processes can generate melts, and thus zircons that would crystallize from them, with low d18O signatures and with compositions and d18OZrc in the range of those from El Teniente (Fig. 12). How O-depleted the calculated zircons are relative to the d18O of a normal mantle zircon depends largely on the compositional and temperature parameters. The Fig. 12. Calculation of d18O for the components involved in experimental dehydration melting of a mafic amphibolite (Wolf & Wyllie, 1994; Getsinger et al., 2009) with a bulk oxygen isotope composition of 5·8ø. This value was assumed to be representative of a mantle-derived basaltic magma (5·5ø) contaminated with 3% subducted sediment (15ø). The d18O of the whole system initial composition, residue, generated melt, and zircon in equilibrium with the latter for each experimental step are plotted against melt wt % SiO2. Calculations for the melt and residue were made following Eiler (2001) assuming a d18O for the whole system and considering the melt as the weighted sum of its normative mineralogy. Oxygen isotope fractionation was calculated using the empirical factors of Zheng (1991, 1993a, 1993b) and the corresponding temperatures for each experimental step. d18OZrc for zircon in equilibrium with each generated melt was calculated according to the relation of this value with the whole-rock d18O and wt % SiO2 determined by Valley et al. (1994). Further details on the calculations are given in the Appendix. 1113 JOURNAL OF PETROLOGY VOLUME 53 lower the melt^residue compositional difference the lower the 18OMelt^Residue and the generated melts will have a d18O signature more similar to that of the residue and the whole system; in other words, progressively more basic melts will have less exotic low-d18O compositions (Fig. 12). Additionally, such melts are overall produced at increasingly higher temperatures. The comparatively lower d18O melts and zircons obtained from the Getsinger et al. (2009) experiments, with respect to those of Wolf & Wyllie (1994), also result from differences in these parameters. Melts similar in composition from both studies (wt % SiO2 67%, Fig. 12) are produced at higher temperatures in the Getsinger et al. (2009) experiments. This causes a significant reduction of the 18OMelt^Residue and thus the melts are comparatively more depleted in their oxygen isotope signature (Fig. 12), an effect that propagates to the higher silica, lower temperature melts. This cannot be associated directly with the effects of melt accumulation and segregation, as the Getsinger et al. (2009) experiments produce siliceous, high-temperature melts following the initial 15% melting of the original alkali basalt. In summary, intermediate to felsic melts generated through dehydration melting processes will most probably have a low-d18O signature, although the extent of depletion is strongly dependent on temperature and compositional parameters during melting. Among equivalent compositions, this will be more pronounced for melts generated at higher temperatures. A comparatively less depleted signature is expected in melts of similar composition but formed through fractional crystallization of more basic precursors derived from dehydration melting processes. Overall, in both experiments and in all melt compositions, the effect on the d18O of the residue is minimal. Thus, unless melt extraction is by a Rayleigh distillation process, no measurable effect is expected on subsequent ascending magmas that may interact with this residue. M A G M AT I C M O D E L A N D F I N A L REMARKS The model described above aims to evaluate the O isotopic composition of melts generated by dehydration melting reactions. It is not considered to be a close approximation to magma genesis at El Teniente as it certainly does not take into account the numerous complexities that can be induced by other processes occurring in a MASH source. Following the original model of Hildreth & Moorbath (1988), we envisage that Cenozoic magmatism in Central Chile originates from a deep lithospheric MASH reservoir, in which ascending subduction-related, mantle-derived magmas initially stall, isotopically homogenize, and differentiate until the resultant H2O-rich residual melts with a lower density are able to continue their ascent to upper crustal levels (Fig. 13). This model relies entirely on the NUMBER 6 JUNE 2012 nearly constant Hf isotopic composition shown by the Cenozoic igneous units of the region, for which calculated depleted mantle model ages of 480^630 Ma indicate a significant crustal residence time. Additionally, the observed Hf isotope signature remains the same within rocks of different composition that were formed at different times, during a period when convergence parameters varied and major changes in upper plate tectonics took place. Residual assemblages, along with the original material in the MASH reservoir, can stabilize garnet through a dehydration melting reaction to produce the fertile magmas later involved in the El Teniente mineralization (Fig. 13). This arises as a consequence of the thermal perturbation induced by the nearly 30 Myr of magmatism that precedes the El Teniente ore deposit formation and the increase in crustal thickness. Although this process produces an H2O-undersaturated melt, greater quantities of H2O can be involved if the heat source efficiently fluxes the source region with H2O (Annen et al., 2006). Additionally, higher H2O contents can also result from mixing with H2O-rich residual melts in the source or in magma reservoirs at upper crustal levels, a most likely scenario for the El Teniente magmas, which are inferred to be related to long-lived magmatic chamber processes (Stern et al., 2010). The current active volcanic zone in Central Chile is located nearly 35 km east of El Teniente, which along with the record of progressively younger igneous units towards the east reflects Cenozoic arc migration along the margin (Stern & Skewes, 1995; Kay et al., 2005). As early as 8 Ma magmatism had already reached the area near the current active volcanic zone (Mun‹oz et al., 2009), whereas volcanic activity was progressively waning in the El Teniente region. Although there is no precise estimation of when arc magmas no longer reached the area beneath El Teniente, by the time of ore deposit formation this activity was probably in decline (Fig. 13). This might have also influenced the fertility of the magmas leaving the MASH source, by increasing the melt component derived from dehydration melting reactions in this domain and decreasing the component derived from primary basalt differentiation. Even though El Teniente is one of the largest porphyry copper deposits known at present, its formation is not unique within the evolution of the Chilean continental margin, which contains several other world-class deposits. These are ultimately linked to a long history of subduction-related magmatism, which, in the appropriate conditions, leads to their formation. However, the specific characteristics that make some areas especially productive in forming numerous world-class deposits in a particular metallogenic epoch is a question yet to be resolved. This is the case for the northern Chile late Eocene^early Oligocene and the Central Chile late Miocene^early Pliocene Cu^Mo belts. Although emplaced at different 1114 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS Fig. 13. Schematic model for the generation of the El Teniente ore deposit fertile magmas within the context of the Cenozoic evolution of the Central Chilean Andes. Magmatism from 28 Ma to 6 Ma occurring in the Western Principal Cordillera proceeded with the establishment of a long-lived lower crustal MASH zone, which became progressively enriched through early differentiation of ascending subduction-related, mantle-derived magmas. Following basin inversion, after 21Ma, coeval progressive crustal shortening and thickening induced dehydration melting of the fertile cumulate residual assemblages of the MASH zone. This occurred upon reaching a critical crustal thickness and was prompted by the perturbed thermal gradient induced by the nearly 30 Myr of preceding magmatism. Coevally, owing to arc migration, the volume of primary magmas reaching the MASH zone beneath El Teniente probably decreased. This may also have influenced the fertility of the magmas, by increasing the melt component derived from dehydration melting and decreasing the component derived from primary basalt differentiation. These magmas, after subsequent ascent and differentiation at upper crustal levels, led to the El Teniente porphyry copper deposit formation. Magmatic activity in the Western Principal Cordillera gradually ceased with progressive eastward arc migration, which is finally fully established in the current location of the volcanic arc along the Eastern Principal Cordillera. Uplift and denudation processes during the construction of the modern Andean orogen ultimately exhumed the deposits to the actual surface exposure. WPC, Western Principal Cordillera; EPC, Eastern Principal Cordillera; FC, Frontal Cordillera; Fr, Foreland; CC, continental crust; OC, oceanic crust; SCLM, sub-continental lithospheric mantle. 1115 JOURNAL OF PETROLOGY VOLUME 53 times and through lithologically different sections of the crust (Mpodozis & Ramos, 1989; Camus, 2003; Ramos et al., 2004), they are formed at the end of periods of similar geological evolution; that is, the time after extended arc-related magmatism, during overall waning of igneous activity and cessation of volcanism, and following stages of strong compressive deformation, crustal thickening, uplift and denudation (e.g. Maksaev & Zentilli, 1988; Richards et al., 2001; Tosdal & Richards, 2001). Overall, porphyry-related magmas share a similar composition. The recurrence of deposit formation in different geological frames but following similar histories indicates that they result from the operation of regional processes rather than singularities. In the context of progressive crustal thickening, dehydration melting of a long-lived MASH reservoir can produce highly fertile magmas and/or enhance the fertility of subduction-related, mantle-derived magmas, which can be further enhanced during middle^upper crustal magma chamber processes. Generation of fertile magmas would reduce the high volumes necessary to source metals into at least the giant deposits (Cline & Bodnar, 1991; Richards, 2005; Stern & Skewes, 2005; Stern et al., 2010). However, this constitutes one favorable, but not a key aspect, of a multi-variable process, as several other appropriate conditions also need to be met for their final formation (e.g. Burnham, 1979; Carroll & Rutherford, 1985; Tosdal & Richards, 2001; Cloos, 2002; Richards, 2003, 2005). In this regard, for example, during the formation of the northern Chile Paleocene^Early Eocene metallogenic belt the crust never reached more than 40 km in thickness (for a review see Camus, 2003) and yet several comparatively minor porphyry copper deposits were formed. Other aspects of the geological evolution during this time differentiate this belt from the more productive late Eocene^early Oligocene and late Miocene^early Pliocene belts. However, they may represent examples of how, given the appropriate conditions, dehydration melting processes can enhance magma fertility in different metallogenic epochs. (3) (4) (5) (6) NUMBER 6 JUNE 2012 main igneous zircon populations in all other compositional parameters. El Teniente zircons qualitatively describe an evolutionary trend towards more differentiated compositions, in terms of higher incompatible element enrichment and increasing cooling rates in the progressively younger intrusions. This trend is accompanied by decreasing crystallization temperatures, as measured by Ti-in-zircon thermometry. Such characteristics in the evolving magmatic system agree well with the evolution of the region during this period as characterized by increased uplift and denudation processes during the constructive period of the Andes (Far|¤ as et al., 2008; Maksaev et al., 2009). Zircon Hf and O isotopic composition are uniform at the grain and sample scale and define a single signature for all El Teniente intrusions. This observation indicates a primary control from the source as opposed to any significant crustal contamination processes involved in magma genesis. Hf isotopic composition is inferred to fingerprint the source and O the isotope composition of the melting process. Cenozoic magmatism in Central Chile, including that at El Teniente, shows a remarkably homogeneous Hf isotopic composition over a period of more than 25 Myr. Throughout this time the continental margin was subjected to different configurations and tectonic regimes, indicating that a stable and long-lived MASH-type reservoir in the deep lithosphere has buffered the observed compositions. Dehydration melting reactions in the fertile MASH source probably occur in response to crustal thickening prompted by the anomalous thermal regime governing this zone as a consequence of the long-lived preceding magmatism. This process, coupled with waning arc activity, is likely to generate fertile magmas and/or enhance the fertility of subductionrelated mantle-derived magmas whose later evolution in the upper crust can lead to the formation of porphyry copper deposits on a regional scale. CONC LUSIONS (1) The study of zircon populations from porphyryrelated igneous rocks provides valuable information on petrogenetic processes that are otherwise obscured by the intensive hydrothermal alteration that characteristically affects these rocks. (2) Zircons from mineralization-related intrusive rocks from the El Teniente porphyry Cu^Mo deposit show morphological and compositional features that indicate crystallization from a series of cogenetic melts. Some grains show evidence of a minor hydrothermal imprint in high U^Th weakly luminescent overgrowth rims, which are indistinguishable from the AC K N O W L E D G E M E N T S We thank the geologists Patricio Zu¤n‹iga, Ricardo Floody and Jose¤ Seguel from the Superintendencia de Geolog|¤ a El Teniente, CODELCO-Chile, for providing mine access and logistical support for the development of this work. Help in organizing visits to El Teniente and assistance during the field work of Marcela Cereceda and Rene¤ Padilla is gratefully acknowledged. We thank Dr G. Yaxley and the technical staff of the Research School of Earth Sciences, Australian National University (RSES-ANU), and technical staff of the Departamento de Geolog|¤ a, Universidad de Chile, for their helpful assistance in 1116 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS sample preparation and during the analytical sessions. We also thank the reviewers J. P. Richards and M. Chiaradia, and the Editor G. Wo«rner, whose comments and discussions have helped us greatly to improve the final paper. FUNDING This work was supported by the Chilean government through the Comisio¤n Nacional de Ciencia y Tecnolog|¤ a^ CONICYT (Anillo ACT-18 project, PBCT program). Additional funding was provided by the Departamento de Postgrado y Post|¤ tulo, Universidad de Chile. This work is part of the PhD thesis of M. Mun‹oz, which was supported by a 4 years grant from CONICYT. R EF ER ENC ES Annen, C. & Sparks, R. S. J. (2002). Effects of repetitive emplacement of basaltic intrusions on thermal evolution and melt generation in the crust. Earth and Planetary Science Letters 202, 937^955. Annen, C., Blundy, J. D. & Sparks, R. S. J. (2006). The genesis of intermediate and silicic magmas in deep crustal hot zones. Journal of Petrology 47, 505^539. Antofagasta plc. (2009). Annual Report and Financial Statements. London: Antofagasta plc, 120 p. Ballard, J. R., Palin, J. M. & Campbell, I. H. (2002). Relative oxidation states of magmas inferred from Ce(IV)/Ce(III) in zircon: applications to porphyry copper deposits of northern Chile. Contributions to Mineralogy and Petrology 144, 347^364. Benisek, A. & Finger, F. (1993). Factors controlling the development of prism faces in granite zircons: a microprobe study. Contributions to Mineralogy and Petrology 114, 441^151. Berger, J., Renaud, C., Lie¤geois, J.-P., Mercier, J.-C. C. & Demaiffe, D. (2009). Dehydration, melting and related garnet growth in deep root of the Amalaoulaou Neoproterozoic magmatic arc (Gourma, NE Mali). Geological Magazine 146, 173^186. Blichert-Toft, J. & Albare'de, F. (1997). The Lu^Hf isotope geochemistry of chondrites and the evolution of the mantle^crust system. Earth and Planetary Science Letters 148, 243^258. Burnham, C. W. (1979). Magmas and hydrothermal fluids. In: Barnes, H. L. (ed.) Geochemistry of Hydrothermal Ore Deposits. New York: John Wiley, pp. 71^136. Campbell, I. H. & Turner, J. S. (1987). A laboratory investigation of assimilation at the top of a basaltic magma chamber. Journal of Geology 95, 155^172. Camus, F. (1975). Geology of El Teniente orebody with emphasis on wall-rock alteration. Economic Geology 70, 1341^1372. Camus, F. (2003). Geolog|¤ a de los Sistemas Porf|¤ ricos en los Andes de Chile. Santiago: Servicio Nacional de Geolog|¤ a y Miner|¤ a, 267 p. Cannell, J. (2004). El Teniente porphyry copper^molybdenum deposit, Central Chile. PhD thesis, University of Hobart, 317 p. Cannell, J., Cooke, D. R., Walshe, J. L. & Stein, H. (2005). Geology, mineralization, alteration, and structural evolution of the El Teniente porphyry Cu^Mo deposit. Economic Geology 100, 979^1003. Cannell, J., Cooke, D. R., Walshe, J. L. & Stein, H. (2007). Geology, mineralization, alteration and structural evolution of the El Teniente porphyry Cu^Mo depositçA reply. Economic Geology 102, 1171^1180. Carroll, M. R. & Rutherford, M. J. (1985). Sulfide and sulfate saturation in hydrous silicate melts. Journal of Geophysical Research 90, C601^C612. Charrier, R., Baeza, O., Elgueta, S., Flynn, J. J., Gana, P., Kay, S. M., Mun‹oz, N., Wyss, A. R. & Zurita, E. (2002). Evidence for Cenozoic extensional basin development and tectonic inversion south of the flat-slab segment, southern Central Andes, Chile (338^368S.L.). Journal of South American Earth Sciences 15, 117^139. Charrier, R., Pinto, L. & Rodr|¤ guez, M. P. (2007). Tectonostratigraphic evolution of the Andean orogen in Chile. In: Moreno, T. & Gibbons, W. (eds) The Geology of Chile. London: Geological Society, pp. 21^114. Charrier, R., Far|¤ as, M. & Maksaev, V. (2009). Evolucio¤n tecto¤nica, pelogeogra¤fica y metaloge¤nica durante el Cenozoico en los Andes de Chile Norte y Central e implicaciones para las regiones adyacentes de Bolivia y Argentina. Revista de la Asociacio¤ n Geolo¤ gica Argentina 65, 5^35. Cherniak, D. J. & Watson, E. B. (2003). Diffusion in zircon. In: Hanchar, J. M. & Hoskin, P. W. O. (eds) Zircon. Mineralogical Society of America and Geochemical Society, Reviews in Mineralogy and Geochemistry 53, 113^143. Chiaradia, M., Merino, D. & Spikings, R. (2009). Rapid transition to long-lived deep crustal magmatic maturation and the formation of giant porphyry-related mineralization (Yanacocha, Peru). Earth and Planetary Science Letters 288, 505^515. Cline, J. S. & Bodnar, R. J. (1991). Can economic porphyry copper mineralization be generated by a typical calc-alkaline melt? Journal of Geophysical Research 96, 8113^8126. Cloos, M. (2002). Bubbling magma chambers, cupolas and porphyry copper deposits. In: Ernst, G. (ed.) Frontiers in Geochemistry: Organic, Solution, and Ore Deposit Geochemistry. International Book Series 6, 191^217. CODELCO. (2010). Annual Memoir. Santiago: CODELCO, 236 p. Core, D. P., Stephen, E. K. & Essene, E. J. (2006). Unusually Cu-rich magmas associated with giant porphyry deposits: evidences from Bingham, Utah. Geology 34, 41^44. Cornejo, P., Tosdal, R. M., Mpodozis, C., Tomlinson, A. J., Rivera, O. & Fanning, C. M. (1997). El Salvador, Chile porphyry copper deposit revisited: geological and geochronologic framework. International Geology Review 39, 22^54. Cuadra, P. (1986). Geocronolog|¤ a K^Ar del yacimiento El Teniente y a¤reas adyacentes. Revista Geolo¤ gica de Chile 27, 3^26. Davidson, J. P., Turner, S. P., Handley, H., Macpherson, C. & Dosseto, A. (2007). Amphibole ‘sponge’ in arc crust? Geology 35, 787^790. Deckart, K., Clark, A. H., Aguilar, R., Vargas, R., Bertens, A., Mortensen, J. & Fanning, M. (2005). Magmatic and hydrothermal chronology of the giant R|¤o Blanco porphyry copper deposit, Central Chile: implications of an integrated U^Pb and 40Ar/39Ar database. Economic Geology 100, 905^934. Deckart, K., Godoy, E., Bertens, A., Jere¤z, D. & Saeed, A. (2010). Barren Miocene granitoids in the central Andean metallogenic belt, Chile: geochemistry and Nd^Hf and U^Pb isotope systematics. Andean Geology 37, 1^31. DePaolo, D. J. (1981). Trace element and isotopic effects of combined wall rock assimilation and fractional crystallization. Earth and Planetary Science Letters 53, 189^202. Eiler, J. M. (2001). Oxygen isotope variations in basaltic lavas and upper mantle rocks. In: Valley, J. W. & Cole, D. (eds) Stable Isotope Geochemistry. Mineralogical Society of America and Geochemical Society, Reviews in Mineralogy and Geochemistry 43, 319^364. Far|¤ as, M., Charrier, R., Carretier, S., Martinod, J., Fock, A., Campbell, D., Ca¤ceres, J. & Comte, D. (2008). Late Miocene high 1117 JOURNAL OF PETROLOGY VOLUME 53 and rapid surface uplift and its erosional response in the Andes of Central Chile (338^358S). Tectonics 27, TC1005. Far|¤ as, M., Comte, D., Charrier, R., Martinod, J., Tassara, A. & Fock, A. (2010). Crustal-scale structural architecture of the Central Chile Andes based on 3D seismic tomography, seismicity, and surface geology: Implications for mountain building in subduction zones. Tectonics 29, TC3006. Faunes, A. (1981). Caracterizacio¤n de la mineralog|¤ a y de la alteracio¤n en un sector del stock tonal|¤ tico del yacimiento El Teniente. BSc thesis, University of Chile, Santiago, 175 p. Ferry, J. M. & Watson, E. B. (2007). New thermodynamic models and revised calibrations for the Ti-in-zircon and Zr-in-rutile thermometers. Contributions to Mineralogy and Petrology 154, 429^437. Fock, A., Charrier, R., Far|¤ as, M. & Mun‹oz, M. (2006). Fallas de vergencia oeste en la Cordillera Principal de Chile Central: inversio¤n de la cuenca de Abanico (338^348S). Revista de la Asociacio¤ n Geolo¤ gica Argentina, Special Publication Series 6, 48^55. Fu, B., Page, F. Z., Cavosie, A. J., Clechenki, C. C., Fournelle, J., Kita, N. T., Lackey, J. S., Wilde, S. A. & Valley, J. W. (2008). Ti-in-zircon thermometry: applications and limitations. Contributions to Mineralogy and Petrology 156, 197^212. Garrido, C. J., Bodinier, J.-L., Burg, J.-P., Zeilinger, G., Hussain, S. S., Dawood, H., Chaudhry, M. N. & Gervilla, F. (2006). Petrogenesis of mafic garnet granulite in the lower crust of the Kohistan paleo-arc complex (Northern Pakistan): implications for intracrustal differentiation of island arcs and generation of continental crust. Journal of Petrology 10, 1873^1914. Getsinger, A., Rushmer, T., Jackson, M. D. & Baker, D. (2009). Generating high Mg-numbers and chemical diversity in tonalite^ trondhjemite^granodiorite (TTG) magmas during melting and melt segregation in the continental crust. Journal of Petrology 50, 1935^1954. Giambiagi, L. B. & Ramos, V. (2002). Structural evolution of the Andes between 33830 and 33845’S, above the transition zone between the flat and normal subduction segment, Argentina and Chile. Journal of South American Earth Sciences 15, 99^114. Gonza¤lez, R. (2006). Petrograf|¤ a, geoqu|¤ mica y microtermometr|¤ a de los intrusivos fe¤lsicos del yacimiento El Teniente. BSc thesis, University of Concepcio¤n, 149 p. Goodge, J. W. & Vervoort, J. D. (2006). Origin of Mesoproterozoic A-type granites in Laurentia: Hf isotope evidence. Earth and Planetary Science Letters 243, 711^731. Hedenquist, J. W. & Lowenstern, J. B. (1994). The role of magmas in the formation of hydrothermal ore deposits. Nature 370, 519^527. Hiess, J., Nutman, A. P., Bennett, V. C. & Holden, P. (2008). Ti-in-zircon thermometry applied to contrasting Archean metamorphic and igneous systems. Chemical Geology 247, 323^338. Hildreth, W. & Moorbath, S. (1988). Crustal contributions to arc magmatism in the Andes of Central Chile. Contributions to Mineralogy and Petrology 98, 455^489. Hitschfeld, M. (2006). Petrograf|¤ a y geoqu|¤ mica de los intrusivos leucocra¤ticos del sector sur este del yacimiento El Teniente, VI Regio¤n, Chile. BSc thesis, University of Concepcio¤n, 129 p. Hoskin, P. W. O. (1998). Minor and trace element analysis of natural zircon (ZrSiO2) by SIMS and laser ablation ICPMS: a consideration and comparison of two broadly competitive techniques. Journal of Trace and MicroprobeTechniques 16, 301^326. Hoskin, P. W. O. & Schaltegger, U. (2003). The composition of zircon and igneous and metamorphic petrogenesis. In: Hanchar, J. M. & Hoskin, P. W. O. (eds) Zircon. Mineralogical Society of America and Geochemical Society, Reviews in Mineralogy and Geochemistry 53, 27^62. Howell, F. H. & Molloy, J. S. (1960). Geology of the Braden orebody, Chile, South America. Economic Geology 55, 863^905. NUMBER 6 JUNE 2012 Huppert, H. E. & Sparks, R. S. J. (1988). The generation of granitic magmas by intrusion of basalt into continental crust. Journal of Petrology 29, 599^624. Ickert, R. B., Hiess, J., Williams, I. S., Holden, P., Ireland, T. R., Lanc, P., Schram, N., Foster, J. J. & Clement, S. W. (2008). Determining high precision, in situ, oxygen isotope ratios with SHRIMP II: analyses of MPI-DING silicate-glass reference materials and zircon from contrasting granites. Chemical Geology 257, 114^128. Ireland, T. R. & Wlotzka, F. (1992). The oldest zircons in the solar system. Earth and Planetary Science Letters 109, 1^10. Isacks, B. L. (1988). Uplift of the Andean plateau and bending of the Bolivian orocline. Journal of Geophysical Research 93, 3211^3231. Jagoutz, O., Muntener, O., Ulmer, P., Pettke, T., Burg, J.-P., Dawood, H. & Hussain, S. (2007). Petrology and mineral chemistry of lower crustal intrusions: the Chilas Complex, Kohistan (NW Pakistan). Journal of Petrology 48, 1895^1953. Kay, S. M. & Mpodozis, C. (2001). Central Andes ore deposits linked to evolving shallow subduction systems and thickening crust. GSA Today 11, 4^9. Kay, S. M. & Mpodozis, C. (2002). Magmatism as a probe to the Neogene shallowing of the Nazca plate beneath the modern Chilean flat-slab. Journal of South American Earth Sciences 15, 39^57. Kay, S. M., Godoy, E. & Kurtz, A. (2005). Episodic arc migration, crustal thickening, subduction erosion, and magmatism in the South^Central Andes. Geological Society of America Bulletin 117, 67^88. Klemm, L. M., Pettke, T., Heinrich, C. A. & Campos, E. (2007). Hydrothermal evolution of the El Teniente deposit, Chile: porphyry Cu^Mo ore deposition from low-salinity magmatic fluids. Economic Geology 102, 1021^1045. Kusakabe, M., Hori, M., Nakagawa, S., Matsushisa, Y., Ojeda, J. M. & Serrano, L. M. (1984). Oxygen and sulfur isotopic compositions of quartz, anhydrite and sulfide minerals from the El Teniente and R|¤o Blanco porphyry copper deposits, Chile. Bulletin of the Geological Survey of Japan 35, 583^614. Kusakabe, M., Hori, M. & Matsushisa,Y. (1990). Primary mineralization^alteration of the El Teniente and R|¤o Blanco porphyry copper deposits, Chile: stable isotopie, fluid inclusion and Mgþ2/ Feþ2/Feþ3 ratios of hydrothermal fluids. In: Herbert, H. K. & Ho, S. E. (eds) Stable Isotopes and Fluid Processes in Mineralization. Perth: University of Western Australia, pp. 244^259. Laursen, J., Scholl, D. W. & Von Huene, R. (2002). Neotectonic deformation of the Central Chile margin: deep water forearc basin formation in response to hot spot ridge and seamount subduction. Tectonics 21, 1038, doi:10.1029/2001TC901023. Lindgren, W. (1933). Mineral Deposits, 4th edn. New York: McGraw^ Hill, 930 p. Lindgren, W. & Bastin, E. S. (1922). Geology of the Braden mine, Rancagua, Chile. Economic Geology 17, 863^905. Lo¤pez, S & Castro, A. (2001). Determination of the fluid-absent solidus and supersolidus phase relationships of MORB-derived amphibolites in the range 4^14 kbar. American Mineralogist 86, 1396^2001. Ludwig, K. R. (2003). Isoplot 3.00. Berkeley Geochronology Center Special Publication 4. Maksaev, V. & Zentilli, M. (1988). Marco metaloge¤nico regional de los megadepo¤sitos de tipo Po¤rfido Cupr|¤ fero del Norte Grande de Chile. In: Proceedings V Congreso Geolo¤ gico Chileno Santiago 1: B181^B212 (Acta). Maksaev, V., Munizaga, F., McWilliams, M., Fanning, C. M., Mathur, R., Ruiz, J. & Zentilli, M. (2004). New chronology for El Teniente, Chilean Andes, from U^Pb, 40Ar/39Ar, Re^Os, and fission-track dating: implications for the evolution of a supergiant 1118 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS porphyry Cu^Mo deposit. Society of Economic Geologists Special Publications 11, 15^54. Maksaev, V., Townley, B., Palacios, C. & Camus, F. (2007). Metallic ore deposits. In: Moreno, T. & Gibbons, W. (eds) The Geology of Chile. London: Geological Society, pp. 180^199. Maksaev, V., Munizaga, F., Zentilli, M. & Charrier, R. (2009). Fission track thermochronology of Neogene plutons in the Principal Andean Cordillera of Central Chile (33^358S): implications for tectonic evolution and porphyry Cu^Mo mineralization. Revista Geolo¤ gica de Chile 36, 153^171. Mathur, R., Ru|¤ z, J. & Munizaga, F. (2000). Relationship between copper tonnage of Chilean base-metal porphyry deposits and Os isotope ratios. Geology 28, 555^558. Mathur, R., Ru|¤ z, J. R. & Munizaga, F. M. (2001). Insights into Andean metallogenesis from the perspective of Re^Os analyses of sulfides. In: Extended Abstracts III South American Symposium on Isotope Geology, Puco¤n, 500^503 (CD). McDonough, D. F. & Sun, S.-s. (1995). The composition of the earth. Chemical Geology 120, 223^253. McKee, E. H. & Noble, D. C. (1989). Cenozoic tectonic events, magmatic pulses, and base- and precious-metal mineralization in the central Andes. In: Ericksen, G. E., Can‹as, M. T. & Reinemund, J. A. (eds) Geology of the Andes and its Relation to Hydrocarborn and Mineral Resources. Earth Science Series, Vol. 11. Houston, TX: Circum-Pacific Council for Energy and Mineral Resources, pp. 189^194. Montecinos, P., Scha«rer, U., Vergara, M. & Aguirre, L. (2008). Lithospheric origin of Oligocene^Miocene magmatism in Central Chile: U^Pb ages and Sr^Pb^Hf isotope composition of minerals. Journal of Petrology 49, 555^580. Mpodozis, C. & Ramos, V. A. (1989). The Andes of Chile and Argentina. In: Ericksen, G. E., Can‹as, M. T. & Reinemund, J. A. (eds) Geology of the Andes and its Relation to Hydrocarbon and Mineral Resources. Earth Science Series, Vol. 11. Houston, TX: Circum-Pacific Council for Energy and Mineral Resources, pp. 59^90. Mungall, J. (2002). Roasting the mantle: slab melting and the genesis of major Au and Au-rich Cu deposits. Geology 30, 915^918. Munizaga, F., Maksaev, V., Fanning, C. M., Giglio, S., Yaxley, G. & Tassinari, C. C. G. (2008). Late Paleozoic^Early Triassic magmatism on the western margin of Gondwana: Collahuasi area, Northern Chile. Gondwana Research 13, 407^427. Mun‹oz, M., Deckart, K., Charrier, R. & Fanning, C. M. (2009). New geochronological data on Neogene^Quaternary intrusive rocks from the high Andes of Central Chile (33815’^34800’S). In: Proceedings XII Congreso Geolo¤ gico Chileno, Santiago, S8^008. Oyarzu¤n, R., Ma¤rquez, A., Lillo, J., Lo¤pez, I. & Rivera, S. (2001). Giant versus small porphyry copper deposits of Cenozoic age in northern Chile: adakitic versus normal calc-alkaline magmatism. Mineralium Deposita 36, 794^798. Pardo-Casas, F. & Molnar, O. (1987). Relative motion of the Nazca (Farallon) and South American plate since Late Cretaceous times. Tectonics 6, 233^248. Perello¤, J., Sillitoe, R. H., Brockway, H., Posso, H. & Mpodozis, C. (2009). Contiguous porphyry Cu^Mo and Cu^Au mineralization at Los Pelambres, central Chile. In: Proceedings XII Congreso Geolo¤ gico Chileno, Santiago, S11^026. Petford, N. & Gallagher, K. (2001). Partial melting of mafic (amphibolitic) lower crust by periodic influx of basaltic magma. Earth and Planetary Science Letters 193, 483^499. Pupin, J. P. (1980). Zircon and granite petrology. Contributions to Mineralogy and Petrology 73, 207^220. Rabbia, O. M. (2002). Cristaloqu|¤ mica de rutilo y anatasa en sistemas de po¤rfidos cupr|¤ feros andinos: evaluacio¤n de su uso como monitores de la actividad de metales en flu|¤ dos hidrotermales corticales. PhD thesis, University of Chile, Santiago, 147 p. Rabbia, O. M., Reich, M., Herna¤ndez, L. B., King, R. & Lo¤pez Escobar, L. (2000). High-Al TTG-like suite at the El Teniente porphyry copper deposit Chile1. In: Proceedings IX Congreso Geolo¤ gico Chileno, Puerto Varas, 326^300. Ramos, V., Zapata, T., Cristallini, E. & Introcaso, A. (2004). The Andean thrust systemçlatitudinal variations in structural styles and orogenic shortening. In: McClay, K. R. (ed.) Thrust Tectonics and Hydrocarborn Systems. AAPG Memoirs 82, 30^50. Rapp, R. P. & Watson, E. B. (1995). Dehydration melting of metabasalt at 8^32 kbar: implications for crust^mantle boundary recycling. Journal of Petrology 36, 891^893. Reich, M. (2001). Estudio petrogra¤fico, mineraloqu|¤ mico y geoqu|¤ mico de los cuerpos intrusivos de Sewell y La Huifa, yacimiento El Teniente, VI Regio¤n, Chile. BS. thesis, University of Concepcio¤n, 104 p. Richards, J. P. (2003). Tectono-magmatic precursors for porphyry Cu^(Mo^Au) deposit formation. Economic Geology 98, 1515^1533. Richards, J. P. (2005). Cumulative factors in the generation of giant calc-alkaline porphyry Cu deposits. In: Porter, T. M. (ed.) Super Porphyry Copper and Gold Deposits: A Global Perspective, Vol. 1. Adelaide: PGC Publishing, pp. 7^25. Richards, J. P. (2009). Postsubduction porphyry Cu^Au and epithermal Au deposits: products of remelting of subduction-modified lithosphere. Geology 37, 247^250. Richards, J. P., Boyce, A. J. & Pringle, M. S. (2001). Geologic evolution of the Escondida area, Northern Chile: a model for spatial an temporal localization of porphyry Cu mineralization. Economic Geology 96, 271^305. Rivera, O. & Falco¤n, M. (2000). Las formaciones Farellones, Coya^ Machal|¤ y Abanico en los alrededores del yacimiento El Teniente: secuencias de cuencas volcano-tecto¤nicas transversales del Oligoceno^Mioceno de Chile central (33845’^34830’ LS). Proceedings IX Congreso Geolo¤ gico Chileno, Puerto Varas 1, 819^823. Rojas, A. (2003). Petrograf|¤ a y geoqu|¤ mica del Po¤rfido Teniente, ubicado en el sector norte del yacimiento El Teniente, Provincia de Cachapoal, VI Regio¤n, Chile. BSc thesis, University of Concepcio¤n, 133 p. Sen, C. & Dunn, T. (1994). Dehydration melting of a basaltic composition amphibolite at 1·5 and 2·0 GPa: implications for the origin of adakites. Contributions to Mineralogy and Petrology 117, 394^409. Shafiei, B., Haschke, M. & Shahabpour, J. (2009). Recycling of orogenic arc crust triggers porphyry Cu mineralization in Kerman Cenozoic arc rocks, southeastern Iran. Mineralium Deposita 44, 265^283. Sillitoe, R. H. (1972). A plate tectonic model for the origin of porphyry copper deposits. Economic Geology 67, 184^197. Skewes, A. & Holmgren, C. (1993). Solevantamiento andino, erosio¤n y emplazamiento de brechas mineralizadas en el depo¤sito de cobre porf|¤ dico Los Bronces, Chile central (338S): Aplicacio¤n de geotermometr|¤ a de inclusiones fluidas. Revista Geolo¤ gica de Chile 20, 71^84. Skewes, A. & Stern, C. R. (2007). Geology, mineralization, and structural evolution of the El Teniente porphyry Cu^Mo depositçA discussion. Economic Geology 102, 1165^1180. Skewes, A., Are¤valo, A., Holmgren, C. & Stern, C. R. (2001). Stable isotope evidence for the formation from magmatic fluids of the mineralized breccias in Los Bronces and El Teniente copper deposits, Central Chile. Extended Abstracts III South American Symposium on Isotope Geology, Puco¤ n 531^534 (CD). Skewes, A., Are¤valo, A., Floody, R. & Zun‹iga, P. (2002). The giant El Teniente breccia deposit: hypogene copper distribution and 1119 JOURNAL OF PETROLOGY VOLUME 53 emplacement. In: Goldfarb, R. J. & Nielson, R. L. (eds) Global Exploration 2002: Integrated Methods of Discovery. Society of Economic Geologists Special Publication 9, 299^332. Skewes, A., Are¤valo, A., Floody, R., Zun‹iga, P. & Stern, C. R. (2005). The El Teniente Megabreccia Deposit, the world’s largest copper deposit. In: Porter, T. M. (ed.) Super Porphyry Copper and Gold Deposits: A Global Perspective, Vol. 1. Adelaide: PGC Publishing, pp. 88^113. So«derlund, U., Patchett, P. J., Vervoort, J. D. & Isachsen, C. E. (2004). The 176Lu decay constant determined by Lu^Hf and U^Pb isotope systematics of Precambrian mafic intrusions. Earth and Planetary Science Letters 219, 311^324. Stern, C. R. (1991). Role of subduction erosion in the generation of Andean magmas. Geology 19, 78^81. Stern, C. R. & Skewes, A. (2005). Origin of giant Miocene and Pliocene Cu^Mo deposits in Central Chile: Role of ridge subduction, decreased subduction angle, and long-lived, batholith size, open-system magma chambers. In: Porter, T. M. (ed.) Super Porphyry Copper and Gold Deposits: A Global Perspective, Vol. 1. Adelaide: PGC Publishing, pp. 65^82. Stern, C. R. & Skewes, M. A. (1995). Miocene to present magmatic evolution at the northern end of the Andean Southern Volcanic Zone, Central Chile. Revista Geolo¤ gica de Chile 22, 261^271. Stern, C. R., Funk, J. A., Skewes, M. A. & Are¤valo, A. (2007). Magmatic anhydrite in plutonic rocks at the El Teniente Cu^Mo deposit, Chile, and the role of sulfur- and copper-rich magmas in its formation. Economic Geology 102, 1335^1344. Stern, C. R., Skewes, A. & Are¤valo, A. (2010). Magmatic evolution of the giant El Teniente Cu^Mo deposit, Central Chile. Journal of Petrology 52, 1591^1617, doi:10.1093/petrology/egq029. Tassara, A., Go«tze, H.-J., Schmidt, S. T. & Hackney, R. (2006). Three-dimensional density model of the Nazca plate and the Andean continental margin. Journal of Geophysical Research 111, B09404. Tosdal, R. M. & Richards, J. P. (2001). Magmatic and structural controls on the development of porphyry Cu Mo Au deposits. Reviews in Economic Geology 14, 157^181. Valley, J. W., Chiarenzelli, J. R. & McLelland, J. M. (1994). Oxygen isotope geochemistry of zircon. Earth and Planetary Science Letters 126, 187^206. Valley, J. W., Kinny, P., Schulze, D. J. & Spicuzza, M. J. (1998). Zircon megacrysts from kimberlite: oxygen isotope variability among mantle melts. Contributions to Mineralogy and Petrology 133, 1^11. van Dongen, M., Weinberg, R. F., Tomkins, A. G., Armstrong, R. A. & Woodhead, J. D. (2010). Recycling of Proterozoic crust in Pleistocene juvenile magma and rapid formation of the Ok Tedi porphyry Cu^Au deposit, Papua New Guinea. Lithos 114, 282^292. Vavra, G. (1994). Systematics of internal zircon morphology in major Variscan granitoids types. Contributions to Mineralogy and Petrology 117, 331^334. Vervoort, J. D. (2010). Hf analysis in zircon by LA-MC-ICPMS: promise and pitfalls. Geological Society of America, Abstracts with programs 42(5), 667. Vervoort, J. D. & Blichert-Toft, J. (1999). Evolution of the depleted mantle: Hf isotope evidence from juvenile rocks through time. Geochimica et Cosmochimica Acta 63, 533^556. Vry, V. H., Wilkinson, J. J., Seguel, J. & Milla¤n, J. (2010). Multistage intrusion brecciation and veining at El Teniente, Chile: evolution of a nested porphyry system. Economic Geology 105, 119^153. Watson, E. B. & Cherniak, D. J. (1997). Oxygen diffusion in zircon. Earth and Planetary Science Letters 148, 527^544. NUMBER 6 JUNE 2012 Wolf, M. B. & Wyllie, P. J. (1993). Garnet growth during amphibolite anatexisçimplications of a garnetiferous restite. Journal of Geology 101, 357^373. Wolf, M. B. & Wyllie, P. J. (1994). Dehydration-melting of amphibolite at 10 kbar: the effects of temperature and time. Contributions to Mineralogy and Petrology 115, 369^383. Woodhead, J. & Hergt, J. (2005). A preliminary appraisal of seven natural zircon reference materials for in situ Hf isotope determination. Geostandards and Geoanalytical Research 29, 183^195. Wyllie, P. J. & Wolf, M. B. (1993). Amphibolite dehydration-melting: sorting out the solidus. In: Pritchard, H. M., Alabaster, T., Harris, N. B. W. & Neary, C. R. (eds) Magmatic Processes and Plate Tectonics. Geological Society, London, Special Publications 76, 405^416. Zhao, Z.-F. & Zheng, Y.-F. (2003). Calculation of oxygen isotope fractionation in magmatic rocks. Chemical Geology 193, 59^80. Zheng, Y.-F. (1991). Calculation of oxygen isotope fractionation in metal oxides. Geochimica et Cosmochimica Acta 55, 2299^2307. Zheng, Y.-F. (1993a). Calculation of oxygen isotope fractionation in anhydrous silicate minerals. Geochimica et Cosmochimica Acta 57, 1079^1091. Zheng, Y.-F. (1993b). Calculation of oxygen isotope fractionation in hydroxyl-bearing silicates. Earth and Planetary Science Letters 120, 247^263. Zheng, Y.-F. & Fu, B. (1998). Estimation of oxygen diffusivity from anion porosity in minerals. Geochemical Journal 32, 71^89. A P P E N D I X : C A L C U L AT I O N O F O I S O T O P E F R AC T I O N AT I O N I N E X P E R I M E N TA L D E H Y D R AT I O N M E LT I N G R E AC T I O N S The model presented in this study to estimate O isotope fractionation in dehydration melting reactions (Fig. 12) follows the calculation approach of Eiler (2001) using the experimental results of Wolf & Wyllie (1994) and Getsinger et al. (2009). According to this approach, the melt is assumed to behave as the weighted sums of its normative constituents. This principle is applied in combination with fractionation factors to calculate the expected isotopic distribution among all coexisting phases in the observed assemblages in the melting experiments. An initial composition of d18OWS ¼ 5·8ø has been used for the whole system, which corresponds to that of a mantle-derived basaltic magma (5·5ø) contaminated with 3% of subducted sediment (15ø; Eiler, 2001). This value has been assumed considering that a long-lived deep MASH zone would ultimately have the O isotope signature of subduction-related, mantle-derived magmas. The experimental results of Wolf & Wyllie (1994) and Getsinger et al. (2009) reported for runs at various temperatures the volume per cent of melt and residue, the melt major element chemical composition, and the volume per cent of the minerals composing the residue. For each experimental run, this information has been used in the calculation of O isotope fractionation in the system as described below, where calculations for the 1120 MUN‹OZ et al. GENESIS OF PORPHYRY Cu DEPOSITS 8508C run of Wolf & Wyllie (1994) are shown as an example (Table A1). (1) Separation of the melt into mineral components by calculating the CIPW normative mineralogy (Table A1). Melts from both experiments considered share most of the normative constituents, which are quartz, albite, anorthite, orthoclase, corundum, hypersthene, ilmenite and magnetite. Melts from the Getsinger et al. (2009) experiments also incorporate apatite. (2) Calculation of mineral-pair oxygen isotope fractionation factors. This has been carried out for a set of minerals that include the melt normative constituents and the residue mineralogy. The fractionation factors for quartz^mineral pairs of Zheng (1991, 1993a, 1993b) have been used, which are expressed in the form (3) Calculation of oxygen isotope fractionation between melt and minerals composing the residue. Considering that 103ln (aMx1^Mx2) Mx1^Mx2, this can be done assuming that the melt behaves as the sum of its normative constituents (Table A1) and using the data from Table A3 (Table A4). (4) Calculation of d 18OMelt according to the following considerations. For the whole system 103 lnðaMx1Mx2 Þ ¼ ðA 106 Þ=T 2 þ ðB 103 Þ=T þ C Table A2: Oxygen isotope fractionation factors for quartz^ mineral pairs (with T in K). They include factors for anhydrous silicate minerals, hydroxyl-bearing silicates and metal oxides (Table A2). For each experimental run a matrix of oxygen isotope fractionation factors has been calculated using these data along with the corresponding temperature (Table A3). Table A1: Main parameters of the 8508C experimental run (No. 150) of Wolf & Wyllie (1994) and melt normative composition Whole system Phases and abundances* Chemical composition Normative composition Melt Grt Opx Hbl Cpx Pl 12·0 4·0 2·7 52·0 11·0 18·3 SiO2 TiO2 Al2O3 FeO MnO MgO CaO Na2O K2O wt % Normative minerals Relative proportion 64·79 0·43 19·07 4·51 0·18 1·51 7·95 1·25 0·29 Qtz Ab An Kfs Crs Hy Ilm Mag 0·35 0·11 0·39 0·02 0·02 0·09 0·01 0·01 ðaÞ where M and R correspond to the proportion of melt and residue, respectively, and M þ R ¼ 1. Mineral A B C Ref.* Ab Kfs An Opxy Cpxz Hbl Py Mag Ilm Crs 0·15 0·16 0·36 0·51 0·56 0·59 0·73 1·22 1·36 2·00 1·39 1·50 2·73 3·45 3·67 3·80 4·30 8·22 8·61 9·94 0·57 0·62 1·14 1·44 1·53 1·59 1·80 4·35 4·57 5·32 1 1 1 1 1 2 1 3 3 3 *Fractionation factors taken from: 1, Zheng (1993a); 2, Zheng (1993b); 3, Zheng (1991). yTaken after hypersthene. zTaken after diopside. Melt vol. % d18 OWS ¼ Md18 OMelt þ Rd18 OResidue Table A3: Matrix of oxygen isotope fractionation factors between quartz and selected minerals at 8508C Qtz Ab Kfs Mineral abbreviations used in this and the following tables: Ab, albite; An, anorthite; Cpx, clinopyroxene; Crs, corundum; Grt, garnet; Hbl, hornblende; Hy, hypersthene; Ilm, ilmenite; Kfs, K-feldspar; Mag, magnetite; Opx, orthopyroxene; Pl, plagioclase; Py, pyrope; Qtz, quartz. *Modes indicated by Wolf & Wyllie (1994) are derived from BSE and SEM-EDS analyses. In this work melt and garnet abundances have been taken from the BSE data, when available, and remaining mineral abundances from the SEM-EDS data recalculated for fitting the whole system at 100%. An Opx Cpx Hbl Py Mag Ilm 1121 Ab Kfs An Opx Cpx Hbl Py Mag Il Crs 0·79 0·84 1·58 2·04 2·18 2·26 2·61 3·94 4·17 5·12 0·06 0·79 1·25 1·40 1·47 1·82 3·15 3·39 4·33 0·73 1·19 1·34 1·42 1·76 3·09 3·33 4·27 0·46 0·61 0·69 1·03 2·36 2·60 3·54 0·15 0·23 0·57 1·90 2·14 3·08 0·08 0·43 1·75 1·99 2·93 0·35 1·67 1·91 2·85 1·33 1·57 2·51 0·24 1·18 0·94 JOURNAL OF PETROLOGY VOLUME 53 Table A4: Oxygen isotope fractionation between melt and minerals composing the residue of the 8508C experimental run of Wolf & Wyllie (1994) Melt Mineral phases in the residue Norm. Hbl An Cpx Opx Py NUMBER 6 JUNE 2012 Using (b) and considering that Melt^MxI ¼ d18OMelt d OMxI equation (a) can be rewritten as X I d18 OMelt MeltMxI d18 OWS ¼ Md18 OMelt þ X X d18 OWS ¼ M þ I d18 OMelt IMeltMxI ðcÞ X 18 18 d OWS ¼ ðM þ RÞd OMelt IMeltMxI X d18 OMelt ¼ d18 OWS þ IMeltMxI 18 min. Qtz 0·78 0·55 0·76 0·70 0·90 Ab 0·16 0·08 0·15 0·13 0·19 An 0·27 0·00 0·24 0·18 0·41 Kfs 0·02 0·01 0·02 0·02 0·03 Crs 0·06 0·08 0·07 0·07 0·06 0·02 0·04 0·01 0·00 0·05 Ilm 0·02 0·02 0·02 0·02 0·01 Mag 0·02 0·03 0·03 0·03 0·02 0·46 1·07 0·92 1·50 Opx Melt–Mx 1·15 Expression (c) allows calculation of the d18OMelt with the assumed d18OWS ¼ 5·8ø and the data reported in Table A4. Thus, a d18OMelt ¼ 6·7ø is obtained for the 8508C experimental run of Wolf & Wyllie (1994). (5) Calculation of d18OResidue ¼ (d18OWS Md18OMelt)/ R, which is 5·7ø for the preceding example. (6) Calculation of d 18O for zircon in equilibrium with the melt (d18OZrc). This has been carried out according to the relation given by Valley et al. (1994) where d18ORock ¼ 0·06(wt % SiO2) 2·25 þ d18OZrc, using for the d18ORock the d18OMelt. Thus a value of d 18OZrc ¼ 5·1ø is obtained for the preceding example. Rd18OResidue can be expressed in terms of the weighted sum of the residue composing minerals as X Id18 OMxI ðbÞ with I being P the corresponding proportion of the mineral MxI and I ¼ R. 1122
© Copyright 2024 Paperzz