18
1 Introduction THE GLOBAL INVENTORY of metals is critically dependent on the inordinately large contributions made by relatively limited numbers of exceptionally endowed deposits, and copper is no exception (Singer, 1995; Laznicka, 1999). Indeed, approxi- mately one-third of the world’s defined copper resources are contributed by just seven districts (Fig. 1), and approximately 2.5% of producing mines currently supply 25% of total copper Chapter 1 Copper Provinces RICHARD H. SILLITOE 27 West Hill Park, Highgate Village, London N6 6ND, England Abstract It has been recognized for the past century that copper deposits, in common with those of many other metals, are heterogeneously concentrated in Earth’s upper crust, resulting in areally restricted copper provinces that were generated during several discrete metallogenic epochs over time intervals of up to several hundred million years. Various segments of circum-Pacific magmatic arcs, for example, have total contained copper contents that differ by two orders of magnitude. Each metallogenic epoch introduced its own deposit type(s), of which porphyry copper (and related skarn), followed by sediment-hosted stratiform copper and then iron oxide copper-gold (IOCG), are globally preeminent. Nonetheless, genesis of the copper provinces remains somewhat enigmatic and a topic of ongoing debate. A variety of deposit-scale geometric and geologic features and factors strongly influence the size and/or grade of porphyry copper, sediment-hosted stratiform copper, and/or IOCG deposits. For example, develop- ment of major porphyry copper deposits/districts is favored by the presence of clustered alteration-mineraliza- tion centers, mafic or massive carbonate host rocks, voluminous magmatic-hydrothermal breccias, low sulfida- tion-state core zones conducive to copper deposition as bornite ± digenite, hypogene and supergene sulfide enrichment, and mineralized skarn formation, coupled with lack of serious dilution by late, low-grade porphyry intrusions and breccias. Furthermore, the copper endowment of all deposit types undoubtedly benefits from optimization of the ore-forming processes involved. Tectonic setting also plays a fundamental role in copper metallogeny. Contractional tectonomagmatic belts, created by flat-slab subduction or, less commonly, arc-continent collision and characterized by crustal thicken- ing and high rates of uplift and exhumation, appear to host most large, high-grade hypogene porphyry copper deposits. Such mature arc crust also undergoes mafic magma input during porphyry copper formation. The premier sediment-hosted stratiform copper provinces were formed in cratonic or hinterland extensional sedi- mentary basins that subsequently underwent tectonic inversion. The IOCG deposits were generated in associ- ation with extension/transtension and felsic intrusions, the latter apparently triggered by deep-seated mafic magmas in either intracratonic or subduction settings. The radically different exhumation rates characteristic of these various tectonic settings account well for the secular distribution of copper deposit types, in particu- lar the youthfulness of most porphyry relative to sediment-hosted stratiform and IOCG deposits. Notwith- standing the importance of these deposit-scale geologic, regional tectonic, and erosion-rate criteria for effec- tive copper deposit formation and preservation, they seem inadequate to explain the localization of premier copper provinces, such as the central Andes, southwestern North America, and Central African Copperbelt, in which different deposit types were generated during several discrete epochs. By the same token, the paucity of copper mineralization in some apparently similar geologic settings elsewhere also remains unexplained. It is proposed here that major copper provinces occur where restricted segments of the lithosphere were pre- disposed to upper-crustal copper concentration throughout long intervals of Earth history. This predisposition was most likely gained during oxidation and copper introduction by subduction-derived fluids, containing met- als and volatiles extracted from hydrated basalts and sediments in downgoing slabs. As a result, superjacent lithospheric mantle and lowermost crust were metasomatized as well as gaining cupriferous sulfide-bearing cumulates during magmatic differentiation—processes that rendered them fertile for tapping during subsequent subduction- or, uncommonly, intraplate extension-related magmatic events to generate porphyry copper and IOCG districts or belts. The fertile lithosphere beneath some accretionary orogens became incorporated during earlier collisional events, commonly during Precambrian times. Relatively oxidized crustal profiles—as opposed to those dominated by reduced, sedimentary material—are also required for effective formation of all major cop- per deposits. Large sedimentary basins underlain by or adjoining oxidized and potentially copper-anomalous crust and filled initially by immature redbed strata containing magmatic arc-derived detritus provide optimal sites for large-scale, sediment-hosted stratiform copper mineralization. Translithospheric fault zones, acting as giant plumbing systems, commonly played a key role in localizing all types of major copper deposits, districts, and belts. These proposals address the long-debated concept of metal inheritance in terms of the fundamental role played by subduction-metasomatized mantle lithosphere and lowermost crust in global copper metallogeny. E-mail: [email protected] © 2012 Society of Economic Geologists, Inc. Special Publication 16, pp. 1–18 Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdf by guest on 05 August 2020

Chapter 1 Copper Provinces

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Chapter 1 Copper Provinces

1

IntroductionTHE GLOBAL INVENTORY of metals is critically dependent onthe inordinately large contributions made by relatively limited

numbers of exceptionally endowed deposits, and copper is noexception (Singer, 1995; Laznicka, 1999). Indeed, approxi-mately one-third of the world’s defined copper resources arecontributed by just seven districts (Fig. 1), and approximately2.5% of producing mines currently supply 25% of total copper

Chapter 1

Copper Provinces

RICHARD H. SILLITOE†

27 West Hill Park, Highgate Village, London N6 6ND, England

AbstractIt has been recognized for the past century that copper deposits, in common with those of many other

metals, are heterogeneously concentrated in Earth’s upper crust, resulting in areally restricted copperprovinces that were generated during several discrete metallogenic epochs over time intervals of up to severalhundred million years. Various segments of circum-Pacific magmatic arcs, for example, have total containedcopper contents that differ by two orders of magnitude. Each metallogenic epoch introduced its own deposittype(s), of which porphyry copper (and related skarn), followed by sediment-hosted stratiform copper and theniron oxide copper-gold (IOCG), are globally preeminent. Nonetheless, genesis of the copper provinces remainssomewhat enigmatic and a topic of ongoing debate.

A variety of deposit-scale geometric and geologic features and factors strongly influence the size and/orgrade of porphyry copper, sediment-hosted stratiform copper, and/or IOCG deposits. For example, develop-ment of major porphyry copper deposits/districts is favored by the presence of clustered alteration-mineraliza-tion centers, mafic or massive carbonate host rocks, voluminous magmatic-hydrothermal breccias, low sulfida-tion-state core zones conducive to copper deposition as bornite ± digenite, hypogene and supergene sulfideenrichment, and mineralized skarn formation, coupled with lack of serious dilution by late, low-grade porphyryintrusions and breccias. Furthermore, the copper endowment of all deposit types undoubtedly benefits fromoptimization of the ore-forming processes involved.

Tectonic setting also plays a fundamental role in copper metallogeny. Contractional tectonomagmatic belts,created by flat-slab subduction or, less commonly, arc-continent collision and characterized by crustal thicken-ing and high rates of uplift and exhumation, appear to host most large, high-grade hypogene porphyry copperdeposits. Such mature arc crust also undergoes mafic magma input during porphyry copper formation. Thepremier sediment-hosted stratiform copper provinces were formed in cratonic or hinterland extensional sedi-mentary basins that subsequently underwent tectonic inversion. The IOCG deposits were generated in associ-ation with extension/transtension and felsic intrusions, the latter apparently triggered by deep-seated maficmagmas in either intracratonic or subduction settings. The radically different exhumation rates characteristicof these various tectonic settings account well for the secular distribution of copper deposit types, in particu-lar the youthfulness of most porphyry relative to sediment-hosted stratiform and IOCG deposits. Notwith-standing the importance of these deposit-scale geologic, regional tectonic, and erosion-rate criteria for effec-tive copper deposit formation and preservation, they seem inadequate to explain the localization of premiercopper provinces, such as the central Andes, southwestern North America, and Central African Copperbelt, inwhich different deposit types were generated during several discrete epochs. By the same token, the paucityof copper mineralization in some apparently similar geologic settings elsewhere also remains unexplained.

It is proposed here that major copper provinces occur where restricted segments of the lithosphere were pre-disposed to upper-crustal copper concentration throughout long intervals of Earth history. This predispositionwas most likely gained during oxidation and copper introduction by subduction-derived fluids, containing met-als and volatiles extracted from hydrated basalts and sediments in downgoing slabs. As a result, superjacentlithospheric mantle and lowermost crust were metasomatized as well as gaining cupriferous sulfide-bearing cumulates during magmatic differentiation—processes that rendered them fertile for tapping during subsequentsubduction- or, uncommonly, intraplate extension-related magmatic events to generate porphyry copper andIOCG districts or belts. The fertile lithosphere beneath some accretionary orogens became incorporated duringearlier collisional events, commonly during Precambrian times. Relatively oxidized crustal profiles—as opposedto those dominated by reduced, sedimentary material—are also required for effective formation of all major cop-per deposits. Large sedimentary basins underlain by or adjoining oxidized and potentially copper-anomalouscrust and filled initially by immature redbed strata containing magmatic arc-derived detritus provide optimalsites for large-scale, sediment-hosted stratiform copper mineralization. Translithospheric fault zones, acting asgiant plumbing systems, commonly played a key role in localizing all types of major copper deposits, districts,and belts. These proposals address the long-debated concept of metal inheritance in terms of the fundamentalrole played by subduction-metasomatized mantle lithosphere and lowermost crust in global copper metallogeny.

† E-mail: [email protected]

© 2012 Society of Economic Geologists, Inc.Special Publication 16, pp. 1–18

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 2: Chapter 1 Copper Provinces

output (M. Harris, Rio Tinto, unpub. comp., 2012). Further-more, large proportions of most major metals, particularlywell exemplified by copper, are concentrated in areally re-stricted provinces (Fig. 1), which were typically assembledduring several discrete metallogenic epochs. At least in thecase of intrusion-related deposits, individual epochs com-monly have durations of ≤10 m.y. (e.g., Sillitoe, 1988). Thisspatial and temporal confinement of copper and other metaldeposits was appreciated by Lindgren (1909) and subsequentpioneers, as reviewed by Turneaure (1955), but has becomemuch better defined over the ensuing century as a result ofnumerous discoveries and geologic advances, particularly direct isotopic dating of ore-related minerals.

Although the fundamental reasons for the development ofthe world’s largest copper deposits and premier copper beltsand provinces are not well understood, this introductory paperexplores some of the more plausible possibilities. The principalcontributors to the global copper inventory, namely porphyryand any associated skarn deposits (~70%), sediment-hostedstratiform deposits (~15%), and, a distant third, iron oxidecopper-gold (IOCG) deposits, are emphasized both herein(Fig. 1) and throughout the rest of this volume. Other relativelyminor copper sources, including magmatic nickel-copper, vol-canogenic massive sulfide (VMS), nonporphyry-related skarn,vein, Chilean manto-type, and carbonatite-hosted deposits

are not specifically discussed, although because of the impor-tance of Noril’sk, Russia (Fig. 1), the first of these sources hasa paper devoted to it (Burrows and Lesher, 2012).

The copper endowment considered herein (≈2,500 millionmetric tons [Mt]; Figs. 1, 2) exceeds the global inventory of1,900 Mt determined by Kesler and Wilkinson (2008), and ismore than four times larger than some other recent estimates(e.g., ~570 Mt; U.S. Geological Survey, 2011, p. 48–49). Al-though only formal resources plus past production are takeninto account, the greater copper tonnage may be attributed tomajor recent expansions of hypogene resources, particularlyin the central Andes (e.g., Sillitoe, 2010a), and application oflower cutoff grades. If probabilistic methodologies employedby Cunningham et al. (2007) and Kesler and Wilkinson (2008)are followed, then at least twice the number of copper de-posits exist (most at greater depths), albeit probably mainlyconfined to the currently defined belts and provinces.

Deposit-Scale Contributions to Large Size and High Grade

For the sake of brevity, only copper deposits and districtsthat attain supergiant status by containing ≥24 Mt (Singer,1995) or ≥25 Mt (Laznicka, 1999) of copper metal are dis-cussed and individually plotted in Figure 1; most exceed the31.1-Mt Cu threshold used to define behemothian deposits

2 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 2

FIG. 1. The world’s supergiant copper deposits and districts (defined as those containing ≥24 Mt [Singer, 1995] to ≥25Mt [Laznicka, 1999] Cu in resources and past production) and preeminent provinces, keyed to deposit types. The newly dis-covered Kamoa deposit in the Central African Copperbelt (Broughton and Rogers, 2010) contains 22 Mt Cu, but is also con-sidered as a supergiant because of the likelihood of further growth. Data compiled from numerous published and unpub-lished sources, including company press releases.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 3: Chapter 1 Copper Provinces

(Clark, 1993). These truly exceptional copper concentrations(Fig. 1) include 14 porphyry ± skarn districts (Pebble, Butte,Bingham, Resolution, Morenci, Chuquicamata, Collahuasi-Quebrada Blanca, Escondida, Los Pelambres-El Pachón, RíoBlanco-Los Bronces, El Teniente, Reko Diq, Oyu Tolgoi, andGrasberg), four and probably five sediment-hosted stratiformdistricts (Lubin-Polkowice, Konkola, Kolwezi, Udokan, andprobably Kamoa), one IOCG deposit (Olympic Dam), andone magmatic nickel-copper district (Noril’sk). Six of theseporphyry copper districts are the principal contributors to thecentral Andean copper province (Figs. 1, 2).

Porphyry copper deposits

Intrusion, host-rock, and alteration-mineralization featuresmay all be discerned as controls on the large size and/or highgrade of the world’s preeminent porphyry copper depositsand districts, as shown schematically in Figure 3.

The presence of several closely spaced, mineralized por-phyry stocks, which define clusters (e.g., Escondida and RekoDiq districts) or alignments (e.g., Collahuasi-Quebrada Blanca,Chuquicamata, Los Pelambres-El Pachón, Río Blanco-LosBronces, and Oyu Tolgoi districts), is an important localizer ofsupergiant copper concentrations. At least 13 discrete de-posits constitute the Reko Diq cluster (Perelló et al., 2008).

Nevertheless, as currently understood, Bingham and Resolu-tion comprise single, zoned deposits rather than clusters oralignments (e.g., Hehnke et al., 2012; Porter et al., 2012).Such clusters and alignments are inferred to occur abovecupolas on the roof zones of parental plutons located atdepths of several kilometers (e.g., Emmons, 1927; Dilles andProffett, 1995; Sillitoe, 2010b). The volume of porphyry in-trusions within the confines of a deposit appears unrelated toits copper content, as emphasized by comparison of thelargely porphyry-hosted Collahuasi deposits with the largelywall rock-hosted El Teniente deposit (e.g., Camus, 2003).Nonetheless, it is critically important that the earliest, gener-ally best-mineralized porphyry phase and its immediate wallrocks remain as a physically coherent entity little diluted bysubsequent pulses of intermineral and, especially, late min-eral porphyry, which typically contain progressively lowercopper contents as they become younger. Such low-gradeporphyry phases are volumetrically restricted in most of thesupergiant copper deposits (Fig. 3).

It is well known that magmatic-hydrothermal breccias (Silli-toe, 1985) in porphyry copper deposits can give rise to substan-tially higher-grade hypogene (and supergene) mineralizationbecause of the greater permeability and resultant fluid focus-ing that they provide (Fig. 3). The prime example of the key

COPPER PROVINCES 3

0361-0128/98/000/000-00 $6.00 3

FIG. 2. Total copper endowment (resources and past production) of different segments of Phanerozoic circum-Pacificmagmatic arcs. Only Paleozoic arc terranes, potentially somewhat more deeply eroded because of their greater antiquity, arepresent in eastern Australia. Note the two orders of magnitude difference among the segments. Data compiled from nu-merous published and unpublished sources, including company press releases.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 4: Chapter 1 Copper Provinces

4 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 4

FIG. 3. Anatomy of a non-eroded, telescoped porphyry copper system showing spatial interrelationships of a centrally located porphyry Cu ± Au ± Mo deposit in a porphyry stock and its immediate host rocks, including overlying high- and intermediate-sulfidation epithermal deposits in and alongside the lithocap environment. Not all depicted features are nec-essarily present in any single district. The legend explains the temporal sequence of rock types, with the porphyry stock pre-dating maar-diatreme emplacement, which, in turn, overlaps lithocap development and related phreatic brecciation. Shallowalteration types generally overprint deeper ones. Circled letters highlight deposit-scale features that can enhance hypogenegrade development: A. Minimal dilution caused by restricted volume of lower grade, intermineral and late-mineral por-phyries; B. Large volume of well-mineralized magmatic-hydrothermal breccia; C. Late-stage, barren diatreme located be-yond rather than within the ore zone; D. Mafic wall rocks induce effective copper precipitation; E. Massive, impermeablecarbonate wall rocks inhibit dispersion of mineralizing fluids and favor internal copper precipitation and grade development;F. Presence of bornite and digenite in the deep, central parts of the potassic zone increases hypogene copper tenor; G. Hypo -gene enrichment by high sulfidation-state sulfide minerals in the roots of the sericitic zone; and H. Skarn development withcopper tenors exceeding those in the adjacent porphyry stock. Modified from Sillitoe (2010b).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 5: Chapter 1 Copper Provinces

role played by magmatic-hydrothermal brecciation in the de-velopment of both deposit size and grade is at Río Blanco-LosBronces, the world’s largest copper district (Fig. 1), where onthe order of 50% of the ore may be breccia hosted (Serranoet al., 1996; Irarrazaval et al., 2010; Toro et al., 2012). How-ever, magmatic-hydrothermal breccias make only relativelyminor contributions to the other premier deposits, with Peb-ble, Bingham Canyon, and Chuquicamata being almost de-void of them (Ossandón et al., 2001; Lang and Gregory, 2012;Porter et al., 2012; Rivera et al., 2012). If present, late, low-grade breccias, particularly diatremes, are best to be situatedbeyond deposits so as not to destroy ore (Fig. 3), althoughsuch destruction can be tolerated in the largest deposits (e.g.,El Teniente; Howell and Molloy, 1960).

Wall-rock composition and permeability seem to be impor-tant controls on large size and hypogene grades exceeding 1%Cu at several of the premier deposits (Sillitoe, 2010b; Fig. 3).Intensely biotitized, ferrous iron-rich wall rocks—a Mesopro-terozoic diabase sill complex at Resolution (Hehnke et al.,2012; Leveille and Stegen, 2012), gabbro and dolerite sillsand dikes at El Teniente (Skewes et al., 2002), and submarinetholeiitic basalt at Oyu Tolgoi (Kirwin et al., 2005; Crane andKavalieris, 2012)—acted as highly effective copper precipi-tants from oxidized magmatic fluids. By the same token,quartzite at Bingham Canyon acted as a poor host (Porter etal., 2012). In contrast, the massively bedded limestone wallrocks at Grasberg may have created a relatively impermeablesleeve that inhibited magmatic fluid escape and promotedgrade development within the porphyry stock (Sillitoe, 1997).A comparable role has also be ascribed to hornfelsed wallrocks, including those above both the East zone at Pebble(Lang and Gregory, 2012) and the world’s largest (17 Mt Cu)skarn copper deposit at Antamina, Peru (Love et al., 2004).

The evolution of alteration, including the accompanyingsulfide minerals, over the lifespans of porphyry copper de-posits profoundly influences grade development and conser-vation (Einaudi et al., 2003). In deposits characterized bymajor copper introduction in the early, high-temperature,potassic stages (e.g., Bingham Canyon, Los Pelambres-ElPachón, El Teniente, Grasberg), the absence of appreciableoverprinting by sericitic or chlorite-sericite alteration, whichcan cause partial or even total copper removal (e.g., Kouz-manov and Pokrovski, 2012), favors conservation of high cop-per tenors; these attain their maxima in the deep, centralparts of systems where low-sulfidation conditions can lead tobornite ± digenite accompanying or even dominating chal-copyrite (Fig. 3; e.g., Bingham Canyon; Porter et al., 2012).Coalescence of several bornite-rich centers favors develop-ment of large orebodies, as at Los Pelambres and El Teniente(Vry et al., 2010; Perelló et al., 2012). Nonetheless, wherecopper deposition is relatively late, sericitic and even, in somecases, chlorite-sericite alteration appears to accompany gradedevelopment (e.g., Oyu Tolgoi; Crane and Kavalieris, 2012).The highest hypogene grades in some major porphyry de-posits are present with sericite ± pyrophyllite ± dickite alter-ation and high sulfidation-state copper minerals, in the tele-scoped roots of advanced argillic lithocaps (Sillitoe, 2010b;Fig. 3), with Butte (Meyer et al., 1968), Chuquicamata (Os-sandón et al., 2001; Rivera et al., 2012), and Resolution(Hehnke et al., 2012) and, to a lesser extent, Escondida

(Hervé et al., 2012) and Pebble East zone (Lang and Gregory,2012) being prominent examples.

Both calcic and magnesian skarns, developed proximallywith respect to porphyry copper stocks emplaced into shelf-carbonate sequences, as in the Grasberg and Bingham dis-tricts, can also give rise to high-grade, hypogene copper min-eralization (Meinert et al., 1997; Leys et al., 2012; Porter etal., 2012). Particularly large tonnages can be developed wherethe receptive carbonate horizons either flank the apices ofporphyry stocks (e.g., Antamina skarn; Love et al., 2004) orparallel the stock contacts (e.g., Ertsberg East skarn, Gras-berg district; Gandler and Kyle, 2008; Leys et al., 2012).

Supergene sulfide enrichment since ~40 Ma is well knownto have been a key process in grade development, particularlyin parts of the central Andes and southwestern North Ameri-can porphyry copper provinces (Sillitoe, 2005, and referencestherein), with Morenci (Leveille and Stegen, 2012), Chuquica-mata (Ossandón et al., 2001; Rivera et al., 2012), and Escon-dida (Hervé et al., 2012) being prime examples. However,particularly in the former region, exploration over the pastdecade or so has outlined far larger hypogene resources be-neath the previously and/or currently mined parts of theCenozoic supergene profiles (e.g., Escondida; Hervé et al.,2012).

Sediment-hosted stratiform copper deposits

The most obvious parameters controlling the size of sedi-ment-hosted stratiform copper deposits, as exemplified bymajor examples in the Central African Copperbelt of Zambiaand Democratic Republic of Congo (DRC), the Kupfer-schiefer province of Poland and Germany, Dzhezkazgan inKazakhstan, and Udokan in Russia (Fig. 1), are geometric anddepend on structurally uninterrupted stratigraphic continuity.The largest deposits have the greatest along-strike and down-dip extents and/or thickest or greatest number of ore hori-zons, as exemplified by the sandstone-, carbonaceous shale-,and dolomite-hosted Lubin, Polkowice, and contiguous de-posits of the Fore-Sudetic Monocline in Poland, which un-derlie an area of ~500 km2 and are mined to a depth of ~1,300m (Oszczepalski, 1999; Borg et al., 2012). The subhorizontal,diamictite-hosted Kamoa deposit in DRC underlies an area ofat least 80 km2 (Broughton and Rogers, 2010). The number ofindividual, ore-bearing sandstone beds approaches 30 over a600-m stratigraphic interval at Dzhezkazgan, thereby ac-counting for its large size (Gablina, 1981; Box et al., 2012). Incontrast, structural repetition of the main mineralizeddolomitic shale horizons, in conjunction with their consider-able strike and dip continuity, contributes to the large size ofthe Kolwezi and Tenke-Fungurumé district deposits, DRC(Hitzman et al., 2012; Schuh et al., 2012).

The controls on ore grade in sediment-hosted stratiformcopper deposits are less readily appreciated, although theyappear to be regional in extent since province-wide averagegrades tend to be broadly similar, in marked contrast to thesituation in porphyry copper and IOCG provinces. Conse-quently, availability of stratigraphic traps (e.g., pinchoutsagainst basement highs, anticlinal crests) and originalamounts and effectiveness of the copper-precipitating reduc-tant in the mineralized horizons, be it in situ (as carbonaceousmatter and/or diagenetic pyrite) or mobile and introduced

COPPER PROVINCES 5

0361-0128/98/000/000-00 $6.00 5

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 6: Chapter 1 Copper Provinces

(e.g., Selley et al., 2005; Hitzman et al., 2012), are seeminglyimportant factors. Nonetheless, the hypogene bornite- andchalcocite-bearing parts of the zoned deposits, especiallywhere veining is well developed, consistently have the highestgrade (e.g., Sillitoe et al., 2010; Schuh et al., 2012). Supergenechalcocite enrichment, suggested by some (e.g., Hitzman etal., 2005, 2012) to have enhanced copper grades in the shallowparts of the Central African Copperbelt deposits, is consideredunimportant. Suppression of the enrichment process is causedby the abundance of carbonate-bearing host rocks and paucityof ore-related pyrite, and is reflected by the ubiquity of high-grade oxide copper mineralization throughout the supergeneprofiles (e.g., Tenke-Fungurumé; Fay and Barton, 2012).

IOCG deposits

Notwithstanding the disparity of deposits assigned to theIOCG category and disaccord over their genesis (e.g., Sillitoe,2003; Barton, 2009; Groves et al., 2010), the paucity of su-pergiant deposits limits assessment of deposit-scale controlson their size and grade. However, the overwhelming domi-nance of Olympic Dam within the IOCG grouping, both interms of size and grade (90 Mt Cu; Fig. 1), must surely be re-lated to the unusually large volume (~5 km3) of the hosthematitic breccias (Ehrig et al., 2012), the products of high-energy hydrothermal fragmentation of granite (Reeve et al.,1990). Identical breccias occur at Carrapateena (Fairclough,2005), 120 km southeast of Olympic Dam, but their far morerestricted volume results in a commensurately smaller de-posit (currently 4.4 Mt Cu).

The world’s second-largest IOCG deposit, Candelaria-Punta del Cobre (~11 Mt Cu) in northern Chile, contains rel-atively minor hydrothermal breccia, thereby showing thatlarge size is by no means exclusively breccia controlled. In-deed, a combination of permeable volcaniclastic rocks, an an-ticlinal fold, faults, and an overlying limestone seal combinedto localize Candelaria-Punta del Cobre (Marschik and Font-boté, 2001). The third largest IOCG deposit, Salobo (offi-cially 7.8 Mt Cu) in the Carajás province (Fig. 1), is differentagain, being an elongate, steeply dipping, structurally con-trolled orebody associated with sheared and highly altered,granulite-facies, siliciclastic metasedimentary rocks (Réquiaand Fontboté, 2000; Xavier et al., 2012; M.W. Hitzman, writ.commun., 2012).

Fluorine-rich ore fluids were proposed as an influentialcontrol on size and grade at Olympic Dam (McPhie et al.,2011), but comparable fluorine enrichment at Carrapateena,Salobo, and elsewhere clearly did not generate the same sizeeffect. The enhanced permeability provided by the brecciahost seems to be a more likely explanation for the high cop-per grades, similar to the porphyry copper environment (seeabove). In common with sediment-hosted stratiform copperdeposits, average grades are also clearly influenced by thecopper contents of the hypogene sulfide species present, asunderscored by the high grades at Olympic Dam where bor-nite and chalcocite are widespread (Reeve et al., 1990; Ehriget al., 2012). However, in common with the sediment-hostedstratiform copper deposits, the overall deficiency of pyrite inIOCG orebodies militates against significant development ofsupergene chalcocite enrichment (e.g., central Andes IOCGdeposits; Sillitoe, 2005).

Summary statement—deposit-scale contributions

Optimization of all aspects of deposit-scale mineralizationprocesses is a likely prerequisite for formation of large, high-grade orebodies (e.g., Richards, 2005, 2011); however, it isapparent in the case of porphyry copper deposits, and proba-bly other ore types too, that no single geologic characteristicor set of characteristics seems able to adequately predict de-posit size or grade (cf. Clark, 1993). Nonetheless, a number ofgeometric, host-rock, brecciation, and alteration-mineraliza-tion features may offer at least partial explanations for thepreeminence of individual porphyry, sediment-hosted strati-form, and IOCG deposits and districts.

Notwithstanding these proposals, it is difficult to see howany of these disparate, deposit-scale parameters or mecha-nisms can satisfactorily explain the origin of exceptionally en-dowed copper belts and provinces. In the case of the centralAndean porphyry copper province, for example, the premierdeposits are assigned quite different key geologic controls ongrade and/or tonnage: for example, biotitized mafic wall rocksat El Teniente, exceptional development of magmatic-hydro-thermal breccias at Río Blanco-Los Bronces, and a combina-tion of hypogene and supergene copper enrichment atChuquicamata. Furthermore, in the case of IOCG deposits,even orebodies within individual provinces display radicallydifferent geologic characteristics (e.g., Carajás; Xavier et al.,2012). Clearly, therefore, these deposit-scale parameters aresecondary to more regionally relevant processes that are re-quired to explain the localization of the world’s premier cop-per provinces.

Tectonic Controls on Copper Belts and ProvincesPorphyry copper deposits

It has long been appreciated that porphyry copper depositsoccur in accretionary orogens formed at sites of subduction ofoceanic lithosphere (Sawkins, 1972; Sillitoe, 1972). However,copper endowment in magmatic arcs around the PacificOcean (Fig. 2), as well as elsewhere, is highly heterogeneous,ranging from nearly 1,000 Mt in the central Andes to <10 Mtknown in several major arc segments, such as the northernand southern Andes, central and southern Mexico and Cen-tral America (excluding Panama), Russian Far East, Japan,and New Zealand. In an attempt to address this distributionalinequality, Sillitoe (1998) proposed that the world’s largestand highest grade hypogene porphyry copper deposits weregenerated in contractional arc settings: a proposition sup-ported by others (Cooke et al., 2005; Perelló, 2006; Tosdal etal., 2009; Loucks, 2012; Leveille and Stegen, 2012; Mpodozisand Cornejo, 2012).

Contractional conditions in accretionary orogens have beenascribed to three main tectonic mechanisms (Fig. 4): (1) rapidadvance of overriding plates coupled with high mantle vis-cosities, causing enhanced interplate coupling and flat-slabsubduction (e.g., van Hunen et al., 2004; Lallemand et al.,2005); (2) subduction of young, buoyant oceanic lithospheresurmounted by plume-generated topographic features, suchas aseismic ridges, seamount chains, and small- to moderate-sized oceanic plateaus, also causing flat-slab subduction (e.g.,Gutscher, 2002; Martinod et al., 2010); and (3) accretion ofexotic terranes, such as island arcs, microcontinental blocks,

6 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 6

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 7: Chapter 1 Copper Provinces

and large oceanic plateaus to convergent margins (e.g., Clooset al., 2005; Shafiei et al., 2009). The first and second mecha-nisms may operate in concert (van Hunen et al., 2004) or, al-ternatively, slab flattening may be a time-dependent processat convergent margins, without the need for subduction oftopographic prominences (Haschke et al., 2002; Lee andKing, 2011). Far more highly endowed porphyry copperprovinces worldwide coincided with contraction triggered byslab flattening than with terrane accretion.

In the case of the central Andes—characterized by con-tractional tectonism for much of the Cenozoic during forma-tion of the world’s premier copper province—the contractionis generally attributed to opening of the South Atlantic Oceanand the consequent westward advance of the South Americanplate relative to the Chile-Peru trench (e.g., Oncken et al.,2006). Sobolev and Babeyko (2005) estimated that 58% of thewestward drift since 35 Ma was accommodated by trenchrollback, 37% by tectonic shortening of the Andean margin,and 5% by subduction erosion of the Andean forearc by thedowngoing Nazca plate (e.g., Rutland, 1971). Furthermore, the

position of the central Andean copper province approximatelycoincides with the middle of the longitudinally extensive An-dean orogen where shortening, including oroclinal bending,of the overriding plate was maximized (Schellart, 2008),rather than within the less-contractional northern and south-ern Andes (Fig. 2). However, in the case of the late Mioceneto Pliocene copper belt of central Chile, subduction of a spe-cific topographic feature, the Juan Fernández aseismic ridge,is widely invoked as at least a contributory cause of the slabflattening, regional shortening, surface uplift, and exhuma-tion of the overriding plate (e.g., Skewes and Stern, 1995).

The contraction in southwestern North America that char-acterized the Laramide (~80–45 Ma) porphyry copperprovince is generally inferred to have been caused by sub-duction of an oceanic plateau embedded in the downgoingFarallon plate (Livacarri et al., 1981; Liu et al., 2010). In con-trast, the Pliocene contractional tectonism in New Guinea,with which the Grasberg porphyry-skarn district and severalsmaller deposits are related, was due to collision of theMelanesian intraoceanic island arc with the leading edge ofthe Australian craton (Cloos et al., 2005; Leys et al., 2012).

The empirical correlation between major porphyry copperemplacement events and contractional tectonism in accre-tionary orogens seems to explain the existence of many of theworld’s premier porphyry copper belts, but lacks a well-founded cause. However, it may be speculated that the con-traction and concomitant shortening, thickening, and uplift ofarc crust inhibit volcanism, favoring mid- to upper-crustalmagma accumulation (Takada, 1994) and the consequentgeneration and focused expulsion of the voluminous copper-bearing magmatic fluids required for porphyry copper gene-sis (Sillitoe, 1998). Certainly, there is clear evidence for surfaceuplift and greatly reduced volcanic activity (and eruption- related magmatic fluid dissipation) in such contractional set-tings (e.g., mid-Eocene–early Oligocene belt of the centralAndes; Mpodozis and Ramos, 1990) as well as localized evi-dence for assembly of large epizonal plutons (e.g., RíoBlanco-Los Bronces district; Toro et al., 2012). Orogen-scalecontraction may also favor development in the uppermostlithospheric mantle or lowermost crust of long-lived and re-peatedly replenished magma chambers, which undergo evac-uation once contraction is relaxed (Loucks, 2012). Certainly,there is emerging evidence for porphyry copper emplace-ment in the central Andes immediately following active re-verse faulting (e.g., Escondida and Los Pelambres; Hervé etal., 2012; Perelló et al., 2012).

In marked contrast, predominantly extensional accretionaryorogens constructed during trench retreat (rollback), such asPaleozoic southeastern Australia (Terra Australis; Collins, 2002),the Paleozoic Altaids of central Asia (Yakubchuk et al., 2005,2012), Mesozoic eastern Asia, including the Yanshanian belt ofsoutheastern China, Korea, Japan, and the Russian Far East(e.g., Zhou et al., 2006), and the Cenozoic Cascades of north-western United States (du Bray and John, 2011), are either onlymodestly endowed with or entirely lack porphyry copper de-posits (cf. Uyeda and Nishiwaki, 1980; Fig. 2); however, the po-tentially greater erosion suffered by the pre-Cenozoic orogensmay be a contributory factor (see below). In contrast to con-tractional arcs, extensional island arcs, particularly intraoceanicarcs during their initial stages of construction, are characterized

COPPER PROVINCES 7

0361-0128/98/000/000-00 $6.00 7

Subductedoceanicplateau

Minimalvolcanism

Minimalvolcanism

Collisionalsuture Subduction

reversal

Detachedslab

Forelandfold-thrust belt

+

+

+

a

b

c

FIG. 4. Models for generation of contractional tectonism in accretionaryorogens: a. Flat-slab subduction developed episodically at advancing subduc-tion margins. b. Flat-slab subduction caused by oceanic plateaus and otherbuoyant objects on the downgoing slab (e.g., central Chile during the lateMiocene-Pliocene); and c. Accretion of exotic terranes (microcontinents, is-land arcs, large oceanic plateaus), which may cause reversal of subductionpolarity (e.g., New Guinea in the late Miocene-Pliocene).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 8: Chapter 1 Copper Provinces

by large-volume volcanism (e.g., Stern and Bloomer, 1992).Such immature arcs lack the large, felsic magma chambers nec-essary for porphyry copper deposit formation, as exemplified bythe apparently barren Izu-Bonin-Mariana, Tonga-Kermadec,and Kurile arcs in the western Pacific Ocean (Fig. 2).

Sediment-hosted stratiform copper deposits

Sediment-hosted stratiform copper deposits formed in cra-tonic, foreland, and hinterland basins at low paleolatitudes(Kirkham, 1989; Hitzman et al., 2010). The principal copperprovinces are hosted by basins in which the receptive strata areboth areally extensive and hydrologically accessible to large vol-umes of mineralizing brines. The basins have different origins:the Central African Copperbelt occupies part of an intracon-tinental rift system, which, farther west in the Damara belt ofNamibia, underwent limited ocean-floor spreading to generatea narrow, Red Sea-type ocean (Hanson, 2003); the Kupfer-schiefer deposits lie within the intracontinental Polish basin,which was initiated by rifting induced by postcollisional oro-genic collapse and transcurrent faulting (van Wees et al., 2000);and the Dzhezkazgan district is hosted by the Chu-Sarysutranspressional hinterland basin that developed at the site of aformer retroarc basin as a consequence of collisional orogeny(Yakubchuk et al., 2012, and references therein). Nonetheless,all are characterized by basal immature redbed siliciclastic(and, in many cases, volcanic) sequences, above or withinwhich thin, reduced horizons induced copper precipitationfrom ascendant, moderate-temperature, oxidized brines (e.g.,Hitzman et al., 2005, 2010; Fig. 5). Deeply penetrating faultsacted as principal conduits for fluid ascent, as documented inthe Kupferschiefer province (Borg et al., 2012, and referencestherein). The Kupferschiefer copper deposits are overlain bymassive, halite-bearing evaporites, which may have contributedto mineralization efficiency by acting as hanging-wall aquitardsthat minimized upward fluid dispersion (Jowett, 1986; SchmidtMumm and Wolfgramm, 2004; Fig. 5). Breccias, includingkilometer-scale blocks, in parts of the Central African Cop-perbelt are interpreted to imply the former presence there ofcomparable, but now-dissolved evaporitic strata, which mayalso have acted as a top seal during mineralization (Jackson et

al., 2003; Selley et al., 2005; Koziy et al., 2009; Hitzman et al.,2012). Evaporites may also have served as sources for thebrines responsible for scavenging, transport, and precipitationof the copper and associated metals (e.g., Koziy et al., 2009).

The exceptional copper contents of the premier sediment-hosted stratiform belts are attributed by some investigators tosemicontinuity or multiplicity of copper introduction events,spanning the commonly protracted (100–300 m.y.) time in-tervals between sedimentation and early diagenesis throughto basin inversion and, in the Central African Copperbelt, theassociated low-grade metamorphism (Michalik and Sawlow-icz, 2001; Schmidt Mumm and Wolfgramm, 2004; Hitzman etal., 2005, 2010, 2012; Haest and Muchez, 2011; Schuh et al.,2012). A variant on this model implicates shorter, but highlyeffective copper-bearing brine expulsion events driven by ei-ther igneous activity related to extension or basin inversion, assupported by several lines of evidence for postlithifactionmineralization (e.g., McGowan et al., 2003; Sillitoe et al.,2010; Symons et al., 2011; Borg et al., 2012; Hitzman et al.,2012). Both models satisfy the observed presence of mineral-ization over appreciable stratigraphic intervals in the CentralAfrican Copperbelt (Hitzman et al., 2012) and elsewhere.

IOCG deposits

Formation of the largest IOCG deposits in the Gawler, cen-tral Andes, and Carajás provinces is synchronous with felsicand subordinate mafic magmatism in extensional or transten-sional settings. In all three provinces, minor intrusions ofmafic composition appear to be the most closely related toIOCG generation (Sillitoe, 2003; Skirrow, 2009; Groves et al.,2010). Regional-scale fault zones or lineaments are obviousdeposit controls (e.g., O’Driscoll, 1985; Sillitoe, 2003; Xavieret al., 2012). The Gawler and Carajás provinces were mostprobably generated in Mesoproterozoic and Neoarcheananorogenic, intracratonic settings, respectively, although Skir-row (2009) discussed evidence for a distal retroarc positionfor the former. In contrast, the Coastal Cordillera of the cen-tral Andes was a retreating accretionary orogen characterizedby trench rollback during Late Jurassic to mid-CretaceousIOCG generation (Sillitoe, 2003). Mantle plume activity may

8 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 8

FIG. 5. Schematic depiction of key lithologic elements of sediment-hosted stratiform copper deposits. Rift-bounding nor-mal faults undergo inversion during contractional tectonism. Residual brines or brines produced by evaporite dissolutionleach copper and associated metals (cobalt or silver) from oxidized basement strata and overlying redbed siliciclastic rocks.Copper-bearing sulfides are precipitated by in situ or mobile reductants present in overlying siliciclastic rocks and carbon-ates as well as locally in the basement rocks. Inspired by Hitzman et al. (2010).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 9: Chapter 1 Copper Provinces

have triggered the Gawler and Carajás magmatism and IOCGmineralization (Groves et al., 2010), but was certainly not afeature of the Mesozoic central Andes.

Notwithstanding the synchronism of the largest IOCG de-posits with magmatism in broadly extensional/transtensionalsettings, the tectonic position of Olympic Dam and Salobo onthe one hand and the much younger Candelaria-Punta delCobre district on the other are markedly different. Hence, itmust be concluded that tectonic setting alone cannot satisfac-torily account for the localization of IOCG deposits.

Summary statement—tectonic controls

Contractional tectonic conditions and the resultant crustalthickening, surface uplift, and exhumation offer a viable ex-planation for the world’s most productive porphyry copperbelts. Basin size, stratigraphic architecture, and brine expul-sion history (during extension and/or subsequent contraction)may, either singly or, more likely, in combination, account forthe world’s principal sediment-hosted stratiform copperprovinces. Major IOCG deposits seem to be products ofcrustal extension/transtension, albeit with indications of aclose link to mafic magmatism.

However, these tectonic models, including their attendantmagmatic or sedimentologic histories, do not seem adequate

to fully explain the premier copper provinces, in which sev-eral metallogenic epochs and several different deposit typesgenerated under different tectonic conditions are commonlyrepresented. Furthermore, in some accretionary orogens, con-tractional tectonic events failed to give rise to known majorporphyry copper deposits, as exemplified by oceanic plateausubduction beneath the northern Andes of Ecuador duringthe Late Cretaceous (Vallejo et al., 2009) and flat-slab sub-duction beneath southern Alaska during the mid- to late Ceno-zoic (Finzel et al., 2011). Similarly, many major redbed- andevaporite-bearing cratonic, foreland, and hinterland basinscontain only minor or even lack sediment-hosted stratiformcopper deposits (e.g., Khorat basin, Thailand, and neighboringcountries; Racey, 2009). And, of course, numerous anoro-genic, intracratonic magmatic provinces and extensional ac-cretionary orogens worldwide lack even minor examples ofIOCG mineralization.

Recurrent Mineralization in Major Copper ProvincesThe world’s major copper provinces tend to be dominated

by a single deposit type, albeit in some cases generated duringtwo or more metallogenic epochs (Figs. 1, 6), but commonlyalso contain subsidiary copper deposits of other types. For example, the pre-eminent central Andes and southwestern

COPPER PROVINCES 9

0361-0128/98/000/000-00 $6.00 9

200 km

200 km

200 km

PERU

MEXICO

USA

BOLIVIA

AR

GE

NT

INA

PacificOcean

PacificOcean

Middle Eocene-early Oligocene

Miocene-earlyPliocene

Miocene

LateCretaceous

Early Cretaceous

Permian

Paleoproterozoic

Late Cretaceous-early Eocene(Laramide)

Late Jurassic

Tengizbasin

Chu-Sarysubasin

LateCarboniferous Devonian-

Carboniferous

Jurassic

20°S

30°S

70°W

50°N

80°E65°E

115°W

35°W

110°W

Paleocene-early Eocene

CH

ILE

a

c

b

PorphyryCopper deposit types

IOCG, manto type + porphyry

Sediment-hostedstratiformVolcanogenic massive sulfide

FIG. 6. Recurrent metallogenic epochs in representative major copper provinces: a. Central Andes (Sillitoe and Perelló,2005); b. Southwestern North America (Titley, 1982); and c. Kazakhstan Altaids (Yakubchuk et al., 2005). Note proximity ofthe different deposit types.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 10: Chapter 1 Copper Provinces

North America copper provinces are characterized by severalintrusion- and volcanic-related metallogenic epochs. Thehighly endowed Miocene to early Pliocene, middle Eocene toearly Oligocene, and Paleocene to early Eocene belts of thecentral Andes—all apparently generated during tectonic con-traction—were preceded by subordinate Late Carboniferousto Triassic porphyry copper, Late Jurassic and Early Creta-ceous IOCG, manto-type copper, and porphyry copper, andLate Cretaceous and Miocene sediment-hosted stratiformcopper epochs that all seem to have formed inextensional/transtensional settings (Sillitoe, 1988; Sillitoe andPerelló, 2005; Fig. 6a). Indeed, the entire post-Paleozoic cen-tral Andean orogen is copper dominated, a fact that led Stoll(1965) to define it as a chalcophile province. Similarly, south-western North America is characterized not only by the syn-contractional Laramide porphyry copper deposits but, as pre-saged by Spurr (1923, Ch. 9, 10), also by Late Jurassic porphyrycopper emplacement at Bisbee (Titley, 1982), Paleoproterozoiccopper-rich volcanogenic massive sulfide formation at Jeromeand elsewhere (Lindberg, 1989), and minor younger coppermineralization (e.g., McLemore, 1996), all plausibly extension-related mineralization events (Fig. 6b). The Altaid provinceof central Asia is made up of several Paleozoic intrusion- andvolcanic-related copper epochs, to which may be added themajor Dzhezkazgan sediment-hosted stratiform copper dis-trict (e.g., Yakubchuk et al., 2005, 2012; Fig. 6b). The Carajásprovince (Fig. 1), part of the Southern Amazon craton ofnortheastern Brazil, is primarily defined by a series of proba-bly Neoarchean IOCG deposits, but also includes greisen-and vein-type copper deposits related to Paleoproterozoic fel-sic intrusions (e.g., Breves; Grainger et al., 2008). Even theCentral African Copperbelt includes structurally localized,carbonate-replacement zinc-copper deposits (e.g., Kipushi,DRC; Haest and Muchez, 2011), and Mesoproterozoic IOCGdeposits share the Gawler province with sediment-hostedstratiform copper deposits of Neoproterozoic age (e.g., MountGunson; Tonkin and Creelman, 1990).

Therefore, it is tempting to think that a yet more funda-mental factor may need to be identified in order to explainthe origin of the world’s premier copper provinces, in whichsuperposition of several copper-forming epochs and domi-nance of copper over other metals (except, perhaps, for iron)seem to be commonplace. Furthermore, given their differentformational ages and preservation potentials, the known de-posits are unlikely to be fully representative of a province’stotal original copper endowment.

Copper Deposit PreservationAttempts to understand the global distribution of copper

deposits are dictated by the exposed and near-surface (≤1.5km) deposits discovered to date. Eroded deposits are neces-sarily ignored, although it is recognized that they are likely tohave been sufficiently numerous, large, and/or high-grade tohave greatly increased the importance of known copperprovinces or even constituted former, now-unsuspectedprovinces. Undiscovered copper deposits pose the same fun-damental problems.

The preservation potential of different types of copper de-posits is clearly not the same. Porphyry copper deposits, par-ticularly their near-surface epithermal manifestations, are

readily lost to erosion because of confinement of most ofthem to shallow (<4 km) crustal levels (e.g., Sawkins, 1972;Sillitoe, 1972; Kesler and Wilkinson, 2006; Fig. 3). This maybe particularly true for the largest deposits, which are ar-guably generated only during rapid uplift and exhumationconsequent upon tectonic contraction (Sillitoe, 1998; seeabove), as well as for gold-rich examples, which tend to format shallower depths than those containing by-product molyb-denum (Murakami et al., 2010). Hence, many major porphyrycopper deposits are Cenozoic, in particular Mio-Pliocene inage (Los Pelambres-El Pachón, Río Blanco-Los Bronces, ElTeniente, Reko Diq, Grasberg), with progressively fewer beinghosted by Mesozoic, Paleozoic, and older accretionary oro-gens (cf. Kesler and Wilkinson, 2006). The oldest exploitedporphyry copper deposit is the highly deformed, Paleo-proterozoic example at Aitik, Sweden, in the Svecofennianorogen (Wanhainen et al., 2006).

Hawkesworth et al. (2009, 2010) proposed that the preser-vation potential of rocks (and any enclosed ore deposits) insubduction or extensional settings is substantially less thanthat of collisional orogens (Fig. 7). Potentially for this reason,the most important Paleozoic porphyry copper province ishosted by the extensional Altaid accretionary orogen of cen-tral Asia (Figs. 1, 6c), which eventually became caught up inthe late Paleozoic to earliest Mesozoic collision of the NorthChina and Tarim cratonic blocks with the southern margin ofthe Siberian craton (Yakubchuk et al., 2005, 2012; Xiao et al.,

10 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 10

Tota

lco

pp

eren

dow

men

t

Retreatingsubductionzones

Advancingsubductionzones

Supercontinent cycle

SUBDUCTION COLLISION BREAKUP

Pre

serv

atio

np

oten

tial

FIG. 7. Hypothetical preservation potential of rocks and contained copperdeposits (blue lines) versus approximate copper endowment (red line) of sub-duction, collisional, and extensional breakup settings. The preservation po-tential of advancing (Andean type, contractional) subduction margins ismuch lower than that of retreating (southwestern Pacific type, extensional)margins characterized by trench rollback. The copper endowment of sub-duction settings (mainly porphyry deposits) is much greater than that of col-lisional (sediment-hosted stratiform deposits and a subset of porphyry deposits,including the Grasberg district) and extensional (some sediment-hostedstratiform and IOCG deposits) settings. The figure implies that porphyry andsediment-hosted stratiform copper deposits in collisional settings and IOCGdeposits in extensional settings tend to be older than porphyry copper andIOCG deposits in subduction settings. The shape of the right side of the cop-per endowment curve would change if sediment-hosted stratiform depositsare considered to form mainly during extension rather than basin inversion.Based on Hawkesworth et al. (2009).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 11: Chapter 1 Copper Provinces

2009). In the Kazakh-Mongol magmatic arc in the Altaids, ad-vanced argillic lithocaps—characteristic of the shallow epi-thermal parts of porphyry copper systems (e.g., Sillitoe,2010b; Fig. 3)—are widely preserved, thereby further em-phasizing the generally shallow erosion level. Typically,preservation is a product of postmineral concealment beneathyounger volcanosedimentary cover (e.g., Oyu Tolgoi; Perellóet al., 2001; Wainwright et al., 2011; Crane and Kavalieris,2012), as it commonly is elsewhere (e.g., Resolution and Peb-ble East; Hehnke et al., 2012; Lang and Gregory, 2012; Lev-eille and Stegen, 2012).

Furthermore, accretionary orogens of classical Andeantype are less well preserved than extensional examples char-acterized by retreating subduction zones and retroarc basindevelopment (Hawkesworth et al., 2009, 2010; Fig. 7). Thus,in the central Andes, the IOCG deposits, formed in a retreat-ing subduction regime, are ~70 m.y. older than the parallelbelt of major porphyry copper deposits, generated and ex-humed during contraction at an advancing subduction margin(Sillitoe, 2003; Sillitoe and Perelló, 2005; Fig. 6a).

In contrast to porphyry copper deposits, most sediment-hosted stratiform and IOCG deposits have far greater preser-vation potential and, as a consequence, tend to be much olderbecause they occur in major sedimentary basins and exten-sional, commonly intraplate environments, respectively. Thepreservation potential of the principal sediment-hosted strat-iform copper provinces was further enhanced by incorpora-tion in collisional orogens. Hence, ancient deposits are farmore common, as exemplified by the Neoarchean age of theCarajás IOCG province and Paleoproterozoic age of the supergiant (26 Mt Cu) sediment-hosted stratiform copper de-posit at Udokan, Russia (Chechetkin et al., 2000; Fig. 1). Fur-thermore, the relatively buoyant nature of the underlyingArchean lithospheric mantle was also a cogent factor incrustal stability and preservation of the Gawler and CarajásIOCG provinces (Groves et al., 2010).

Fundamental Controls of Major Copper ProvincesThe limited number, spatial restriction, and multistage ori-

gin of the world’s premier copper provinces hint at more fun-damental reasons than simply tectonic/tectonomagmatic set-ting to explain their localization. In the case of porphyrycopper deposits, for example, the reasonably well-understoodmagmatic and hydrothermal processes involved—as summa-rized by Sillitoe (2010b), Richards (2011), Audétat and Simon(2012), and Kouzmanov and Pokrovski (2012)—fail to satis-factorily explain the two orders of magnitude difference inthe copper endowment of the various circum-Pacific (Fig. 2)and other youthful arc segments worldwide, since neitherdramatically different degrees of deposit exposure (except,perhaps, for Paleozoic eastern Australia) nor exploration ef-fectiveness (except, perhaps, for the northern Andes andRussian Far East) can be realistically called upon. Thereforeit is difficult to escape the conclusion that limited volumes ofEarth’s upper crust are predisposed to recurrent copper con-centration, on time scales of tens to hundreds of millions ofyears and, potentially, as long as 2,000 m.y. It is proposed herethat lithosphere chemistry, perhaps most importantly copperavailability and redox state, is the fundamental control favor-ing repetitive copper mineralization: a proposal that builds on

long-debated concepts of metal inheritance and provinciality(e.g., Noble, 1970; Krauskopf, 1971; Routhier, 1976; Titley,2001).

Porphyry copper deposits

As reviewed by Richards (2003, 2011) and Audétat andSimon (2012), porphyry copper magma generation may beinitiated by hydration and partial melting of peridotite in as-thenospheric mantle wedges above subduction zones: aprocess induced by upward transfer of aqueous fluid and/orhydrous melt from downgoing oceanic slabs (e.g., Schmidtand Poli, 2005; Dreyer et al., 2010; Fig. 8). Subducted ser-pentinite may contribute a significant proportion of the re-quired water (Hattori and Guillot, 2003).

The slab-derived fluids introduce sulfur, chlorine, carbonspecies, and incompatible elements and, at least during thePhanerozoic, progressively oxidize the mantle wedges (e.g.,Kelley and Cottrell, 2009; Evans and Tomkins, 2011). Ascentof the hydrous basaltic melts from mantle wedges causes progressive metasomatism and oxidation of the overlyinglithospheric mantle (e.g., Peslier et al., 2002) as well as un-derplating and radically modifying the lower crust (Fig. 8).Furthermore, magmatic differentiation generates cumulatespotentially enriched in copper-bearing sulfides in the vicinityof the lithospheric mantle-crust boundary (e.g., Richards,2009, 2011; Lee et al., 2012). Hence, long-standing accre-tionary orogens—such as the central Andes where subductionwas initiated >500 m.y. ago (e.g., Bahlburg et al., 2009;Mpodozis and Cornejo, 2012)—and those subjected to multi-stage subduction histories generally should be characterizedby relatively oxidized crust-mantle systems as well as by hav-ing been subjected to the greatest incompatible element andvolatile input and concentration. Effective magmatic trans-port of copper and gold through the mantle and crust is fa-vored by relatively oxidized conditions, otherwise such metalsbecome immobilized as a result of retention or sequestrationby magmatic sulfide minerals and/or melts (e.g., Mungall,2002; Richards, 2011; Evans and Tomkins, 2011; Audétat andSimon, 2012). Hence, relatively oxidized rather than reducedcrustal profiles (and contained aqueous fluids) characterizemajor porphyry copper provinces, as emphasized by Keithand Swan (1995) for southwestern North America. The abil-ity of the lower and middle crust to influence magma chem-istry is underscored by the ubiquitous presence of inheritedmagmatic zircon grains in porphyry copper stocks (e.g.,Richards et al., 1999; van Dongen et al., 2010). Translithos-pheric fault systems, some localized by ancient collisional su-tures bounding accreted terranes, have the potential to focusthe ascent of the mantle- and lower crust-derived magmasand contained metals into the upper crust (e.g., Bingham,Chuquicamata, Oyu Tolgoi, Grasberg).

Although still unquantified, much of the copper and associ-ated metals incorporated in the mantle wedge-derived basalticmagmas were supplied by the fluids ascending from downgo-ing slabs and, hence, ultimately by hydrated oceanic lithos-phere and overlying sediments (e.g., Sillitoe, 1972; Noll et al.,1996; Hattori, 2007; Fig. 8). Irrespective of precisely wherethe bulk of the copper is concentrated in the oceanic crust(Cathles, 2011; Hannington, 2011), its distribution in sheeteddike complexes and overlying basaltic lavas (few hundred

COPPER PROVINCES 11

0361-0128/98/000/000-00 $6.00 11

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 12: Chapter 1 Copper Provinces

ppm), contained VMS deposits (up to several %), basalpelagic sediments (~800 ppm), and ferromanganese nodulesand crusts (~4,500 ppm) is undoubtedly highly heteroge-neous at the scale of ocean basins (e.g., Morgan, 2000; Li andSchoonmaker, 2005). Furthermore, fluid and contained metalrelease from downgoing slabs is highly complex because of itsdependence on the interplay between thermomechanicalproperties and reaction paths (Schmidt and Poli, 2005) andthe strong likelihood that a substantial proportion is sub-ducted deeper into the mantle. Therefore the metal flux fromsubducted slabs to mantle wedges and their superjacentlithospheric mantle and lowermost crust may vary consider-ably both geographically and through time. Ascendantbasaltic magmas may obtain additional metals from metasom-atized lithospheric mantle and mafic lower crust (Shafiei etal., 2009), which also seem likely to be characterized by highlyheterogeneous redox and metal distribution patterns inher-ited not only from earlier subduction, but also as a result ofrifting, mantle plume, and accretion events dating back as faras the Precambrian (e.g., Griffin et al., 2003; Clowes et al.,2010). Indeed, pre-Phanerozoic lithospheric mantle mayhave been particularly well fertilized because the slab-derivedfluids at that time seem likely to have been less oxidized(Evans and Tomkins, 2011). Subduction-zone flattening andcrustal thickening, which appear to have accompanied gene-sis of many major porphyry copper belts (see above), cancause almost complete elimination of mantle wedges, aprocess which may enhance the metasomatism of lithosphericmantle and lower crust as well as their contributions to mag-matism just prior to eventual onset of amagmatic conditions(Kay et al., 1999; Haschke et al., 2002; Lee et al., 2012), as

well as subsequently (Humphreys et al., 2003). Metasoma-tized lithospheric mantle and mafic lowermost crust may alsodirectly contribute partial melts capable of forming porphyrycopper deposits after collision has shut down subduction(Richards, 2009, 2011), a situation that can also cause delam-ination (gravity-induced foundering) of potentially fertilelithospheric material into the asthenosphere (e.g., Grasberg;Cloos et al., 2005; Fig. 9a).

Using evidence furnished by mafic enclaves in the porphyrycopper-related stock, Core et al. (2006) proposed that theBingham Canyon deposit and, by extrapolation, porphyrycopper belts elsewhere may be the results of partial meltingof previously copper- and sulfur-enriched lower crust orlithospheric mantle. Based on strontium and neodymium iso-tope ratios of coeval volcanic rocks (Waite et al., 1997) andlead isotope ratios of fluid inclusions in hydrothermal quartzveinlets (Pettke et al., 2010), the subcontinental lithosphericmantle, subduction-metasomatized and accreted during thePaleoproterozoic (Whitmeyer and Karlstrom, 2007), is per-haps the more likely source of the magmas and associatedcopper (and gold and molybdenum) at Bingham Canyon.Similar Paleoproterozoic lithosphere may also have been theultimate copper source for the Laramide porphyry copper de-posits of southwestern North America (e.g., Bouse et al.,1999; Humphreys et al., 2003; Pettke et al., 2010; Leveilleand Stegen, 2012).

The fundamental role played by fertile, subduction-meta-somatized, oxidized, and copper-enriched mantle and lower-most crust in the generation of porphyry copper belts andprovinces is underscored by two additional observations.First, many magmatic arcs that lack appreciable porphyry

12 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 12

Pelagicsediment

Oxidizedcrust

Magmachamber

Asthenosphere

Lithosphericmantle

Partialmelting

Convectivemantle flow

Oxidized+copper-rich

Cumulates

Oxidizedfluids fromslab

Oceaniclithosphere

FIG. 8. Schematic section of a convergent margin, showing ascent of oxidized, copper-bearing fluids from the subductedoceanic plate into the asthenospheric wedge, where relatively low-degree partial melting produces basaltic magma. Thebasaltic magma ascends through and causes further partial melting of the mantle lithosphere and lower crust, which locallymay have been oxidized and copper enriched during previous subduction episodes/cycles. The resulting fertile basalticmagma may underplate the crust or pass directly into the lower crust where it evolves to an andesitic composition by theMASH (mixing, assimilation, storage, and homogenization) process (Richards, 2003, 2011, and references therein). Cuprif-erous sulfide-bearing cumulates can also result from magmatic differentiation in the upper lithospheric mantle and lowercrust. Relatively oxidized (as opposed to reduced) crustal profiles maximize delivery of copper and other chalcophile metalsto ore-depositional sites in the upper crust. Major porphyry copper provinces are believed to develop above the oxidized andcopper-enriched zones of lithospheric mantle and lowermost crust, which, in some cases, may have been accreted duringmuch earlier collisional orogeny.

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 13: Chapter 1 Copper Provinces

copper mineralization have been shown by isotopic studies topossess only limited mantle contributions and to have formedmainly from recycled crustal material. The Mesozoic arcs ofthe Russian Far East, southeastern China, Korea, and Japanare prime examples (e.g., Jahn, 2010). The crust in these re-gions is dominated by thick, reduced sedimentary packages,which therefore favor ilmenite-series magmatism inimical to

porphyry copper formation (Sato, 2012, and referencestherein). Second, episodic high-flux magmatic events in An-dean-type arcs, resulting from enhanced crustal melting, giverise to voluminous felsic ignimbrite eruptions (de Silva et al.,2006; DeCelles et al., 2009), but few, if any, porphyry copperdeposits. Hattori and Keith (2001), Audétat and Simon(2012), and Loucks (2012) emphasized the critical role of ju-venile mafic melts introduced into midcrustal or even deeperfelsic magma chambers for effective porphyry copper genesis.

Sediment-hosted stratiform copper deposits

It is now commonly accepted that the voluminous brinesresponsible for sediment-hosted stratiform copper mineral-ization in the Central African Copperbelt and Kupferschieferprovince may have circulated through fault arrays in underly-ing basement rocks as well as the redbed sedimentary se-quences that floor the copper-mineralized sedimentary basins(e.g., Blundell et al., 2003; Schmidt Mumm and Wolfgramm,2004; Koziy et al., 2009; Hitzman et al., 2010; Borg et al.,2012). Indeed, the basement itself also locally underwentmineralization in the Central African Copperbelt (e.g.,Lumwana deposit, Zambia; Bernau et al., 2007; Hitzman etal., 2012).

The immediate basement to the Dzhezkazgan province(Fig. 6c) seems likely to be relatively oxidized and potentiallycopper mineralized, given that it comprises accreted early Pa-leozoic magmatic arc terranes and the western portion of aDevonian arc. Furthermore, both the Devonian and Car-boniferous porphyry copper-bearing arcs probably shed detri-tus into the Chu-Sarysu basin, the host to the Dzhezkazgandistrict (Figs. 6c, 9b). Oxidized crust is also a strong possibil-ity in the Central African Copperbelt and Kupferschieferprovince because both are partly underlain by arc terranesand derivative (meta)sedimentary rocks of Paleoproterozoicand Paleozoic age, respectively (Oszczepalski, 1999; Rainaudet al., 2005; Borg et al., 2012; Schuh et al., 2012). Severallow-grade porphyry copper deposits and associated skarnsoccur in the Paleozoic basement of southwestern Poland(Haranczyk, 1980) but, although minor copper occurrencesare widespread in the pre-Katangan basement of the CentralAfrican Copperbelt in Zambia (Pienaar, 1961), it remains tobe determined if they are broadly contemporaneous withtheir host rocks or were introduced during passage of brinesthrough the basement at the time of the stratigraphicallyhigher, sediment-hosted stratiform copper mineralization, asappears to be the case at Lumwana (see above).

IOCG deposits

By analogy with porphyry copper deposits, it may be spec-ulated that the mafic magmas related temporally to majorIOCG deposits may also have tapped preexisting segments ofoxidized and copper-enriched subcontinental lithosphericmantle located on the margins of Archean cratons (Graingeret al., 2008; Groves et al., 2010), prior to ascent through dom-inantly oxidized crustal rocks via deeply penetrating faultzones (e.g., Olympic Dam; O’Driscoll, 1985; Ehrig et al.,2012). The mafic magmas may have been introduced by man-tle plumes (Groves et al., 2010; Fig. 9c). Certainly, in the caseof the Olympic Dam deposit, some of the neodymium in thehematite-rich ore was derived from the same source as that in

COPPER PROVINCES 13

0361-0128/98/000/000-00 $6.00 13

a

b

c

Magmachamber

Magma chamber

Crustalextension

Tectonically thickenedcontinental crustCollisional

suture

Asthenosphere

Upwellingasthenosphere

Forelandbasin

Lithospheric mantle

Asthenosphere

Asthenosphere

MantleplumeMagma generation

in metasomatizedlithospheric mantle

Detached slabsinking into

asthenosphere

Collisionalsuture

Erosion ofarc detritus Hinterland

extensional basin

Thickenedcontinental

crustPluton

Metasomatizedlithospheric

mantle

Metasomatizedlithospheric

mantle

Magma generationin metasomatizedlithospheric mantle

Lithosphericmantle

Cumulates

Delaminatedsinking

into asthenosphere

lithosphericmantle

FIG. 9. Cartoons of additional geotectonic scenarios that could be influen-tial in major copper deposit/province formation: a. Formation of porphyrycopper magma by small-degree partial melting of metasomatized (and possi-bly cumulate-bearing) lithospheric mantle and lower crust during asthenos-phere upwelling after subduction ceases, with delaminated lithospheric man-tle sinking into the asthenosphere (e.g., Richards, 2011); b. Contribution oferosional products of copper-bearing magmatic arc to sediment-hosted strat-iform copper deposit formation in extensional hinterland basin, followingcontinental collision (and slab detachment), as applicable to Dzhezkazgan;and c. Intraplate formation of IOCG-related magma from copper-enrichedlithospheric mantle and lowermost crust during mantle-plume activity (e.g.,Groves et al., 2010).

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 14: Chapter 1 Copper Provinces

synmineral, alkaline mafic/ultramafic dikes within the ore-body, implying a mantle source for both and, arguably, muchof the associated copper (Johnson and McCulloch, 1995;Ehrig et al., 2012). Indeed, availability of such fertile lithos-pheric mantle may be a fundamental requirement for IOCGgeneration in intraplate settings (Fig. 9c), in which mostanorogenic magmatic provinces are devoid of copper (andgold) mineralization. If the existence of subjacent fertilelithospheric mantle and lowermost crust was a prerequisitefor formation of the central Andean copper province, as hy-pothesized above, then the transition from IOCG dominancein the Mesozoic to porphyry copper dominance in the Ceno-zoic may be somehow related to the temporally coincidentchange from extensional to contractional tectonic conditionsinstigated by opening of the South Atlantic Ocean (seeabove).

ConclusionsSeveral recent studies of major copper deposits and

provinces have identified the potentially important metallo-genic role played by the lithospheric mantle in both accre-tionary orogens and anorogenic, intraplate settings (Graingeret al., 2008; Sillitoe, 2008; Bierlein et al., 2009; Shafiei et al.,2009; Groves et al., 2010; Pettke et al., 2010). Although por-phyry copper genesis is commonly triggered by subduction-induced partial melting of asthenospheric mantle wedges,previously oxidized and metasomatized lithospheric mantleand overlying lowermost crust appear to have contributed im-portantly to the development of most major porphyry copperdeposits, belts, and provinces. This fertile lithospheric mantleand lowermost crust, commonly parts of accreted Precam-brian terranes, may contribute supplementary copper to as-cending asthenospheric magmas, both during subductionand, less commonly, intraplate extension, as well as undergo-ing partial melting in their own right. Shallower crust, al-though necessarily oxidized, seems to play a subordinate rolein the provision of copper, except in the case of extensionalsedimentary basins where subjacent magmatic arc terranes aswell as the redbed detritus derived from them are the likelysources of the copper in sediment-hosted stratiform deposits.Recurrent tapping of oxidized and potentially copper-en-riched domains of the lithospheric mantle and derivative copper-enriched crust during several discrete metallogenicepochs has the potential to account for the origin of theworld’s major copper provinces. However, in marked contrastto intraplate IOCG provinces, mantle plume-related, mag-matic nickel-copper deposits, including major, copper-richexamples like Noril’sk (Fig. 1), probably derive their metalsdirectly from the lower asthenospheric mantle (Maier andGroves, 2011; Burrows and Lesher, 2012), although this couldbe copper enriched locally as a result of deep cycling of eithersubducted slabs or delaminated lithospheric mantle (e.g.,Walter et al., 2011; Fig. 9a, b). Notwithstanding the funda-mental role proposed here for lithospheric processes, opti-mization of ore formation at both the deposit and districtscales as well as tectonic setting—for example, regional con-traction and consequent crustal thickening during porphyrycopper formation—can contribute to size and grade en-hancement of individual deposits and belts and, as a result, tooverall copper endowment.

In summary, the world’s premier copper provinces may beconsidered as the cumulative products of a well-orchestratedseries of copper concentration processes: copper precipita-tion on and beneath the ocean floor, at and beyond mid-oceanridges, and its eventual subduction at accretionary margins;extraction of copper from the subducted slabs and its intro-duction into superjacent asthenospheric mantle wedges; het-erogeneous copper accumulation and short- or long-termstorage in the lithospheric mantle and lowermost crust; re-current incorporation of copper into ascendant magmas; andcopper ore formation in association with upper-crustal intru-sions and sedimentary basins. In the final analysis, however,the key role proposed here for localized copper enrichmentof the lithospheric mantle and lowermost crust is simply a lat-ter-day elaboration of what Spurr (1923, p. 431) termed “aheterogeneous stable under-earth......a rich storehouse ofmetals.”

AcknowledgmentsI thank Rio Tinto Exploration for honoring me with this

dedicated volume, the Society of Economic Geologists for ac-cepting it as a Special Publication, and the editors—JeffHedenquist, Mike Harris, and Francisco Camus—for theirhard work during volume planning and manuscript reviewand revision as well as for inviting me to contribute this in-troductory paper. George Steele provided me with sage ad-vice during volume conception, and Alice Bouley, Brian Hoal,Christine Horrigan, Mabel Peterson, Stuart Simmons, andVivian Smallwood assisted the editors with the planning andpublication processes.

I am also deeply grateful to the numerous geologists whoover the years have shared their knowledge of the copper de-posits and provinces briefly referred to in this introductorypaper, during consulting assignments and the occasional minetour. The authors of the other 21 papers—all of them cited inthis introduction—are also thanked for helping to make this alandmark volume on the major copper deposits and districtsof the world. Manuscript reviews by Mike Harris, Keiko Hat-tori, Jeff Hedenquist, Pepe Perelló, Stuart Simmons and, onbehalf of the Society of Economic Geologists, Murray Hitz-man, John Thompson, and Dick Tosdal greatly improved thispaper.

REFERENCESAudétat, A., and Simon, A.C., 2012, Magmatic controls on porphyry copper

genesis: Society of Economic Geologists Special Publication 16, p. 553–572.Bahlburg, H., Vervoort, J.D., Du Frane, S.A., Bock, B., Augustsson, C., and

Reimann, C., 2009, Timing of crust formation and recycling in accretionaryorogens: Insights learned from the western margin of South America:Earth-Science Reviews, v. 97, p. 215–241.

Barton, M.D., 2009, IOCG deposits—a Cordilleran perspective, in Williams,P.J., et al., eds., Smart science for exploration and mining: Biennial Meet-ing of the Society for Geology Applied to Mineral Deposits (SGA), 10th,Townsville, Proceedings, v. 1, p. 4–7.

Bernau, R., Richards, M., Roberts, S., and Boyce, A., 2007, The Lumwanabasement hosted Cu-Co (U) deposits, NW Zambia, in Andrew C.J., et al.,eds., Digging deeper: Biennial Meeting of the Society for Geology Appliedto Mineral Deposits (SGA), 9th, Dublin, 2007, Proceedings, v. 2, p.1003–1006.

Bierlein, F.P., Groves, D.I., and Cawood, P.A., 2009, Metallogeny of accre-tionary orogens—The connection between lithospheric processes andmetal endowment: Ore Geology Reviews, v. 36, p. 282–292.

Blundell, D.J., Karnkowski, P.H., Alderton, D.H.M., Oszczepalski, S., andKucha, H., 2003, Copper mineralization of the Polish Kupferschiefer: A

14 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 14

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 15: Chapter 1 Copper Provinces

proposed basement fault-fracture system of fluid flow: Economic Geology,v, 98, p. 1487–1495.

Borg, G., Piestrzynski, A., Bachmann, G.H., Püttmann, W., Walther, S., andFiedler, M., 2012, An overview of the European Kupferschiefer deposits:Society of Economic Geologists Special Publication 16, p. 455–486.

Bouse, R.M., Ruiz, J., Titley, S.R., Tosdal, R.M., and Wooden, J.L., 1999, Leadisotope compositions of Late Cretaceous and early Tertiary igneous rocksand sulfide minerals in Arizona; implications for the sources of plutons andmetals in porphyry copper deposits: Economic Geology, v. 94, p. 211–244.

Box, S.E., Syusyura, B., Seltmann, R., Creaser, R.A., Dolgopolova, A., andZientek, M.L., 2012, Dzhezkazgan and associated sandstone copper de-posits of the Chu-Sarysu basin, central Kazakhstan: Society of EconomicGeologists Special Publication 16, p. 303–328.

Broughton, D.W., and Rogers, T., 2010, Discovery of the Kamoa copper de-posit, Central African Copperbelt, D.R.C.: Society of Economic GeologistsSpecial Publication 15, p. 287–297.

Burrows, D.R., and Lesher, C.M., 2012, Copper-rich magmatic Ni-Cu-PGEdeposits: Society of Economic Geologists Special Publication 16, p.515–552.

Camus, F., 2003, Geología de los sistemas porfíricos en los Andes de Chile:Santiago, Chile, Servicio Nacional de Geología y Minería, 267 p.

Cathles, M.L., 2011, What processes at mid-ocean ridges tell us about vol-canogenic massive sulfide deposits: Mineralium Deposita, v. 46, p.639–657.

Chechetkin, V.S., Yurgenson, G.A., Narkelyun, L.F., Trubachev, A.I., and Sa-likhov, V.S., 2000, Geology and ores of the Udokan copper deposit (review):Russian Geology and Geophysics, v. 41, p. 733–745.

Clark, A.H., 1993, Are outsize porphyry copper deposits either anatomicallyor environmentally distinctive?: Society of Economic Geologists SpecialPublication 2, p. 213–283.

Cloos, M., Sapiie, B., van Ufford, A.Q., Weiland, R.J., Warren, P.Q., andMcMahon, T.P., 2005, Collisional delamination in New Guinea: The geo-tectonics of subducting slab breakoff: Geological Society of America Spe-cial Paper 400, 51 p.

Clowes, R.M., White, D.J., and Hajnal, Z., 2010, Mantle heterogeneities andtheir significance: results from Lithoprobe seismic reflection and refrac-tion–wide-angle reflection studies: Canadian Journal of Earth Sciences, v.47, p. 409–443.

Collins, W.J., 2002, Nature of extensional accretionary orogens: Tectonics, v.21, p. 6-1–6-12.

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

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

Crane, D., and Kavalieris, I., 2012, Geologic overview of the Oyu Tolgoi por-phyry Cu-Au-Mo deposits, Mongolia: Society of Economic Geologists Spe-cial Publication 16, p. 187–213.

Cunningham, C.G., Singer, D.A., Zappettini, E.O., Vivallo, W., Celado, C.M.,Quispe, J., Briskey, J.A., Sutphin, D.M., Gajardo, M., Diaz, A., Portigliati,C., Berger, V.I., Carrasco, R., and Schulz, K.J., 2007, A preliminary quanti-tative mineral resource assessment of undiscovered porphyry copper re-sources in the Andes Mountains of South America: Society of EconomicGeologists, Newsletter 71, p. 1, 8–13.

DeCelles, G., Ducea, M.N., Kapp, P., and Zandt, G., 2009, Cyclicity inCordilleran orogenic systems: Nature Geoscience, v. 2, p. 251–257.

de Silva, S., Zandt, G., Trumbull, R., Viramonte, J.G., Salas, G., and Jiménez,N., 2006, Large ignimbrite eruptions and volcano-tectonic depressions inthe Central Andes: A thermomechanical perspective: Geological Society,London, Special Publication 269, p. 47–63.

Dilles, J.H., and Proffett, J.M., 1995, Metallogenesis of the Yeringtonbatholith: Arizona Geological Society Digest 20, p. 306–315.

Dreyer, B.M., Morris, J.D., and Gill, J.B., 2010, Incorporation of subductedslab-derived sediment and fluid in arc magmas: B-Be-10Be–εNd systemat-ics of the Kurile convergent margin, Russia: Journal of Petrology, v. 51, p.1761–1782.

du Bray, E.A., and John, D.A., 2011, Petrologic, tectonic, and metallogenicevolution of the Ancestral Cascades magmatic arc, Washington, Oregon,and northern California: Geosphere, v. 7, p. 1102–1133.

Ehrig, K., McPhie, J., and Kamenetsky, V., 2012, Geology and mineralogicalzonation of the Olympic Dam iron oxide Cu-U-Au-Ag deposit, South Aus-tralia: Society of Economic Geologists Special Publication 16, p. 237–267.

Einaudi, M.T., Hedenquist, J.W., and Inan, E.E., 2003, Sulfidation state offluids in active and extinct hydrothermal systems: Transitions from por-phyry to epithermal environments: Society of Economic Geologists SpecialPublication 10, p. 285–313.

Emmons, W.H., 1927, Relations of metalliferous lode systems to igneous in-trusives: Transactions of the American Institute of Mining and Metallurgi-cal Engineers, v. 74, p. 29–70.

Evans, K.-A., and Tomkins, A.-G., 2011, The relationship between subduc-tion zone redox budget and arc magma fertility: Earth and Planetary Sci-ence Letters, v. 308, p. 401–409.

Fairclough, M., 2005, Geological and metallogenic setting of the Carrap-ateena FeO-Cu-Au prospect—a PACE success story: MESA Journal, v. 38,p. 4–7. Adelaide, Department of Primary Industries and Resources SouthAustralia.

Fay, I., and Barton, M.D., 2012, Alteration and ore distribution in the Pro-terozoic Mines Series, Tenke-Fungurume Cu-Co district, Democratic Re-public of Congo: Mineralium Deposita, v. 47, p. 501–519.

Finzel, E.S., Trop, J.M., Ridgway, K.D., and Enkelmann, E., 2011, Upperplate proxies for flat-slab subduction processes in southern Alaska: Earthand Planetary Science Letters, v. 303, p. 348–360.

Gablina, I.F., 1981, New data on formation conditions of the Dzhezkazgancopper deposit: International Geology Review, v. 23, p. 1303–1311.

Gandler, L.M., and Kyle, J.R., 2008, Stratigraphic controls of calc-silicate al-teration and copper-gold mineralisation of the Deep Mill Level zone skarn,Ertsberg district, Papua, Indonesia: Pacrim 2008 Conference, Gold Coast,Queensland, 2008, Proceedings: Melbourne, Australasian Institute of Min-ing and Metallurgy, p. 313–317.

Grainger, C.J., Groves, D.I., Tallarico, F.H.B., and Fletcher, I.R., 2008, Met-allogenesis of the Carajás Mineral Province, Southern Amazon Craton,Brazil: Varying styles of Archean through Paleoproterozoic to Neoprotero-zoic base- and precious-metal mineralisation: Ore Geology Reviews, v. 33,p. 451–489.

Griffin, W.L., O’Reilly, S.Y., Natapov, L.M., and Ryan, C.G., 2003, The evo-lution of lithospheric mantle beneath the Kalahari Craton and its margins:Lithos, v. 71, p. 215–241.

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

Gutscher, M-A., 2002, Andean subduction styles and their effect on thermalstructure and interplate coupling: Journal of South American Earth Sci-ences, v. 15, p. 3–10.

Haest, M., and Muchez, P., 2011, Stratiform and vein-type deposits in thePan-African orogen in central and southern Africa: Evidence for multi-phase mineralisation: Geologica Belgica, v. 14, p. 23–44.

Hannington, M., 2011, Comments on ‘‘What processes at mid-ocean ridgestell us about volcanogenic massive sulfide deposits’’ by L.M. Cathles: Min-eralium Deposita, v. 46, p. 659–663.

Hanson, R.E., 2003, Proterozoic geochronology and tectonic evolution ofsouthern Africa: Geological Society, London, Special Publication 206, p.427–463.

Haranczyk, C., 1980, Palaeozoic porphyry copper deposits in Poland, inJankovic, S., and Sillitoe, R.H., eds., European copper deposits: Society ofGeology Applied to Mineral Deposits (SGA) Special Publication 1, p.89–95.

Haschke, M., Scheuber, E., Günther, A., and Reutter, A.J., 2002, Evolution-ary cycles during the Andean orogeny: repeated slab breakoff and flat sub-duction: Terra Nova, v. 14, p. 49–56.

Hattori, K.H., 2007, Sulfur and copper in magmatic arcs: Sources and link-ages [abs.]: American Geophysical Union Fall Meeting, San Francisco,2007, Abstract V33F-01, 1 p.

Hattori, K., and Guillot, S., 2003, Volcanic fronts form as a consequence ofserpentinite dehydration in the forearc mantle wedge: Geology. v. 31, p.525–528.

Hattori, K.H., and Keith, J.D., 2001, Contribution of mafic melt to porphyrycopper mineralization: evidence from Mount Pinatubo, Philippines, andBingham Canyon, Utah, USA: Mineralium Deposita, v. 36, p. 799–806.

Hawkesworth, C., Cawood, P., Kemp, T., Storey, C., and Dhuime, B., 2009,A matter of preservation: Science, v. 323, p. 49–50.

Hawkesworth, C.J., Dhuime, B., Pietranik, A.B., Cawood, P.A., Kemp, A.I.S.,and Storey, C.D., 2010, The generation and evolution of the continentalcrust: Journal of the Geological Society, London, v. 167, p. 229–248.

COPPER PROVINCES 15

0361-0128/98/000/000-00 $6.00 15

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 16: Chapter 1 Copper Provinces

Hehnke, C., Ballantyne, G., Martin, H., Hart, W., Schwarz, A., and Stein, H.,2012, Geology and exploration progress at the Resolution porphyry Cu-Modeposit, Arizona: Society of Economic Geologists Special Publication 16, p.147–166.

Hervé, M., Sillitoe, R.H., Wong, C., Fernández, P., Crignola, F., Ipinza, M.,and Urzúa, F., 2012, Geologic overview of the Escondida porphyry copperdistrict, northern Chile: Society of Economic Geologists Special Publica-tion 16, p. 55–78.

Hitzman, M., Kirkham, R., Broughton, D., Thorson, J., and Selley, D., 2005,The sediment-hosted stratiform copper ore system: Society of EconomicGeologists 100th Anniversary Volume, p. 609–642.

Hitzman, M.W., Selley, D., and Bull., S., 2010, Formation of sedimentaryrock-hosted stratiform copper deposits through Earth history: EconomicGeology, v. 105, p. 627–639.

Hitzman, M.W., Broughton, D., Selley, D., Woodhead, J., Wood, D., andBull., S., 2012, The Central African Copperbelt: Diverse stratigraphic,structural, and temporal settings in the world’s largest sedimentary copperdistrict: Society of Economic Geologists Special Publication 16, p. 487–514.

Howell, F.H., and Molloy, S.J., 1960, Geology of the Braden orebody, Chile,South America: Economic Geology, v. 70, p. 863–905.

Humphreys, E., Hessler, E., Dueker, K., Farmer, G.L., Erslev, E., and At-water, T., 2003, How Laramide-age hydration of North American lithos-phere by the Farallon slab controlled subsequent activity in the westernUnited States: International Geology Review, v. 45, p. 575–595.

Irarrazaval, V., Sillitoe, R.H., Wilson, A.J., Toro, J.C., Robles, W., and Lyall,G.D., 2010, Discovery history of a giant, high-grade, hypogene porphyrycopper-molybdenum deposit at Los Sulfatos, Los Bronces-Río Blanco dis-trict, central Chile: Society of Economic Geologists Special Publication 15,p. 253–269.

Jackson, M.P.A., Warin, O.N., Woad, G.M., and Hudec, M.R., 2003, Neo-proterozoic allochthonous salt tectonics during the Lufilian orogeny in theKatangan Copperbelt, central Africa: Geological Society of America Bul-letin, v. 115, p. 314–330.

Jahn, B.-M., 2010, Accretionary orogen and evolution of the Japanese is-lands—implications from a Sr-Nd isotopic study of the Phanerozoic grani-toids from SW Japan: American Journal of Science, v. 310, p. 1210–1249.

Johnson, J.P., and McCulloch, M.T., 1995, Sources of mineralising fluids forthe Olympic Dam deposit (South Australia): Sm-Nd isotopic constraints:Chemical Geology, v. 121, p. 177–199.

Jowett, E.C., 1986, Genesis of Kupferschiefer Cu-Ag deposits by convectiveflow of Rotliegende brines during Triassic rifting: Economic Geology, v. 81,p. 1823–1837.

Kay, S.M., Mpodozis, C., and Coira, B., 1999, Neogene magmatism, tecton-ism, and mineral deposits of the central Andes (22° to 33°S latitude): Soci-ety of Economic Geologists Special Publication 7, p. 27–59.

Keith, S.B., and Swan, M.M., 1995, Tectonic setting, petrology, and genesisof the Laramide porphyry copper cluster of Arizona, Sonora, and NewMexico: Arizona Geological Society Digest 20, p. 339–346.

Kelley, K.A., and Cottrell, E., 2009, Water and the oxidation state of subduc-tion zone magmas: Science, v. 325, p. 605–607.

Kesler, S.E., and Wilkinson, B.H., 2006, The role of exhumation in the tem-poral distribution of ore deposits: Economic Geology, v. 101, p. 919–922.

——2008, Earth’s copper resources estimated from tectonic diffusion of por-phyry copper deposits: Geology, v. 36, p. 255–258.

Kirkham, R.V., 1989, Distribution, settings, and genesis of sediment-hostedstratiform copper deposits: Geological Association of Canada Special Paper36, p. 3–38.

Kirwin, D.J., Forster, C.N., Kavalieris, I., Crane, D., Orssich, C., Panther, C.,Garamjav, D., Munkhbat, T.O., and Niislelkhuu, 2005, The Oyu Tolgoi cop-per-gold porphyry deposits, South Gobi, Mongolia, in Seltmann, R., Gerel,O., and Kirwin, D.J., eds., Geodynamics and metallogeny of Mongolia withspecial emphasis on copper and gold deposits: Society of Economic Geol-ogists–International Association for the Genesis of Ore Deposits Field Trip,2005, London, Centre for Russian and Central EurAsian Mineral Studies,Natural History Museum, IAGOD Guidebook Series 11, p. 155–168.

Kouzmanov, K., and Pokrovski, G.S., 2012, Hydrothermal controls on metaldistribution in porphyry Cu (-Mo-Au) deposits: Society of Economic Geol-ogists Special Publication 16, p. 573–618.

Koziy, L., Bull, S., Large, R., and Selley, D., 2009, Salt as a fluid driver, andbasement as a metal source, for stratiform sediment-hosted copper de-posits: Geology, v. 37, p. 1107–1110.

Krauskopf, K.B., 1971, The source of ore metals: Geochimica et Cos-mochimica Acta, v. 35, p. 643–659.

Lallemand, S., Heuret, A., and Boutelier, D., 2005, On the relationships be-tween slab dip, back-arc stress, upper plate absolute motion, and crustal na-ture in subduction zones: Geochemistry, Geophysics, Geosystems, v. 6,Q09006, 18 p.

Lang, J.R., and Gregory, M.J., 2012, Magmatic-hydrothermal-structural evo-lution of the giant Pebble porphyry Cu-Au-Mo deposit with implicationsfor exploration in southwest Alaska: Society of Economic Geologists Spe-cial Publication 16, p. 167–185.

Laznicka, P., 1999, Quantitative relationships among giant deposits of metals:Economic Geology, v. 94, p. 455–473.

Lee, C., and King, S.D., 2011, Dynamic buckling of subducting slabs recon-ciles geological and geophysical observations: Earth and Planetary ScienceLetters, v. 312, p. 360–370.

Lee, C.-T.A., Luffi, P., Chin, E.J., Bouchet, R., Dasgupta, R., Morton, D.M.,Le Roux, V., Yin, Q.-Z., and Jin, D., 2012, Copper systematics in arc magmasand implications for crust-mantle differentiation: Science, v. 336, p. 64–68.

Leveille, R.A., and Stegen, R.J., 2012, The southwestern North America por-phyry copper province: Society of Economic Geologists Special Publication16, p. 361–401.

Leys, C.A., Cloos, M., New, B.T.E., and MacDonald, G.D., 2012, Copper-gold ± molybdenum deposits of the Ertsberg-Grasberg district, Papua, In-donesia: Society of Economic Geologists Special Publication 16, p. 215–235.

Li, Y.-H., and Schoonmaker, J.E., 2005, Chemical composition and mineral-ogy of marine sediments: Treatise on Geochemistry, v. 7, p. 1–35.

Lindberg, P.A., 1989, Precambrian ore deposits of Arizona: Arizona Geolog-ical Society Digest, v. 17, p. 187–210.

Lindgren, W., 1909, Metallogenetic epochs: Economic Geology, v. 4, p.409–420.

Liu, L., Gurnis, M., Seton, M., Saleeby, J., Müller, R.D., and Jackson, J.M.,2010, The role of oceanic plateau subduction in the Laramide orogeny: Na-ture Geoscience, v. 3, p. 353–357.

Livaccari, R.F., Burke, K., and Sengör, A.M.C., 1981, Was the Laramideorogeny related to subduction of an oceanic plateau?: Nature, v. 289, p.276–278.

Loucks, R., 2012, Chemical characteristics, geodynamic settings, and petro-genesis of copper-ore-forming arc magmas: Perth, Centre for ExplorationTargeting, University of Western Australia, CET Quarterly Newsletter 19,p. 1–10.

Love, D.A., Clark, A.H., and Glover, J.K., 2004, The lithologic, stratigraphic,and structural setting of the giant Antamina copper-zinc skarn deposit, An-cash, Peru: Economic Geology, v. 99, p. 887–916.

Maier, W.D., and Groves, D.I., 2011, Temporal and spatial controls on theformation of magmatic PGE and Ni-Cu deposits: Mineralium Deposita, v.46, p. 841–857.

Marschik, R., and Fontboté, L., 2001, The Candelaria-Punta del Cobre ironoxide Cu-Au (-Zn-Ag) deposits, Chile: Economic Geology, v. 96, p.1799–1826.

Martinod, J., Husson, L., Roperch, P., Guillaume, B., and Espurt, N., 2010,Horizontal subduction zones, convergence velocity and the building of theAndes: Earth and Planetary Science Letters, v. 299, p. 299–309.

McGowan, R.R., Roberts, S., Foster, R.P., Boyce, A.J., and Coller, D., 2003,Origin of the copper-cobalt deposits of the Zambian Copperbelt: An epige-netic view from Nchanga: Geology, v. 31, p. 497–500.

McLemore, V.T., 1996, Copper in New Mexico: New Mexico Geology, v. 18,p. 25–36.

McPhie, J., Kamenetsky, V., Allen, S., Ehrig, K., Agangi, A., and Bath, A.,2011, The fluorine link between a supergiant ore deposit and a silicic largeigneous province: Geology, v. 39, p. 1003–1006.

Meinert, L.D., Hefton, K.K., Mayes, D., and Tasiran, I., 1997, Geology, zona-tion, and fluid evolution of the Big Gossan Cu-Au skarn deposit, Ertsbergdistrict, Irian Jaya: Economic Geology, v. 92, p. 509–534.

Meyer, C., Shea, E.P., Goddard, C.C., Jr., and Staff, 1968, Ore deposits atButte, Montana, in Ridge, J.D., ed., Ore deposits of the United States,1933-67, v. 2: New York, American Institute of Mining, Metallurgical, andPetroleum Engineers, p. 1373–1416.

Michalik, M., and Sawlowicz, Z., 2001, Multi-stage and long-term origin ofthe Kupferschiefer copper deposits in Poland, in Piestrzynski, A., et al., eds.,Mineral deposits at the beginning of the 21st century: Joint Biennial Meet-ing of the Society for Geology Applied to Mineral Deposits (SGA) and So-ciety of Economic Geologists, 6th, Kraków, 2001, Proceedings, p. 235–238.

Morgan, C.J., 2000, Resource estimates of the Clarion-Clipperton man-ganese nodule deposits, in Cronan, D.S., ed., Handbook of marine mineraldeposits: Boca Raton, Florida, CRC Press, p. 145–170.

16 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 16

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 17: Chapter 1 Copper Provinces

Mpodozis, C., and Cornejo, P., 2012, Cenozoic tectonics and porphyry cop-per systems of the Chilean Andes: Society of Economic Geologists SpecialPublication 16, p. 329–360.

Mpodozis, C., and Ramos, V., 1990, The Andes of Chile and Argentina: Cir-cum-Pacific Council for Energy and Mineral Resources Earth Science Se-ries, v. 11, p. 59–90.

Mungall, J.E., 2002, Roasting the mantle: Slab melting and the genesis ofmajor Au and Au-rich Cu deposits: Geology, v. 30, p. 915–918.

Murakami, H., Seo, J.H., and Heinrich, C.A., 2010, The relation betweenCu/Au ratio and formation depth of porphyry-style Cu-Au ± Mo deposits:Mineralium Deposita, v. 45, p. 11–21.

Noble, J.A., 1970, Metal provinces of the western United States: GeologicalSociety of America Bulletin, v. 81, p. 1607–1624.

Noll, P.D., Jr., Newsom, H.E., Leeman, W.P., and Ryan, J.G., 1996, The roleof hydrothermal fluids in the production of subduction zone magmas: Evi-dence from siderophile and chalcophile trace elements and boron:Geochimica et Cosmochimica Acta, v. 60, p. 587–611.

O’Driscoll, E.S.T., 1985, The application of lineament tectonics in the dis-covery of the Olympic Dam Cu-Au-U deposit at Roxby Downs, South Aus-tralia: Global Tectonics and Metallogeny, v. 3, no. 1, p. 43–57.

Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P., and Schemmann, K.,2006, Deformation of the Central Andean upper plate system—facts, fic-tion, and constraints for plateau models, in Oncken, O., et al., eds., TheAndes: Active subduction orogeny: Berlin Heidelberg, Springer-Verlag, p.3–27.

Ossandón, G., Fréraut, R., Gustafson, L.B., Lindsay, D.D., and Zentilli, M.,2001, Geology of the Chuquicamata mine: A progress report: EconomicGeology, v. 96, p. 249–270.

Oszczepalski, S., 1999, Origin of the Kupferschiefer polymetallic mineraliza-tion in Poland: Mineralium Deposita, v. 34, p. 599–613.

Perelló, J., 2006, Metallogeny of major copper deposits and belts in theNorth and South American Cordillera [abs.]: Backbone of the Americas.Patagonia to Alaska, GSA Specialty Meeting No. 2, Mendoza, Argentina,2006, Abstracts with Programs: Asociación Geológica Argentina and Geo-logical Society of America, p. 83–84.

Perelló, J., Cox, D., Garamjav, D., Sanjdorj, S., Diakov, S., Schissel, D.,Munkhbat, T.-O., and Gonchig, O., 2001, Oyu Tolgoi, Mongolia: Siluro-De-vonian porphyry Cu-Au-(Mo) and high-sulfidation Cu mineralization witha Cretaceous chalcocite blanket: Economic Geology, v. 96, p. 1407–1438.

Perelló, J., Razique, A., Schloderer, J., and Rehman, A.-U., 2008, The Cha-gai porphyry copper belt, Baluchistan province, Pakistan: Economic Geol-ogy, v. 103, p. 1583–1612.

Perelló, J., Sillitoe, R.H., Mpodozis, C., Brockway, H., and Posso, H., 2012,Geologic setting and evolution of the porphyry copper-molybdenum andcopper-gold deposits at Los Pelambres, central Chile: Society of EconomicGeologists Special Publication 16, p. 79–104.

Peslier, A.H., Francis, D, and Ludden, J., 2002, The lithospheric mantle be-neath continental margins: Melting and melt–rock reaction in CanadianCordillera xenoliths: Journal of Petrology, v. 43, p. 2013–2047.

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

Pienaar, P., 1961, Mineralization in the basement complex, in Mendelsohn,F., ed., Geology of the Northern Rhodesian Copperbelt: London, Mac-Donald & Co. (Publishers) Ltd., p. 30–41.

Porter, J.P., Schroeder, K., and Austin, G., 2012, Geology of the BinghamCanyon porphyry Cu-Mo-Au deposit, Utah: Society of Economic Geolo-gists Special Publication 16, p. 127–146.

Racey, A., 2009, Mesozoic red bed sequences from SE Asia and the signifi-cance of the Khorat Group of NE Thailand: Geological Society, London,Special Publications 315, p. 41–67.

Rainaud, C., Master, S., Armstrong, R.A., and Robb, L.J., 2005, Geochronol-ogy and nature of the Palaeoproterozoic basement of the Central AfricanCopperbelt (Zambia and the Democratic Republic of Congo), with re-gional implications: Journal of African Earth Sciences, v. 42, p. 1–31.

Reeve, J.S., Cross, K.C., Smith, R.N., and Oreskes, N., 1990, Olympic Damcopper-uranium-gold-silver deposit: Australasian Institute of Mining andMetallurgy Monograph 14, v. 2, p. 1009–1035.

Réquia, K.C.M., and Fontboté, L., 2000, The Salobo iron oxide copper-golddeposit, Carajás, Northern Brazil, in Porter, T.M., ed., Hydrothermal ironoxide copper-gold and related deposits: A global perspective: Adelaide,Australian Mineral Foundation, p. 225–236.

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

——2005, Cumulative factors in the generation of giant calc-alkaline por-phyry Cu deposits, in Porter, T.M., ed., Super porphyry copper & gold de-posits: A global perspective, v. 1: Adelaide, PGC Publishing, p. 7–25.

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

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

Richards, J.P., Noble, S.R., and Pringle, M.S., 1999, A revised late Eoceneage for porphyry Cu magmatism in the Escondida area, northern Chile:Economic Geology, v. 94, p. 1231–1248.

Rivera, S.L., Alcota, H., Proffett, J., Díaz, J., Leiva, G., and Vergara, M.,2012, Update of the geologic setting and porphyry Cu-Mo deposits of theChuquicamata district, northern Chile: Society of Economic GeologistsSpecial Publication 16, p. 19–54.

Routhier, P., 1976, A new approach to metallogenic provinces: The exampleof Europe: Economic Geology, v. 71, p. 803–811.

Rutland, R.W.R., 1971, Andean orogeny and ocean floor spreading: Nature,v. 233, p. 252–255.

Sato, K., 2012, Sedimentary crust and metallogeny of granitoid affinity: Impli-cations from the geotectonic histories of the circum-Japan Sea region, cen-tral Andes and southeastern Australia: Resource Geology, v. 62, p. 329–351.

Sawkins, F.J., 1972, Sulfide ore deposits in relation to plate tectonics: Journalof Geology, v. 80, p. 377–397.

Schellart, W.P., 2008, Overriding plate shortening and extension above sub-duction zones: A parametric study to explain formation of the Andes Moun-tains: Geological Society of America Bulletin, v. 120, p. 1441–1454.

Schmidt, M.W., and Poli, S., 2005, Generation of mobile components duringsubduction of oceanic crust: Treatise on Geochemistry, v. 3, p. 567–591.

Schmidt Mumm, A., and Wolfgramm, M., 2004, Fluid systems and mineral-ization in the north German and Polish basin: Geofluids, v. 4, p. 315–328.

Schuh, W., Leveille, R., Fay, I., and North, R., 2012, Geology of the Tenke-Fungurume sediment-hosted strata-bound copper-cobalt district, Katanga,Democratic Republic of Congo: Society of Economic Geologists SpecialPublication 16, p. 269–301.

Selley, D., Broughton, D., Scott, R., Hitzman, M., Bull, S., Large, R., Mc-Goldrick, P., Croaker, M., Pollington, N., and Barra, F., 2005, A new lookat the geology of the Zambian Copperbelt: Economic Geology 100th An-niversary Volume, p. 965–1000.

Serrano, L., Vargas, R., Stambuk, V., Aguilar, C., Galeb, M., Holmgren, C.,Contreras, A., Godoy, S., Vela, I., Skewes, A.M., and Stern, C.R., 1996, Thelate Miocene to early Pliocene Río Blanco-Los Bronces copper deposit,central Chilean Andes: Society of Economic Geologists Special Publication5, p. 119–129.

Shafiei, B., Haschke, M., and Shahabpour, J., 2009, Recycling of orogeniccrust triggers porphyry Cu mineralization in Kerman Cenozoic arc rocks,southeastern Iran: Mineralium Deposita, v. 44, p. 265–283.

Sillitoe, R.H., 1972, A plate tectonic model for the origin of porphyry copperdeposits: Economic Geology, v. 67, p. 184–197.

——1985, Ore-related breccias in volcanoplutonic arcs: Economic Geology,v. 80, p. 1467–1514.

——1988, Epochs of intrusion-related copper mineralization in the Andes:Journal of South American Earth Sciences, v. 1, p. 89–108.

——1997, Characteristics and controls of the largest porphyry copper-goldand epithermal gold deposits in the circum-Pacific region: Australian Jour-nal of Earth Sciences, v. 44, p. 373–388.

——1998, Major regional factors favouring large size, high hypogene grade,elevated gold content and supergene oxidation and enrichment of porphyrycopper deposits, in Porter, T.M., ed., Porphyry and hydrothermal copperand gold deposits: A global perspective: Adelaide, Australian MineralFoundation, p. 21–34.

——2003, Iron oxide-copper-gold deposits: an Andean view: MineraliumDeposita, v. 38, p. 787–812.

——2005, Supergene oxidized and enriched porphyry copper and relateddeposits: Economic Geology 100th Anniversary Volume, p. 723–768.

——2008, Major gold deposits and belts of the North and South AmericanCordillera: Distribution, tectonomagmatic settings, and metallogenic consi-derations: Economic Geology, v. 103, p. 663–687.

——2010a, Exploration and discovery of base- and precious-metal depositsin the circum-Pacific region—a 2010 perspective: Resource Geology Spe-cial Issue 22, 139 p.

COPPER PROVINCES 17

0361-0128/98/000/000-00 $6.00 17

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020

Page 18: Chapter 1 Copper Provinces

——2010b, Porphyry copper systems: Economic Geology, v. 105, p. 3–41.Sillitoe, R.H., and Perelló, J., 2005, Andean copper province: Tectonomag-

matic settings, deposit types, metallogeny, exploration, and discovery: Eco-nomic Geology 100th Anniversary Volume, p. 845–890.

Sillitoe, R.H., Perelló, J., and García, A., 2010, Sulfide-bearing veinletsthroughout the stratiform mineralization of the Central African Copper-belt: Temporal and genetic implications: Economic Geology, v. 105, p.1361–1368.

Singer, D.A., 1995, World class base and precious metal deposits–A quanti-tative analysis: Economic Geology, v. 90. p. 88–104.

Skewes, M.A., and Stern, C.R., 1995, Genesis of the giant Late Miocene toPliocene copper deposits of central Chile in the context of Andean mag-matic and tectonic evolution: International Geology Review, v. 37, p.893–909.

Skewes, M.A., Arévalo, A., Floody, R., Zuñiga, P.H., and Stern, C.R., 2002,The giant El Teniente breccia deposit: Hypogene copper distribution andemplacement: Society of Economic Geologists Special Publication 9, p.299–332.

Skirrow, R.G., 2009, “Hematite-group” IOCG±U ore systems: tectonic set-tings, hydrothermal characteristics, and Cu-Au and U mineralizing processes:Geological Association of Canada Short Course Notes 20, p. 39–57.

Sobolev, S.V., and Babeyko, A.Y., 2005, What drives orogeny in the Andes?:Geology, v. 33, p. 617–620.

Spurr, J.E., 1923, The ore magmas: New York, McGraw-Hill Book Company,915 p.

Stern, R.J., and Bloomer, S.H., 1992, Subduction zone infancy: Examplesfrom the Eocene Izu-Bonin-Mariana and Jurassic California arcs: Geologi-cal Society of America Bulletin, v. 104, p. 1621–1636.

Stoll, W.C., 1965, Metallogenetic provinces of South America: Mining Mag-azine, v. 112, p. 22–33, 90–99.

Symons, D.T.A., Kawasaki, K., Walther, S., and Borg, G., 2011, Paleomag-netism of the Cu–Zn–Pb-bearing Kupferschiefer black shale (Upper Per-mian) at Sangerhausen, Germany: Mineralium Deposita, v. 46, p. 137–152.

Takada, A., 1994, The influence of regional stress and magmatic input onstyles of monogenetic and polygenetic volcanism: Journal of GeophysicalResearch, v. 99, B7, p. 13,563–13,573.

Titley, S.R., 1982, Geologic setting of porphyry copper deposits: Southeast-ern Arizona, in Titley, S.R., ed., Advances in geology of the porphyry cop-per deposits, southwestern North America: Tucson, University of ArizonaPress, p. 37–58.

——2001, Crustal affinities of metallogenesis in the American Southwest:Economic Geology, v. 96, p. 1323–1342.

Tonkin, D.G., and Creelman, R.A., 1990, Mount Gunson copper deposits:Australasian Institute of Mining and Metallurgy Monograph 14, v. 2, p.1037–1043.

Toro, J.C., Ortúzar, J., Zamorano, J., Cuadra, P., Hermosilla, J., and Spröhnle,C., 2012, Protracted magmatic-hydrothermal history of the Río Blanco-LosBronces district, central Chile: Development of world’s greatest knownconcentration of copper: Society of Economic Geologists Special Publica-tion 16, p. 105–126.

Tosdal, R.M., Dilles, J.H., and Cooke, D.R., 2009, From source to sinks inauriferous magmatic-hydrothermal porphyry and epithermal deposits: Ele-ments, v. 5, p. 289–295.

Turneaure, F.S., 1955, Metallogenetic provinces and epochs: Economic Ge-ology 50th Anniversary Volume, Part I, p. 38–98.

U.S. Geological Survey, 2011, Mineral commodity summaries 2011: Reston,Virginia, U. S. Geological Survey, 198 p.

Uyeda, S., and Nishiwaki, C., 1980, Stress field, metallogenesis and mode ofsubduction: Geological Association of Canada Special Paper 20, p.323–340.

Vallejo, C., Winkler, W., Spikings, R.A., Luzieux, L., Heller, F., and Bussy, F.,2009, Mode and timing of terrane accretion in the forearc of the Andes inEcuador: Geological Society of America Memoir 204, p. 197–216.

van Dongen, M., Weinberg, R.F., Tomkins, A.G., Armstrong, R.A., andWoodhead, J.D., 2010, Recycling of Proterozoic crust in Pleistocene juve-nile magma and rapid formation of the Ok Tedi porphyry Cu–Au deposit,Papua New Guinea: Lithos, v. 114, p. 282–292.

van Hunen, J., van den Berg, A.P., and Vlaar, N.J., 2004, Various mechanismsto induce present-day shallow flat subduction and implications for theyounger Earth: A numerical parameter study: Physics of the Earth andPlanetary Interiors, v. 146, p. 179–194.

van Wees, J-D., Stephenson, R.A., Ziegler, P.A., Bayer, U., McCann, T.,Dadlez, R., Gaupp, R., Narkiewicz, M., Bitzer, F., and Scheck, M., 2000,On the origin of the Southern Permian Basin, Central Europe: Marine andPetroleum Geology, v. 17, p. 43–59.

Vry, V.H., Wilkinson, J.J., Seguel, J., and Millán, J., 2010, Multistage intru-sion, brecciation, and veining at El Teniente, Chile: Evolution of a nestedporphyry system: Economic Geology, v. 105, p. 119–153.

Wainwright, A.J., Tosdal, R.M., Forster, C.N., Kirwin, D.J., Lewis, P.D., andWooden, J.L., 2011, Devonian and Carboniferous arcs of the Oyu Tolgoiporphyry Cu-Au district, South Gobi region, Mongolia: Geological Societyof America Bulletin, v. 123, p. 306–328.

Waite, K.A., Keith, J.D., Christiansen, E.H., Whitney, J.A., Hattori, K.,Tingey, D.G., and Hook, C.J., 1997, Petrogenesis of the volcanic and intru-sive rocks associated with the Bingham Canyon porphyry Cu-Au-Mo de-posit, Utah: Society of Economic Geologists Guidebook, v. 29, p. 69–90.

Walter, M.J., Kohn, S.C., Araujo, D., Bulanova, G.P., Smith, C.B., Gaillou,E., Wang, J., Steele, A., and Shirey, S.B., 2011, Deep mantle cycling ofoceanic crust: Evidence from diamonds and their mineral inclusions: Sci-ence, v. 334, p. 54–57.

Wanhainen, C., Billström, K., and Martinsson, O., 2006, Age, petrology andgeochemistry of the porphyritic Aitik intrusion, and its relation to the dis-seminated Aitik Cu-Au-Ag deposit, northern Sweden: GFF [Geological So-ciety of Sweden], v. 128, p. 273–286.

Whitmeyer, S.J., and Karlstrom, K.E., 2007, Tectonic model for the Protero-zoic growth of North America: Geosphere, v. 3, p. 22–59.

Xavier, R.P., Monteiro, L.V.S., Moreto, C.P.N., Pestilho, A.L.S., Coelho deMelo, G.H., Delinardo da Silva, M.A., Aires, B., Ribeiro, C., and Freitas eSilva, F.H., 2012, The iron oxide copper-gold systems of the Carajás min-eral province, Brazil: Society of Economic Geologists Special Publication16, p. 433–454.

Xiao, W.J., Windley, B.F., Huang, B.C., Han, C.M., Yuan, C., Chen, H.L.,Sun, M., Sun, S., and Li, J.L., 2009, End-Permian to mid-Triassic termina-tion of the accretionary processes of the southern Altaids: implications forthe geodynamic evolution, Phanerozoic continental growth, and metal-logeny of Central Asia: International Journal of Earth Sciences, v. 98, p.1189–1217.

Yakubchuk, A.S., Shatov, V.V., Kirwin, D., Edwards, A., Tomurtogoo, O.,Badarch, G., and Buryak, V.A., 2005, Gold and base metal metallogeny ofthe central Asian orogenic supercollage: Economic Geology 100th Anniver-sary Volume, p. 1035–1068.

Yakubchuk, A., Degtyarev, K., Maslennikov, V., Wurst, A., Stekhin, A., andLobanov, K., 2012, Tectonomagmatic settings, architecture, and metal-logeny of the Central Asian copper province: Society of Economic Geolo-gists Special Publication 16, p. 403–432.

Zhou, X., Sun, T., Shen, W., and Shu, L., 2006, Petrogenesis of Mesozoicgranitoids and volcanic rocks in South China: A response to tectonic evolu-tion: Episodes, v. 29, p. 26–33.

18 RICHARD H. SILLITOE

0361-0128/98/000/000-00 $6.00 18

Downloaded from https://pubs.geoscienceworld.org/books/chapter-pdf/3811772/9781629490410_ch01.pdfby gueston 05 August 2020