26
Review Magmatic to hydrothermal metal uxes in convergent and collided margins Jeremy P. Richards Dept. of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta, Canada T6G 2E3 abstract article info Article history: Received 1 December 2010 Received in revised form 18 May 2011 Accepted 19 May 2011 Available online 27 May 2011 Keywords: Porphyry deposit Epithermal deposit Subduction Post-subduction Magmatichydrothermal uid Ore formation Metals such as Cu, Mo, Au, Sn, and W in porphyry and related epithermal mineral deposits are derived predominantly from the associated magmas, via magmatichydrothermal uids exsolved upon emplacement into the mid- to upper crust. Four main sources exist for magmas, and therefore metals, in convergent and collided plate margins: the subducting oceanic plate basaltic crust, subducted seaoor sediments, the asthenospheric mantle wedge between the subducting and overriding plates, and the upper plate lithosphere. This paper rstly examines the source of normal arc magmas, and concludes that they are predominantly derived from partial melting of the metasomatized mantle wedge, with possible minor contributions from subducted sediments. Although some metals may be transferred from the subducting slab via dehydration uids, the bulk of the metals in the resultant magmas are probably derived from the asthenospheric mantle. The most important contributions from the slab from a metallogenic perspective are H 2 O, S, and Cl, as well as oxidants. Partial melting of the subducted oceanic crust and/or sediments may occur under some restricted conditions, but is unlikely to be a widespread process (in Phanerozoic arcs), and does not signicantly differ metallogenically from slab-dehydration processes. Primary, mantle-derived arc magmas are basaltic, but differ from mid-ocean ridge basalt in having higher water contents (~10× higher), oxidation states (~2 log f O2 units higher), and concentrations of incompatible elements and other volatiles (e.g., S and Cl). Concentrations of chalcophile and siderophile metals in these partial melts depend critically on the presence and abundance of residual sulde phases in the mantle source. At relatively high abundances of suldes thought to be typical of active arcs where f S2 and f O2 are high (magma/sulde ratio = 10 2 10 5 ), sparse, highly siderophile elements such as Au and PGE will be retained in the source, but magmas may be relatively undepleted in abundant, moderately chalcophile elements such as Cu (and perhaps Mo). Such magmas have the potential to form porphyry Cu ± Mo deposits upon emplacement in the upper crust. Gold-rich porphyry deposits would only form where residual sulde abundance was very low (magma/sulde ratio N 10 5 ), perhaps due to unusually high mantle wedge oxidation states. In contrast, some porphyry Mo and all porphyry SnW deposits are associated with felsic granitoids, derived primarily from melting of continental crust during intra-plate rifting events. Nevertheless, mantle-derived magmas may have a role to play as a heat source for anatexis and possibly as a source of volatiles and metals. In post-subduction tectonic settings Tulloch and Kimbrough, 2003, such as subduction reversal or migration, arc collision, continentcontinent collision, and post-collisional rifting, a subducting slab source no longer exists, and magmas are predominantly derived from partial melting of the upper plate lithosphere. This lithosphere will have undergone signicant modication during the previous subduction cycle, most importantly with the introduction of large volumes of hydrous, mac (amphibolitic) cumulates residual from lower crustal differentiation of arc basalts. Small amounts of chalcophile and siderophile element-rich suldes may also be left in these cumulates. Partial melting of these subduction-modied sources due to post-subduction thermal readjustments or asthenospheric melt invasion will generate small volumes of calc-alkaline to mildly alkaline magmas, which may redissolve residual suldes. Such magmas have the potential to form Au-rich as well as normal Cu ± Mo porphyry and epithermal Au systems, depending on the amounts of sulde present in the lower crustal source. Alkalic-type epithermal Au deposits are an extreme end-member of this range of post-subduction deposits, formed from subduction-modied mantle sources in extensional or transtensional environments. Ore formation in porphyry and related epithermal environments is critically dependent on the partitioning of metals from the magma into an exsolving magmatichydrothermal uid phase. This process occurs most efciently at depths greater than ~6 km, within large mid- to upper crustal batholithic complexes fed by arc or post-subduction magmas. Under such conditions, metals will partition efciently into a single-phase, supercritical aqueous uid (~213 wt.% NaCl equivalent), which may exist as a separate volatile plume or as bubbles entrained in buoyant magma. Focusing of upward ow of bubbly magma and/or uid into the apical regions of the batholithic complex forms cupolas, which represent high mass- and heat-ux channelways Ore Geology Reviews 40 (2011) 126 Tel.: +1 780 492 3430; fax: +1 780 492 2030. E-mail address: [email protected]. 0169-1368/$ see front matter © 2011 Elsevier B.V. All rights reserved. doi:10.1016/j.oregeorev.2011.05.006 Contents lists available at ScienceDirect Ore Geology Reviews journal homepage: www.elsevier.com/locate/oregeorev

Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Embed Size (px)

Citation preview

Page 1: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Ore Geology Reviews 40 (2011) 1–26

Contents lists available at ScienceDirect

Ore Geology Reviews

j ourna l homepage: www.e lsev ie r.com/ locate /oregeorev

Review

Magmatic to hydrothermal metal fluxes in convergent and collided margins

Jeremy P. Richards ⁎Dept. of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta, Canada T6G 2E3

⁎ Tel.: +1 780 492 3430; fax: +1 780 492 2030.E-mail address: [email protected].

0169-1368/$ – see front matter © 2011 Elsevier B.V. Aldoi:10.1016/j.oregeorev.2011.05.006

a b s t r a c t

a r t i c l e i n f o

Article history:Received 1 December 2010Received in revised form 18 May 2011Accepted 19 May 2011Available online 27 May 2011

Keywords:Porphyry depositEpithermal depositSubductionPost-subductionMagmatic–hydrothermal fluidOre formation

Metals such as Cu, Mo, Au, Sn, and W in porphyry and related epithermal mineral deposits are derivedpredominantly from the associatedmagmas, viamagmatic–hydrothermal fluids exsolved upon emplacement intothemid- to upper crust. Fourmain sources exist formagmas, and thereforemetals, in convergent and collided platemargins: the subducting oceanic plate basaltic crust, subducted seafloor sediments, the asthenospheric mantlewedge between the subducting and overriding plates, and the upper plate lithosphere. This paper firstly examinesthe source of normal arc magmas, and concludes that they are predominantly derived from partial melting of themetasomatized mantle wedge, with possible minor contributions from subducted sediments. Although somemetals may be transferred from the subducting slab via dehydration fluids, the bulk of the metals in the resultantmagmas are probably derived from the asthenospheric mantle. The most important contributions from the slabfromametallogenicperspective areH2O, S, andCl, aswell as oxidants. Partialmeltingof the subductedoceanic crustand/or sediments may occur under some restricted conditions, but is unlikely to be a widespread process (inPhanerozoic arcs), and does not significantly differ metallogenically from slab-dehydration processes.Primary, mantle-derived arc magmas are basaltic, but differ frommid-ocean ridge basalt in having higher watercontents (~10× higher), oxidation states (~2 log fO2 units higher), and concentrations of incompatible elementsand other volatiles (e.g., S and Cl). Concentrations of chalcophile and siderophile metals in these partial meltsdepend critically on the presence and abundance of residual sulfide phases in the mantle source. At relativelyhigh abundances of sulfides thought to be typical of active arcs where fS2 and fO2 are high (magma/sulfideratio=102–105), sparse, highly siderophile elements such as Au and PGE will be retained in the source, butmagmas may be relatively undepleted in abundant, moderately chalcophile elements such as Cu (and perhapsMo). Suchmagmas have the potential to form porphyry Cu±Modeposits upon emplacement in the upper crust.Gold-rich porphyry deposits would only form where residual sulfide abundance was very low (magma/sulfideratio N105), perhaps due to unusually high mantle wedge oxidation states.In contrast, some porphyry Mo and all porphyry Sn–W deposits are associated with felsic granitoids, derivedprimarily from melting of continental crust during intra-plate rifting events. Nevertheless, mantle-derivedmagmas may have a role to play as a heat source for anatexis and possibly as a source of volatiles and metals.In post-subduction tectonic settings Tulloch and Kimbrough, 2003, such as subduction reversal or migration, arccollision, continent–continent collision, and post-collisional rifting, a subducting slab source no longer exists,andmagmas are predominantly derived from partialmelting of the upper plate lithosphere. This lithospherewillhave undergone significant modification during the previous subduction cycle, most importantly with theintroduction of large volumes of hydrous, mafic (amphibolitic) cumulates residual from lower crustaldifferentiation of arc basalts. Small amounts of chalcophile and siderophile element-rich sulfidesmay also be leftin these cumulates. Partial melting of these subduction-modified sources due to post-subduction thermalreadjustments or asthenospheric melt invasion will generate small volumes of calc-alkaline to mildly alkalinemagmas, which may redissolve residual sulfides. Such magmas have the potential to form Au-rich as well asnormal Cu±Moporphyry and epithermal Au systems, depending on the amounts of sulfide present in the lowercrustal source. Alkalic-type epithermal Au deposits are an extreme end-member of this range of post-subductiondeposits, formed from subduction-modified mantle sources in extensional or transtensional environments.Ore formation in porphyry and related epithermal environments is critically dependent on the partitioning ofmetals from the magma into an exsolving magmatic–hydrothermal fluid phase. This process occurs mostefficiently at depths greater than ~6 km, within large mid- to upper crustal batholithic complexes fed by arc orpost-subduction magmas. Under such conditions, metals will partition efficiently into a single-phase,supercritical aqueous fluid (~2–13 wt.% NaCl equivalent), which may exist as a separate volatile plume or asbubbles entrained in buoyant magma. Focusing of upward flow of bubbly magma and/or fluid into the apicalregions of the batholithic complex forms cupolas, which represent high mass- and heat-flux channelways

l rights reserved.

Page 2: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

2 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

towards the surface. Cupolas may be self-organizing to the extent that once formed, further magma and fluidflow will be enhanced along the weakened and heated axes. Cupolas may form initially as breccia pipes byvolatile phase (rather than magma) reaming-out of extensional structures in the brittle cover rocks, to befollowed immediately by magma injection to form cylindrical plugs or dikes.Cupola zonesmay extend to surface,wheremagmas andfluids vent as volcanic products and fumaroles. Betweenthe surface and the underlyingmagma chamber, a very steep thermal gradient exists (700°–800 °C over b5 km),which is the primary cause of vertical focusing of oremineral deposition. The bulk ofmetals (Cu±Mo±Au) thatformsporphyry ore bodies areprecipitatedover anarrowtemperature interval between~425° and320 °C,whereisotherms in the cupola zone rise to within ~2 km of the surface. Over this temperature range, four importantphysical and physicochemical factors act to maximize ore mineral deposition: (1) silicate rocks transition fromductile to brittle behavior, thereby greatly enhancing fracture permeability and enabling a threefold pressuredrop; (2) silica shows retrograde solubility, thereby further enhancing permeability and porosity for oredeposition; (3) Cu solubility dramatically decreases; and (4) SO2 dissolved in the magmatic–hydrothermal fluidphase disproportionates to H2S and H2SO4, leading to sulfide and sulfate mineral deposition and the onset ofincreasingly acidic alteration.The bulk of the metal flux into the porphyry environment may be carried by moderately saline supercriticalfluids or vapors, with a volumetrically lesser amount by saline liquid condensates. However, these vapors rapidlybecome dilute at lower temperatures and pressures, such that they lose their capacity to transport metals aschloride complexes. They retain significant concentrations of sulfur species, however, and bisulfide complexingof Cu and Au may enable their continued transport into the epithermal regime. In the high-sulfidationepithermal environment, intense acidic (advanced-argillic) alteration is caused by the flux of highly acidicmagmatic volatiles (H2SO4, HCl) in this vapor phase. Ore formation, however, is paragenetically late, andmay belocated in these extremely altered and leached cap rocks largely because of their high permeability and porosity,rather than there being any direct genetic connection. Ore-forming fluids, where observed, are low- tomoderate-salinity liquids, and are thought to represent later-stage magmatic–hydrothermal fluids that haveascended along shallower (cooler) geothermal gradients that either do not, or barely, intersect the liquid–vaporsolvus. Such fluids “contract” from the original supercritical fluid or vapor to the liquid phase. Brief intersectionof the liquid–vapor solvus may be important in shedding excess chloride and chloride-complexed metals (suchas Fe), so that bisulfide-complexed metals remain in solution. Such a restrictive pressure–temperature path islikely to occur only transiently during the evolution of amagmatic–hydrothermal system,whichmay explain therarity of high-sulfidation Cu–Au ore deposits, despite the ubiquitous occurrence of advanced-argillic alteration inthe lithocaps above porphyry-type systems.

© 2011 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32. Magma generation in convergent and collided margins: geochemical characteristics and partitioning of metals . . . . . . . . . . . . . . . . . . 3

2.1. Slab dehydration and asthenospheric melting in subduction zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32.1.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2. Sediment dehydration and/or melting in subduction zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52.2.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

2.3. Oceanic slab melting in subduction zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62.3.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.4. Supra-subduction zone lithospheric melting: the MASH process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62.4.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82.4.2. Sources of Mo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.5. Lithospheric melting during post-subduction events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82.5.1. Behavior of metals in subduction-modified sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

2.6. Crustal melting during post-collisional stress relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102.6.1. Sources of metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

3. Behavior of metals during magma fractionation and fluid exsolution in the upper crust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113.1. Partitioning of metals from magma into exsolving hydrothermal fluid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

4. Magmatic–hydrothermal ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134.1. Porphyry Cu ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134.2. Epithermal Cu–Au ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

4.2.1. High-sulfidation epithermal Cu–Au deposits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154.2.2. Low-sulfidation epithermal Au deposits (including alkalic-type deposits) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

5. Summary and conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185.1. Sources of magmas and metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185.2. Porphyry and epithermal ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

Page 3: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

3J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

1. Introduction

The question of the source of various elements in convergent andcollidedmarginmagmas has challenged geologists for decades. Igneouspetrologists seek to understand the petrogenesis of such magmasthrough geochemical and isotopic tracing,whereas economic geologistsare generally more interested in the source of potentially valuableelements suchasCu,Mo, Sn,W,Au, andplatinumgroupelements (PGE),which may ultimately be found in intrusion-related hydrothermaldeposits.

Igneous petrologists are broadly in agreement that arc magmas areprimarily derived from hydrous melting of the asthenospheric mantlewedge above subducting plates, but melts from the subducted oceaniccrust (including sediments) and the upper plate lithosphere may alsobe involved to varying degrees.

Economic geologists are also broadly in agreement that ore-formingelements are partitioned from such magmas into an exsolving volatilephase upon emplacement in the upper crust, and may then beprecipitated from thesefluids during cooling, fluidmixing, andwallrockreaction processes in porphyry-type and related epithermal mineraldeposits. However, these process theories do not address where themetals originally came from, nor why porphyry deposits vary sowidelyin their metal contents (from Au-rich, through Cu±Mo±Au, to Mo-only deposits, with Sn–W deposits forming a distinct variant).

In addition to subduction-related calc-alkaline magmas, a diversesuite of calc-alkaline to alkaline magmas is generated in post-subduction and collisional tectonic settings, and these magmaticsystems may also generate porphyry and epithermal ore deposits.Such systems raise an additional set of petrogenetic and metallogenicquestions.

It is the intent of this paper to merge these different geologicalperspectives on magmagenesis and metallogeny in order to discussprimary metal fluxes in convergent and collisional margins in terms ofigneous petrogenetic and magmatic–hydrothermal processes. Theultimate metal inventory and metal ratios in any given porphyry orrelated deposit is secondarily controlled by late-stage magmatic andshallow crustal processes. These processes are examined, closingwith areview of fluid and metal sources and behavior in related epithermalenvironments.

2. Magma generation in convergent and collided margins:geochemical characteristics and partitioning of metals

Most magmas erupted through or emplaced within the Earth's crustare not primary magmas (in the sense of being chemically unmodifiedsince extraction from their source), andmost are not even primitive (inthe sense of being relatively unevolved; Hildreth and Moorbath, 1988;Leeman, 1983; Neuendorf et al., 2005; Smith et al., 2010; Thirlwall et al.,1996). Except formagmasproduced and erupted in extensional tectonicregimes (where rapid ascent to the surface is facilitated by normalfaulting), most deeply-derived magmas undergo some degree offractionation and crustal contamination during their passage towardsthe surface. It is therefore challenging to isolate geochemical andisotopic characteristics ofmagma source regions from the effects of laterprocesses (Davidson, 1996). Magmas erupted through mature conti-nental crust are the most difficult to fingerprint uniquely in terms ofsource characteristics because wallrock assimilation and fractionalcrystallization (AFC; DePaolo, 1981) are ubiquitous and commonlyextensiveprocesses thatwill significantlymodify bulk rock geochemicaland isotopic compositions; and yet, these are also the magmas that aremost commonly associated with porphyry- and epithermal-typemineral deposits. The difficulty in constraining source characteristicsin such magmas is perhaps responsible for the plethora of theories thathave beenproposed for the origin of ore-formingmagmas in convergentmargin settings, ranging from the melting of subducting oceanic crustand/or seafloor sediments, through melting of subduction metasoma-

tized asthenospheric or lithospheric mantle, to melting of underplatedor primitive lower crustal rocks, and even melting of evolved crustalrocks in the case of some felsic porphyry Mo and Sn–W magmas.

Therefore, rather than start by trying to identify a unique sourcefor the typical intermediate-to-felsic calc-alkaline magmas that areassociated with ore deposits in mature convergent margins, I beginthis review by focusing on the much better constrained primitiveisland arc environment, where the effects of fractionation and crustalcontamination, particularly by continentally derived materials, areminimized, and processes in mantle source regions can be moreclearly defined.

2.1. Slab dehydration and asthenospheric melting in subduction zones

There is a general consensus that, with the exception of youngoceanic lithosphere (b25 m.y.-old; Defant and Drummond, 1990) orplate edges (Yogodzinski et al., 2001), basaltic oceanic crust undergoeslow-temperature, high pressure metamorphism upon subduction,which releasesfluids through a series of prograde dehydration reactionsto form anhydrous eclogite (Fig. 1). Water and other volatilecomponents and solutes (including S and Cl) were originally incorpo-rated into oceanic crustal and upper mantle rocks during oxidizingseafloor alteration, generatinghydrousminerals suchas serpentine, talc,amphibole, micas, chlorite, zoisite, chloritoid, and lawsonite. Variousexperimental studies have shown that these minerals undergodehydration reactions over a depth range extending to ~100 km,corresponding to the blueschist–eclogite transition in crustal rocks;serpentine (antigorite) and the 10 Å equivalent of chlorite may extendthis range to200 km(e.g., Dvir et al., 2011; Forneris andHolloway, 2003;Fumagalli and Poli, 2005; Poli and Schmidt, 2002; Schmidt and Poli,1998; Ulmer and Trommsdorff, 1995). Below these depths, theanhydrous eclogitic crust is essentially infusible, and the dense slabcontinues its descent into the mantle without melting. See reviews ofthis subject by Richards (2003, 2005, and references therein).

Numerous studies have explored the character of the fluids that arereleased from the dehydrating slab, because they are thought to accountfor the unique geochemical character of subduction-related magmasduring later partial melting in the metasomatized asthenosphericmantle wedge (located between the downgoing slab and the upperplate). Slab-derived fluids are thought to be water-rich at these depths,and to carry significant amounts of other volatile components such as Cland S. For example, salinities in the range of 4–10 wt.% NaCl have beeninferred from primitive basalt or melt inclusion studies (de Hoog et al.,2001; Kent et al., 2002; Portnyagin et al., 2007; Wallace, 2005;Wysoczanski et al., 2006), and salinities of 0.4–2 wt.% NaCl equivalentwere measured in fluid inclusions from high-pressure rocks thought torepresent subductedoceanicmantle (Scambelluri et al., 2004). At higherpressures (~6 GPa) and greater depths (~175 km) there may no longerbe a physical distinction between solute-rich aqueous liquids andhydrous silicate melts, and the fluid may be supercritical in nature (e.g.,Kessel et al., 2005a,b), but the role of such deeply released fluids insubduction zone magmatism is unclear (see discussion in Richards andKerrich, 2007).

In addition to volatiles, water-soluble large ion lithophile elements(LILE: K, Rb, Cs, Ca, Sr, andBa, andU), andB, Pb, As, andSb (Breeding et al.,2004; Hattori et al., 2005; Hattori and Guillot, 2003; Kogiso et al., 1997;Manning, 2004; Tatsumi et al., 1986) are thought to bemobilized into theforearc mantle wedge by dehydration fluids, which may then beconvected by corner-flow into the sub-arc melting zone (Fig. 1). Thesefluid–mobile components are also characteristically enriched in arcmagmas (e.g., typical ranges of: 1–3 wt.% H2O, 500–2000 ppm Cl, 900–2500 ppm S; Davidson, 1996; Gill, 1981; Noll et al., 1996; Sobolev andChaussidon, 1996;Wallace, 2005; Portnyagin et al., 2007),which is takenas evidence of aqueousfluidmetasomatism of themantlewedgemagmasource. Silica may also be significantly mobilized in these slab fluids(Aerts et al., 2010;Manning, 2004), aswell as Tl and Cu (Noll et al., 1996;

Page 4: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Mantle corner flow

Sea level

Asthenosphere

Volcanic arc

Talc

Serpentine

Chlorite +serpentine

Chlorite

1000°C

600°C

Oce

an

ic m

an

tle

lith

osp

he

reOceanic crust

Dehydration of oceanic

lithosphere1000°C

Amphibole

ZoisiteCtd

Oceanic crust

Partial melting

Mantlelithosphere

Chlorite–10Å phase

+ serpentine

0 km

100 km

200 km

1400°C

1400°C

600°C

1000°C

Eclogite

Sediment

Metasomatizedasthenosphere

Fig. 1. Structure and processes beneath an oceanic island arc (sources: Tatsumi and Eggins, 1995; Schmidt and Poli, 1998; Winter, 2001; Poli and Schmidt, 2002; Fumagalli and Poli,2005). Primary hydrous basaltic arc magmas are derived from partial melting of the metasomatized asthenospheric mantle wedge. Mineral zones shown in the subducting plateindicate lower limits of stability of hydrous phases in the basaltic oceanic crust and peridotitic mantle lithosphere. Abbreviation: Ctd=chloritoid.

4 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Stolper and Newman, 1994). The normally relatively incompatible highfield strength elements (HFSE) Ti, Nb, and Ta are not significantly fluidsoluble under subduction zone conditions, and are retained in mineralssuch as rutile either in the slab or the mantle wedge (Audétat andKeppler, 2005; Brenan et al., 1994; Green and Adam, 2003). Arc magmasderived from these sources therefore show characteristic negativeanomalies for these three elements on mantle-normalized spiderdiagrams, but display enrichments in most other incompatible elements(Foley et al., 2000; Gill, 1981; Klemme et al., 2005; Ryerson andWatson,1987; Schmidt et al., 2004; Schmidt et al., 2009).

Hydrous metasomatism of the mid-ocean-ridge basalt (MORB)-depleted asthenospheric mantle wedge causes partial melting bylowering the solidus of peridotite (Arculus, 1994; Kushiro et al., 1968;Stolper andNewman, 1994). This occurs either throughdirect infiltrationmetasomatism by slab-derived fluids percolating into the hot inner zoneof themantlewedge (Bourdonet al., 2003;Groveet al., 2006;Kelley et al.,2010; Peacock, 1993), or by convective corner-flow mixing of metaso-matized peridotite into these hotter central regions (Fig. 1; Schmidt andPoli, 1998; Tatsumi, 1986; Wysoczanski et al., 2006).

Partial melting of hydrated peridotite under these conditions inthe mantle wedge generates high-Mg basalts (Greene et al., 2006;Pichavant et al., 2002; Smith et al., 2010). Such arc basalts aredistinguished from MORB by higher contents of incompatibleelements (as noted above) and water (up to 6 wt.% H2O; Cervantesand Wallace, 2003; Grove et al., 2003; Pichavant et al., 2002; Sobolevand Chaussidon, 1996). Critically, Hamada and Fujii (2008) andZimmer et al. (2010) report that a water content of 2 wt.% separates“dry” tholeiitic (olivine+plagioclase/orthopyroxene) from “wet”calc-alkaline (clinopyroxene+magnetite) magmatic fractionationtrends.

Arc basalts are also characterized by distinctly higher oxidationstates than MORB (up to 2 log units above the fayalite–magnetite–quartz buffer: ΔFMQ+2; Ballhaus, 1993; Brandon and Draper, 1996;

Parkinson and Arculus, 1999; Rowe et al., 2009). The relatively highoxidation state of arc magmas is a critical factor in their subsequentmetallogeny, and originates from oxidative seafloor alteration of theoceanic plate (Staudigel et al., 1996), transmitted into themantlewedgeby the metasomatic fluid flux (Brandon and Draper, 1996; Kelley andCottrell, 2009; Malaspina et al., 2009).

2.1.1. Behavior of metalsMost base and precious metals would be expected to have at least

moderate solubilities in the hot, relatively oxidized, saline aqueousfluids exsolved from the downgoing slab. In particular, as noted inSection 2.1, Pb, As, and Sb are stronglymobilized by such fluids, possiblyalong with Tl and Cu (Noll et al., 1996). The behavior of highlysiderophile elements (HSE) such asAu andPGE is lesswell knownunderthese conditions, but studies of metasomatized mantle xenoliths fromisland arc lavas suggest that Au, Re, and the Pd-group elements(including Pt) are mobilized into the mantle wedge during subductionmetasomatism (Dale et al., 2009; Kepezhinskas et al., 2002; McInnes etal., 1999; Sun et al., 2004a; Widom et al., 2003).

The volumetric extent and efficiency of mobilization of metals intothemantlewedge by this process are unknown, butfluidmetasomatismclearly represents one viable mechanism for metal transfer into arcmagma sources.

The behavior of chalcophile and siderophile metals during subse-quent partial melting of the metasomatized mantle wedge dependscritically on oxidation state (fO2) and sulfur fugacity (fS2), because theseparameters control the stability and abundance of sulfide phases. Goldand PGE partition strongly into sulfide phases relative to silicate melts,but Cu to a somewhat lesser degree (Campbell and Naldrett, 1979;Peach et al., 1990), so if sulfides are abundant in the magma sourceregion, partial melts will be depleted in Au and PGE relative to Cu(Fig. 2). Under the high fO2 and fS2 conditions of the supra-subductionzonemantlewedge, thebulk of the sulfurfluxwill likely consist of SO2 or

Page 5: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

5J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

sulfate, dissolved first in slab fluids and then magma (Carroll andRutherford, 1985; Jenner et al., 2010; Jugo et al., 2005a). However,because of the equilibria betweenvarious sulfur species, at high fS2 somecondensed sulfide phases will likely also be present (McInnes et al.,2001). Consequently, Richards (2009) suggested that normal arcmagmas will be minimally depleted in Cu (due to its higher abundanceand moderate chalcophile nature) relative to sparse, highly siderophileelements suchasAu and PGE,whichwill be strongly retained in residualsulfide phases in the source region (e.g., Hamlyn et al., 1985; MitchellandKeays, 1981; Peachet al., 1990). Thismay explain theCu-rich natureof typical arc-related porphyry deposits (relative to Au and PGE;Richards, 2005). In contrast, Au-rich porphyry deposits may requireatypical subduction-related or collisional tectonic settings and petro-genetic processes, which act to destabilize residual sulfide phases andrender Au incompatible (e.g., Jégo et al., 2010; Richards, 1995, 2009;Sillitoe, 2000; Solomon, 1990; Wyborn and Sun, 1994; see Section 2.5).The low abundances of PGE in many arc-related ore deposits suggest afurther separation of these elements from Au and Cu, perhaps throughthe formation of residual platinoid alloy phases (e.g., Barnes et al., 1985;Borisov and Palme, 1997; Kepezhinskas et al., 2002; Peach et al., 1990)or Cr-spinels (into which Ir-group PGE strongly partition; Hattori et al.,2010; Righter et al., 2004).

2.2. Sediment dehydration and/or melting in subduction zones

Seafloor sediments on the surface of the downgoing slab areanother potential source of metasomatic contributions to the mantlewedge. Much of this sedimentary material will be scraped off at thetrench to form an accretionary prism, but varying amounts may alsobe subducted, depending on the degree of coupling between theupper and lower plates, and also the sediment input load (Fig. 1). Suchsediments will be water-rich and pelitic in bulk composition, and thusare more likely to undergo partial melting under subduction zoneconditions than basaltic oceanic crust (Hermann and Spandler, 2008).

Au

(ppb

)C

u (p

pm)

R = (mass of silicate melt)/(mass of sulfide)

Cu maximized inmagma (R ≥ 103)

Au maximized inmagma (R ≥ 106)

0.1

1

10

104

105

106

10–5

1 10 104 105 106 107 108

Cu in sulfide

Cu in magma

Au in sulfide

Au in magma

Sulfide/silicate melt partitioncoefficients:D(Cu) = 103

D(Au) = 105

Metal concentrations inmagma in absence of sulfide:Cu = 50 ppmAu = 5 ppb

Au depleted in magma

Au enriched in

residual sulfide

Porphyry Cupotential arc

magmas

Porphyry Cu-Aupotentialmagmas

102 103

102

103

10–4

10–3

10–2

5 ppb Au

50 ppm Cu

Fig. 2. Concentrations of Cu and Au in silicate magma and coexisting sulfide liquid as afunction of R=[mass of silicate melt]/[mass of sulfide melt] (Campbell and Naldrett,1979; diagram modified from Richards, 2005, 2009). At R-factors below ~102, magmaswill be depleted in both Cu and Au. At R-factors between ~102–105, magmas will bedepleted in Au but essentially undepleted in Cu (porphyry Cu-potential magmas). At R-factors N105, magmas will be undepleted in Au and Cu (porphyry Cu–Au–potentialmagmas). A corollary of this diagram is that arc magmatismwill leave small amounts ofrelatively Au-rich sulfide in the mantle source or lithosphere during fractionation,which can be remelted during post-subduction tectonomagmatic processes, to formsmall-volume, alkaline, porphyry Au-potential magmas.

Nevertheless, Aizawa et al. (1999); Dreyer et al. (2010); and Duggenet al. (2007) have suggested that dehydration is the principal processaffecting sediments down to depths of ~100 km (i.e., to below thevolcanic arc), with melting only occurring significantly at greaterdepths when temperatures exceed ~800 °C (possibly reflected in thegeochemistry of some back-arc magmas).

Sediment contributions to the source of arc magmas have been thesubject of numerous studies, with the least ambiguous evidencecoming from island arcs (e.g., MacDonald et al., 2000; Thirlwall et al.,1996; Wysoczanski et al., 2006). In continental arcs, it can be difficultto distinguish between chemical and isotopic signatures fromsubducted continent-derived sediment versus crustal contaminationduring magma ascent (e.g., Hildreth and Moorbath, 1988; Kemp et al.,2007): both sources will contribute incompatible elements andcrustal isotopic values to primary mantle-derived arc magmas(Breeding et al., 2004).

Trace elements commonly used as indicators of sediment contribu-tions to arcmagmas are Ba, B, Be, Th, andPb (Dreyer et al., 2010; Johnsonand Plank, 1999), and Ba/La and Th/La ratios can be used as ameasure ofsediment versusmantle source components (Plank, 2005;Walker et al.,2001). In particular, the cosmogenic radioisotope 10Be can be used as atracer of recent (b10 Ma) introduction of sediments into arc magmasources (Dreyer et al., 2010; Morris et al., 1990). However, although aclear sediment-derived isotopic signature can be observed in manyisland arc systems, the volumetric contribution of sediments to islandarc magmas seems to be relatively minor (Hawkesworth et al., 1994;Kilian and Behrmann, 2003; Poli and Schmidt, 2002; Stern et al., 2006).

One additional element that may be added to the mantle wedgefrom subducted sediments is sulfur, as suggested by the positive δ34Scompositions of arc magmas (de Hoog et al., 2001), which are similarto those of seafloor sediments (Alt et al., 1993). Analyses of glassinclusions in olivine from primitive arc magmas reveal concentrationsof up to 2900 ppm S (de Hoog et al., 2001), and Jugo et al. (2005b)measured experimental concentrations of up to 1.5 wt.% S in oxidizedarc basalts. These high sulfur contents have great significance for thebehavior of chalcophile and siderophile metals (see Sections 2.1.1,2.5.1, and 3).

2.2.1. Behavior of metalsLead, which is significantly enriched in the continental crust (and

crustally-derived sediments) relative to themantle, is the onlymetal forwhich a clear sedimentary source can be inferred in some island arcmagmas, because it can be readily identified by its radiogenic isotopiccomposition compared with depleted mantle sources. However, incontinental arcs, distinguishing a subducted sediment source ofradiogenic Pb from crustal contamination during magma ascent isvery difficult (e.g., Barreiro, 1984; Chiaradia et al., 2004; Kontak et al.,1990). This has led to diverging opinions: for example, Aitcheson et al.(1995);Hildreth andMoorbath (1988); James (1982); Kay et al. (1999);and Tilton et al. (1981) concluded that the bulk of the radiogenic Pb incentral Andean magmas comes from upper plate crustal sources,whereas Macfarlane (1999); McNutt et al. (1979); Mukasa et al.(1990); and Sillitoe and Hart (1984) preferred a subducted sedimentsource for Pb in some Andean igneous rocks and ore deposits.

Inferring a seafloor sediment source for othermetallic componentsin arc magmas (and related ore deposits) is much more speculative,and generally involves the subduction of metal-rich manganesenodules or even massive sulfide deposits. The latter, however, likelyoxidize and disperse geologically rapidly after formation on theseafloor (e.g., Edwards, 2004; Herzig et al., 1991) unless quicklyburied by lava; theymight thus only be expected to be subductedwithvery young oceanic crust. Few studies have specifically invoked asubducted sediment source for ore metals other than a component ofPb, and the majority of authors have concluded that such a source iseither unnecessary or unproven (e.g., Burnham, 1981; Chiaradia et al.,2004; de Hoog et al., 2001; Fontboté et al., 1990).

Page 6: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

6 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

2.3. Oceanic slab melting in subduction zones

The question of whether the downgoing oceanic slab, or morespecifically the basaltic oceanic crust, melts during subduction hasprompted lively debate recently, not only in the petrology literature(e.g., Conrey, 2002; Defant and Kepezhinskas, 2001, 2002; Garrison andDavidson, 2003), but also amongst economic geologists because of thesuggestion that “slabmelts”might in somewaybeuniquely fertile for laterporphyry ore formation (e.g., Mungall, 2002; Oyarzun et al., 2001, 2002;Sajona and Maury, 1998; Thiéblemont et al., 1997; for contrary opinions,see: Rabbia et al., 2002; Richards, 2002; Richards and Kerrich, 2007).

The idea that the hydrated basaltic ocean crust might melt duringsubduction was an early assumption of the plate tectonic revolution,because it seemed conveniently to explain the relatively felsiccomposition of arc magmas (as opposed to basalts formed by meltingof peridotitic mantle). The theory was given credence by theexperiments of Rapp et al. (1991) and Rapp and Watson (1995), whoshowed that melting of amphibolite under upper mantle conditions(1025 °C and 0.8–1.6 GPa) could generate an intermediate compositiontonalite–trondhjemitemelt, not dissimilar to anarc andesite. Defant andDrummond (1990) termed the products of subducted slab melting“adakites”, after a single anomalous lava flow on Adak Island in theAleutians described by Kay (1978). Because garnet should be present inthe eclogitic source of these magmas, Defant and Drummond (1990)argued that such melts could be distinguished by anomalously lowconcentrations of heavy rare earth elements (HREE) and Y (which arecompatiblewithgarnet) relative to light rare earthelements (LREE), andhigh concentrations of Sr (because of the absence of plagioclase at suchdepths). Thus, slab melts, or adakites, could be distinguished on Sr/Yversus Y, or La/Yb versus Yb diagrams from normal arc magmas formedin the absence of garnet.

However, despite the theoretical possibility of melting subductedoceanic crust, most thermal models of subduction zones indicate thattemperatures in the slab do not normally reach melting conditions(N800 °C) prior to dehydration and eclogitization (Fig. 1), which wouldrender the slab infusible (e.g., Davies and Stevenson, 1992; Peacock,1996; Poli and Schmidt, 2002). Thus, Defant and Drummond (1990)proposed that slab melting might be restricted to the subduction ofyoung (≤25 m.y. old) and therefore still hot oceanic crust, and Peacocket al. (1994) were even more restrictive (b5 m.y. old). Other relativelyuncommon scenarios that might result in slab melting include shallowor stalled subduction (whereby the slab has more time to heat up atshallow depths; Gutscher et al., 2000; Peacock et al., 1994), ridgesubduction (Guivel et al., 2003; Kay et al., 1993), or edge-melting ofdetached slabs or slab windows (Haschke and Ben-Avraham, 2005;Thorkelson and Breitsprecher, 2005; Yogodzinski et al., 2001).

A further complication is added by the fact that most adakitesdescribed in the petrology literature are not in fact primary slabmelts,but are substantially evolved, having reacted or hybridized with theasthenosphere during ascent (and likely also the upper platelithosphere). This modification to the adakite slab-melting model isrequired to explain the high contents of MgO, Ni, and Cr present insome adakites relative to expected levels for hydrated basalt partialmelts (Defant and Kepezhinskas, 2001; Drummond et al., 1996;Martin, 1999; Martin et al., 2005; Yogodzinski et al., 1995).

Direct evidence for slabmelting is lacking, but supra-subduction zonexenoliths from the Tabar-Lihir-Tanga-Feni (Papua New Guinea), Philip-pine, and Patagonian arcs preserve hydrous, silica-rich glass inclusionsthat are thought to represent migrating slab melts (respectively: Kilianand Stern, 2002; McInnes and Cameron, 1994; Schiano et al., 1995). Theglass inclusions characterized by Schiano et al. (1995) were calc-alkalinein composition, with high incompatible element and low Ti, Nb, and Ycontents, high LREE/HREE ratios, and homogenization temperatures of~920 °C. They thus compositionally resemble melts that would bepredicted to form from slabmelting, and appear to provide evidence thatthis process occurs at least locally where conditions permit.

The chemical difference between slab dehydration and slabmelting would seem to be rather small, given that both mediawould be enriched in volatiles, incompatible elements, and silica.Indeed, as noted in Section 2.1, there may well be a continuumbetween silica-rich aqueous fluids and aqueous silicate melts atgreater depths in subduction zones (Kawamoto, 2006; Kessel et al.,2005a,b; Manning, 2004; Portnyagin et al., 2007). This likely explainswhy the debate between slab melting and dehydration is somewhatintractable, and mostly hinges on subtle trace element characteristics.

2.3.1. Behavior of metalsSlab melts are predicted to be volatile-rich (including H2O, S, and

Cl) and oxidized, and thus, like hydrous slab fluids, would be expectedto be able to transport base and precious metals at least to somedegree. However, analyses of such metals (except iron) are notreported in most melt inclusion studies (e.g., Kilian and Stern, 2002;McInnes and Cameron, 1994; Schiano et al., 1995), so there are nodirect constraints on the capacity of suchmelts to transfer chalcophileand siderophile metals from the slab to the mantle wedge.

In a study of metasomatized mantle xenoliths from a submarinevolcano near Lihir Island, Papua New Guinea, McInnes et al. (1999)concluded that enrichments in Cu, Au, and PGEwere caused by slab fluidmetasomatism, rather thanmelts. In contrast, Kepezhinskas et al. (2002)measured the concentrations of Au and PGEs in mantle xenoliths fromthe Kamchatka arc, and suggested that a fluid-transported componentcould be distinguished froma slabmelt component by co-enrichments inPGE and high field strength elements (HFSE) in the latter, because of thelow capacity of aqueous fluids to carry HFSE. Intriguingly, they noted nosuch correlation betweenHFSE-enrichments andAu, and concluded that,whereas PGEmight be transported into themantle wedge by both fluidsand melts, Au was likely only carried by hydrous fluids.

Two theoretical studies have proposed that slab melts should beunusually effective as metal-transporting and ore-forming agents.Oyarzun et al. (2001) argued that slab melts should be unusuallyoxidized and rich in H2O and SO2 (relative to normal arc magmasderived by asthenospheric partial melting), although no evidence wasgiven for this assertion. Such magmas, they argued, should beparticularly suitable for the formation of magmatic–hydrothermalporphyry copper deposits upon emplacement in the upper crust.

Mungall (2002) presented a theoretical model for oxidation of themantle wedge by Fe3+-rich slab melts to the point of complete sulfidedestruction, thereby rendering chalcophile and siderophile elementsincompatible in mantle phases, and free to partition into silicatemelts. He argued that ferric iron is a much stronger oxidant than slab-derivedwater, and that slab melts should be rich in Fe3+ generated byoxidative seafloor alteration. Thus, slab melts might be uniquelyfavorable for the subsequent generation of metal-rich, and particu-larly Au-rich, magmas derived from the mantle wedge.

Mungall's (2002)modelmay have applicability for less commonAu-rich porphyry deposits formed in atypical subduction zone settings thatmight cause slab melting, but does not seem well suited to explainregular arc porphyry Cu deposits. In either case, metals are envisaged tobe sourced from themantle wedge, not the slab. In contrast, Oyarzun etal.'s (2001) model does not address the source of metals, and is at rootbased on the assumption that slab melts are uniquely more H2O- andSO2-rich, and more oxidized than normal arc magmas, leading tospecific ore depositional processes rather than source processes. It is notclear that these assumptions are justified, and some have argued thatslabmelts might in fact be relatively reducing because of the additionalpresence of organic-rich sediment melts (Wang et al., 2007a).

2.4. Supra-subduction zone lithospheric melting: the MASH process

Hydrous basaltic magmas generated in themantlewedgewill havetemperatures in excess of 1000 °C (Eggins, 1993; Grove et al., 2006;MacDonald et al., 2000), and perhaps as high as 1350 °C (Schmidt and

Page 7: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

7J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Poli, 1998; Tatsumi, 2003). Because their densities will be lower thanthe mantle but not the crust (Herzberg et al., 1983), they will tend torise from their asthenospheric source region and penetrate themantlelithosphere, but pool at the crust/mantle density barrier (“level ofneutral buoyancy”: Fig. 3; Fyfe, 1992; Hildreth, 1981). Here, if the fluxof magma and heat is maintained and supplemented by the latentheat of crystallization as the magma begins to crystallize, hightemperatures can be brought to bear on lower crustal rocks that willcause partial melting (Annen et al., 2006; Bergantz and Dawes, 1994;Huppert and Sparks, 1988; Klepeis et al., 2003; Petford and Gallagher,2001; Rushmer, 1993). Hildreth and Moorbath (1988) suggested thatit is the interaction between this hot, hydrous basalt flux from thesubduction zone and felsic crustal partial melts that gives rise to theuniform composition of andesites in continental volcanic arcs, by aprocess they dubbed melting–assimilation–storage–homogenization(MASH). In a refinement of this model, Annen et al. (2006) referred tothe region of magma–crust interaction as a “hot zone” (Fig. 3).Because garnet is a product of such lower crustal fractionation andpartial melting processes (Alonso-Perez et al., 2009; Berger et al.,2009; Dufek and Bergantz, 2005; Garrido et al., 2006; Hansen et al.,2002; Klepeis et al., 2003; Rushmer, 1993; Wolf and Wyllie, 1994),derivative calc-alkaline magmas may display trace element compo-sitions that resemble adakites (see Section 2.3) but which areunrelated to slab melting (Klepeis et al., 2003; Macpherson et al.,

Volcanic arc

Sea level

1000°C

Oceanicmantle

lithosphere

M

Feeder

600°C

Dehydrating oceanic crust

EHydrothermalalteration P

Fig. 3. Schematic section through a continental arc, showing the development of a MASH or “hbuoyancy, differentiate, and interact with crustal rocks and melts. Evolved, less dense, andesitbuoyancy to form batholithic complexes. Along with volcanic structures, porphyry and epithexsolvedmagmatic fluids ascend, cool, and interact with near-surface upper crustal rocks. ModAnnen et al. (2006), and Sillitoe (2010).

2006; Richards and Kerrich, 2007; Tulloch and Kimbrough, 2003).Such common processes, affecting batches of magma crystallizing andfractionating at different crustal depths (e.g., Annen et al., 2006) areentirely consistent with petrological observations in arc volcanicsystems where “adakite-like” (i.e., high-Sr/Y) andesitic lavas may beinterlayered with “normal” andesites in a single volcano, and do notrequire a fundamental change in magma source 100 km below thevolcano (e.g., Feeley and Davidson, 1994; Grunder et al., 2008;Richards et al., 2006a).

Once these hybrid magmas reach basaltic andesitic to andesiticcompositions, their densities are low enough to allow them to risethrough the lower continental crust (Herzberg et al., 1983), but theytend to stall again at a second density barrier in the middle to uppercrust below light supracrustal rocks. This is the level (5–10 km) atwhichlarge arc batholiths will form if the flux of magma is sustained, and isalso the level atwhichevolved felsicmelts andvolatiles are accumulated(Fig. 3; see Richards, 2003, and references therein). These volatiles drivebuoyant, bubbly, evolved magma upwards into the cover rocks to formsubvolcanic stocks and dikes, or explosive volcanic eruptions if theyreach surface. The volatiles may also separate from the magma flux toform a separate fluid plume, which ultimately vents at the surface(fumaroles) butmay also formporphyry- and epithermal-type depositsin the hypabyssal and near-surface environment (see discussion ofthese processes in Sections 3 and 4).

1400°C

600°C

1000°C

Continental crust

Metasomatizedasthenosphere

1400°C

Subcontinentalmantle lithosphereantle corner flow

Asthenosphere

0 km

50 km

100 km

Mid- to upper crustalbatholithic complex

Lower crustalMASH or “hot zone”

dikes

pithermal depositsorphyry deposits

Partial melting

ot zone” at the base of the crust where basaltic arc magmas pool at their level of neutralic magmas rise into the mid-to-upper crust where they pool at their new level of neutralermal deposits may form at shallower levels above these batholithic complexes whereified from Richards (2003, 2005); sources: Hildreth andMoorbath (1988), Winter (2001),

Page 8: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

8 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

2.4.1. Behavior of metalsSome of the clearest evidence for the involvement of crustal rocks in

continental arcmagmas comes fromPb isotopic data (e.g.,Wörner et al.,1992), although as noted in Section 2.2, there may be ambiguitybetween lower crustalmelting and themelting of subducted continent-derived sediments. Lead and uranium are much more abundant in thebulk continental crust (11 ppm Pb, 1.3 ppm U; Rudnick and Gao, 2003)or lower continental crust (4 ppm Pb, 0.2 ppm U; Rudnick and Gao,2003) than the primitive mantle (b0.2 ppm Pb, b0.02 ppm U; Taylorand McLennan, 1985; Sun and McDonough, 1989), MORB (0.3 ppm Pb,0.05 ppm U; Sun and McDonough, 1989), or typical low-K mafic arcandesites (b1.8 ppm Pb, b0.2 ppm U; Gill, 1981), so it only requiressmall amounts of contamination by radiogenic crustal lead tosignificantly modify a magma's Pb isotopic composition. Thus, it is notclear that a particularly large amount of Pb in arc magmas is sourcedfrom crustal rocks versus the subduction zone, and Macfarlane et al.(1990) have argued that the crustal contribution is minimal in CentralAndean magmas and ores. Moreover, porphyry-type systems aretypically not Pb-rich, except in late-stage skarns and distal veinswhere some of the Pbmay have been derived from local host rocks (e.g.,Mukasa et al., 1990).

Most researchers have assumed that, because of the higherconcentrations of Cu in primitive andesites (145 ppm; Gill, 1981)compared with the bulk continental crust (27 ppm; Rudnick and Gao,2003), the bulk of Cu in porphyry-type deposits is mantle-derived. Asimilar assumption is made for Au, although the primitive mantle andcontinental crust actually have comparable concentrations (1–3 ppbAu; Rudnick and Gao, 2003; Taylor and McLennan, 1985). Moreover,porphyry Cu–(Au) deposits are found in association with mantle-derived arc magmas worldwide, regardless of crustal type (oceanic orcontinental) or thickness (Kesler, 1973), so a crustal heritage for thesemetals does not appear to be critical. Nevertheless, a lower crustalsource, perhaps hybridized with mantle-derived magmas, has beenproposed by Bouse et al. (1999) for both magmas and metals in theLaramide porphyry systems of Arizona. Moreover, Titley (1987, 2001)has specifically suggested that Au and Ag are crustally derived in arange of ore deposits including porphyries in southwestern USA,because the ratios of these elements correlate closely with twodistinct basement domains in this region. Given that Au is notespecially enriched in the mantle (see above), and that Ag is in factmore abundant in the crust than the mantle (80 ppb in the bulkcontinental crust, versus b19 ppb in the primitive mantle; Taylor andMcLennan, 1985), a crustal source for at least some proportion ofthese minor metals, and especially Ag, in arc magma-related systemsmay be reasonable. However, it seems unlikely that this argument canbe extended to copper, except perhaps on the margins.

2.4.2. Sources of MoMolybdenum occurs in varying amounts in porphyry-type deposits,

ranging from trace levels (b0.01 wt.%Mo) in porphyry Cu–(Mo)deposits,where itmay not even be recovered as a byproduct, to being themain orecomponent (up to 0.3 wt.% Mo) in porphyry Mo deposits (Seedorff et al.,2005;Westra and Keith, 1981). At theMo-rich end of the spectrum, thereare two clearly different tectonomagmatic associations, only one ofwhichis directly related to subduction: calc-alkaline porphyry Mo deposits aregenerally relatively low grade (0.1–0.02 wt.% Mo; Carten et al., 1993),whereas intra-cratonic rift-related deposits associated with high-silica,fluorine-rich, peraluminous granitoids are relatively high grade (0.1–0.3 wt.%Mo;e.g., “Climax-type”deposits; Cartenet al., 1993;KirkhamandSinclair, 1996; Sinclair, 2007; Stein, 1988; White et al., 1981).

Kesler (1973) noted a general association (with exceptions) ofporphyry Cu–Au deposits in island arcs, and porphyry Cu–Mo depositsin continental arcs, and it is clear that the peraluminous felsic rocksassociated with rift-related Climax-type porphyry Mo deposits areprimarily of continental crustal origin (Farmer and DePaolo, 1984;Stein, 1988). This has led to one view that Mo might be predominantly

derived from continental crustal sources (Farmer and DePaolo, 1984;Stein, 1988; Klemm et al., 2008; White et al., 1981). On the other hand,minor amounts of Mo do occur in some island arc-related porphyrydeposits where no continental crustal sources are inferred (Westra andKeith, 1981), so amantle (subduction zone) source for at least someMocannot be excluded. Moreover, Blevin and Chappell (1992) and Blevinet al. (1996) have demonstrated a continuum from Cu–Au depositsassociated with unevolved, mafic I-type granitoids to W–Mo depositsassociated with cogenetic, evolved granites in eastern Australia,suggesting a common, magmatic source for all of these elements.

A complication is introduced in the Climax-type deposits, becausealthough the immediate source of the Mo-bearing fluids is felsicmagma of clear crustal origin, many deposits also show a close geneticassociation with mafic alkaline magmas, which may have introducedvolatiles, S, and possibly Mo into the evolved felsic magma chamber(Audétat, 2010; Carten et al., 1993; Keith et al., 1986, 1998). Keith etal. (1997), Hattori and Keith (2001), and Maughan et al. (2002) havealso suggested that injections of mafic alkaline magmas into theevolving Bingham Canyon magmatic system may have given rise tothe unusually large size and high grades of this porphyry Cu–Mo–Audeposit. Along the same lines, Pettke et al. (2010) have proposed thatthe unusual Cu–Mo–Au endowment of the southwestern USA (e.g.,the giant Bingham, Butte, Climax, Henderson, and Questa porphyryCu–Mo–Au and porphyry–Mo deposits) reflects Cenozoic remobiliza-tion of Proterozoic subduction-metasomatized subcontinental mantlelithosphere (see Section 2.5).

Thus, at this time there is no consensus regarding the crustalversusmantle origin ofmolybdenum in porphyry deposits, although itis clear that the highest grade porphyry Mo deposits are formed inintra-plate continental settings, and if a mantle source is important inthese cases, it is not directly related to subduction activity but ratherto rifting or reactivation of previously subduction-enriched litho-spheric sources.

2.5. Lithospheric melting during post-subduction events

A number of mineral deposits with broad similarities to thoseformed by subduction-related processes are also found in post-subduction tectonic settings, such as subduction reversal or migration,arc collision, continent–continent collision, and post-collisional rifting.They include porphyry Sn–W, Mo, Cu–Mo, and Cu–Au deposits andepithermal Au deposits, and in many cases are only known not to bedirectly related to subduction because of precise geochronology andplate tectonic reconstructions that place their formation after subduc-tion has demonstrably ceased. Associated magmas are typically calc-alkaline, but tend towards somewhat more alkaline compositionscompared with normal arc magmas (Richards, 2009).

In complex accretionary arcs, it can be very difficult to ascribe anygiven pluton (and any associated mineral deposits) to a particularsubduction or collisional event, because subduction commonly con-tinues after collision, albeit normally with a shift in the locus ofmagmatism. However, in continent–continent collision zones or wherearc collision terminates subduction, there can be greater certainty aboutthe timing of cessation of subduction magmatism. Consequently, it is incollisional orogens such as theNeo-Tethyan belt of southeastern Europeand southernAsia that someof the clearest examples of post-subductionmagmatism and mineralization are found. These include, from east towest, theMioceneGangdese porphyry Cu–Mobelt in the Tibetanorogen(Hou et al., 2006, 2009; Yanget al., 2009), theMioceneKermanporphyryCu–Mo belt in southeastern Iran (Shafiei et al., 2009), the Miocene SariGunay epithermal Au deposit in northwestern Iran (Richards et al.,2006b), theEoceneÇöpler epithermalAudeposit in southeasternTurkey(Keskin et al., 2008; Kuscu et al., 2010), the Pliocene Kisladag porphyryAu deposit in western Turkey, the Miocene Skouries porphyry Cu–Au–PGEdeposit inGreece (Economou-Eliopoulos andEliopoulos, 2000), and

Page 9: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

9J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

the Roşia Montană epithermal Au deposit in Romania (Manske et al.,2006; Neubauer et al., 2005).

Similarly, in the southwest Pacific ocean, accurate plate tectonicreconstructions permit the identification of a number of post-subduction porphyry and epithermal deposits, such as the Grasbergporphyry Cu–Au deposit in Papua, Indonesia (Cloos et al., 2005;Paterson and Cloos, 2005), the Ok Tedi porphyry Cu–Au deposit (vanDongen et al., 2010) and the Porgera alkalic-type epithermal Au depositin mainland Papua New Guinea (Richards et al., 1990; Richards andKerrich, 1993), the Lihir alkalic-type epithermal Au deposit on LihirIsland, PapuaNewGuinea (Carman, 2003;Kennedyet al., 1990), and theEmperor alkalic-type epithermal Au deposit in Fiji (Gill and Whelan,1989; Setterfield et al., 1992). (For reviews of alkalic-type epithermaldeposits, see Jensen and Barton, 2000 and Richards, 1995).

Because subduction has ceased in these regions, a fresh supply offluids, volatiles, and other slab-derived components to the mantlewedge no longer exists. Nevertheless, the broad geochemical similarityof many of these magmas to normal arc magmas, including theirhydrous and generally oxidized nature, suggests some link tosubduction metasomatism. Consequently, many researchers haveimplicated upper-plate lithospheric sources, modified by earliersubduction-related fluids and/or hydrous melts (e.g., Clemens et al.,2009; Cloos et al., 2005; Guo et al., 2007; Harris et al., 1986; Johnsonet al., 1978; Pearce et al., 1990; Pettke et al., 2010; Richards, 2009).Previously subduction-modified asthenosphere is unlikely to be a viablesource except for a short period after subduction has ceased (e.g.,Richards et al., 1990; Solomon, 1990), because such material will bequickly dispersed by mantle convection.

The key to all of these models is subduction-derived water, which ismost likely stored in amphibolitic cumulates, residual from the earlierarcmagmaflux and located in thedeep crust ormantle lithosphere (e.g.,Claeson and Meurer, 2004; Davidson et al., 2007; DeBari and Coleman,1989; Jagoutz et al., 2009; Larocque and Canil, 2010; Müntener andUlmer, 2006; Tiepolo and Tribuzio, 2008). Water lowers the solidus ofsilicate assemblages, and will lead to the formation of hydrous partialmelts during pro-grade metamorphism or mafic melt invasion (Beardand Lofgren, 1991; Rushmer, 1991; Wolf and Wyllie, 1994).

Thermal rebound in thickened orogenic crust, delamination ofsub-continental mantle lithosphere, or post-collisional rifting (withingress of asthenospheric melts into the lower crust in the last twocases) can all cause small-volume partial melting of arc-metasoma-tized lithosphere and/or hydrous lower crustal cumulates (Fig. 4;Brown, 2010; Clemens et al., 2009; Harris et al., 1986; Richards, 2009).Such melts, being derived from subduction-modified sources, willshare many of the characteristic geochemical features of arc magmas,including their relatively high water contents and oxidation states,and potentiallymetal contents (see Section 2.5.1). The smaller volumeof partial melting to be expected in such tectonic settings will givethese magmas a somewhat more alkaline composition than arcmagmas (e.g., Clemens et al., 2009), and will also mean that largebatholithic complexes are unlikely to be formed, consistent with thegenerally smaller and more isolated occurrence of such post-subduction magmatic systems (compared with arc-related Cordille-ran batholiths, or collisional S-type batholiths; Pitcher, 1997).

Because these post-subduction magmas are derived from amphibo-litic sources in which garnet (±titanite) is likely also present, andbecause their hydrous nature will suppress plagioclase fractionation(similar to other hydrous arc magmas), they may be characterized byelevated Sr/Y and La/Yb ratios; that is, they may display adakite-liketrace element characteristics.

2.5.1. Behavior of metals in subduction-modified sourcesAs discussed in Section 2.1, arc magmas are characterized by high

fO2 and fS2 relative to normal melts from MORB-depleted astheno-sphere. Consequently, such magmas may be sulfide-saturated (atoxidation states up to ΔFMQ+2.3; Jugo, 2009) despite sulfur being

predominantly present in the melt as sulfate or SO2 (Carroll andRutherford, 1985). Xenoliths from supra-subduction zone mantle(McInnes et al., 1999) and samples of mafic cumulates from lowercrustal arc roots (Fig. 5; Canil et al., 2010; Greene et al., 2006; Jagoutzet al., 2007) reveal the common presence of small amounts of sulfide,typically trapped as inclusions in silicate phases (suggesting a primarymagmatic rather than secondary hydrothermal origin).

Hamlyn et al. (1985), Richards (1995, 2009), Solomon (1990), andWyborn and Sun (1994) have explored the role of residual sulfidephases on the metal content of fractionating magmas, and also ofpartial melts formed during later, post-subduction melting events.The high partition coefficients for chalcophile and highly siderophileelements (HSE) between sulfide phases and silicate melt mean thatsuch metals should be strongly partitioned into any coexisting sulfidephases (Campbell and Naldrett, 1979; Peach et al., 1990). As shown inFig. 2, at high abundances of sulfide relative to silicate melt (low R-factor; Campbell and Naldrett, 1979), the melts will be depleted in allof these chalcophile and siderophile elements. In contrast, atintermediate abundances of sulfide (intermediate R-factor), onlyoriginally sparse HSE will show significant depletions. This ledRichards (2005, 2009) to propose that small amounts of sulfide leftbehind as residual phases from fractionation of arc magmas in thedeep lithosphere (or asthenosphere) will not significantly depletethose magmas in relatively abundant chalcophile elements such as Cuand Mo, but might significantly deplete them in highly siderophileelements such as Au. This would give rise to magmas with relativelyhigh Cu/Au ratios (which might form Cu-rich porphyry deposits), butwould leave a residue of potentially HSE-rich sulfides in the mantleand/or lower crustal amphibolitic cumulate arc roots.

As noted in Section2.5, subduction-modified asthenospheric sourceswill be rapidly convected away when subduction ceases, and so couldonly contribute to immediately post-subduction magmatism. Incontrast, deep crustal amphibolites are preserved in the lithosphere,andwill be susceptible to partialmeltingat any later timedue to thermalreboundor reheating by invading asthenosphericmelts. Under lower fS2post-subduction conditions (a flux of S from the subduction zone is nolonger present), any residual sulfide phases would likely dissolve intothe S-undersaturated silicatemelt, carrying their metal loadswith them(e.g., Ackerman et al., 2009). Richards (2009) proposed that this mightexplain the occurrence of Au-rich porphyry and related epithermalsystems in some post-subduction settings, such as the alkalic-typeepithermalAu deposits of the SWPacific, and various post-collisional Audeposits in the Balkans–Turkey–Iran Neotethyan belt. Pettke et al.(2010) have proposed a similar model for giant porphyry Cu–Mo–Audeposits in the southwestern USA.

This model can also explain the occurrence of Au–poor porphyryCu–(Mo) deposits in post-subduction settings, such as those in Tibetand Iran (Hou et al., 2009; Shafiei et al., 2009; Wang et al., 2007b), theonly difference being that in this case larger proportions of sulfidemay have fractionated out from the original arc magmas in the deepcrust. Such sulfides would have retained significant amounts of thesubduction flux of Cu and Mo, but HSE would be diluted to lowconcentrations by the greater volume of sulfide (low R-factor; Fig. 2).Second-stage melts from such cumulate sources would therefore beCu–(Mo)-rich, but not necessarily Au-rich.

A control on these two scenarios (abundant Cu-rich residual sulfideversus sparse but HSE-rich residual sulfide, or low versus high R-factor)might be the average oxidation state and sulfur fugacity of the generativesubduction system. Inmore oxidized or S-poor systems, smaller volumesof HSE-rich sulfide would exsolve from the silicate melt (high R-factor;Campbell andNaldrett, 1979),whereas in less oxidized or S-rich systems,larger volumes of Cu-rich but HSE-poor sulfide would exsolve (low R-factor; Fig. 2). In particular, the proportion of sulfides exsolving from arcmagmas may be very sensitive to small changes in their oxidation state,because of the rapid change from sulfide to sulfate dominance inmagmatic systems between ΔFMQ+1 and +2 (Jugo et al., 2010). The

Page 10: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Fig. 4. Post-subduction tectonic environments conducive to the formation of porphyry and epithermal deposits by remobilization of previously subduction-modified lithosphere(modified from Richards, 2009). (a) Porphyry Cu±Mo deposits formed in normal arc settings; a continental arc is shown, but similar processes can occur inmature island arcs. (b–d)During post-subduction tectonic processes, previously subduction-modified sub-continental lithospheric mantle (SCLM) or lower crustal hydrous cumulate zones residual fromprevious arc magmatism (black layer) may undergo small-volume partial melting. Such magmas may remobilize Au as well as Cu±Mo left behind in residual sulfide phases by arcmagmatism, leading to the potential formation of porphyry Cu±Au±Mo and alkalic-type epithermal Au deposits. Magmas may be characterized by high Sr/Y and La/Yb ratios dueto the presence of hornblende (±garnet, titanite) in the amphibolitic lower crustal source rocks. See text for discussion.

10 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

oxidation state of the mantle wedge will depend on the character of theflux from the subducting slab (e.g., a higher proportion of subductedorganic-rich sediment would lead to lower oxidation states; Wang et al.,2007a); this property is therefore likely to have a characteristic averagevalue along any given arc at any particular period of time.

Variations in oxidation state over typical ranges for arc magmas(ΔFMQ=0 to +2; Ballhaus, 1993; Brandon and Draper, 1996; Blatterand Carmichael, 1998; Malaspina et al., 2009; Parkinson and Arculus,1999; Rowe et al., 2009) will not greatly affect the potential to form syn-subduction porphyry Cu–(Mo) deposits, but might control the Cu/HSEratio in later magmas formed by post-subduction melting of thesesulfide-bearing residues. Specifically, Au-rich (low Cu/Au) post-subduc-tion porphyries might form in settings where previous arc magmatismwas relatively oxidized (sparse but HSE-rich sulfide residue), whereasAu-poor (high Cu/Au) post-subduction porphyries might form whereprevious arcmagmatismwas relatively reduced (more abundant Cu-richsulfide residue). Such a mechanism might also explain why coeval beltsof porphyry deposits tend to have characteristic Cu/Au ratios.

Finally, partial melting of predominantly reduced, sulfide-rich crustalrocks in orogenic settings may lead to chalcophile and siderophileelement-depleted, but potentially lithophile element-rich S-typemagmas (see Section 2.6).

2.6. Crustal melting during post-collisional stress relaxation

Collisional orogens commonly undergo crustal thickening followedby extensional or transpressional collapse. Bimodal magmatism is

characteristic of such tectonic settings, resulting from partial melting ofpelitic protoliths in the deep crust triggered by the heat from upwellingasthenospheric melts (Hildreth, 1981). Peraluminous S-type granites(Chappell and White, 1974) are subsequently emplaced as largebatholith complexes in the mid- to upper orogenic crust (e.g., theHercynian peraluminous granites of Europe; Barbarin, 1996; Clemens,2003; Darbyshire and Shepherd, 1994; Harris et al., 1986; Wyllie et al.,1976). These granites tend to be enriched in lithophile rather thanchalcophile elements, reflecting their crustal origins, and may generatemagmatic–hydrothermal deposits containingSn,W,U,Mo, REE, Li, Be, B,and F.

2.6.1. Sources of metalsTin and especially tungsten commonly accompany molybdenum in

porphyry deposits as trace metals and byproducts, but they also form aclass of porphyries on their own, associated with S-type granites incontinental orogens (Hart et al., 2005; Ishihara, 1981; Ishihara andMurakami, 2006; Kerrich and Beckinsale, 1988; Kirkham and Sinclair,1996; Lehmann, 1982). The Hercynian tin granites of Europe, and theBolivian and SE Asian tin belts are examples of such deposits, withmineralization occurring in skarns and greisens around the graniteintrusions, and to a lesser extent as internal stockworks anddisseminations (e.g., Černý et al., 2005; Meinert et al., 2005). As withthe source magmas, metals in these deposits appear to be predomi-nantly of crustal origin (Hedenquist and Lowenstern, 1994). Forexample, in a recent assessment of the source of Sn in the Cornubianbatholith of SW England, Williamson et al. (2010) concluded that all of

Page 11: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Chalcopyrite

Chalcopyrite

Pyrite

Pyrrhotite

(a)

(b)

Fig. 5. Reflected light photomicrographs of sulfide inclusions in amphibole-rich lowercrustal arc cumulates from: (a) the Talkeetna arc, Alaska; (b) the Bonanza arc,Vancouver Island, Canada (samples courtesy A. Greene and D. Canil, respectively).

11J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

the Sn could have been extracted from the crustally-derived granites.Uranium is also significantly enriched in crustal rocks versus themantle(0.91 ppm versus ~0.02 ppm, respectively; Taylor and McLennan,1985), and so is unlikely to have a mantle source in such deposits.However, Dietrich et al. (1999) have suggested a possible role formantle-derived magmas in triggering volatile (andmetal) release fromevolved, felsic magmas in the Bolivian tin belt, andWalshe et al. (2011)have identified a mantle Nd isotopic signature in tin granites fromeastern Australia.

In contrast, in the case of W skarns associated with Mo mineraliza-tion in calc-alkaline I-type magmas, a shared mantle origin with Momight be indicated (e.g., Newberry and Swanson, 1986), consistentwiththe similar siderophile tendencies of these two elements, and theirposition in the periodic table (group VIB).

3. Behavior of metals during magma fractionation and fluidexsolution in the upper crust

Key to the formation of magmatic–hydrothermal deposits ofchalcophile and siderophile elements in the upper crust is the lack ofsignificant saturation with and loss of sulfide phases prior to aqueousvolatile exsolution from a cooling magma (Candela, 1989b, 1992;Candela and Holland, 1986; Candela and Piccoli, 2005; Richards, 1995;Richards and Kerrich, 1993; Spooner, 1993). As discussed in Sections2.1.1 and 2.5.1, chalcophile and siderophile elements partition

strongly into sulfide phases exsolving or crystallizing from silicatemelts. Thus, if extensive fractionation and removal of magmaticsulfide phases were to occur, the remaining silicate melt would bestrongly depleted in these elements (Jugo et al., 1999; Lynton et al.,1993). This is in essence the model for formation of orthomagmaticsulfide deposits from relatively reduced mafic magmas (e.g., Naldrett,1989), and perhaps explains the deficiency of chalcophile elementssuch as Cu and Au in more reduced, evolved, Sn–W-bearing S-typemagmas (Blevin and Chappell, 1992; Hedenquist and Lowenstern,1994).

In more oxidized subduction-related or post-subduction magmas,although sulfide saturation likely occurs at some point during theirevolution (as indicated by the common presence of sparse sulfideinclusions in phenocrysts in arc volcanic rocks as well as lower crustalarc cumulates; e.g., Burnham, 1979; Halter et al., 2002; Hattori, 1997;Keith et al., 1997; Stavast et al., 2006; Fig. 5), it may not occur to theextent that sulfides physically separate from the magma, or at leastnot in large amounts. Instead, small sulfide droplets or crystals may beentrained in magma ascending buoyantly through the crust (e.g.,Bockrath et al., 2004; Tomkins andMavrogenes, 2003) or as inclusionsin silicate phenocrysts, and will not be substantially lost to the overallmagma flux. Indeed, several authors have argued that pre-concen-tration of ore metals in magmatic sulfide phases may be an importantstep in porphyry metallogenesis (e.g., Jenner et al., 2010). In thesemodels, sulfide phases subsequently break down due to changes inoxidation state and sulfur fugacity in response to volatile exsolutionand magnetite crystallization upon emplacement in the upper crust,thereby rendering metals available for redissolution in the volatilephase (e.g., Cygan and Candela, 1995; Halter et al., 2002, 2005; Jugo etal., 1999; Keith et al., 1997; Stavast et al., 2006). Other authors haveargued that this process, while it may occur, is not critical to metalbehavior in magmatic–hydrothermal systems, and that direct parti-tioning from the silicate melt to the hydrothermal fluid phase is thedominant mechanism (e.g., Audétat and Pettke, 2006; Lynton et al.,1993; Simon et al., 2008; Sun et al., 2004b). As noted in Section 2.5.1,Richards (2009) suggested that separation of small amounts of sulfidefrom arcmagmas at depth in lower crustal MASH cumulate zonesmayprovide a source of metals (especially HSE) for later post-subductionmagmas, but that this process may not substantially affect the Cucontent of the original arc magmas.

Regardless of the exact role of magmatic sulfides, the ultimaterelationship is that of partitioning of metals between silicate melts andexsolving hydrothermal fluids, with sulfides as a possible intermediarystep. Our current understanding of these partitioning processes is nowquite advanced following several decades of hydrothermal experiments(e.g., Candela and Piccoli, 1995) and more recently the advance ofquantitative single fluid andmelt inclusion analysis (e.g., Heinrich et al.,2003a,b), which has enabled direct measurement of metal contents inore-forming fluids andmelts. In the following sections, I review some ofthe key factors in metal solvation and transport in magmatic–hydrothermal fluids.

3.1. Partitioning of metals from magma into exsolving hydrothermalfluid

Subduction-relatedmagmas commonly contain at least 4 wt.% H2Oduring crustal ascent, as evidenced by the presence of hornblende andbiotite phenocrysts in many andesitic volcanic rocks and arc plutons(Burnham, 1979; Naney, 1983; Rutherford and Devine, 1988). Inresponse to the decreasing solubility of water in silicate melts aspressure decreases, such hydrous magmas inevitably exsolve anaqueous volatile phase upon emplacement at shallow crustal levels oron eruption (Burnham, 1979, 1997; Eichelberger, 1995). This processhas in the past rather confusingly been called first and second boiling(although neither process is technically “boiling”), the first eventoccurring during ascent and depressurization of the magma, and the

Page 12: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

2000

1500

1000

500

0900

800700

600500

400300

200100 0

2040

6080

100

P (

bars

)

Wt.%

NaC

l

0

V+L

V-L isotherm (vapour)

V-L isotherm (liquid)

Critical / boiling curve

H2O critical point

Halite saturation

Salinity isopleth

Supercritical fluid

Liquid path

Vapour path

Coexisting vapour+liquid

Shallowmagmatic fluid

exsolution

L+Halite

V+Halite

Supercriticalfluid intersects

2-phase surface

Deepsupercriticalmagmaticfluid exsolution(10 wt.% NaCl)

Porphyry

(a)

2

1

Liquid phaseprogressivelyseparates fromvapour

2000

1500

1000

500

0

900800

700600

500400

300200

100 020

4060

8010

0

T (°C)

T (°C)

P (

bars

)

Wt.%

NaC

l

0

V+L

L+Halite

V+Halite

V-L isotherm (vapour)

V-L isotherm (liquid)

Critical / boiling curve

H2O critical point

Halite saturation

Salinity isopleth

Supercritical fluid

Liquid path

Vapour path

Contraction path

Coexisting vapour+liquid

Deepsupercriticalmagmaticfluid exsolution(10 wt.% NaCl)

Vapour phasedeparts from

2-phase surface

Supercriticalfluid neverintersects2-phasesurface

Supercriticalfluid intersects

2-phase surface

Low tomoderate

salinityliquids

4 3

Epithe

rmal

(b)

Liquid phaseprogressivelyseparates fromvapour

Fig. 6. P–T–XNaCl phase diagram,modified fromDriesner and Heinrich (2007), illustratingfluid pathways for amagmatic–hydrothermal fluid exsolvedwith an initial bulk salinity of10 wt.% NaCl. (a) Early, high thermal gradient fluids exsolved from deeply (path 1) andshallowly (path 2) emplaced magmas (porphyry environment). (b) Late, low thermalgradient fluids exsolved from deeply emplaced magma (high-sulfidation epithermalenvironment). Path 3 illustrates the supercritical fluid contraction path proposed byHedenquist et al. (1998), whereas path 4 illustrates a slightly steeper thermal gradientwith brief intersection of the 2-phase (L–V) field, as proposed by Heinrich et al. (2004).Note that in the two-phasefield, thedense saline liquidphase progressively separates fromthe vapour phase, and the two-phase pathways shown do not represent a closed system.See text for discussion.

12 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

second occurring after emplacement as crystallization progressivelyincreases the concentration of volatiles in the residual melt (Candela,1989a). In reality, volatile exsolution is probably a more-or-lesscontinuous process during arc magma ascent and cooling, starting atdepth with the exsolution of relatively insoluble CO2 (Blundy et al.,2010; Holloway, 1976; Lowenstern, 2001; Wallace, 2005). However,the bulk of the magmatic water content likely exsolves relativelyrapidly as the magma approaches its solvus, at depths of ~5–10 km,depending on magma composition and water content (Burnham,1979).

The composition of this exsolved magmatic volatile phase isdominantly aqueous, containing sulfur species (predominantly SO2 inoxidized systems, but also some reduced S species), CO2, NaCl, KCl, HCl,andmetal chlorides. The exact composition depends onmany variables,including the depth of exsolution and the magma composition(especially the initialmagmatic Cl/H2O ratio and alkali content; Candela,1989c; Candela and Piccoli, 2005; Cline and Bodnar, 1991; Webster,1992), but typical estimates for a single-phase supercritical fluidexsolved at depths below the H2O–NaCl solvus are ~2–13 wt.% NaClequivalent (average 5 wt.% NaCl equivalent) with minor CO2 (Audétatand Pettke, 2003; Audétat et al., 2008; Burnham, 1979; Candela, 1989c;Hedenquist et al., 1998; John, 1991; Redmond et al., 2004), and up to1.3 wt.% Cu and 0.3 wt.% Fe (Klemmet al., 2007; Rusk et al., 2004, 2008;Sawkins and Scherkenbach, 1981). These high observed metal solubil-ities are consistent with or exceed experimental observations andtheoretical predictions based on chloride complexing alone (e.g.,Candela and Holland, 1984, 1986), and suggest that other volatileligands such as sulfide speciesmay enhance the solubility of chalcophileelements such as Cu and Au in high temperature aqueous fluids (e.g.,Heinrich et al., 1992; Pokrovski et al., 2005, 2008; Seo et al., 2009; Simonet al., 2006; Zajacz et al., 2008, 2011).

Shallowly emplaced magmas will exsolve fluids under pressure–temperature (P–T) conditions that lie within the two-phase liquid–vapor field for the bulk fluid composition, resulting in immediateformation of an immiscible low salinity vapor and high salinity brine(Fig. 6a).

The critical point in the H2O–NaCl system (which is commonlyused as a proxy for magmatic fluids) occurs between ~1.0 and 1.4 kbfor fluid temperatures between 600° and 800 °C (the typicaltemperature range for fluids exsolved from intermediate to felsicmagmas) (Pitzer and Pabalan, 1986; Sourirajan and Kennedy, 1962),which is equivalent to depths of between ~3 and 5 km at lithostaticpressures. Most porphyry deposits and their host plutons areemplaced at depths between 1 and 6 km (Seedorff et al., 2005), sofluids exsolving directly from magmas at these depths will typicallyform immiscible liquid and vapor plumes (e.g., Henley and McNabb,1978; Nash, 1976). However, the bulk of the fluids and metals inporphyry deposits are likely initially sourced at deeper levels (5–10 km, as noted above) from larger volumes of magma in mid- toupper crustal batholithic complexes (Candela and Piccoli, 2005; Cloos,2001; Damon, 1986; John, 1991; Richards, 2003, 2005; Shinohara andHedenquist, 1997), at which depths the fluids will be supercritical. Asthese deep fluids rise into shallow cupola zones extending above themain batholith, they will likely intersect the solvus on the vapor side,and will begin to condense a dense saline brine (Ahmad and Rose,1980; Figs. 6a and 7).

Supercritical fluids are highlymobile (e.g., Coumou et al., 2008; Dunnand Hardee, 1981; Norton and Dutrow, 2001) and behave differently interms of magma–fluid partitioning compared to fluids exsolved atshallower depths in the two-phase field. In particular, Henley andMcNabb (1978) suggested that the higher density and viscosity of salinebrine condensates might restrict their flow, leaving them as a denseresidual liquid in the deeper parts of evolving magmatic hydrothermalsystems (see also: Lewis and Lowell, 2009;White et al., 1971). The lowerdensity vapor or supercritical fluid would be expected to be highlyupwardlymobile, aswell as larger in termsof bothvolumeandmass than

the brine phase (assuming an initial bulk salinity below ~20 wt.% NaCl),and sohasmuchgreater potential as anefficient transportingmediumforore components. However, until fairly recently, it was assumed that lowsalinity vaporswould not have the capacity to dissolve large quantities ofbasemetals as chloride complexes, and the brine phasewas therefore thefavored ore-forming medium (e.g., Bodnar and Beane, 1980; Cline andBodnar, 1991; Eastoe, 1982; Hedenquist and Richards, 1998; Moore andNash, 1974; Nash, 1976; Shinohara, 1994; Williams et al., 1995).Observations of chalcopyrite crystals trapped in some vapor-rich fluidinclusions (e.g., Bodnar and Beane, 1980) were explained by some asproducts of heterogeneous trapping (see discussion in Mavrogenes and

Page 13: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Fig. 7. Schematic cross-section through a typical coupled arc batholith–cupola–volcanic system, with associated porphyry Cu±Au and linked high sulfidation Cu–Au epithermaldeposits. Also shown are the thermal structure, fluid flow pathways and characteristics during the main stage of hydrothermal activity, and overlapping hydrothermal alterationzones. Propylitic alteration by circulating heated groundwaters can be assumed to affect all the supracrustal rocks in the field of view, with greatest intensity (epidote, actinolite)close to the intrusions, fading to background distally. Modified from Richards (2005); sources: Sillitoe (1973, 2010), Dilles (1987), Shinohara and Hedenquist (1997), Hedenquist etal. (1998), and Fournier (1999).

13J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Bodnar, 1994). The advent of quantitative analysis of single fluidinclusions by synchrotron X-ray microprobe, proton-induced X-rayemission spectroscopy (PIXE), and laser ablation-inductively coupledplasma-mass spectrometry (LA-ICP-MS) revealed that, contrary to theseearlier assumptions, considerable amounts ofmetal, including CuandAu,were indeed consistently present in some vapor-rich inclusions (Audétatet al., 2000; Harris et al., 2003; Heinrich et al., 1992, 1999; Klemm et al.,2007; Lowenstern et al., 1991; Simon et al., 2005, 2006; Ulrich et al.,2001), and subsequent work has shown that much of this metal contentis transported as sulfide complexes rather than as (or in addition to)chloride complexes (Cauzid et al., 2007; Heinrich et al., 1992, 1999;Pokrovski et al., 2005, 2008; Seoet al., 2009; Simonet al., 2006;Zajacz andHalter, 2009; Zajacz et al., 2010).

Early resistance to these ideas was, I believe, at least in part due toconfusions of terminology: most papers dealing with this subject havereferred tometal solubility and transport in a vapor phase, suggesting tothe unwary a lowdensity, low salinity gas,whereas for themost part thefluids in question were either single-phase supercritical fluids, orrelatively highdensity vapors just below their critical points. Under suchhigh P–T conditions, vapors are almost as saline as the initial single-phase fluid from which they evolved, and therefore they still containplenty of chloride for base metal complexing. But perhaps moreimportantly, as noted above, these vapors will also contain a highconcentration of volatile sulfur species, which now appear to beessential for the efficient solvation of chalcophile elements under highP–T conditions.When combinedwith the highermass proportion of thevapor phase (versus brine condensate) and its highmobility, amodel fortransport of the bulk of the metal flux in porphyries by relatively densevapors or supercritical fluids is now well established (Klemm et al.,2007; Landtwing et al., 2010; Williams-Jones and Heinrich, 2005).

The rapid reduction in the efficiency of transport of metals as thevapor plume rises, cools, and becomes less dense by brine condensation,

may explain precipitation of the bulk of Cu, Mo, and some Au over arelatively narrow temperature interval between 425°–320 °C (Hemley etal., 1992;Klemmet al., 2007; Landtwing et al., 2005). At shallower depthsand lower temperatures, the vapor phase may become too dilute totransport significant amounts of base metals as chloride complexes, butmay continue to carry some metals such as Au, Cu, As, and Sb as sulfidecomplexes (Deditius et al., 2009; Simon et al., 2006, 2007), eventuallyeither venting them to the surface in high-temperature fumaroles (e.g.,Chaplygin et al., 2007; Hedenquist et al., 1994a; Symonds et al., 1987;Tarana et al., 1995; Tessalina et al., 2008) or precipitating them in thenear-surfacehigh-sulfidationepithermal environment (see Section4.2.1;Deditius et al., 2009; Hedenquist et al., 1993, 1994b; Heinrich et al., 1999,2004; Larocque et al., 2008; Murakami et al., 2010; Pudack et al., 2009;Williams-Jones and Heinrich, 2005).

4. Magmatic–hydrothermal ore formation

The focus of this paper is on the flux of metals in subduction-related magmatic systems, but this would be of little practical interestif that flux did not ultimately lead to ore formation. Thus far, we havefocused on the importance of firstly not losing significant amounts ofmetal to a fractionating or residual sulfide phase, and then efficientlypartitioning those metals into a highly mobile aqueous fluid phase.What subsequently happens to that fluid phase dictates whethereconomic concentrations of metals are precipitated (grade), whereasthe scale of the magmatic and derivative hydrothermal systemcontrols the total amount of metals precipitated (tonnage).

4.1. Porphyry Cu ore formation

In a landmark paper, Cline and Bodnar (1991) presented a modelfor the evolution of magmatic-hydrothermal systems from initial

Page 14: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

14 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

aqueous fluid exsolution to cooling and mineral precipitation. A keyfinding of this work was that, although there are many variables thatcan affect the specific evolutionary path of a given system, aneconomic porphyry Cu deposit can potentially be formed from quitesmall volumes of typical andesitic arc magma. For example, theauthors showed that 15–90 km3 of andesitic magma initially contain-ing 50 ppm Cu, 2.5 wt.% H2O, and with a Cl/H2O ratio=1, is sufficientto generate a 250 Mt deposit with a grade of 0.75 wt.% Cu. The range ofrequired magma volumes reflects variables such as the depth ofmagma emplacement and the compatibility of Cu with fractionatingmineral phases (larger volumes are required if Cu is compatible withearly fractionating silicate, oxide, or sulfide minerals). Details of themodel, later updated in Cline (1995), show that magmas emplaced atmoderate depths (1.0 to 2.0 kb, equivalent to depths of ~4–8 km)most efficiently partition Cu into early saturating saline fluids,whereas Cu and Cl are only released from shallowly emplaced(0.5 kb, or ~2 km) magmas during late stages of crystallization (seealso Candela, 1989b). (Note that these depths reflect the locus of fluidand metal exsolution from the magma under lithostatic pressureconditions, as opposed to the depth of subsequent hydrothermalmetal deposition, which will be at shallower levels and likely underhydrostatic pressure conditions; Fournier, 1999.) In large measure,these differences reflect the change in properties of the magmatic–hydrothermal fluid, which will separate initially as a moderatelysaline supercritical fluid in deeper systems (Fig. 6a, path 1), but asimmiscible vapor and brine at shallower levels (Fig. 6a, path 2). Cline(1995) concluded that optimum conditions for porphyry Cu oreformation are obtained where magma containing ~4 wt.% H2O andwith a high initial Cl/H2O (typical of most arc magmas) is emplaced atmoderate crustal depths, thereby maximizing the efficiency of Cupartitioning into an early saline fluid phase.

The results of these studies indicate that no special conditions ormagmatic metal enrichment are required to form even large porphyryCu deposits. Instead, the question is rather one of process efficiency:where (what depth) and how are metalliferous fluids released andchanneled? This finding is of fundamental importance from anexploration perspective, because it shifts the focus from seekinganomalously metal-rich magmas (which have proven elusive; Jenneret al., 2010) to searching for optimal ore depositional settings inotherwise normal tectonomagmatic environments (e.g., Tosdal andRichards, 2001).

Cloos (2001), Shinohara et al. (1995), and Shinohara andHedenquist(1997) considered the question of process efficiency from theperspective of physical separation and focusing of volatile releasefrom the magma. Cloos (2001) suggested that the classic cupola shape(Norton, 1982) of porphyry systems above larger batholithic magmachambers reflects convective circulation of bubbly magma into theshallowapical parts of these systems (at 1–3 kmdepth), where volatilesphysically separate from the melt and coalesce to form a discrete fluid-filled cupola. The now-dense, volatile-depleted magma forms adownward return flow to complete the convective cycle. In contrast,Shinohara and Hedenquist (1997) envisaged fluids physically separat-ing from the magma at greater depth within the underlying magmachamber, and rising as a discrete plume up apical channelways formedby fractures and dikes in the brittle carapace (Fig. 7).

A key aspect of both of thesemodels is that volatile separation, andCu partitioning, occurs from a much larger volume of magma(emplaced at deeper levels) than that preserved and commonlyvisible within the shallow-level ore body. In Cloos's (2001) model,volatile-rich, bubbly magma rising from the underlying batholithconvects through the cupola zone where it releases its fluids, whereasin Shinohara and Hedenquist's (1997) model, vesiculation andconvective circulation occur in the underlying magma chamber itself,and fluids are released as a plume into the base of the apical dikesystem. Combining these models with (Cline, 1995; Cline andBodnar's, 1991) calculations suggests that maximum ore-forming

efficiency in porphyry systems is likely achieved where volatilesaturation occurs in large (≥100 km3) mid- to upper crustal magmachambers at depths ≥6 km, containing moderately hydrous (N4 wt.%H2O) and Cl-rich magmas. These fluids either rise as bubbly magma oras a separate volatile plume into the apical parts of the systemwhere decreasing pressure and temperature cause deposition of Cuand Mo±Au (Candela, 1989b; Shinohara et al., 1995).

Focusing of magma ascent and fluid flow into narrow apicalregions, or cupolas, is likely to be a function of structure in the brittlerocks overlying the batholithic system (Tosdal and Richards, 2001,and references therein). Shallow crustal magma emplacement willcause extensional doming in the cover rocks, with dilational faultzones providing high-permeability pathways for fluid and magmaascent (Burnham, 1979). Evidence from the dike emplacementliterature suggests that such fractures may first be opened andpropagated by volatile pressure, and only later filled by more viscousmagma (Burnham, 1979; Carrigan et al., 1992; Rubin, 1995). Thisraises the intriguing possibility that the cylindrical shapes of manyporphyry stocks may have arisen first as breccia pipes or diatremesbored out by rapidly escaping volatiles, only later to be back-filledwith porphyritic magma (Fig. 8; Norton and Cathles, 1973; see alsoFig. 2 in Anderson et al., 2009, and Fig. 8 in Sillitoe, 2010, and Fig. 17 inVry et al., 2010).

To some extent, focusing of magma and fluid flow along narrowconduits may be self-organizational, because once initial channelways have developed, they will represent high-permeability path-ways and are likely to thermally weaken the wall rocks and promotefurther fracturing and channeling.

Narrow focusing of apical fracturing and subsequent fluid flow arelikely to be critical to the formation of high grade porphyry deposits(e.g., El Teniente, Chile; Vry et al., 2010), while multiple breccia/intrusive events can potentially increase tonnage (provided that laterevents do not destroy earlier mineralization).

Given suitable fluid flow focusing, three further factors combine tocause maximum efficiency of metal deposition within relatively smallvolumes in porphyry Cu±Mo±Au deposits. All three effects arerelated to the steep temperature gradient in the cupola zone (Fig. 7),and they therefore control the vertical range of ore deposition. On theother hand, fluid focusing controls the lateral extent of mineralization.In combination, the highest grades will occur where ore deposition isboth focused laterally and restricted vertically.

The first factor is that Cu solubility (as chloride species) decreasesdramatically as fluids cool through the temperature interval ~400° to300 °C (Crerar and Barnes, 1976; Hemley et al., 1992; Klemm et al.,2007; Landtwing et al., 2005; Xiao et al., 1998). Given the very sharptemperature gradient implied by near-surface emplacement ofmagma (Fig. 7), this temperature interval will correspond to a narrowdepth range, likely with 1 or 2 km of the surface (although it mayextend to greater depths with time as the magmatic–hydrothermalsystem begins to cool). This depth range is typical for ore formation inmany porphyry systems.

The second factor is that SO2 dissolved in the magmatic–hydrothermal fluid phase progressively disproportionates to H2Sand H2SO4 as the fluid cools below ~400 °C (Holland, 1965; Kusakabeet al., 2000; Reeves et al., 2010; Sakai and Matsubaya, 1977):

4SO2 + 4H2O ⇔ H2S + 3H2SO4: ð1Þ

This reaction generates both hydrogen sulfide, which initiatesabundant precipitation of sulfide minerals (i.e., chalcopyrite, pyrite,molybdenite), and also sulfuric acid, which causes early deposition oflarge volumes of anhydrite in the potassic alteration zone, andprogressively increasing degrees of hydrolytic alteration (an initialshift from feldspar-stable potassic alteration, to muscovite/sericite-stable phyllic alteration). Consequently, the bulk of Cu-sulfide miner-alization occurs at the low-temperature, late-stage end of the potassic

Page 15: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Porphyryintrusion

Mixed magmatic-hydrothermal

breccia

Cuspateintrusiveborder

Porphyryintrudingcomagmaticbreccia

Cuspateporphyry

clastMixed magmatic-

hydrothermalbreccia

Hydrothermalquartz fillingcavity

Limestonewallrock

clast

Porphyrydike

(a) (b)

(c)

~20 cm

Fig. 8. Photographs of porphyritic magma invading co-magmatic hydrothermal breccia pipe (a, c) and breccia vein/dike (b) from the Pachapaqui Ag–Cu–Pb–Zn deposit, Péru. Thebreccia consists of fragments of porphyry magma and country rock; note scalloped, cuspate margins of the porphyry body in (a, b) and porphyry clasts in (c), indicating that themagma was molten at the time of breccia formation and subsequent intrusion into the breccias. In (b), large vuggy spaces are partially filled with hydrothermal quartz, reflecting therole of fluids in breccia formation. Scale in (c) is in centimeters.

15J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

alteration phase, just prior to the onset of phyllic alteration (i.e.,corresponding to the “ore shell” of the classic Lowell and Guilbert, 1970,porphyry Cu model).

The third factor is the increased permeability over this tempera-ture range caused by a combination of the transition in silicate rocksfrom ductile to brittle behavior at temperatures (between ~400°–350 °C; i.e., the brittle–ductile transition; Fig. 7; Cathles, 1991;Fournier, 1999; Landtwing et al., 2005), and a window of retrogradesilica solubility (between ~550–350 °C; Fournier, 1985). Not only dothese processes result in the formation of open-standing brittle veinsand porosity, thereby facilitating rapid upward fluid flow andwallrock permeation (e.g., the crackle breccias, stockworks, anddisseminated mineralization textures so characteristic of porphyry Cudeposits), but this transition also represents the boundary betweenlithostatic and hydrostatic fluid pressures (a pressure differential of~3×). Sudden depressurization of fluids across this boundary can beexpected to have major effects on fluid properties, including phaseseparation (Fig. 6) and consequent changes in metal solubility (e.g.,Landtwing et al., 2005, 2010; Murakami et al., 2010).

In combination, these four factors, (1) spatial focusingoffluidflow innarrow cupolas, (2) reduction of metal solubility, (3) increaseddissolved sulfideactivity, and (4)permeability increasedue to transitionfrom ductile to brittle fracturing and retrograde silica solubility (with a

large pressure drop), serve to narrowly focus Cu-sulfide mineralizationboth laterally and vertically within cupola zones above large mid- toupper crustal batholithic complexes. The most likely reasons forotherwise prospective porphyry systems to be unproductive will beeither a failure to focus fluid flow, or simply insufficient fluid supply(likely due to an insufficiently large underlying magmatic system).

4.2. Epithermal Cu–Au ore formation

4.2.1. High-sulfidation epithermal Cu–Au depositsAs noted in Section 3.1, although the bulk of Cu (andMo) appears to

beprecipitated over the temperature interval 425°–320 °C, someCu andother metals such as Au, Sb, and As may remain in solution as sulfidecomplexes, to be carried into the shallow epithermal regime. Observa-tions from rare fluid inclusions in high-sulfidation epithermal Cu–Audeposits suggest that the ore-forming fluid was a low- to moderate-salinity liquid (0.2 to 4.5 wt.%NaCl equivalent; Hedenquist et al., 1994b;Mancano and Campbell, 1995), which paragenetically post-datesadvanced argillic alteration formed by highly acidic magmatic gasses(Arribas, 1995; Hedenquist et al., 1994b; Stoffregen, 1987). Thisapparent inconsistency has been explained by Hedenquist et al.(1998), Heinrich et al. (2004), and Heinrich (2005) in terms ofcontraction of a moderate salinity supercritical magmatic fluid or

Page 16: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

16 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

vapor by rapid cooling during ascent such that it does not touch, orbarely touches, the two-phase solvus (Figs. 6b and 9, paths 3 and 4,respectively). Because of the curvature of the solvus crest (critical curve)to lower salinities at low temperatures and pressures, this moderatelysaline fluid will lie on the liquid side of the solvus at shallow depths,although it may have contracted from what was originally a vapor orsupercritical fluid phase at higher temperatures and depths.

Heinrich et al. (2004) specifically suggested that brief intersectionwith the solvus at high pressures and temperatures (Figs. 6b and 9, path4)might be an importantway to shed chloride-complexed componentssuchas Fe fromthe vapor in a brine condensate, leavingbisulfide ligandsfree to bond with Cu and Au in the residual vapor and transport it toshallower (epithermal) levels. Thismodel requires that the vapor leavesthe two-phase surface again by coolingat pressure, therebypassingoverthe crest of the solvus (the critical curve; (Figs. 6b and 9, path 4); uponsubsequent ascent and depressurization, this fluid will be a liquid, asdescribed above. Heinrich et al. (2004) argued that the Fe-rich brinecondensation step is essential for retention of Au (and Cu) in the vaporphase, because otherwise Fe will tend to precipitate as Cu–Fe-sulfideminerals and strip the fluid of bisulfide ligands, thus causing Au to co-precipitate at depth (a possible mechanism for the formation ofporphyry Cu–Au deposits). However, this condensation step mustthen be followed by cooling while still at depth, in order to “lift” thevapor phaseoff the two-phase surface, and allow it to contract to a liquid(Figs. 6b and 9, path 4).

Epithermal

V-L solvus (vapour)

V-L solvus (liquid)

Brit

tle(h

ydro

stat

ic)

Ductile

(lithostatic)

30°C

/km

geo

ther

m

V

Main-stagebrittle

porphyryveins

T (°C)0 100 200 300 400 500

P (

bars

)

500

0

1000

1500

2000

0 100 200 300 400 500

Fig. 9. Pressure (depth)–temperature section through the H2O–NaCl phase diagram, with vaDriesner and Heinrich, 2007). Red curves indicate that the solvus phase is a vapor, blue ccomposition. Also shown are the wet granite and granodiorite solidi (Burnham, 1979), andgradients (Barbier, 2002; Goff et al., 1992; Noorollahi et al., 2007). Depths are indicated forranges for supercritical magmatic fluid exsolution, early high-temperature porphyry veins, lafluid P–T paths are shown corresponding to: (1) a typical porphyry-forming fluid path; (2contraction path of Hedenquist et al. (1998); and (4) the contraction with minor brine con

A more detailed analysis of the phase diagram shown in Fig. 6reveals that a typical magmatic fluid of 2–13 wt.% NaCl would have tofirst intersect the solvus at temperatures above ~400°–500 °C for thismodel to work optimally, otherwise it would fall on the liquid side ofthe two-phase field, and would become more saline by boiling off adilute vapor (Figs. 9 and 10). Assuming that this did indeed happen,then the smaller (by mass) vapor component would have to leave thetwo-phase surface again at pressure before cooling below ~400 °C,otherwise its salinity rapidly falls to sub-weight percent levels atlower temperatures and lower pressures (Figs. 6 and 10), which areinconsistent with fluid inclusion evidence for low tomoderately salineliquids in high-sulfidation epithermal ore formation (Hedenquist etal., 1994b;Mancano and Campbell, 1995). Thus, the P–T trajectory of afluid that satisfies Heinrich et al.'s (2004) model for high-sulfidationepithermal Cu–Aumineralization (Figs. 6b and 9, path 4) is somewhatunique, and may occur only rarely or fleetingly during the waningstages of a cooling magmatic–hydrothermal system.

A more normal, or early ascent pathway for a magmatic–hydrothermal fluid would be for it to rise more-or-less is enthalpicallyor quasi-adiabatically along a steep P–T gradient (e.g., Hemley andHunt, 1992; Henley and Hughes, 2000; Wood and Spera, 1984), andtherefore to dive deeply into the two-phase field and separate into anincreasingly dilute vapor phase and a saline brine (Figs. 6 and 9, path1), or even to boil dry to halite plus vapor (Figs. 6 and 9, path 2). Suchfluid pathways might be consistent with the widespread and intense

sudil

oseti

narg

teW

sudil

oseti

roid

onar

gte

W

Earlyporphyry

veins(2-phase

fluid)

Deepmagmatic

fluidexsolution

3

41 1 wt.%

5 wt.%

10 w

t.%

20 w

t.%

300°C/km

geotherm

+Halite

V/L+Halite

Shallowmagmatic

fluidexsolution

2

Deepsingle-phasefluid

Depth (km

)

1 (hyd)

2 (hyd)1 (lith)

2 (lith)

8 (hyd) 3 (lith)

4 (lith)

5 (lith)

6 (lith)

7 (lith)

8 (lith)

3 (hyd)

4 (hyd)

5 (hyd)

6 (hyd)

7 (hyd)

9 (hyd)

10 (hyd)

0600 700 800 900 1000

600 700 800 900 1000

11 (hyd)

12 (hyd)

13 (hyd)

14 (hyd)

15 (hyd)

pour–liquid (V–L) solvi drawn for 1, 5, 10, and 20 wt.% NaCl (data from Driesner, 2007;urves that it is a liquid; the transition point corresponds to the critical point for thataverage crustal (30 °C/km) and high (300 °C/km, near active volcanism) geothermal

hydrostatic (hyd) and lithostatic (lith) pressure conditions. Typical temperature–depthter main-stage brittle porphyry veins, and epithermal mineralization are indicated. Four) a shallow high-temperature path boiling to dryness (V+Halite field); (3) the deepdensation path of Heinrich et al. (2004).

Page 17: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

Crit

ical

cur

ve

L+H

(40

0°C

)V+H (400°C)

V+L(700°C)

V

wt.% NaCllog (wt.% NaCl)

P (

bars

)

scalechange

100

0

200

300

400

500

600

700

800

900

1000

1100

1200

1300700°C

600°C

500°C

400°C

Singlephasemag-maticfluid

V+L(600°C)

V+L(400°C)

V+L(500°C)

V+H (700°C)

V+H (600°C)

V+H (500°C)

L+H

(50

0°C

)

L+H

(70

0°C

)L+H

(60

0°C

)

Critical Pof H2O L

(400

°C)

L (5

00°C

)

L (7

00°C

)

L (6

00°C

)-8 -7 -6 -5 -4 -3 -2 -1 0/1 10 20 30 40 50 60 70 80 90 100

100

0

200

300

400

500

600

700

800

900

1000

1100

1200

1300-8 -7 -6 -5 -4 -3 -2 -1 0 10 20 30 40 50 60 70 80 90 100

V-L solvus (liquid)

V-L solvus (vapour)

V-H solvus (vapour)

Fig. 10. Pressure–XNaCl section through the H2O–NaCl phase diagram, with vapor–liquid (V–L) solvi drawn at various temperatures (data from Driesner, 2007; Driesner and Heinrich,2007). Red curves indicate that the solvus phase is a vapor, blue curves that it is a liquid; below the vapor–halite (V–H) solvus, vapor curves are shown in orange. Note the scalechange on the salinity axis at 1 wt.% NaCl, in order to illustrate the extremely low salinity of low-temperature vapor phases. The range of salinity for typical deeply exsolved single-phase (supercritical) magmatic fluids (2–13 wt.% NaCl) is shown in gray. Fluids that intersect the V–L solvus above ~400°–500 °C will be moderate-density vapors and will condensea small amount of dense, saline liquid; fluids that intersect the solvus below this temperature will be liquids and will boil off a dilute, low-density vapor phase.

17J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

acidic (advanced argillic) alteration commonly found at shallow levelsabove porphyry systems, which is caused by acidic gasses (H2SO4,HCl) condensing from a low-density vapor plume. As Heinrich et al.(2004) noted, such acidic vapors do not appear to transport Au (or Cu)effectively because bisulfide (HS–) ligands are hydrolized to H2S. Thismay explain why advanced argillic alteration caps are commonlybarren (in terms of Au and Cu) and merely generate permeability thatpotentially focuses later ore-forming fluid flow. Thus, mineralizationmay only occur where later moderate salinity liquids have followed ahigher pressure, rather specialized cooling path, as described above(Heinrich et al., 2004).

4.2.2. Low-sulfidation epithermal Au deposits (including alkalic-typedeposits)

Although low-sulfidation epithermal Au deposits are commonlyfound in volcanic terrains, their link to magmatism is more tenuousthan high-sulfidation deposits, and there is commonly evidence for agreater involvement of meteoric groundwater in their formation thanmagmatic fluids (e.g., Faure et al., 2002; Field and Fifarek, 1985; Healdet al., 1987). Nevertheless, alkalic-type epithermal Au deposits, whichare mineralogically similar to low-sulfidation deposits (adularia andsericite — or roscoelite [vanadium mica] — are stable), do show astrong temporal and genetic relationship to alkalic magmas, typicallyin relatively small and isolated intrusive complexes located in back-arc or post-subduction settings (e.g., Jensen and Barton, 2000; Kelley

et al., 1998; Müller and Groves, 1993; Mutschler et al., 1985; Richards,1995; Thompson et al., 1985).

Fluids in these alkalic-type deposits are commonly low- tomoderate salinity (0–10 wt.% NaCl equiv.), low temperature (typically≤250 °C) liquids, with evidence for decompressional boiling or fluidmixing as the prime ore depositional mechanism in high gradebreccias and veins (Jensen and Barton, 2000; Richards, 1995). Stableisotopic compositions of these fluids are generally ambiguous, andpermit interpretations of the involvement of either isotopicallyexchanged meteoric waters or magmatic fluids, or both (Ahmad etal., 1987; Carman, 2003; Richards, 1995; Richards and Kerrich, 1993;Ronacher et al., 2004; Scherbarth and Spry, 2006; Zhang and Spry,1994). Similar stable isotopic data from other low-sulfidation depositsare commonly interpreted to reflect a meteoric fluid source because ofthe absence of clearly coeval magmatism (as noted above). However,the close link with magmatism in alkalic-type systems suggests that amagmatic fluid source is more likely (e.g., Carman, 2003; Scherbarthand Spry, 2006; Simmons and Brown, 2007), and the vaporcontraction model described in Section 4.2.1 could explain theobserved characteristics of these ore fluids (i.e., direct contraction toa moderate salinity liquid from a high temperature magmatic fluid atdepth). Because of the association of these deposits with relativelysmall intrusive complexes, and therefore a smaller crustal thermalanomaly, a shallower fluid P–T path with cooling at depth is morelikely, consistent with a model of vapor contraction.

Page 18: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

18 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

5. Summary and conclusions

5.1. Sources of magmas and metals

Magmatic–hydrothermal porphyry Cu±Mo±Au, Au,Mo, and Sn–Wdeposits (and related epithermal Au deposits), derive their metals fromtheir associatedmagmas.With the exception of porphyry Sn–Wdepositsthat are associated with crustally derived S-type granites, mostother deposits in this grouping are formed by calc-alkaline to mildlyalkaline I-type granitoids directly or indirectly related to subduction.Sources of these magmas include subduction-metasomatized astheno-spheric mantle wedge, basaltic oceanic crust and/or seafloor sediments,and, in post-subduction settings, subduction-modified upper platelithosphere. The majority of normal arc porphyry systems are generatedfrom hydrous mantle wedge melts that have interacted to varyingdegrees with the upper plate lithosphere during passage towards thesurface. Assimilation and fractional crystallization processes fundamen-tally change the composition of the primary basaltic arc magmas tointermediate calc-alkaline compositions, with fractionated trace elementpatterns and evolved (crustal) isotopic signatures.

Seafloor sediments and basaltic oceanic crust only melt underunusually hot subduction zone conditions or locally at plate edges, anddespite widespread claims of the identification of slab melts (adakites)in the literature based on high Sr/Y and La/Yb ratios in some evolvedgranitoids, this process is unlikely to be a major contributor to arcmagmatism and metallogeny. Rather, these subtle trace element ratioscan be readily explained by fractionation of hornblende±titanite andresidual garnet, and suppression of plagioclase fractionation fromwater-rich mantle wedge basaltic magmas. These trace elementcharacteristics are therefore an indicator of high magmatic watercontent rather than being a source signature, and this likely explains thecommon association of high-Sr/Y (i.e., hydrous) magmas with mag-matic–hydrothermal ore deposits — without magmatic water, suchdeposits cannot form.

Potential sources of metals in arc magmas include the oceanic crustand sediments (via dehydration fluids or melting), the mantle wedge,and theupper plate crust. Fluid–mobile elements suchasK,Rb, Cs, Ca, Sr,Ba, U, B, Pb, As, Sb, Tl, and possibly Cu, Au, Re, and the Pd-groupelements, alongwith large amounts ofH2O, Cl, and S, arefluxed from thedehydrating subducting slab into the mantle wedge, causing metaso-matism and partialmelting by lowering the peridotite solidus. Althoughthere is some evidence for slab-derived fluid contributions of chalco-phile and highly siderophile elements (Cu, Au, PGE) to the mantlewedge, it is not clear that this is a necessarymetallogenic step, the uppermantle already containing significant amounts of these elements. Likely,a more important control on the metal content of subsequent partialmelts is the abundance and stability of residual sulfide phases in theasthenospheric mantle source. Under the high fO2 and fS2 conditions ofarcmagmatism, sulfur will be dominantly present as sulfate and sulfate,but saturation in small amounts of sulfide phases is also likely. Thesesulfide phases will tend to deplete the magma in highly siderophileelements (Au and PGE), but will not be present in sufficient volume tosignificantly deplete themagma inmore abundant chalcophile elementssuch as Cu and Mo. Such magmas therefore have the potentialsubsequently to form porphyry Cu–Mo deposits. Thus, Cu (and perhapsMo) are thought to be predominantly derived from the mantle, plus orminus contributions from the subducting slab. Gold and silverpresent inminor amounts in such deposits may also be derived from subductionsources, although there is some evidence for additional contributionsfrom the upper plate crust (especially for Ag, and also perhaps Mo).

Arc-like magmas and related porphyry and epithermal ore depositsalso occur in post-subduction tectonic settings, such as subductionreversal or migration, arc collision, continent–continent collision, andpost-collisional rifting. They are distinguished from normal subduction-related suites by slightly higher magmatic alkali (K2O and Na2O)contents, and by the occurrence of Au-rich deposits (although normal

porphyryCu–Modeposits can alsooccur). Suchmagmatic–metallogenicsystems are thought to form by remelting of previously subduction-modified upper plate lithosphere, and in particular the lower crustalamphibolitic cumulate roots of former arc magmatic complexes.Remelting can be triggered by crustal thickening and thermal reboundfollowing arc or continent collision, delamination of sub-continentalmantle lithosphere causing direct exposure of the lower crust toasthenospheric temperatures and melts, and asthenospheric upwellingduring rifting of former arc crust. Sparse sulfide phases in these arccumulates, residual from fractionation of previous arc magmas, willlikely be rich in chalcophile and highly siderophile elements. Duringlow-volumemelting under relatively low fS2 conditions (in the absenceof a flux of S from active subduction), these sulfide phases will likelyredissolve in the mildly alkaline partial melt, and may provide a sourcefor Au-rich (±PGE) post-subduction porphyry Cu–Au and epithermalAu systems: examples include the Roşia Montană, Skouries, Kisladag,Çöpler, and Sari Gunay porphyry Cu–Au and epithermal Au deposits inthe Neo-Tethyan belt of Romania, Greece, Turkey, and Iran, and theGrasberg, Ok Tedi, Porgera, Lihir, and Emperor porphyry Cu–Au andepithermal Au deposits in the southwest Pacific. However, goldenrichments may not occur where more abundant sulfides werepresent in the former arc complex, leading to more “normal” porphyryCu±Mo systems: examples include the Kerman porphyry Cu belt ofcentral Iran, and the Gangdese porphyry Cu belt of Tibet.

Porphyry Mo and Sn–W deposits associated with felsic magmas incontinental interiors are thought to form mainly by partial melting ofcontinental crust during rifting to form S-type, lithophile element-richgranitic magmas. A role for mafic, mantle-derivedmagmas is suggestedby the common association with such rocks, but their role may bepredominantly as a heat source for crustal melting and a trigger forvolatile saturation and eruption, rather than as a unique source ofmetals.

5.2. Porphyry and epithermal ore formation

In the porphyry and epithermal ore depositional environment, acritical role is played by aqueous fluids exsolving from hydrousmagmas emplaced in themid- to upper crust. The P–T–X properties ofmagmatic hydrothermal fluids, approximated by the H2O–NaClsystem, combinedwith volatile solubility in intermediate compositionmagmas, suggest that fluid saturation occurs at depths of 5–10 km inthe batholithic roots of arc magmatic systems. Volatile exsolutionmaylead to the upward propagation of buoyant bubbly magma as dikesand stocks intruded into the overlying shallow crust (with or withoutsubsequent eruption at surface), and/or the rapid ascent of a separatevolatile plume. Structural focusing and chanelling of these evolvedmagmas and fluids creates a cupola zone characterized by highthermal gradients and fluid flux. Metals (Cu, Mo, Au), which partitionstrongly into the saline (2–13 wt.% NaCl equivalent) and S-richmagmatic hydrothermal phase at high P and T in the underlyingbatholithic magma chamber, experience rapid reduction in solubilityas these fluids ascend, depressurize, and cool, with the bulk of Cu andMo being precipitated over a temperature range of 425°–320 °C at 1–6 km (commonly ≤2 km) depth.

This depth–temperature interval is critical because it also repre-sents: (1) the upward transition from ductile to brittle behavior in thecover rocks (~400°–350 °C), which facilitates fracturing and rapid fluiddepressurization; (2) a window of retrograde silica solubility (~550°–350 °C), which enhances permeability and porosity for ore deposition;and (3) the temperature range (b400 °C) over which SO2 in the fluidphase begins to disproportionate to H2S and H2SO4, which causesprecipitation of sulfideminerals (chalcopyrite, pyrite,molybdenite, rarebornite). This reaction also generates increasingly acidic fluids, leadingto the characteristic progression from feldspar-stable alteration assem-blages (potassic), through muscovite/sericite-stable assemblages

Page 19: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

19J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

(phyllic), to clay- (argillic) and alunite-stable assemblages (advancedargillic).

Depending on the depth of exsolution, the initialmagmatic fluidwilleither be a supercriticalfluid (below~6 kmdepth) orwill exist as a two-phase moderate salinity vapor and high density brine. As the plumeascends, it will intersect its solvus (in the case of a deeply exsolvedsupercritical fluid), and the vapor phase will become progressively lesssaline through brine condensation. At the level of porphyry oreformation, both the brine and vapor phase may contribute to metaltransport and deposition, although there is increasing evidence for theimportance of the vapor phase as a large-volume, highly upwardlymobile transportationmedium. However, as this vapor phase continuesto ascend, it will rapidly decrease in salinity, such that chloride-complexed metals are unlikely to be transported by such fluids toshallow epithermal levels. It will also increase in acidity, therebyreducing the solubility of bisulfide-complexed metals such as Au (andperhaps also Cu) by protonation of HS– to H2S. This acidic vapor phase isresponsible for the extreme acid leaching in advanced argillic lithocapsabove porphyry systems.

High-sulfidation epithermal Cu–Au deposits are hosted by thesehighly permeable advanced argillic alteration zones, but appear to havebeen formed by later, moderately saline (0.2 to 4.5 wt.% NaClequivalent), less acidic liquids. Two models have been proposed toexplain the origin of these paragenetically late mineralizing fluids interms of contraction of a single phase (supercritical) magmatic fluid,with (Heinrich et al., 2004) or without (Hedenquist et al., 1998) briefintersection of the solvus. Because of the topology of the P–T–XNaCl

phase diagram, such fluids contract to a liquid phase upon cooling atpressure. Thus, these authors propose that high-sulfidation epithermalCu–Au deposits may be formed directly from late-stage magmatichydrothermal fluids.

Although low sulfidation epithermal Au deposits have not beendiscussed in detail in this paper, because most such deposits are notdirectly related to porphyry-type magmatic–hydrothermal systems,the vapor contraction mechanism might have applicability to alkalic-type epithermal gold deposits, which do show a close geneticrelationship to post-subduction alkalic magmas.

Acknowledgments

I would like to thank Nigel Cook and Timothy Horscroft for invitingme to submit this paper. Reviewpapers, by their nature, drawheavily onthe work of others, and I would particularly like to acknowledge thefollowing people who have influenced my thinking on this subject: P.Candela, J. Cline, J. Dilles, J. Hedenquist, C. Heinrich, R. Sillitoe, and R.Tosdal. An anonymous reviewer is thanked for helpful and constructivecomments. Thisworkwas funded by aDiscovery Grant from theNaturalSciences and Engineering Research Council of Canada.

References

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

Aerts, M., Hack, A.C., Reusser, E., Ulmer, P., 2010. Assessment of the diamond-trapmethod for studying high-pressure fluids and melts and an improved freezingstage design for laser ablation ICP-MS analysis. American Mineralogist 95,1523–1526.

Ahmad, S.N., Rose, A.W., 1980. Fluid inclusions in porphyry and skarn ore at Santa Rita,New Mexico. Economic Geology 75, 229–250.

Ahmad, M., Solomon, M., Walshe, J.L., 1987. Mineralogical and geochemical studies ofthe Emperor gold telluride deposit, Fiji. Economic Geology 82, 345–370.

Aitcheson, S.J., Harmon, R.S., Moorbath, S., Schneider, A., Soler, P., Soria-Escalante, E.,Steele, G., Swainbank, I., Wörner, G., 1995. Pb isotopes define basement domains ofthe Altiplano, central Andes. Geology 23, 555–558.

Aizawa, Y., Tatsumi, Y., Yamada, H., 1999. Element transport by dehydration ofsubducted sediments: implications for arc and ocean island magmatism. Island Arc8, 38–46.

Alonso-Perez, R., Müntener, O., Ulmer, P., 2009. Igneous garnet and amphibolefractionation in the roots of island arcs: experimental constraints on andesiticliquids. Contributions to Mineralogy and Petrology 157, 541–558.

Alt, J.C., Shanks, W.C., Jackson, M.C., 1993. Cycling of sulfur in subduction zones: thegeochemistry of sulfur in the Mariana Island Arc and back-arc trough. Earth andPlanetary Science Letters 119, 477–494.

Anderson, E.D., Atkinson Jr., W.W., Marsh, T., Iriondo, A., 2009. Geology andgeochemistry of the Mammoth breccia pipe, Copper Creek mining district,southeastern Arizona: evidence for a magmatic–hydrothermal origin. MineraliumDeposita 44, 151–170.

Annen, C., Blundy, J.D., Sparks, R.S.J., 2006. The genesis of intermediate and silicicmagmas in deep crustal hot zones. Journal of Petrology 47, 505–539.

Arculus, R.J., 1994. Aspects of magma genesis in arcs. Lithos 33, 189–208.Arribas Jr., A., 1995. Characteristics of high-sulfidation epithermal deposits, and their

relation to magmatic fluid. In: Thompson, J.F.H. (Ed.), Magmas, fluids, and oredeposits: Mineralogical Association of Canada Short Course, 23, pp. 419–454.

Audétat, A., 2010. Source and evolution of molybdenum in the porphyry Mo(–Nb)deposit at Cave Peak, Texas. Journal of Petrology 51, 1739–1760.

Audétat, A., Keppler, H., 2005. Solubility of rutile in subduction zone fluids, asdetermined by experiments in the hydrothermal diamond anvil cell. Earth andPlanetary Science Letters 232, 393–402.

Audétat, A., Pettke, T., 2003. The magmatic-hydrothermal evolution of two barrengranites: a melt and fluid inclusion study of the Rito del Medio and Canada Pinabeteplutons in northern New Mexico (USA). Geochimica et Cosmochimica Acta 67,97–121.

Audétat, A., Pettke, T., 2006. Evolution of a porphyry-Cu mineralized magma system atSanta Rita, New Mexico (USA). Journal of Petrology 47, 2021–2046.

Audétat, A., Günther, D., Heinrich, C.A., 2000. Magmatic–hydrothermal evolution in afractionating granite: a microchemical study of the Sn–W–F-mineralized MoleGranite (Australia). Geochimica et Cosmochimica Acta 64, 3373–3393.

Audétat, A., Pettke, T., Heinrich, C.A., Bodnar, R.J., 2008. The composition of magmatic-hydrothermal fluids in barren and mineralized intrusions. Economic Geology 103,877–908.

Ballhaus, C., 1993. Redox states of lithospheric and asthenospheric upper mantle.Contributions to Mineralogy and Petrology 114, 331–348.

Barbarin, B., 1996. Genesis of the two main types of peraluminous granitoids. Geology24 (4), 295–298.

Barbier, E., 2002. Geothermal energy technology and current status: an overview.Renewable and Sustainable Energy Reviews 6, 3–65.

Barnes, S.-J., Naldrett, A.J., Gorton, M.P., 1985. The origin of the fractionation ofplatinum-group elements in terrestrial magmas. Chemical Geology 53, 303–323.

Barreiro, B.A., 1984. Lead isotopes and Andean magmagenesis. In: Harmon, R.S.,Barreiro, B.A. (Eds.), Andean Magmatism Chemical and Isotopic Constraints.Cheshire, Shiva, Nantwich, pp. 21–30.

Beard, J.S., Lofgren, G.E., 1991. Dehydration melting and water-saturated melting ofbasaltic and andesitic greenstones and amphibolites at 1, 3, and 6.9 kb. Journal ofPetrology 32, 365–401.

Bergantz, G.W., Dawes, R., 1994. Aspects of magma generation and ascent incontinental lithosphere. In: Ryan, M.P. (Ed.), Magmatic Systems. AcademicPress, San Diego, pp. 291–317.

Berger, J., Caby, R., Liégeois, J.-P., Mercier, J.-C.C., Demaiffe, D., 2009. Dehydration,melting and related garnet growth in the deep root of the AmalaoulaouNeoproterozoic magmatic arc (Gourma, NE Mali). Geological Magazine 146,173–186.

Blatter, D.L., Carmichael, I.S.E., 1998. Hornblende peridotite xenoliths from centralMexico reveal the highly oxidized nature of subarc upper mantle. Geology 26,1035–1038.

Blevin, P.L., Chappell, B.W., 1992. The role of magma sources, oxidation states andfractionation in determining the granite metallogeny of eastern Australia. Trans-actions of the Royal Society of Edinburgh, Earth Sciences 83, 305–316.

Blevin, P.L., Chappell, B.W., Allen, C.M., 1996. Intrusive metallogenic provinces ineastern Australia based on granite source and composition. Geological Society ofAmerica, Special Paper 315, 281–290.

Blundy, J., Cashman, K.V., Rust, A., Witham, F., 2010. A case for CO2-rich arc magmas.Earth and Planetary Science Letters 290, 289–301.

Bockrath, C., Ballhaus, C., Holzheid, A., 2004. Fractionation of the platinum-groupelements during mantle melting. Science 305, 1951–1953.

Bodnar, R.J., Beane, R.E., 1980. Temporal and spatial variations in hydrothermal fluidcharacteristics during vein filling in preore cover overlying deeply buried porphyrycopper-type mineralization at Red Mountain, Arizona. Economic Geology 75,876–893.

Borisov, A., Palme, H., 1997. Experimental determination of the solubility of platinum insilicate melts. Geochimica et Cosmochimica Acta 61, 4349–4357.

Bourdon, B., Turner, S., Dosseto, A., 2003. Dehydration and partial melting in subductionzones: constraints from U-series disequilibria. Journal Geophysical Research 108,B6. doi:10.1029/2002JB001839.

Bouse, R.M., Ruiz, J., Titley, S.R., Tosdal, R.M., Wooden, J.L., 1999. Lead isotopecompositions of Late Cretaceous and Early Tertiary igneous rocks and sulfideminerals in Arizona: implications for the sources of plutons and metals in porphyrycopper deposits. Economic Geology 94, 211–244.

Brandon, A.D., Draper, D.S., 1996. Constraints on the origin of the oxidation state ofmantle overlying subduction zones: an example from Simcoe, Washington, USA.Geochimica et Cosmochimica Acta 60, 1739–1749.

Breeding, C.M., Ague, J.J., Bröcker, M., 2004. Fluid–metasedimentary rock interactions insubduction-zone mélange: implications for the chemical composition of arcmagmas. Geology 32, 1041–1044.

Page 20: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

20 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Brenan, J.M., Shaw, H.F., Phinney, D.L., Ryerson, F.J., 1994. Rutile–aqueous fluidpartitioning of Nb, Ta, Hf, Zr, U and Th: implications for high field strength elementdepletions in island arc basalts. Earth and Planetary Science Letters 128, 327–339.

Brown, M., 2010. Melting of the continental crust during orogenesis: the thermal,rheological, and compositional consequences of melt transport from lower toupper continental crust. Canadian Journal of Earth Sciences 47, 655–694.

Burnham, C.W., 1979. Magmas and hydrothermal fluids, In: Barnes, H.L. (Ed.),Geochemistry of Hydrothermal Ore Deposits, 2nd edition. John Wiley and Sons,New York, pp. 71–136.

Burnham, C.W., 1981. Convergence and mineralization — is there a relation? Geol. Soc.America Memoir 154, 761–768.

Burnham, C.W., 1997. Magmas and hydrothermal fluids, In: Barnes, H.L. (Ed.),Geochemistry of Hydrothermal Ore Deposits, 3rd edition. John Wiley and Sons,New York, pp. 63–123.

Campbell, I.H., Naldrett, A.J., 1979. The influence of silicate:sulfide ratios on thegeochemistry of magmatic sulfides. Economic Geology 74, 1503–1506.

Candela, P.A., 1989a. Felsic magmas, volatiles, and metallogenesis. In: Whitney, J.A.,Naldrett, A.J. (Eds.), Ore Deposition Associated with Magmas: Soc. Econ. Geol.,Reviews in Economic Geology, 4, pp. 223–233.

Candela, P.A., 1989b. Calculation of magmatic fluid contributions to porphyry-type oresystems: predicting fluid inclusion chemistries. Geochemical Journal 23, 295–305.

Candela, P.A., 1989c. Magmatic ore-forming fluids: thermodynamic and mass transfercalculationsofmetal concentrations. In:Whitney, J.A.,Naldrett,A.J. (Eds.),OreDepositionAssociatedwithMagmas: Soc. Econ.Geol., Reviews inEconomicGeology, 4, pp. 203–221.

Candela, P.A., 1992. Controls on ore metal ratios in granite-related ore systems: anexperimental and computational approach. Transactions of the Royal Society ofEdinburgh, Earth Sciences 83, 317–326.

Candela, P.A., Holland, H.D., 1984. The partitioning of copper and molybdenumbetween silicate melts and aqueous fluids. Geochimica et Cosmochimica Acta 48,373–380.

Candela, P.A., Holland, H.D., 1986. A mass transfer model for copper and molybdenumin magmatic hydrothermal systems: the origin of porphyry-type ore deposits.Economic Geology 81, 1–19.

Candela, P.A., Piccoli, P.M., 1995. Model ore—metal partitioning frommelts into vapor andbapor/brine mixtures. In: Thompson, J.F.H. (Ed.), Magmas, Fluids, and Ore Deposits. :Short Course Series, 23. Mineralogical Association of Canada, pp. 101–127. ch. 5.

Candela, P.A., Piccoli, P.M., 2005. Magmatic processes in the development of porphyry-type ore systems. In: Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P.(Eds.), Economic Geology 100th Anniversary Volume. Society of EconomicGeologists, Littleton, CO, pp. 25–37.

Canil, D., Styan, J., Larocque, J., Bonnet, E., Kyba, J., 2010. Thickness and composition ofthe Bonanza arc crustal section, Vancouver Island, Canada. Geological Society ofAmerica, Bulletin 122, 1094–1105.

Carman, G.D., 2003. Geology, mineralization, and hydrothermal evolution of theLadolam gold deposit, Lihir Island, Papua New Guinea. In: Simmons, S.F., Graham, I.(Eds.), Volcanic, Geothermal, and Ore-Forming Fluids: Rulers and Witnesses ofProcesses Within the Earth. Special Publication, No. 10. Society of EconomicGeologists, pp. 247–284.

Carrigan, C.R., Schubert, G., Eichelberger, J.C., 1992. Thermal and dynamical regimes ofsingle- and two-phase magmatic flow in dikes. Journal of Geophysical Research 97(17), 392 377-17.

Carroll, M.R., Rutherford, M.J., 1985. Sulfide and sulfate saturation in hydrous silicatemelts. Proceedings of the Fifteenth Lunar and Planetary Science Conference, Part 2:Journal of Geophysical Research, 90, pp. C601–C612. Supplement.

Carten, R.B., White, W.H., Stein, H.J., 1993. High-grade granite-related molybdenumsystems. In: Kirham, R.V., Sinclair,W.D., Thorpe, R.I., Duke, J.M. (Eds.), Mineral DepositModeling. Special Paper, 40. Geological Association of Canada, pp. 521–554.

Cathles, L.M., 1991. The importance of the 350 °C isotherm in ore-forming hydrother-mal systems. Geological Society of America, Annual Meeting, October 21–24, 1991.San Diego, CA, Abstracts with Programs 23, p. A21.

Cauzid, J., Philippot, P., Martinez-Criado, G., Ménez, B., Labouré, S., 2007. Contrasting Cu-complexing behaviour in vapour and liquid fluid inclusions from the Yankee Lodetin deposit, Mole Granite, Australia. Chemical Geology 246, 39–54.

Černý, P., Blevin, P.L., Cuney, M., London, D., 2005. Granite-related ore deposits. In:Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), EconomicGeology 100th Anniversary Volume. Society of Economic Geologists, Littleton, CO,pp. 337–370.

Cervantes, P., Wallace, P.J., 2003. Role of H2O in subduction-zone magmatism: newinsights from melt inclusions in high-Mg basalts from central Mexico. Geology 31,235–238.

Chaplygin, I.V.,Mozgova, N.N.,Mokhov, A.V., Koporulina, E.V., Bernhardt, H.-J., Bryzgalov,I.A., 2007. Minerals of the system ZnS–CdS from fumaroles of the Kudriavy Volcano,Iturup Island, Kuriles, Russia. Canadian Mineralogist 45, 709–722.

Chappell, B.W., White, A.J.R., 1974. Two contrasting granite types. Pacific Geology 8,173–174.

Chiaradia, M., Fontboté, L., Paladines, A., 2004. Metal sources in mineral deposits andcrustal rocks of Ecuador (1°N–4°S): a lead isotope synthesis. Economic Geology 99,1085–1106.

Claeson, D.T., Meurer, W.P., 2004. Fractional crystallization of hydrous basaltic arc-typemagmas and the formation of amphibole-bearing gabbroic cumulates. Contributionsto Mineralogy and Petrology 147, 288–304.

Clemens, J.D., 2003. S-type granitic magmas — petrogenetic issues, models andevidence. Earth-Science Reviews 61, 1–18.

Clemens, J.D., Darbyshire, D.P.F., Flinders, J., 2009. Sources of post-orogenic calcalkalinemagmas: the Arrochar and Garabal Hill–Glen Fyne complexes, Scotland. Lithos 112,524–542.

Cline, J.S., 1995. Genesis of porphyry copper deposits: the behavior of water, chloride, andcopper in crystallizing melts. In: Pierce, F.W., Bolm, J.G. (Eds.), Porphyry CopperDeposits of the American Cordillera: Arizona Geological Society Digest, 20, pp. 69–82.

Cline, J.S., Bodnar, R.J., 1991. Can economic porphyry copper mineralization begenerated by a typical calc-alkaline melt? Journal of Geophysical Research 96,8113–8126.

Cloos, M., 2001. Bubbling magma chambers, cupolas, and porphyry copper deposits.International Geology Review 43, 285–311.

Cloos, M., Sapiie, B., van Ufford, A.Q., Weiland, R.J., Warren, P.Q., McMahon, T.P., 2005.Collisional delamination in New Guinea: the geotectonics of subducting slabbreakoff. Special Paper, 400. Geological Society of America. 51 pp.

Comment on Defant, M.J., Kepezhinskas, P., 2001Conrey, R.M., 2002. Adakites: a reviewof slab melting over the past decade and the case for a slab-melt component in arcs[EOS, Trans. AGU 82, 65–69]. EOS, Trans. AGU 83, 256.

Coumou, D., Driesner, T., Heinrich, C.A., 2008. Heat transport at boiling, near-criticalconditions. Geofluids 8, 208–215.

Crerar, D.A., Barnes, H.L., 1976. Ore solution chemistry V. Solubilities of chalcopyrite andchalcocite assemblages in hydrothermal solution at 200 °C to 350 °C. EconomicGeology 71, 772–794.

Cygan, G.L., Candela, P.A., 1995. Preliminary study of gold partitioning among pyrrhotite,pyrite, magnetite, and chalcopyrite in gold-saturated chloride solutions at 600 to700 °C, 140 MPa (1400 bars). In: Thompson, J.F.H. (Ed.), Magmas, Fluids, and OreDeposits. : Short Course Series, 23. Mineralogical Association of Canada, pp. 129–137.

Dale, C.W., Burton, K.W., Pearson, D.G., Gannoun, A., Alard, O., Argles, T.W., Parkinson,I.J., 2009. Highly siderophile element behaviour accompanying subduction ofoceanic crust: whole rock and mineral-scale insights from a high-pressure terrain.Geochimica et Cosmochimica Acta 73, 1394–1416.

Damon, P.E., 1986. Batholith-volcano coupling in the metallogeny of porphyry copperdeposits. In: Friedrich, G.H. (Ed.), Geology and Metallogeny of Copper Deposits.Springer-Verlag, Berlin, pp. 216–234.

Darbyshire, D.P.F., Shepherd, T.J., 1994. Nd and Sr isotope constraints on the origin ofthe Cornubian batholith, SW England. Journal of the Geological Society, London151, 795–802.

Davidson, J.P., 1996. Deciphering mantle and crustal signatures in subduction zonemagmatism. In: Bebout, G.E., Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction: Topto bottom. : Geophysical Monograph, 96. American Geophysical Union, pp. 251–262.

Davidson, J., Turner, S., Handley, H., Macpherson, C., Dosseto, A., 2007. Amphibole“sponge” in arc crust? Geology 35, 787–790.

Davies, J.H., Stevenson, D., 1992. Physical model of source region of subduction zonevolcanics. Journal of Geophysical Research 97, 2037–2070.

de Hoog, J.C.M., Mason, P.R.D., van Bergen, M.J., 2001. Sulfur and chalcophile elementsin subduction zones: constraints from a laser ablation ICP-MS study of meltinclusions from Galunggung Volcano, Indonesia. Geochimica et Cosmochimica Acta65, 3147–3164.

DeBari, S.M., Coleman, R.G., 1989. Examination of the deep levels of an island arc:evidence from the Tonsina ultramafic–mafic assemblage, Tonsina, Alaska. Journalof Geophysical Research 94, 4373–4391.

Deditius, A.P., Utsunomiya, S., Ewing, R.C., Chryssoulis, S.L., Venter, D., Kesler, S.E., 2009.Decoupled geochemical behavior of As and Cu in hydrothermal systems. Geology37, 707–710.

Defant, M.J., Drummond, M.S., 1990. Derivation of some modern arc magmas bymelting of young subducted lithosphere. Nature 347, 662–665.

Defant,M.J., Kepezhinskas, P., 2001. Adakites: a reviewof slabmeltingover thepast decadeand the case for a slab-melt component in arcs. EOS, Trans. AGU 82 (65), 68–69.

Defant, M.J., Kepezhinskas, P., 2002. Reply to Comment by Conrey, R. [Adakites: areview of slab melting over the past decade and the case for a slab-melt componentin arcs: EOS, Trans. AGU 82, 65–69]. EOS, Trans AGU 83, 256–257.

DePaolo, D.J., 1981. Trace element and isotopic effects of combined wallrockassimilation and fractional crystallization. Earth and Planetary Science Letters 53,189–202.

Dietrich, A., Lehmann, B., Wallianos, A., Traxel, K., Palacios, C., 1999. Magma mixing inBolivian tin porphyries. Die Naturwissenschaften 86, 40–43.

Dilles, J.H., 1987. Petrology of the Yerington batholith, Nevada: evidence for evolutionof porphyry copper ore fluids. Economic Geology 82, 1750–1789.

Dreyer, B.M., Morris, J.D., Gill, J.B., 2010. Incorporation of subducted slab-derivedsediment and fluid in arc magmas: B–Be–10Be–εNd systematics of the Kurileconvergent margin, Russia. Journal of Petrology 51, 1761–1782.

Driesner, T., 2007. The system H2O–NaCl. Part II: Correlations for molar volume,enthalpy, and isobaric heat capacity from 0 to 1000 °C, 1 to 5000 bar, and 0 to 1XNaCl. Geochimica et Cosmochimica Acta 71, 4902–4919.

Driesner, T., Heinrich, C.A., 2007. The system H2O–NaCl. Part I: correlation formulae forphase relations in temperature–pressure–composition space from 0 to 1000 °C,0 to 5000 bar, and 0 to 1 XNaCl. Geochimica et Cosmochimica Acta 71, 4880–4901.

Drummond, M.S., Defant, M.J., Kepezhinskas, P.K., 1996. Petrogenesis of slab-derivedtrondhjemite–tonalite–dacite/adakite magmas. Special Paper, 315. GeologicalSociety of America, pp. 205–215.

Dufek, J., Bergantz, G.W., 2005. Lower crustal magma genesis and preservation: astochastic framework for the evaluation of basalt–crust interaction. Journal ofPetrology 46, 2167–2195.

Duggen, S., Portnyagin, M., Baker, J., Ulfbeck, D., Hoernle, K., Garbe-Schönberg, D.,Grassineau,N., 2007.Drastic shift in lavageochemistry in thevolcanic-front to rear-arcregion of the SouthernKamchatkan subduction zone: evidence for the transition fromslab surface dehydration to sediment melting. Geochimica et Cosmochimica Acta 71,452–480.

Dunn, J.C., Hardee, H.C., 1981. Superconvecting geothermal zones. Journal ofVolcanology and Geothermal Research 11, 189–201.

Page 21: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

21J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Dvir, O., Pettke, T., Fumagalli, P., Kessel, R., 2011. Fluids in the peridotite–water systemup to 6 GPa and 800 °C: new experimental constrains on dehydration reactions.Contributions to Mineralogy and Petrology 161, 829–844.

Eastoe, C.J., 1982. Physics andchemistryof thehydrothermal systemat thePangunaporphyrycopper deposit, Bougainville, Papua New Guinea. Economic Geology 77, 127–153.

Economou-Eliopoulos, M., Eliopoulos, D.G., 2000. Palladium, platinum and goldconcentration in porphyry copper systems of Greece and their genetic significance.Ore Geology Reviews 16, 59–70.

Edwards, K.J., 2004. Formation and degradation of seafloor hydrothermal sulfidedeposits. Geological Society of America Special Papers 379, 83–96.

Eggins, S.M., 1993. Origin and differentiation of picritic arc magmas, Ambae (Aoba),Vanuatu. Contributions to Mineralogy and Petrology 114, 79–100.

Eichelberger, J.C., 1995. Silicic volcanism: ascent of viscous magmas from crustalreservoirs. Annual Review Earth Planet Science 23, 41–63.

Farmer, G.L., DePaolo, D.J., 1984. Origin of Mesozoic and Tertiary granite in the westernUnited States and implications for pre-Mesozoic crustal structure, 2. Nd and Srisotopic studies of unmineralized and Cu- and Mo-mineralized granite in thePrecambrian craton. Journal of Geophysical Research 89, 10141–10160.

Faure, K., Matsuhisa, Y., Metsugi, H., Mizota, C., Hayashi, S., 2002. The Hishikari Au–Agepithermal deposit, Japan: oxygen and hydrogen isotope evidence in determiningthe source of paleohydrothermal fluids. Economic Geology 97, 481–498.

Feeley, T.C., Davidson, J.P., 1994. Petrology of calc-alkaline lavas at Volcán Ollagüe andthe origin of compositional diversity at Central Andean stratovolcanoes. Journal ofPetrology 35, 1295–1340.

Field, C.W., Fifarek, R.H., 1985. Light stable-isotope systematics in the epithermalenvironment. In: Berger, B.R., Bethke, P.M. (Eds.), Geology and Geochemistry ofEpithermal Systems. Reviews in Economic Geology, 2. Society of EconomicGeologists, pp. 99–128.

Foley, S.F., Barth, M.G., Jenner, G.A., 2000. Rutile/melt partition coefficients for traceelements and an assessment of the influence of rutile on the trace elementcharacteristics of subduction zone magmas. Geochimica et Cosmochimica Acta 64,933–938.

Fontboté, L., Gunnesch, K.A., Baumann, A., 1990. Metal sources in stratabound oredeposits in the Andes (Andean Cycle) — lead isotopic constraints. In: Fontboté, L.,Amstutz, G.C., Cardozo, M., Cedillo, E., Frutos, J. (Eds.), Stratabound Ore Deposits inthe Andes. Springer-Verlag, Berlin, pp. 759–773.

Forneris, J.F., Holloway, J.R., 2003. Phase equilibria in subducting basaltic crust:implications for H2O release from the slab. Earth and Planetary Science Letters 214,187–201.

Fournier, R.O., 1985. The behavior of silica in hydrothermal solutions. In: Berger, B.R.,Bethke, P.M. (Eds.), Reviews in Economic Geology, 2. Society of EconomicGeologists, pp. 45–61.

Fournier, R.O., 1999. Hydrothermal processes related to movement of fluid from plasticinto brittle rock in the magmatic–epithermal environment. Economic Geology 94,1193–1212.

Fumagalli, P., Poli, S., 2005. Experimentally determined phase relations in hydrousperidotites to 6.5 GPa and their consequences on the dynamics of subductionzones. Journal of Petrology 46, 555–578.

Fyfe, W.S., 1992. Magma underplating of continental crust. Journal VolcanologyGeothermal Research 50, 33–40.

Garrido, C.J., Bodinier, J.-L., Burg, J.-P., Zeilinger, G., Hussain, S.S., Dawood, H., Chaudhry,M.N., Gervilla, F., 2006. Petrogenesis of mafic garnet granulite in the lower crust ofthe Kohistan paleo-arc complex (northern Pakistan): implications for intra-crustaldifferentiation of island arcs and generation of continental crust. Journal ofPetrology 47, 1873–1914.

Garrison, J.M., Davidson, J.P., 2003. Dubious case for slab melting in the Northernvolcanic zone of the Andes. Geology 31, 565–568.

Gill, J.B., 1981. Orogenic Andesites and Plate Tectonics. Springer-Verlag, New York.390 pp.

Gill, J., Whelan, P., 1989. Postsubduction ocean island alkali basalts in Fiji. Journal ofGeophysical Research 94, 4579–4588.

Goff, S.J., Goff, F., Janik, C.J., 1992. Tecuamburro volcano, Guatemala: explorationgeothermal gradient drilling and results. Geothermics 21, 483–502.

Green, T.H., Adam, J., 2003. Experimentally-determined trace element characteristics ofaqueous fluid from partially dehydrated mafic oceanic crust at 3.0 GPa, 650–700 °C.European Journal of Mineralogy 15, 815–830.

Greene, A.R., Debari, S.M., Kelemen, P.B., Blusztajn, J., Clift, P.D., 2006. A detailedgeochemical study of island arc crust: the Talkeetna Arc section, south-centralAlaska. Journal of Petrology 47, 1051–1093.

Grove, T.L., Elkins-Tanton, L.T., Parman, S.W., Chatterjee, N., Müntener, O., Gaetani, G.A.,2003. Fractional crystallization and mantle-melting controls on calc-alkalinedifferentiation trends. Contributions to Mineralogy and Petrology 145, 515–533.

Grove, T.L., Chatterjee, N., Parman, S.W., Médard, E., 2006. The influence of H2O onmantle wedge melting. Earth and Planetary Science Letters 249, 74–89.

Grunder, A.L., Klemetti, E.W., Feeley, T.C., McKee, C.M., 2008. Eleven million years of arcvolcanism at the Aucanquilcha Volcanic Cluster, northern Chilean Andes:implications for the life span and emplacement of plutons. Transactions: EarthSciences, 97. Royal Society of Edinburgh, pp. 415–436.

Guivel, C., Lagabrielle, Y., Bourgois, J., Martin, H., Arnaud, N., Fourcade, S., Cotten, J.,Maury, R.C., 2003. Very shallow melting of oceanic crust during spreading ridgesubduction: origin of near-trench Quaternary volcanism at the Chile Triple Junction.Journal of Geophysical Research 108 (B7), 2345. doi:10.1029/2002JB002119.

Guo, F., Wilson, M., Li, C., 2007. Post-collisional adakites in south Tibet: products ofpartial melting of subduction-modified lower crust. Lithos 96, 205–224.

Gutscher, M.-A., Maury, R., Eissen, J.-P., Bourdon, E., 2000. Can slabmelting be caused byflat subduction? Geology 28, 535–538.

Halter, W.E., Pettke, T., Heinrich, C.A., 2002. The origin of Cu/Au ratios in porphyry-typeore deposits. Science 296, 1844–1846.

Halter, W.E., Heinrich, C.A., Pettke, T., 2005. Magma evolution and the formation ofporphyry Cu–Au ore fluids: evidence from silicate and sulfide melt inclusions.Mineralium Deposita 39, 845–863.

Hamada, M., Fujii, T., 2008. Experimental constraints on the effects of pressure and H2Oon the fractional crystallization of high-Mg island arc basalt. Contributions toMineralogy and Petrology 155, 767–790.

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

Hansen, J., Skjerlie, K.P., Pedersen, R.B., De La Rosa, J., 2002. Crustal melting in the lowerparts of island arcs: an example from the Bremanger Granitoid Complex, westNorwegian Caledonides. Contributions to Mineralogy and Petrology 143, 316–335.

Harris, N.B.W., Pearce, J.A., Tindle, A.G., 1986. Geochemical characteristics of collisionzone magmatism. Special Publication, 19. Geological Society, London, pp. 67–81.

Harris, A.C., Kamenetsky, V.S., White, N.C., van Achterbergh, E., Ryan, C.G., 2003. Meltinclusions in veins: linking magmas and porphyry Cu deposits. Science 302,2109–2111.

Hart, C.J.R., Mair, J.L., Goldfarb, R.J., Groves, D.I., 2005. Source and redox controls onmetallogenic variations in intrusion-related ore systems, Tombstone-TungstenBelt, Yukon Territory. In: Ishihara, S., Stephens, W.E., Harley, S.L., Arima, M.,Nakajima, T. (Eds.), Fifth Hutton Symposium; the Origin of Granites and RelatedRocks. Special Paper, 389. Geological Society of America, pp. 339–356.

Haschke, M., Ben-Avraham, Z., 2005. Adakites from collision-modified lithosphere.Geophysical Research Letters 32 (15), L15302. doi:10.1029/2005GL023468.

Hattori, K.H., 1997. Occurrence and origin of sulfide and sulfate in the 1991 MountPinatubo eruption products. In: Newhall, C.G., Punongbayan, R.S. (Eds.), Fire andMud: Eruptions and Lahars of Mount Pinatubo, Philippines. University ofWashington Press, Seattle, pp. 807–824.

Hattori, K., Guillot, S., 2003. Volcanic fronts form as a consequence of serpentinedehydration in the forearc mantle wedge. Geology 31, 525–528.

Hattori, K.H., Keith, J.D., 2001. Contribution of mafic melt to porphyry coppermineralization: evidence from Mount Pinatubo, Philippines, and Bingham Canyon,Utah, USA. Mineralium Deposita 36, 799–806.

Hattori, K., Takahashi, Y., Guillot, S., Johanson, B., 2005. Occurrence of arsenic (V) inforearc mantle serpentinites based on X-ray absorption spectroscopy study.Geochimica et Cosmochimica Acta 69, 5585–5596.

Hattori, K., Wallis, S., Enami, M., Mizukami, T., 2010. Subduction of mantle wedgeperidotites: evidence from the Higashi-akaishi ultramafic body in the Sanbagawametamorphic belt. Island Arc 19, 192–207.

Hawkesworth, C.J., Gallagher, K., Hergt, J.M., McDermott, F., 1994. Destructive platemargin magmatism: geochemistry and melt generation. Lithos 33, 169–188.

Heald, P., Foley, N.K., Hayba, D.O., 1987. Comparative anatomy of volcanic-hostedepithermal deposits: acid-sulfate and adularia-sericite types. Economic Geology 82,1–26.

Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation ofhydrothermal ore deposits. Nature 370, 519–527.

Hedenquist, J.W., Richards, J.P., 1998. The influence of geochemical techniques on thedevelopment of genetic models for porphyry copper deposits. In: Richards, J.P.,Larson, P.B. (Eds.), Techniques in Hydrothermal Ore Deposits Geology: Reviews inEconomic Geology, 10, pp. 235–256. ch. 10.

Hedenquist, J.W., Simmons, S.F., Giggenbach, W.F., Eldridge, C.S., 1993. White Island,New Zealand, volcanic–hydrothermal system represents the geochemical environ-ment of high-sulfidation Cu and Au ore deposition. Geology 21, 731–734.

Hedenquist, J.W., Aoki, M., Shinohara, H., 1994a. Flux of volatiles and ore-formingmetals from the magmatic–hydrothermal system of Satsuma Iwojima volcano.Geology 22, 585–588.

Hedenquist, J.W., Matsuhisa, Y., Izawa, E., White, N.C., Giggenbach, W.F., Aoki, M.,1994b. Geology, geochemistry, and origin of high sulfidation Cu–Au mineralizationin the Nansatsu district, Japan. Economic Geology 89, 1–30.

Hedenquist, J.W., Arribas Jr., A., Reynolds, J.R., 1998. Evolution of an intrusion-centeredhydrothermal system: Far Southeast–Lepanto porphyry and epithermal Cu–Audeposits, Philippines. Economic Geology 93, 373–404.

Heinrich, C.A., 2005. The physical and chemical evolution of low-salinity magmaticfluids at the porphyry to epithermal transition: a thermodynamic study.Mineralium Deposita 39, 864–889.

Heinrich, C.A., Ryan, G.G., Mernagh, T.P., Eadington, P.J., 1992. Segregation of ore metalsbetween magmatic brine and vapor: a fluid inclusion study using PIXEmicroanalysis. Economic Geology 87, 1566–1583.

Heinrich, C.A., Günther, D., Audétat, A., Ulrich, T., Frischknecht, R., 1999. Metalfractionation between magmatic brine and vapor, determined by microanalysis offluid inclusions. Geology 27, 755–758.

Heinrich, C.A., Halter, W., Klemm, L., Landtwing, M.R., Pettke, T., 2003a. Laser ablationmicro-analysis of fluid and melt inclusions: towards understanding the process ofporphyry-style ore formation. Applied Earth Science (Transactions, Institutions ofMining and Metallurgy, Section B) 112, 185–186.

Heinrich, C.A., Pettke, T., Halter, W.E., Aigner-Torres, M., Audétat, A., Günther, D.,Hattendorf, B., Bleiner, D., Guillong, M., Horn, I., 2003b. Quantitative multi-elementanalysis of minerals, fluid and melt inclusions by laser-ablation inductively-coupled-plasma mass-spectrometry. Geochimica et Cosmochimica Acta 67,3473–3497.

Heinrich, C.A., Dreisner, T., Steffánson, A., Seward, T.M., 2004. Magmatic vaporcontraction and the transport of gold from the porphyry environment toepithermal ore deposits. Geology 32, 761–764.

Page 22: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

22 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Hemley, J.J., Hunt, J.P., 1992. Hydrothermal ore-forming processes in the light of studiesin rock-buffered systems: II. Some general geologic applications. Economic Geology87, 23–43.

Hemley, J.J., Cygan, G.L., Fein, J.B., Robinson, G.R., d'Angelo, W.M., 1992. Hydrothermalore-forming processes in the light of studies in rock-buffered systems: I. Iron–copper–zinc–lead sulfide solubility relations. Economic Geology 87, 1–22.

Henley, R.W., Hughes, G.O., 2000. Underground fumaroles: “excess heat” effects in veinformation. Economic Geology 95, 453–466.

Henley, R.W., McNabb, A., 1978. Magmatic vapor plumes and ground–water interactionin porphyry copper emplacement. Economic Geology 73, 1–20.

Hermann, J., Spandler, C.J., 2008. Sediment melts at sub-arc depths: an experimentalstudy. Journal of Petrology 49, 717–740.

Herzberg, C.T., Fyfe, W.S., Carr, M.J., 1983. Density constraints on the formation of thecontinental Moho and crust. Contributions to Mineralogy and Petrology 84, 1–5.

Herzig, P.M., Hannington, M.D., Scott, S.D., Maliotis, G., Rona, P.A., Thompson, G., 1991.Gold-rich sea-floor gossans in the Troodos Ophiolite and on the Mid-Atlantic Ridge.Economic Geology 86, 1747–1755.

Hildreth, W., 1981. Gradients in silicic magma chambers: implications for lithosphericmagmatism. Journal of Geophysical Research 86 (10), 192 153-10.

Hildreth, W., Moorbath, S., 1988. Crustal contributions to arc magmatism in the Andesof central Chile. Contributions to Mineralogy and Petrology 98, 455–489.

Holland, H.D., 1965. Some applications of thermochemical data to problems of oredeposits II. Mineral assemblages and the composition of ore forming fluids.Economic Geology 60, 1101–1166.

Holloway, J.R., 1976. Fluids in the evolution of granitic magmas: consequences of finiteCO2 solubility. Geological Society of America Bulletin 87, 1513–1518.

Hou, Z.Q., Zeng, P.S., Gao, Y.F., Dong, F.L., 2006. Himalayan Cu–Mo–Au mineralization inthe eastern Indo-Asian collision zone: constraints from Re–Os dating ofmolybdenite. Mineralium Deposita 41, 33–45.

Hou, Z., Yang, Z., Qu, X., Meng, X., Li, Z., Beaudoin, G., Rui, Z., Gao, Y., Zaw, K., 2009. TheMiocene Gangdese porphyry copper belt generated during post-collisionalextension in the Tibetan Orogen. Ore Geology Reviews 36, 25–51.

Huppert, H.E., Sparks, R.S.J., 1988. The generation of granitic magmas by intrusion ofbasalt into continental crust. Journal of Petrology 29, 599–624.

Ishihara, S., 1981. The granitoid series and mineralization. Economic Geology 75thAnniversary Volume, pp. 458–484.

Ishihara, S., Murakami, H., 2006. Fractionated ilmenite-series granites in southwest Japan:source magma for REE–Sn–W mineralizations. Resource Geology 56, 245–256.

Jagoutz, O., Muntener, O., Ulmer, P., Pettke, T., Burg, J.-P., Dawood, H., Hussain, S., 2007.Petrology and mineral chemistry of lower crustal intrusions: the Chilas Complex,Kohistan (NW Pakistan). Journal of Petrology 48, 1895–1953.

Jagoutz, O.E., Burg, J.-P., Hussain, S., Dawood, H., Pettke, T., Iizuka, T., Maruyama, S.,2009. Construction of the granitoid crust of an island arc part I: geochronologicaland geochemical constraints from the plutonic Kohistan (NW Pakistan). Contri-butions to Mineralogy and Petrology 158, 739–755.

James, D.E., 1982. A combined O, Sr, Nd, and Pb isotopic and trace element study ofcrustal contamination in central Andean lavas, I. Local geochemical variations.Earth and Planetary Science Letters 57, 47–62.

Jégo, S., Pichavant, M., Mavrogenes, J.A., 2010. Controls on gold solubility in arcmagmas: an experimental study at 1000 °C and 4 kbar. Geochimica et Cosmochi-mica Acta 74, 2165–2189.

Jenner, F.E., O'Neill, H., St, C., Arculus, R.J., Mavrogenes, J.A., 2010. The magnetite crisis inthe evolution of arc-related magmas and the initial concentration of Au, Ag and Cu.Journal of Petrology 51, 2445–2464.

Jensen, E.P., Barton, M.D., 2000. Gold deposits related to alkaline magmatism. In:Hagemann, S.G., Brown, P.E. (Eds.), Gold in 2000: Reviews in Economic Geology, 13,pp. 279–314.

John, D.A., 1991. Evolution of hydrothermal fluids in the Alta Stock, Central WasatchMountains, Utah. U.S. Geological Survey, Bulletin 1977. 51 pp.

Johnson, M.C., Plank, T., 1999. Dehydration and melting experiments constrain the fate ofsubducted sediments. Geochemistry Geophysics Geosystems 1, 1007. doi:10.1029/1999GC000014.

Johnson, R.W., Mackenzie, D.E., Smith, I.E.M., 1978. Delayed partial melting ofsubduction-modified mantle in Papua New Guinea. Tectonophysics 46, 197–216.

Jugo, P.J., 2009. Sulfur content at sulfide saturation in oxidized magmas. Geology 37,415–418.

Jugo, P.J., Candela, P.A., Piccoli, P.M., 1999. Magmatic sulfides and Au:Cu ratios inporphyry deposits: an experimental study of copper and gold partitioning at850 °C, 100 MPa in a haplogranitic melt–pyrrhotite–intermediate solid solution–gold metal assemblage, at gas saturation. Lithos 46, 573–589.

Jugo, P.J., Luth, R.W., Richards, J.P., 2005a. Experimental data on the speciation of sulfur as afunction of oxygen fugacity in basaltic melts. Geochimica et Cosmochimica Acta 69,497–503.

Jugo, P.J., Luth, R.W., Richards, J.P., 2005b. An experimental study of the sulfur content inbasaltic melts saturated with immiscible sulfide or sulfate liquids at 1300 °C and1.0 GPa. Journal of Petrology 46, 783–798.

Jugo, P.J., Wilke, M., Botcharnikov, R.E., 2010. Sulfur K-edge XANES analysis of naturaland synthetic basaltic glasses: implications for S speciation and S content asfunction of oxygen fugacity. Geochimica et Cosmochimica Acta 74, 5926–5938.

Kawamoto, T., 2006. Hydrous phases and water transport in the subducting slab.Reviews in Mineralogy and Geochemistry 62, 273–289.

Kay, R.W., 1978. Aleutian magnesian andesites: melts from subducted Pacific oceancrust. Journal of Volcanology and Geothermal Research 4, 117–132.

Kay, S.M., Ramos, V.A., Marquez, M., 1993. Evidence in Cerro Pampa volcanic rocks forslab-melting prior to ridge–trench collision in southern South America. Journal ofGeology 101, 703–714.

Kay, S.M., Mpodozis, C., Coira, B., 1999. Neogene magmatism, tectonism, and mineraldeposits of the Central Andes (22° to 33°S latitude). In: Skinner, B.J. (Ed.), Geologyand Ore Deposits of the Central Andes. : Special Publication, No. 7. Society ofEconomic Geologists, pp. 27–59.

Keith, J.D., Shanks III, W.C., Archibald, D.A., Farrar, E., 1986. Volcanic and intrusivehistory of the Pine Grove porphyry molybdenum system, southwestern Utah.Economic Geology 81, 553–577.

Keith, J.D., Whitney, J.A., Hattori, K., Ballantyne, G.H., Christiansen, E.H., Barr, D.L.,Cannan, T.M., Hook, C.J., 1997. The role of magmatic sulfides and mafic alkalinemagmas in the Bingham and Tintic mining districts, Utah. Journal of Petrology 38,1679–1690.

Keith, J.D., Christiansen, E.H., Maughan, D.T., Waite, K.A., 1998. The role of mafic alkalinemagmas in felsic porphyry-Cu and Mo systems. In: Lentz, D.R. (Ed.), MineralizedIntrusion-Related Skarn Systems: Mineralogical Association of Canada Short CourseSeries, 26, pp. 211–243.

Kelley, K.A., Cottrell, E., 2009. Water and the oxidation state of subduction zonemagmas. Science 325, 605–607.

Kelley, K.D., Romberger, S.B., Beaty, D.W., Pontius, J.A., Snee, L.W., Stein, H.J., Thompson,T.B., 1998. Geochemical and geochronological constraints on the genesis of Au–Tedeposits at Cripple Creek, Colorado. Economic Geology 93, 981–1012.

Kelley, K.A., Plank, T., Newman, S., Stolper, E.M., Grove, T.L., Parman, S., Hauri, E., 2010.Mantle melting as a function of water content beneath the Mariana Arc. Journal ofPetrology 51, 1711–1738.

Kemp, A.I.S., Hawkesworth, C.J., Foster, G.L., Paterson, B.A., Woodhead, J.D., Hergt, J.M.,Gray, C.M., Whitehouse, M.J., 2007. Magmatic and crustal differentiation history ofgranitic rocks from Hf–O isotopes in zircon. Science 315, 980–983.

Kennedy, A.K., Hart, S.R., Frey, F.A., 1990. Composition and isotopic constraints on thepetrogenesis of alkaline arc lavas: Lihir Island, Papua New Guinea. Journal ofGeophysical Research 95, 6929–6942.

Kent, A.J.R., Peate, D.W., Newman, S., Stolper, E.M., Pearce, J.A., 2002. Chlorine insubmarine glasses from the Lau Basin: seawater contamination and constraints onthe composition of slab-derived fluids. Earth and Planetary Science Letters 202,361–377.

Kepezhinskas, P., Defant, M.J., Widom, E., 2002. Abundance and distribution of PGE andAu in the island-arc mantle: implications for sub-arc metasomatism. Lithos 60,113–128.

Kerrich, R., Beckinsale, R.D., 1988. Oxygen and strontium isotopic evidence for theorigin of granites in the tin belt of southeast Asia. In: Taylor, R.P., Strong, D.F. (Eds.),Recent Advances in the Geology of Granite-Related Mineral Deposits: CanadianInstitute of Mining and Metallurgy, Special Volume 39, pp. 115–123.

Keskin, M., Genç, Ş.C., Tüysüz, O., 2008. Petrology and geochemistry of post-collisionalMiddle Eocene volcanic units in North-Central Turkey: evidence for magmageneration by slab breakoff following the closure of the Northern Neotethys Ocean.Lithos 104, 267–305.

Kesler, S.E., 1973. Copper, molybdenum and gold abundances in porphyry copperdeposits. Economic Geology 68, 106–112.

Kessel, R., Schmidt, M.W., Ulmer, P., Pettke, P., 2005a. Trace element signature ofsubduction-zone fluids, melts and supercritical liquids at 120–180 km depth.Nature 437, 724–727.

Kessel, R., Ulmer, P., Pettke, P., Schmidt, M.W., Thompson, A.B., 2005b. The water–basaltsystem at 4 to 6 GPa: phase relations and second critical endpoint in a K-freeeclogite at 700 to 1400 °C. Earth and Planetary Science Letters 237, 873–892.

Kilian, R., Behrmann, J.H., 2003. Geochemical constraints on the sources of SouthernChile Trench sediments and their recycling in arc magmas of the Southern Andes.Journal of the Geological Society 160, 57–70.

Kilian, R., Stern, C.R., 2002. Constraints on the interaction between slab melts and themantle wedge from adakitic glass in peridotite xenoliths. European Journal ofMineralogy 14, 25–36.

Kirkham, R.V., Sinclair, W.D., 1996. Porphyry copper, gold, molybdenum, tungsten, tin,silver. In: Eckstrand, O.R., Sinclair, W.D., Thorpe, R.I. (Eds.), Geology of Canadianmineral deposits. : Geology of Canada, 8. Geological Survey of Canada, pp. 421–446.

Klemm, L.M., Pettke, T., Heinrich, C.A., Campos, E., 2007. Hydrothermal evolution of theEl Teniente deposit, Chile: porphyry Cu–Mo ore deposit from low-salinitymagmatic fluids. Economic Geology 102, 1021–1045.

Klemm, L.M., Pettke, T., Heinrich, C.A., 2008. Fluid and source magma evolution of theQuesta porphyry Mo deposit, NewMexico, USA. Mineralium Deposita 43, 533–552.

Klemme, S., Prowatke, S., Hametner, K., Gunther, D., 2005. Partitioning of trace elementsbetween rutile and silicate melts: implications for subduction zones. Geochimica etCosmochimica Acta 69, 2361–2371.

Klepeis, K.A., Clarke, G.L., Rushmer, T., 2003. Magma transport and coupling betweendeformation and magmatism in the continental lithosphere. GSA Today 13, 4–11.

Kogiso, T., Tatsumi, Y., Nakano, S., 1997. Trace element transport during dehydrationprocesses in the subducted oceanic crust: 1. Experiments and implications for theorigin of ocean island basalts. Earth and Planetary Science Letters 148, 193–205.

Kontak, D.J., Cumming, G.L., Krstic, D., Clark, A.H., Farrar, E., 1990. Isotopic compositionof lead in ore deposits of the Cordillera Oriental, southeastern Peru. EconomicGeology 85, 1584–1603.

Kusakabe, M., Komoda, Y., Takano, B., Abiko, T., 2000. Sulfur isotopic effects in thedisproportionation reaction of sulfur dioxide in hydrothermal fluids: implicationsfor the ™34S variations of dissolved bisulfate and elemental sulfur from activecrater lakes. Journal of Volcanology and Geothermal Research 97, 287–307.

Kuscu, I., Kuscu, G.G., Tosdal, R.M., Ulrich, T.D., Friedman, R., 2010. Magmatism in thesoutheastern Anatolian orogenic belt: transition from arc to post-collisional setting inan evolving orogen. Special Publications, 340. Geological Society, London, pp. 437–460.

Kushiro, I., Syono, Y., Akimoto, S., 1968. Melting of a peridotite nodule at high pressuresand high water pressures. Journal of Geophysical Research 73, 6023–6029.

Page 23: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

23J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Landtwing, M.R., Pettke, T., Halter, W.E., Heinrich, C.A., Redmond, P.B., Einaudi, M.T.,Kunze, K., 2005. Copper deposition during quartz dissolution by cooling magmatic–hydrothermal fluids: the Bingham porphyry. Earth and Planetary Science Letters235, 229–243.

Landtwing, M.R., Furrer, C., Redmond, P.B., Pettke, T., Guillong, M., Heinrich, C.A., 2010.The Bingham Canyon porphyry Cu–Mo–Au deposit III. Zoned copper–gold oredeposition by magmatic vapor expansion. Economic Geology 105, 91–118.

Larocque, J., Canil, D., 2010. The role of amphibole in the evolution of arc magmas andcrust: the case from the Jurassic Bonanza arc section, Vancouver Island, Canada.Contributions to Mineralogy and Petrology 159, 475–492.

Larocque, A.C.L., Stimac, J.A., Siebe, C., Greengrass, K., Chapman, R., Mejia, S.R., 2008.Deposition of a high-sulfidation Au assemblage from amagmatic volatile phase, VolcánPopocatépetl, Mexico. Journal of Volcanology and Geothermal Research 170, 51–60.

Leeman, W.P., 1983. The influence of crustal structure on compositions of subduction-related magmas. Journal Volcanology and Geothermal Research 18, 561–588.

Lehmann, B., 1982. Metallogeny of tin: magmatic diferentiation versus geologicalheritage. Economic Geology 77, 50–59.

Lewis, K.C., Lowell, R.P., 2009. Numerical modeling of two-phase flow in the NaCl–H2Osystem:2.Examples. Journal ofGeophysical Research114. doi:10.1029/2008JB00603016 pp., B08204.

Lowell, J.D., Guilbert, J.M., 1970. Lateral and vertical alteration-mineralization zoning inporphyry copper ore deposits. Economic Geology 65, 373–408.

Lowenstern, J.B., 2001. Carbon dioxide in magmas and implications for hydrothermalsystems. Mineralium Deposita 36, 490–502.

Lowenstern, J.B., Mahood, G.A., Rivers, M.L., Sutton, S.R., 1991. Evidence for extremepartitioning of copper into a magmatic vapor phase. Science 252, 1405–1409.

Lynton, S.J., Candela, P.A., Piccoli, P.M., 1993. An experimental study of the partitioningof copper between pyrrhotite and a high silica rhyolitic melt. Economic Geology 88,901–915.

MacDonald, R., Hawkesworth, C.J., Heath, E., 2000. The Lesser Antilles volcanic chain: astudy in arc magmatism. Earth-Science Reviews 49, 1–76.

Macfarlane, A.W., 1999. Isotopic studies of northern Andean crustal evolution and oremetal sources. In: Skinner, B.J. (Ed.), Geology and Ore Deposits of the Central Andes. :Special Publication, No. 7. Society of Economic Geologists, pp. 195–217.

Macfarlane, A.W., Marcet, P., LeHuray, A.P., Petersen, U., 1990. Lead isotope provinces ofthe Central Andes inferred from ores and crustal rocks. Economic Geology 85,1857–1880.

Macpherson, C.G., Dreher, S.T., Thirlwall, M.F., 2006. Adakites without slab melting:high pressure differentiation of island arc magma, Mindanao, the Philippines. Earthand Planetary Science Letters 243, 581–593.

Malaspina, N., Poli, S., Fumagalli, P., 2009. The oxidation state of metasomatized mantlewedge: insights from C–O–H-bearing garnet peridotite. Journal of Petrology 50,1533–1552.

Mancano, D.P., Campbell, A.R., 1995. Microthermometry of enargite-hosted fluid in-clusions from the Lepanto, Philippines, high-sulfidation CuAu deposit. Geochimica etCosmochimica Acta 59, 3909–3916.

Manning, C.E., 2004. The chemistry of subduction-zone fluids. Earth and PlanetaryScience Letters 223, 1–16.

Manske, S.L., Hedenquist, J.W., O'Connor, G., Tămaş, C., Cauuet, B., Leary, S., Minut, A.,2006. Roşia Montană, Romania: Europe's largest gold deposit January Society ofEconomic Geologists Newsletter 64 (1), 9–15.

Martin, H., 1999. Adakitic magmas: modern analogues of Archaean granitoids. Lithos46, 411–429.

Martin, H., Smithies, R.H., Rapp, R., Moyen, J.-F., Champion, D., 2005. An overview ofadakite, tonalite–trondhjemite–granodiorite (TTG), and sanukitoid: relationshipsand some implications for crustal evolution. Lithos 79, 1–24.

Maughan, D.T., Keith, J.D., Christiansen, E.H., Pulsipher, T., Hattori, K., Evans, N.J., 2002.Contributions from mafic alkaline magmas to the Bingham porphyry Cu–Au–Modeposit, Utah, USA. Mineralium Deposita 37, 14–37.

Mavrogenes, J.A., Bodnar, R.J., 1994. Hydrogen movement into and out of fluidinclusions in quartz; experimental evidence and geologic implications. Geochimicaet Cosmochimica Acta 58, 141–148.

McInnes, B.I.A., Cameron, E.M., 1994. Carbonated, alkaline hybridizingmelts from a sub-arc environment: mantle wedge samples from the Tabar-Lihir-Tanga-Feni arc,Papua New Guinea. Earth and Planetary Science Letters 122, 125–141.

McInnes, B.I.A.,McBride, J.S., Evans,N.J., Lambert, D.D., Andrew,A.A., 1999.Osmium isotopeconstraints on ore metal recycling in subduction zones. Science 286, 512–516.

McInnes, B.I.A., Gregoire, M., Binns, R.A., Herzig, P.M., Hannington, M.D., 2001. Hydrousmetasomatism of oceanic sub-arc mantle, Lihir, Papua New Guinea: petrology andgeochemistry of fluid-metasomatised mantle wedge xenoliths. Earth and PlanetaryScience Letters 188, 169–183.

McNutt, R.H., Clark, A.H., Zentilli, M., 1979. Lead isotopic compositions of Andeanigneous rocks, Latitudes 26° to 29° S: petrologic and metallogenic implications.Economic Geology 74, 827–837.

Meinert, L.D., Dipple, G.M., Nicolescu, S., 2005. World skarn deposits. In:Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), EconomicGeology 100th Anniversary Volume. Society of Economic Geologists, Littleton,CO, pp. 299–336.

Mitchell, R.H., Keays, R.R., 1981. Abundance and distribution of gold, palladium andiridium in some spinel and garnet lherzolites: implications for the nature and originof precious metal-rich intergranular components in the upper mantle. Geochimicaet Cosmochimica Acta 45, 2425–2442.

Moore, W.J., Nash, J.T., 1974. Alteration and fluid inclusion studies of the porphyrycopper ore body at Bingham, Utah. Economic Geology 69, 631–645.

Morris, J.D., Leeman, W.P., Tera, F., 1990. The subducted component in island arc lavas:constraints from Be isotopes and B–Be systematics. Nature 344, 31–36.

Mukasa, S.B., Vidal, C.E., Injoque-Espinoza, J., 1990. Pb isotope bearing on themetallogenesis of sulfide ore deposits in central and southern Peru. EconomicGeology 85, 1438–1446.

Müller, D., Groves, D.I., 1993. Direct and indirect associations between potassic igneousrocks, shoshonites and gold–copper deposits. Ore Geology Reviews 8, 383–406.

Mungall, J.E., 2002. Roasting the mantle: slab melting and the genesis of major Au andAu-rich Cu deposits. Geology 30, 915–918.

Müntener, O., Ulmer, P., 2006. Experimentally derived high-pressure cumulates fromhydrous arc magmas and consequences for the seismic velocity structure of lowerarc crust. Geophysical Research Letters 33, L21308. doi:10.1029/2006GL027629.

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

Mutschler, F.E., Griffin, M.E., Stevens, D.S., Shannon, S.S., 1985. Precious metal depositsrelated to alkaline rocks in the North American Cordillera— an interpretive review.Transactions, 88. Geological Society of South Africa, pp. 355–377.

Naldrett, A.J., 1989. Sulfide melts—crystallization temperatures, solubilities in silicatemelts, and Fe, Ni, and Cu partitioning between basaltic magmas, and olivine. In:Whitney, J.A., Naldrett, A.J. (Eds.), Ore Deposition Associated with Magmas.Reviews in Economic Geology, 4. Society of Economic Geologists, pp. 5–20. ch. 2,.

Naney, M.T., 1983. Phase equilibria of rock-forming ferromagnesian silicates in graniticsystems. American Journal of Science 283, 993–1033.

Nash, J.T., 1976. Fluid-inclusion petrology — data from porphyry copper deposits andapplications to exploration. Professional Paper, 907-D. U.S. Geological Survey. 16 pp.

Neubauer, F., Lips, A., Kouzmanov, K., Lexa, J., Ivascanu, P., 2005. Subduction, slabdetachment and mineralization: the Neogene in the Apuseni Mountains andCarpathians. Ore Geology Reviews 27, 13–44.

Neuendorf, K.K.E., Mehl Jr., J.P., Jackson, J.A. (Eds.), 2005. Glossary of Geology, 5thedition. American Geological Institute, Alexandria, Virginia. 800 pp.

Newberry, R.J., Swanson, S.E., 1986. Scheelite skarn granitoids: an evaluation of theroles of magmatic source and process. Ore Geology Reviews 1, 57–81.

Noll, P.D., Newsom, H.E., Leeman, W.P., Ryan, J.G., 1996. The role of hydrothermal fluidsin the production of subduction zone magmas: evidence from siderophile andchalcophile trace elements and boron. Geochim. Cosmochim. Acta 60, 587–611.

Noorollahi, Y., Itoi, R., Fujii, H., Tanaka, T., 2007. GIS model for geothermal resourceexploration in Akita and Iwate prefectures, northern Japan. Computers andGeosciences 33, 1008–1021.

Norton, D.L., 1982. Fluid and heat transport phenomena typical of copper-bearingpluton environments. In: Titley, S.R. (Ed.), Advances in Geology of Porphyry CopperDeposits of Southwestern North America. Tuscon, Univ, Arizona Press, pp. 59–72.

Norton, D.L., Cathles, L.M., 1973. Breccia pipes, products of exsolved vapor frommagmas. Economic Geology 68, 540–546.

Norton, D.L., Dutrow, B.L., 2001. Complex behavior of magma–hydrothermal processes:role of supercritical fluid. Geochimica et Cosmochimica Acta 65, 4009–4017.

Oyarzun, R., Márquez, A., Lillo, J., López, I., Rivera, S., 2001. Giant versus small porphyrycopper deposits of Cenozoic age in northern Chile: adakitic versus normal calc-alkaline magmatism. Mineralium Deposita 36, 794–798.

Oyarzun, R., Márquez, A., Lillo, J., López, I., Rivera, S., 2002. Reply to discussion on “Giantversus small porphyry copper deposits of Cenozoic age in northern Chile: adakiticversus normal calc-alkaline magmatism” by Oyarzun et al. (Mineralium Deposita36: 794–798, 2001). Mineralium Deposita 37, 791–794.

Parkinson, I.J., Arculus, R.J., 1999. The redox state of subduction zones: insights fromarc-peridotites. Chemical Geology 160, 409–423.

Paterson, J.T., Cloos, M., 2005. Grasberg porphyry Cu–Au deposit, Papua, Indonesia: 1.Magmatic history. In: Porter, T.M. (Ed.), Super Porphyry Copper and Gold Deposits:A Global Perspective, 2. Porter Geoscience Consulting Publishing, perspectiveLin-den Park, South Australia, pp. 313–329.

Peach, C.L., Mathez, E.A., Keays, R.R., 1990. Sulfide melt–silicate melt distributioncoefficients for noble metals and other chalcophile elements as deduced fromMORB: implications for partial melting. Geochimica et Cosmochimica Acta 54,3379–3389.

Peacock, S.M., 1993. Large-scale hydration of the lithosphere above subducting slabs.Chemical Geology 108, 49–59.

Peacock, S.M., 1996. Thermal andpetrologic structure of subduction zones. In: Bebout, G.E.,Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction: Top to Bottom. : GeophysicalMonograph, 96. American Geophysical Union, Washington, DC, pp. 119–133.

Peacock, S.M., Rushmer, T., Thompson, A.B., 1994. Partial melting of subducting oceaniccrust. Earth and Planetary Science Letters 121, 227–244.

Pearce, J.A., Bender, J.F., De Long, S.E., Kidd,W.S.F., Low, P.J., Güner, Y., Saroglu, F., Yilmaz, Y.,Moorbath, S., Mitchell, J.G., 1990. Genesis of collision volcanism in eastern Anatolia,Turkey. Journal of Volcanology and Geothermal Research 44, 189–229.

Petford, N., Gallagher, K., 2001. Partial melting of mafic (amphibolitic) lower crust byperiodic influx of basaltic magma. Earth and Planetary Science Letters 193,483–499.

Pettke, T., Oberli, F., Heinrich, C.A., 2010. The magma and metal source of giantporphyry-type ore deposits, based on lead isotope microanalysis of individual fluidinclusions. Earth and Planetary Science Letters 296, 267–277.

Pichavant, M., Mysen, B.O., Macdonald, R., 2002. Source and H2O content of high-MgOmagmas in island arc settings: an experimental study of a primitive calc-alkalinebasalt from St. Vincent, Lesser Antilles arc. Geochimica et Cosmochimica Acta 66,2193–2209.

Pitcher, W.S., 1997. The Nature and Origin of Granite, 2nd edition. Chapman and Hall,London. 387 pp.

Pitzer, K.S., Pabalan, R.T., 1986. Thermodynamics of NaCl in steam. Geochimica etCosmochimica Acta 50, 1445–1454.

Plank, T., 2005. Constraints from thorium/lanthanum on sediment recycling at subductionzones and the evolution of the continents. Journal of Petrology 46, 921–944.

Page 24: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

24 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Pokrovski, G.S., Roux, J., Harrichoury, J.-C., 2005. Fluid density control on vapor–liquidpartitioning of metals in hydrothermal systems. Geology 33, 657–660.

Pokrovski, G.S., Borisova, A.Y., Harrichoury, J.-C., 2008. The effect of sulfur on vapor–liquid fractionation of metals in hydrothermal systems. Earth and Planetary ScienceLetters 266, 345–362.

Poli, S., Schmidt, M.W., 2002. Petrology of subducted slabs. Annual Reviews of Earth andPlanetary Sciences 30, 207–235.

Portnyagin, M., Hoernle, K., Plechov, P., Mironov, N., Khubunaya, S., 2007. Constraints onmantle melting and composition and nature of slab components in volcanic arcsfrom volatiles (H2O, S, Cl, F) and trace elements in melt inclusions from theKamchatka Arc. Earth and Planetary Science Letters 255, 53–69.

Pudack, C., Halter, W.E., Heinrich, C.A., Pettke, T., 2009. Evolution of magmatic vapor togold-rich epithermal liquid: the porphyry to epithermal transition at Nevados deFamatina, Northwest Argentina. Economic Geology 104, 449–477.

Rabbia, O.M., Hernández, L.B., King, R.W., López-Escobar, L., 2002. Discussion on “Giantversus small porphyry copper deposits of Cenozoic age in northern Chile: adakiticversus normal calc-alkaline magmatism” by Oyarzun et al. (Mineralium Deposita36: 794–798, 2001). Mineralium Deposita 37, 791–794.

Rapp, R.P., Watson, E.B., 1995. Dehydration melting of metabasalt at 8–32 kbar:implications for continental growth and crust–mantle recycling. Journal ofPetrology 36, 891–931.

Rapp, R.P., Watson, E.B., Miller, C.F., 1991. Partial melting of amphibolite/eclogite andthe origin of Archean trondhjemites and tholeiites. Precambrian Research 51,1–25.

Redmond, P.B., Einaudi, M.T., Inan, E.E., Landtwing, M.R., Heinrich, C.A., 2004. Copperdeposition by fluid cooling in intrusion-centered systems: new insights from theBingham porphyry ore deposit, Utah. Geology 32, 217–220.

Reeves, E.P., Seewald, J.S., Saccocia, P., Walsh, E., Bach, W., Craddock, P.R., Shanks, W.C.,Sylva, S.P., Pichler, T., Rosner, M., 2010. Geochemistry of hydrothermal fluids fromthe PACMANUS, Northeast Pual and Vienna Woods hydrothermal fields, ManusBasin, Papua New Guinea. Geochimica et Cosmochimica Acta 75, 1088–1123.

Richards, J.P., 1995. Alkalic-type epithermal golddeposits— a review. In: Thompson, J.F.H.(Ed.), Magmas, Fluids, and Ore Deposits. Short Course Series, 23. MineralogicalAssociation of Canada, pp. 367–400. ch. 17.

Richards, J.P., 2002. Discussion of “Giant versus small porphyry copper deposits ofCenozoic age in northern Chile: adakitic versus normal calc-alkaline magmatism”

by Oyarzun et al. (Mineralium Deposita 36: 794–798, 2001). Mineralium Deposita37, 788–790.

Richards, J.P., 2003. Tectono-magmatic precursors for porphyry Cu–(Mo–Au) depositformation. Economic Geology 96, 1515–1533.

Richards, J.P., 2005. Cumulative factors in the generation of giant calc-alkaline porphyryCu deposits. In: Porter, T.M. (Ed.), Super Porphyry Copper and Gold Deposits: AGlobal Perspective, 1. Porter Geoscience Consulting Publishing, Linden Park, SouthAustralia, pp. 7–25.

Richards, J.P., 2009. Postsubduction porphyry Cu–Au and epithermal Au deposits:products of remelting of subduction-modified lithosphere. Geology 37, 247–250.

Richards, J.P., Kerrich, R., 1993. The Porgera gold mine, Papua New Guinea: magmatic–hydrothermal to epithermal evolution of an alkalic-type precious metal deposit.Economic Geology 88, 1017–1052.

Richards, J.P., Kerrich, R., 2007. Adakite-like rocks: their diverse origins andquestionable role in metallogenesis. Economic Geology 102, 537–576.

Richards, J.P., Chappell, B.W., McCulloch, M.T., 1990. Intraplate-type magmatism in acontinent–island-arc collision zone: Porgera intrusive complex, Papua NewGuinea.Geology 18, 958–961.

Richards, J.P., Ullrich, T., Kerrich, R., 2006a. The Late Miocene–Quaternary Antofallavolcanic complex, southern Puna, NW Argentina: protracted history, diversepetrology, and economic potential. Journal of Volcanology and GeothermalResearch 152, 197–239.

Richards, J.P., Wilkinson, D., Ullrich, T., 2006b. Geology of the Sari Gunay epithermalgold deposit, northwest Iran. Economic Geology 101, 1455–1496.

Righter, K., Campbell, A.J., Humayun, M., Hervig, R.L., 2004. Partitioning of Ru, Rh, Pd, Re,Ir, and Au between Cr-bearing spinel, olivine, pyroxene and silicate melts.Geochimica et Cosmochimica Acta 68, 867–880.

Ronacher, E., Richards, J.P., Reed, M.H., Bray, C.J., Spooner, E.T.C., Adams, P.D., 2004.Characteristics and evolution of the hydrothermal fluid in the North Zone high-grade area, Porgera gold deposit, Papua New Guinea. Economic Geology 99,843–867.

Rowe, M.C., Kent, A.J.R., Nielsen, R.L., 2009. Subduction influence on oxygen fugacityand trace and volatile elements in basalts across the Cascade Volcanic Arc. Journalof Petrology 50, 61–91.

Rubin, A.M., 1995. Propagation of magma-filled cracks. Annual Review of Earth &Planetary Sciences 23, 287–336.

Rudnick, R.L., Gao, S., 2003. Composition of the continental crust. In: Rudnick, R.L. (Ed.),Treatise on Geochemistry 3: The Crust. Elsevier, Amsterdam, pp. 1–64.

Rushmer, T., 1991. Partial melting of two amphibolites: contrasting experimentalresults under fluid-absent conditions. Contributions to Mineralogy and Petrology107, 41–59.

Rushmer, T., 1993. Experimental high-pressure granulites: some applications to naturalmafic xenolith suites and Archean granulite terranes. Geology 21, 411–414.

Rusk, B.G., Reed, M.H., Dilles, J.H., Klemm, L.M., Heinrich, C.A., 2004. Compositions ofmagmatic hydrothermal fluids determined by LA-ICP-MS of fluid inclusions fromthe porphyry copper–molybdenum deposit at Butte, MT. Chemical Geology 210,173–199.

Rusk, B.G., Reed, M.H., Dilles, J.H., 2008. Fluid inclusion evidence for magmatic–hydrothermal fluid evolution in the porphyry copper–molybdenum deposit atButte, Montana. Economic Geology 103, 307–334.

Rutherford, M.J., Devine, J.D., 1988. The May 18, 1980, eruption of Mount St. Helens. 3.Stability and chemistry of amphibole in the magma chamber. Journal ofGeophysical Research 93 (11), 959 949-11.

Ryerson, F.J., Watson, E.B., 1987. Rutile saturation in magmas: implications for Ti–Nb–Ta depletion in island-arc basalts. Earth and Planetary Science Letters 86, 225–239.

Sajona, F.G., Maury, R.C., 1998. Association of adakites with gold and coppermineralization in the Philippines. Comptes rendus de l'Académie des sciences,Série II, Sciences de la terre et des planètes 326, 27–34.

Sakai, H., Matsubaya, O., 1977. Stable isotopic studies of Japanese geothermal systems.Geothermics 5, 97–124.

Sawkins, F.J., Scherkenbach, D.A., 1981. High copper content of fluid inclusions in quartzfrom northern Sonora: implications for ore-genesis theory. Geology 9, 37–40.

Scambelluri, M., Müntener, O., Ottolini, L., Pettke, T.P., Vannuccie, R., 2004. The fate of B,Cl and Li in the subducted oceanic mantle and in the antigorite breakdown fluids.Earth and Planetary Science Letters 222, 217–234.

Scherbarth, N.L., Spry, P.G., 2006. Mineralogical, petrological, stable isotope, and fluidinclusion characteristics of the Tuvatu Gold–Silver Telluride Deposit, Fiji:comparisons with the Emperor Deposit. Economic Geology 101, 135–158.

Schiano, P., Clocchiatti, R., Shimizu, N., Maury, R.C., Jochum, K.P., Hofmann, A.W., 1995.Hydrous, silica-rich melts in the sub-arc mantle and their relationship with eruptedarc lavas. Nature 277, 595–600.

Schmidt, M.W., Poli, S., 1998. Experimentally based water budgets for dehydrating slabsand consequences for arc magma generation. Earth and Planetary Science Letters163, 361–379.

Schmidt, M.W., Dardon, A., Chazot, G., Vannucci, R., 2004. The dependence of Nb and Tarutile-melt partitioning on melt composition and Nb/Ta fractionation duringsubduction processes. Earth and Planetary Science Letters 226, 415–432.

Schmidt, A., Weyer, S., John, T., Brey, G.P., 2009. HFSE systematics of rutile-bearingeclogites: new insights into subduction zone processes and implications for theearth's HFSE budget. Geochimica et Cosmochimica Acta 73, 455–468.

Seedorff, E., Dilles, J.H., Proffett Jr., J.M., Einaudi, M.T., Zurcher, L., Stavast, W.J.A.,Johnson, D.A., Barton, M.D., 2005. Porphyry deposits: characteristics and origin ofhypogene features. Economic Geology 100th Anniversary Volume, pp. 251–298.

Seo, J.H., Guillong, M., Heinrich, C.A., 2009. The role of sulfur in the formation ofmagmatic–hydrothermal copper–gold deposits. Earth and Planetary ScienceLetters 282, 323–328.

Setterfield, T.N., Mussett, A.E., Oglethorpe, R.D.J., 1992. Magmatism and associatedhydrothermal activity during the evolution of the Tavua caldera: 40Ar/39Ar datingof volcanic, intrusive, and hydrothermal events. Economic Geology 87, 1130–1140.

Shafiei, B., Haschke, M., Shahabpour, J., 2009. Recycling of orogenic arc crust triggersporphyry Cu mineralization in Kerman Cenozoic arc rocks, southeastern Iran.Mineralium Deposita 44, 265–283.

Shinohara, H., 1994. Exsolution of immiscible vapor and liquid phases from acrystallizing silicate melt: implications for chlorine and metal transport. Geochi-mica et Cosmochimica Acta 58, 5215–5221.

Shinohara, H., Hedenquist, J.W., 1997. Constraints on magma degassing beneath the FarSoutheast porphyry Cu–Au deposit, Philippines. Journal of Petrology 38, 1741–1752.

Shinohara, H., Kazahaya, K., Lowenstern, J.B., 1995. Volatile transport in a convectingmagma column: implications for porphyry Mo mineralization. Geology 23,1091–1094.

Sillitoe, R.H., 1973. The tops and bottoms of porphyry copper deposits. EconomicGeology 68, 799–815.

Sillitoe, R.H., 2000. Gold-rich porphyry deposits: descriptive and genetic models andtheir role in exploration and discovery. Reviews in Economic Geology 13,315–345.

Sillitoe, R.H., 2010. Porphyry copper systems. Economic Geology 105, 3–41.Sillitoe, R.H., Hart, S.R., 1984. Lead-isotope signatures of porphyry copper deposits in

oceanic and continental settings, Colombian Andes. Geochimica et CosmochimicaActa 48, 2135–2142.

Simmons, S.F., Brown, K.L., 2007. The flux of gold and related metals through a volcanicarc, Taupo Volcanic Zone, New Zealand. Geology 35, 1099–1102.

Simon, A.C., Frank, M.R., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2005. Goldpartitioning in melt–vapor–brine systems. Geochimica et Cosmochimica Acta 69,3321–3335.

Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2006. Copperpartitioning in a melt–vapor–brine–magnetite–pyrrhotite assemblage. Geochimicaet Cosmochimica Acta 70, 5583–5600.

Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2007. The partitioningbehavior of As and Au in S-free and S-bearing magmatic assemblages. Geochimicaet Cosmochimica Acta 71, 1764–1782.

Simon, A.C., Candela, P.A., Piccoli, P.M., Mengason, M., Englander, L., 2008. The effect ofcrystal-melt partitioning on the budgets of Cu, Au, and Ag. American Mineralogist93, 1437–1448.

Sinclair, W.D., 2007. Porphyry deposits. In: Goodfellow, W.D. (Ed.), Mineral Deposits ofCanada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution ofGeological Provinces, and Exploration Methods. Geological Association of Canada,St. John's, Newfoundland, pp. 223–243.

Smith, I.E.M., Stewart, R.B., Price, R.C., Worthington, T.J., 2010. Are arc-type rocks theproducts of magma crystallisation? Observations from a simple oceanic arcvolcano: Raoul Island, Kermadec Arc, SW Pacific. Journal of Volcanology andGeothermal Research 190, 219–234.

Sobolev, A., Chaussidon, M., 1996. H2O concentrations in primary melts from supra-subduction zones andmid-ocean ridges: implications for H2O storage and recyclingin the mantle. Earth and Planetary Science Letters 137, 45–55.

Solomon, M., 1990. Subduction, arc reversal, and the origin of porphyry copper–golddeposits in island arcs. Geology 18, 630–633.

Page 25: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

25J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Sourirajan, S., Kennedy, G.C., 1962. The system H2O–NaCl at elevated temperatures andpressures. American Journal of Science 260, 115–141.

Spooner, E.T.C., 1993. Magmatic sulphide/volatile interaction as a mechanism forproducing chalcophile element enriched, Archean Au–quartz, epithermal Au–Agand Au skarn hydrothermal ore fluids. Ore Geology Reviews 7, 359–379.

Staudigel, H., Plank, T., White, B., Schmincke, H.-U., 1996. Geochemical fluxes during seaflooralteration of the basaltic upper oceanic crust: DSDP sites 417 and 418. In: Bebout, G.E.,Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction Top to Bottom. GeophysicalMonograph, 96. American Geophysical Union, Washington, DC, pp. 19–38.

Stavast, W.J.A., Keith, J.D., Christiansen, E.H., Dorais, M.J., Tingey, D., 2006. The fate ofmagmatic sulfides during intrusion or eruption, Bingham and Tintic districts, Utah.Economic Geology 101, 329–345.

Stein, H.J., 1988. Genetic traits of climax-type granites and molybdenum mineraliza-tion, Colorado Mineral Belt. In: Taylor, R.P., Strong, D.F. (Eds.), Recent Advances inthe Geology of Granite-Related Mineral Deposits, Special Volume 39. CanadianInstitute of Mining and Metallurgy, pp. 394–401.

Stern, R.J., Kohut, E., Bloomer, S.H., Leybourne, M., Fouch, M., Vervoort, J., 2006.Subduction factory processes beneath the Guguan cross-chain, Mariana Arc: no rolefor sediments, are serpentinites important? Contributions to Mineralogy andPetrology 151, 202–221.

Stoffregen, R., 1987. Genesis of acid-sulfate alteration and Au–Cu–Ag mineralization atSummitville, Colorado. Economic Geology 82, 1575–1591.

Stolper, E., Newman, S., 1994. The role of water in the petrogenesis of Mariana troughmagmas. Earth and Planetary Science Letters 121, 293–325.

Sun, S.-s., McDonough,W.F., 1989. Chemical and isotopic systematics of oceanic basalts:implications for mantle composition and processes. In: Saunders, A.D., Norry, M.J.(Eds.), Magmatism in the Ocean Basins: Geological Society of London SpecialPublication, No. 42, pp. 313–345.

Sun, W., Bennett, V.C., Kamenetsky, V.S., 2004a. The mechanism of Re enrichment in arcmagmas: evidence from Lau Basin basaltic glasses and primitive melt inclusions.Earth and Planetary Science Letters 222, 101–114.

Sun, W.D., Arculus, R.J., Kamenetsky, V.S., Binns, R.A., 2004b. Release of gold-bearingfluids in convergent margin magmas prompted by magnetite crystallization.Nature 431, 975–978.

Symonds, R.B., Rose, W.I., Reed, M.H., Lichte, F.E., Finnegan, D.L., 1987. Volatilization,transport and sublimation of metallic and non-metallic elements in hightemperature gases at Merapi Volcano, Indonesia. Geochimica et CosmochimicaActa 51, 2083–2101.

Tarana, Y.A., Hedenquist, J.W., Korzhinsky, M.A., Tkachenko, S.I., Shmulovich, K.I., 1995.Geochemistry of magmatic gases from Kudryavy volcano, Iturup, Kuril Islands.Geochimica et Cosmochimica Acta 59, 1749–1761.

Tatsumi, Y., 1986. Formation of the volcanic front in subduction zones. GeophysicalResearch Letters 17, 717–720.

Tatsumi, Y., 2003. Some constraints on arc magma genesis. In: Eiler, J. (Ed.), Inside theSubduction Factory. : Geophysical Monograph, 138. American Geophysical Union,Washington, DC, pp. 277–292.

Tatsumi, Y., Eggins, S., 1995. Subduction Zone Magmatism. Blackwell, Oxford. 213 pp.Tatsumi, Y., Hamilton, D.L., Nesbitt, R.W., 1986. Chemical characteristics of fluid phase

released from a subducted lithosphere and the origin of arcmagmas: evidence fromhigh pressure experiments and natural rocks. Journal of Volcanology andGeothermal Research 29, 293–309.

Taylor, S.R., McLennan, S.M., 1985. The continental crust: its composition and evolution:an examination of the geochemical record preserved in sedimentary rocks.Blackwell Scientific 312p.

Tessalina, S.G., Yudovskaya, M.A., Chaplygin, I.V., Birck, J.-L., Capmas, F., 2008. Sources ofunique rhenium enrichment in fumaroles and sulphides at Kudryavy volcano.Geochimica et Cosmochimica Acta 72, 889–909.

Thiéblemont, D., Stein, G., Lescuyer, J.-L., 1997. Gisements épithermaux et porphyr-iques: la connexion adakite. C.R. Acad. Sci. Paris, Sciences de la terre et des planètes/Earth and Planetary Sciences 325, 103–109.

Thirlwall, M.F., Graham, A.M., Arculus, R.J., Harmon, R.S., Macpherson, C.G., 1996.Resolution of the effects of crustal assimilation, sediment subduction, and fluidtransport in island arc magmas: Pb–Sr–Nd–O isotope geochemistry of Grenada,Lesser Antilles. Geochimica et Cosmochimica Acta 60, 4785–4810.

Thompson, T.B., Trippel, A.D., Dwelley, P.C., 1985. Mineralized veins and breccias of theCripple Creek District, Colorado. Economic Geology 80, 1669–1688.

Thorkelson, D.J., Breitsprecher, K., 2005. Partial melting of slab window margins:genesis of adakitic and non-adakitic magmas. Lithos 79, 25–41.

Tiepolo, M., Tribuzio, R., 2008. Petrology and U–Pb zircon geochronology of amphibole-rich cumulates with sanukitic affinity from Husky Ridge (Northern Victoria Land,Antarctica): insights into the role of amphibole in the petrogenesis of subduction-related magmas. Journal of Petrology 49, 937–970.

Tilton, G.R., Pollak, R.J., Clark, A.H., Robertson, R.C.R., 1981. Isotopic composition of Pb incentral Andean ore deposits. Geological Society of America, Memoir 154, 791–816.

Titley, S.R., 1987. The crustal heritage of silver and gold ratios in Arizona ores.Geological Society of America, Bulletin 99, 814–826.

Titley, S.R., 2001. Crustal affinities of metallogenesis in the American Southwest.Economic Geology 96, 1323–1342.

Tomkins, A.G., Mavrogenes, J.A., 2003. Generation of metal-rich felsic magmas duringcrustal anatexis. Geology 31, 765–768.

Tosdal, R.M., Richards, J.P., 2001. Magmatic and structural controls on the developmentof porphyry Cu±Mo±Au deposits. In: Richards, J.P., Tosdal, R.M. (Eds.), StructuralControls on Ore Genesis. : Reviews in Economic Geology, 14. Society of EconomicGeologists, pp. 157–181.

Tulloch, A.J., Kimbrough, D.L., 2003. Paired plutonic belts in convergent margins and thedevelopment of high Sr/Y magmatism: Peninsular Ranges batholith of Baja–

California and Median batholith of New Zealand. Special Paper, 374. GeologicalSociety of America, pp. 1–21.

Ulmer, P., Trommsdorff, V., 1995. Serpentine stability to mantle depths and subduction-related magmatism. Science 268, 858–861.

Ulrich, T., Günther, D., Heinrich, C.A., 2001. The evolution of a porphyry Cu–Au deposit,based on LA-ICP-MS analysis of fluid inclusions: Bajo de la Alumbrera, Argentina.Economic Geology 96, 1743–1774.

van Dongen, M., Weinberg, R.F., Tomkins, A.G., Armstrong, R.A., Woodhead, J.D.,2010. Recycling of Proterozoic crust in Pleistocene juvenile magma and rapidformation of the Ok Tedi porphyry Cu–Au deposit, Papua New Guinea. Lithos114, 282–292.

Vry, V.H., Wilkinson, J.J., Seguel, J., Millan, J., 2010. Multistage intrusion, brecciation, andveining at El Teniente, Chile: evolution of a nested porphyry system. EconomicGeology 105, 119–153.

Walker, J.A., Patino, L.C., Carr, M.J., Feigenson, M.D., 2001. Slab control over HFSEdepletions in central Nicaragua. Earth and Planetary Science Letters 192, 533–543.

Wallace, P.J., 2005. Volatiles in subduction zone magmas: concentrations and fluxesbased on melt inclusion and volcanic gas data. Journal of Volcanology andGeothermal Research 140, 217–240.

Walshe, J.L., Solomon, M., Whitford, D.J., Sun, S.-S., Foden, J.D., 2011. The role of themantle in the genesis of tin deposits and tin provinces of eastern Australia.Economic Geology 106, 297–305.

Wang, J., Hattori, K.H., Kilian, R., Stern, C.R., 2007a. Metasomatism of sub-arc mantleperidotites below southernmost South America: reduction of fO2 by slab-melt.Contributions to Mineralogy and Petrology 153, 607–624.

Wang, Q., Wyman, D.A., Xu, J.-F., Zhao, Z.-H., Jian, P., Zi, F., 2007b. Partial melting ofthickened or delaminated lower crust in the middle of Eastern China: implicationsfor Cu–Au mineralization. Journal of Geology 115, 149–161.

Webster, J.D., 1992. Water solubility and chlorine partitioning in Cl-rich graniticsystems: effects of melt composition at 2 kbar and 800 °C. Geochimica etCosmochimica Acta 56, 679–687.

Westra, G., Keith, S.B., 1981. Classification and genesis of stockwork molybdenumdeposits. Economic Geology 76, 844–873.

White, D.E., Muffler, L.J.P., Truesdell, A.H., 1971. Vapor-dominated hydrothermalsystems compared with hot-water systems. Economic Geology 66, 75–97.

White, W.H., Bookstrom, A.A., Kamilli, R.J., Gangster, M.W., Smith, R.P., Ranta, D.E.,Steininger, R.C., 1981. Character and origin of climax-type molybdenum deposits.Economic Geology 75th Anniversary volume, pp. 270–316.

Widom, E., Kepezhinskas, P., Defant, M., 2003. The nature of metasomatism in the sub-arc mantle wedge: evidence from Re–Os isotopes in Kamchatka peridotitexenoliths. Chemical Geology 196, 283–306.

Williams, T.J., Candela, P.A., Piccoli, P.M., 1995. The partitioning of copper betweensilicate melts and two-phase aqueous fluids: an experimental investigation at1 kbar, 800 °C and 0.5 kbar, 850 °C. Contributions to Mineralogy and Petrology 121,388–399.

Williams-Jones, A.E., Heinrich, C.A., 2005. Vapor transport of metals and the formationof magmatic–hydrothermal ore deposits. Economic Geology 100, 1287–1312.

Williamson, B.J., Müller, A., Shail, R.K., 2010. Source and partitioning of B and Sn in theCornubian batholith of southwest England. Ore Geology Reviews 38, 1–8.

Winter, J.D., 2001. An introduction to igneous and metamorphic petrology: UpperSaddle River. Prentice-Hall Inc., New Jersey. 697 p.

Wolf, M.B., Wyllie, P.J., 1994. Dehydration-melting of amphibolite at 10 kbar — theeffects of temperature and time. Contributions to Mineralogy and Petrology 115,369–383.

Wood, S.A., Spera, F.J., 1984. Adiabatic decompression of aqueous solutions:applications to hydrothermal fluid migration in the crust. Geology 12, 707–710.

Wörner, G., Moorbath, S., Harmon, R.S., 1992. Andean Cenozoic volcanic centers reflectbasement isotopic domains. Geology 20, 1103–1106.

Wyborn, D., Sun, S.-s., 1994. Sulphur-undersaturated magmatism — a key factor forgenerating magma-related copper–gold deposits. AGSO Research Newsletter 21,7–8.

Wyllie, P.J., Huang, W.-L., Stern, C.R., Maaløe, S., 1976. Granitic magmas: possible andimpossible sources, water contents, and crystallization sequences. CanadianJournal of Earth Sciences 13, 1007–1019.

Wysoczanski, R.J., Wright, I.C., Gamble, J.A., Hauri, E.H., Luhr, J.F., Eggins, S.M., Handler,M.R., 2006. Volatile contents of Kermadec Arc–Havre Trough pillow glasses:fingerprinting slab-derived aqueous fluids in the mantle sources of arc and back-arc lavas. Journal of Volcanology and Geothermal Research 152, 51–73.

Xiao, Z., Gammons, C.H., Williams-Jones, A.E., 1998. Experimental study of copper(I)chloride complexing in hydrothermal solutions at 40 to 300 °C and saturated watervapor pressure. Geochimica et Cosmochimica Acta 62, 2949–2964.

Yang, Z., Hou, Z., White, N.C., Chang, Z., Li, Z., Song, Y., 2009. Geology of the post-collisional porphyry copper–molybdenum deposit at Qulong, Tibet. Ore GeologyReviews 36, 133–159.

Yogodzinski, G.M., Kay, R.W., Volynets, O.N., Koloskov, A.V., Kay, S.M., 1995. Magnesianandesite in the western Aleutian Komandorsky region: implications for slabmelting and processes in the mantle wedge. Geological Society of America Bulletin107, 505–519.

Yogodzinski, G.M., Lees, J.M., Churikova, T.G., Dorendorf, F., Wöerner, G., Volynets, O.N.,2001. Geochemical evidence for the melting of subducting oceanic lithosphere atplate edges. Nature 409, 500–504.

Zajacz, Z., Halter, W., 2009. Copper transport by high temperature, sulfur-rich magmaticvapor: evidence from silicatemelt and vapor inclusions in a basaltic andesite from theVillarrica volcano (Chile). Earth and Planetary Science Letters 282, 115–121.

Zajacz, Z., Halter, W.E., Pettke, T., Guillong, M., 2008. Determination of fluid/meltpartition coefficients by LA-ICPMS analysis of co-existing fluid and silicate melt

Page 26: Ore Geology Reviews - Eötvös Loránd Universitylrg.elte.hu/oktatas/Szubdukcio es kopenyek PhD/Kovacs Ivett... · Review Magmatic to hydrothermal metal fluxes in convergent and

26 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

inclusions: controls on element partitioning. Geochimica et Cosmochimica Acta 72,2169–2197.

Zajacz, Z., Seo, Z.H., Candela, P.A., Piccoli, P.M., Heinrich, C.A., Guillong, M., 2010. Alkalimetals control the release of gold from volatile-rich magmas. Earth and PlanetaryScience Letters 297, 50–56.

Zajacz, Z., Seo, J.H., Candela, P.A., Piccoli, P.M., Tossell, J.A., 2011. The solubility of copper inhigh-temperature magmatic vapors: a quest for the significance of various chlorideand sulfide complexes. Geochimica et Cosmochimica Acta 75, 2811–2827.

Zhang, X., Spry, P.G., 1994. Petrological, mineralogical, fluid inclusion, and stableisotope studies of the Gies gold–silver telluride deposit, Judith Mountains,Montana. Economic Geology 89, 602–627.

Zimmer, M.M., Plank, T., Hauri, E.H., Yogodzinski, G.M., Stelling, P., Larsen, J., Singer, B.,Jicha, B., Mandeville, C., Nye, C.J., 2010. The role of water in generating the calc-alkaline trend: new volatile data for Aleutian magmas and a new Tholeiitic Index.Journal of Petrology 51, 2411–2444.