You are on page 1of 26

Ore Geology Reviews 40 (2011) 1–26

Contents lists available at ScienceDirect

Ore Geology Reviews


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / o r e g e o r ev

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

a r t i c l e i n f o a b s t r a c t

Article history: Metals such as Cu, Mo, Au, Sn, and W in porphyry and related epithermal mineral deposits are derived
Received 1 December 2010 predominantly from the associated magmas, via magmatic–hydrothermal fluids exsolved upon emplacement into
Received in revised form 18 May 2011 the mid- to upper crust. Four main sources exist for magmas, and therefore metals, in convergent and collided plate
Accepted 19 May 2011
margins: the subducting oceanic plate basaltic crust, subducted seafloor sediments, the asthenospheric mantle
Available online 27 May 2011
wedge between the subducting and overriding plates, and the upper plate lithosphere. This paper firstly examines
Keywords:
the source of normal arc magmas, and concludes that they are predominantly derived from partial melting of the
Porphyry deposit metasomatized mantle wedge, with possible minor contributions from subducted sediments. Although some
Epithermal deposit metals may be transferred from the subducting slab via dehydration fluids, the bulk of the metals in the resultant
Subduction magmas are probably derived from the asthenospheric mantle. The most important contributions from the slab
Post-subduction from a metallogenic perspective are H2O, S, and Cl, as well as oxidants. Partial melting of the subducted oceanic crust
Magmatic–hydrothermal fluid and/or sediments may occur under some restricted conditions, but is unlikely to be a widespread process (in
Ore formation Phanerozoic arcs), and does not significantly 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 fO2 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 sulfide phases in the mantle source. At relatively
high abundances of sulfides thought to be typical of active arcs where fS2 and fO2 are high (magma/sulfide
ratio = 102–105), 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 sulfide abundance was very low (magma/sulfide
ratio N 105), 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, 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, continent–continent 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 significant modification during the previous subduction cycle, most importantly with the
introduction of large volumes of hydrous, mafic (amphibolitic) cumulates residual from lower crustal
differentiation of arc basalts. Small amounts of chalcophile and siderophile element-rich sulfides may also be left
in these cumulates. Partial melting of these subduction-modified 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 sulfides. 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 sulfide 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-modified 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 magmatic–hydrothermal fluid phase. This process occurs most
efficiently 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 efficiently into a single-phase,
supercritical aqueous fluid (~2–13 wt.% NaCl equivalent), which may exist as a separate volatile plume or as
bubbles entrained in buoyant magma. Focusing of upward flow of bubbly magma and/or fluid into the apical
regions of the batholithic complex forms cupolas, which represent high mass- and heat-flux channelways

⁎ Tel.: + 1 780 492 3430; fax: + 1 780 492 2030.


E-mail address: Jeremy.Richards@ualberta.ca.

0169-1368/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.oregeorev.2011.05.006
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 fluid
flow will be enhanced along the weakened and heated axes. Cupolas may form initially as breccia pipes by
volatile phase (rather than magma) reaming-out of extensional structures in the brittle cover rocks, to be
followed immediately by magma injection to form cylindrical plugs or dikes.
Cupola zones may extend to surface, where magmas and fluids vent as volcanic products and fumaroles. Between
the surface and the underlying magma chamber, a very steep thermal gradient exists (700°–800 °C over b 5 km),
which is the primary cause of vertical focusing of ore mineral deposition. The bulk of metals (Cu ± Mo± Au) that
forms porphyry ore bodies are precipitated over a narrow temperature interval between ~425° and 320 °C, where
isotherms in the cupola zone rise to within ~2 km of the surface. Over this temperature range, four important
physical and physicochemical factors act to maximize ore mineral deposition: (1) silicate rocks transition from
ductile to brittle behavior, thereby greatly enhancing fracture permeability and enabling a threefold pressure
drop; (2) silica shows retrograde solubility, thereby further enhancing permeability and porosity for ore
deposition; (3) Cu solubility dramatically decreases; and (4) SO2 dissolved in the magmatic–hydrothermal fluid
phase disproportionates to H2S and H2SO4, leading to sulfide and sulfate mineral deposition and the onset of
increasingly acidic alteration.
The bulk of the metal flux into the porphyry environment may be carried by moderately saline supercritical
fluids or vapors, with a volumetrically lesser amount by saline liquid condensates. However, these vapors rapidly
become dilute at lower temperatures and pressures, such that they lose their capacity to transport metals as
chloride complexes. They retain significant concentrations of sulfur species, however, and bisulfide complexing
of Cu and Au may enable their continued transport into the epithermal regime. In the high-sulfidation
epithermal environment, intense acidic (advanced-argillic) alteration is caused by the flux of highly acidic
magmatic volatiles (H2SO4, HCl) in this vapor phase. Ore formation, however, is paragenetically late, and may be
located 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- to
moderate-salinity liquids, and are thought to represent later-stage magmatic–hydrothermal fluids that have
ascended along shallower (cooler) geothermal gradients that either do not, or barely, intersect the liquid–vapor
solvus. Such fluids “contract” from the original supercritical fluid or vapor to the liquid phase. Brief intersection
of the liquid–vapor solvus may be important in shedding excess chloride and chloride-complexed metals (such
as Fe), so that bisulfide-complexed metals remain in solution. Such a restrictive pressure–temperature path is
likely to occur only transiently during the evolution of a magmatic–hydrothermal system, which may explain the
rarity of high-sulfidation Cu–Au ore deposits, despite the ubiquitous occurrence of advanced-argillic alteration in
the lithocaps above porphyry-type systems.
© 2011 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2. Magma generation in convergent and collided margins: geochemical characteristics and partitioning of metals . . . . . . . . . . . . . . . . . . 3
2.1. Slab dehydration and asthenospheric melting in subduction zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2. Sediment dehydration and/or melting in subduction zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.2.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.3. Oceanic slab melting in subduction zones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.3.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4. Supra-subduction zone lithospheric melting: the MASH process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.4.1. Behavior of metals. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.4.2. Sources of Mo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.5. Lithospheric melting during post-subduction events . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.5.1. Behavior of metals in subduction-modified sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.6. Crustal melting during post-collisional stress relaxation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.6.1. Sources of metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3. Behavior of metals during magma fractionation and fluid exsolution in the upper crust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.1. Partitioning of metals from magma into exsolving hydrothermal fluid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4. Magmatic–hydrothermal ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.1. Porphyry Cu ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.2. Epithermal Cu–Au ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2.1. High-sulfidation epithermal Cu–Au deposits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2.2. Low-sulfidation epithermal Au deposits (including alkalic-type deposits) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
5. Summary and conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.1. Sources of magmas and metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
5.2. Porphyry and epithermal ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 3

1. Introduction tized asthenospheric or lithospheric mantle, to melting of underplated


or primitive lower crustal rocks, and even melting of evolved crustal
The question of the source of various elements in convergent and rocks in the case of some felsic porphyry Mo and Sn–W magmas.
collided margin magmas has challenged geologists for decades. Igneous Therefore, rather than start by trying to identify a unique source
petrologists seek to understand the petrogenesis of such magmas for the typical intermediate-to-felsic calc-alkaline magmas that are
through geochemical and isotopic tracing, whereas economic geologists associated with ore deposits in mature convergent margins, I begin
are generally more interested in the source of potentially valuable this review by focusing on the much better constrained primitive
elements such as Cu, Mo, Sn, W, Au, and platinum group elements (PGE), island arc environment, where the effects of fractionation and crustal
which may ultimately be found in intrusion-related hydrothermal contamination, particularly by continentally derived materials, are
deposits. minimized, and processes in mantle source regions can be more
Igneous petrologists are broadly in agreement that arc magmas are clearly defined.
primarily derived from hydrous melting of the asthenospheric mantle
wedge above subducting plates, but melts from the subducted oceanic 2.1. Slab dehydration and asthenospheric melting in subduction zones
crust (including sediments) and the upper plate lithosphere may also
be involved to varying degrees. There is a general consensus that, with the exception of young
Economic geologists are also broadly in agreement that ore-forming oceanic lithosphere (b25 m.y.-old; Defant and Drummond, 1990) or
elements are partitioned from such magmas into an exsolving volatile plate edges (Yogodzinski et al., 2001), basaltic oceanic crust undergoes
phase upon emplacement in the upper crust, and may then be low-temperature, high pressure metamorphism upon subduction,
precipitated from these fluids during cooling, fluid mixing, and wallrock which releases fluids through a series of prograde dehydration reactions
reaction processes in porphyry-type and related epithermal mineral to form anhydrous eclogite (Fig. 1). Water and other volatile
deposits. However, these process theories do not address where the components and solutes (including S and Cl) were originally incorpo-
metals originally came from, nor why porphyry deposits vary so widely rated into oceanic crustal and upper mantle rocks during oxidizing
in their metal contents (from Au-rich, through Cu± Mo ± Au, to Mo- seafloor alteration, generating hydrous minerals such as serpentine, talc,
only deposits, with Sn–W deposits forming a distinct variant). amphibole, micas, chlorite, zoisite, chloritoid, and lawsonite. Various
In addition to subduction-related calc-alkaline magmas, a diverse experimental studies have shown that these minerals undergo
suite of calc-alkaline to alkaline magmas is generated in post- dehydration reactions over a depth range extending to ~100 km,
subduction and collisional tectonic settings, and these magmatic corresponding to the blueschist–eclogite transition in crustal rocks;
systems may also generate porphyry and epithermal ore deposits. serpentine (antigorite) and the 10 Å equivalent of chlorite may extend
Such systems raise an additional set of petrogenetic and metallogenic this range to 200 km (e.g., Dvir et al., 2011; Forneris and Holloway, 2003;
questions. Fumagalli and Poli, 2005; Poli and Schmidt, 2002; Schmidt and Poli,
It is the intent of this paper to merge these different geological 1998; Ulmer and Trommsdorff, 1995). Below these depths, the
perspectives on magmagenesis and metallogeny in order to discuss anhydrous eclogitic crust is essentially infusible, and the dense slab
primary metal fluxes in convergent and collisional margins in terms of continues its descent into the mantle without melting. See reviews of
igneous petrogenetic and magmatic–hydrothermal processes. The this subject by Richards (2003, 2005, and references therein).
ultimate metal inventory and metal ratios in any given porphyry or Numerous studies have explored the character of the fluids that are
related deposit is secondarily controlled by late-stage magmatic and released from the dehydrating slab, because they are thought to account
shallow crustal processes. These processes are examined, closing with a for the unique geochemical character of subduction-related magmas
review of fluid and metal sources and behavior in related epithermal during later partial melting in the metasomatized asthenospheric
environments. mantle wedge (located between the downgoing slab and the upper
plate). Slab-derived fluids are thought to be water-rich at these depths,
2. Magma generation in convergent and collided margins: and to carry significant amounts of other volatile components such as Cl
geochemical characteristics and partitioning of metals and S. For example, salinities in the range of 4–10 wt.% NaCl have been
inferred from primitive basalt or melt inclusion studies (de Hoog et al.,
Most magmas erupted through or emplaced within the Earth's crust 2001; Kent et al., 2002; Portnyagin et al., 2007; Wallace, 2005;
are not primary magmas (in the sense of being chemically unmodified Wysoczanski et al., 2006), and salinities of 0.4–2 wt.% NaCl equivalent
since extraction from their source), and most are not even primitive (in were measured in fluid inclusions from high-pressure rocks thought to
the sense of being relatively unevolved; Hildreth and Moorbath, 1988; represent subducted oceanic mantle (Scambelluri et al., 2004). At higher
Leeman, 1983; Neuendorf et al., 2005; Smith et al., 2010; Thirlwall et al., pressures (~6 GPa) and greater depths (~175 km) there may no longer
1996). Except for magmas produced and erupted in extensional tectonic be a physical distinction between solute-rich aqueous liquids and
regimes (where rapid ascent to the surface is facilitated by normal hydrous silicate melts, and the fluid may be supercritical in nature (e.g.,
faulting), most deeply-derived magmas undergo some degree of Kessel et al., 2005a,b), but the role of such deeply released fluids in
fractionation and crustal contamination during their passage towards subduction zone magmatism is unclear (see discussion in Richards and
the surface. It is therefore challenging to isolate geochemical and Kerrich, 2007).
isotopic characteristics of magma source regions from the effects of later In addition to volatiles, water-soluble large ion lithophile elements
processes (Davidson, 1996). Magmas erupted through mature conti- (LILE: K, Rb, Cs, Ca, Sr, and Ba, and U), and B, Pb, As, and Sb (Breeding et al.,
nental crust are the most difficult to fingerprint uniquely in terms of 2004; Hattori et al., 2005; Hattori and Guillot, 2003; Kogiso et al., 1997;
source characteristics because wallrock assimilation and fractional Manning, 2004; Tatsumi et al., 1986) are thought to be mobilized into the
crystallization (AFC; DePaolo, 1981) are ubiquitous and commonly forearc mantle wedge by dehydration fluids, which may then be
extensive processes that will significantly modify bulk rock geochemical convected by corner-flow into the sub-arc melting zone (Fig. 1). These
and isotopic compositions; and yet, these are also the magmas that are fluid–mobile components are also characteristically enriched in arc
most commonly associated with porphyry- and epithermal-type magmas (e.g., typical ranges of: 1–3 wt.% H2O, 500–2000 ppm Cl, 900–
mineral deposits. The difficulty in constraining source characteristics 2500 ppm S; Davidson, 1996; Gill, 1981; Noll et al., 1996; Sobolev and
in such magmas is perhaps responsible for the plethora of theories that Chaussidon, 1996; Wallace, 2005; Portnyagin et al., 2007), which is taken
have been proposed for the origin of ore-forming magmas in convergent as evidence of aqueous fluid metasomatism of the mantle wedge magma
margin settings, ranging from the melting of subducting oceanic crust source. Silica may also be significantly mobilized in these slab fluids
and/or seafloor sediments, through melting of subduction metasoma- (Aerts et al., 2010; Manning, 2004), as well as Tl and Cu (Noll et al., 1996;
4 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

Volcanic arc
Sea level 0 km
Sediment
Oceanic crust Oceanic crust
Chlorite + Mantle
Oceanic mantle

serpentine lithosphere
lithosphere

Talc 600°C
Chlorite
Am
p
De hib 1000
hy ole Man °C
dr tle c
lith atio orne
r flo
os n o w
ph f o Zo Partial melting
Asthenosphere er ce is 100 km
e a ite 140
60 ni Metasomatized 0°C
0° c C
C C td asthenosphere
hl
10 o
00 + rite
°C se –1
rp 0Å
en p

10
tin ha E

00
e se c 14

°C
lo 00
gi
te °C
Se
rp
en
tin
e

200 km

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 plate
indicate lower limits of stability of hydrous phases in the basaltic oceanic crust and peridotitic mantle lithosphere. Abbreviation: Ctd = chloritoid.

Stolper and Newman, 1994). The normally relatively incompatible high Parkinson and Arculus, 1999; Rowe et al., 2009). The relatively high
field strength elements (HFSE) Ti, Nb, and Ta are not significantly fluid oxidation state of arc magmas is a critical factor in their subsequent
soluble under subduction zone conditions, and are retained in minerals metallogeny, and originates from oxidative seafloor alteration of the
such as rutile either in the slab or the mantle wedge (Audétat and oceanic plate (Staudigel et al., 1996), transmitted into the mantle wedge
Keppler, 2005; Brenan et al., 1994; Green and Adam, 2003). Arc magmas by the metasomatic fluid flux (Brandon and Draper, 1996; Kelley and
derived from these sources therefore show characteristic negative Cottrell, 2009; Malaspina et al., 2009).
anomalies for these three elements on mantle-normalized spider
diagrams, but display enrichments in most other incompatible elements 2.1.1. Behavior of metals
(Foley et al., 2000; Gill, 1981; Klemme et al., 2005; Ryerson and Watson, Most base and precious metals would be expected to have at least
1987; Schmidt et al., 2004; Schmidt et al., 2009). moderate solubilities in the hot, relatively oxidized, saline aqueous
Hydrous metasomatism of the mid-ocean-ridge basalt (MORB)- fluids exsolved from the downgoing slab. In particular, as noted in
depleted asthenospheric mantle wedge causes partial melting by Section 2.1, Pb, As, and Sb are strongly mobilized by such fluids, possibly
lowering the solidus of peridotite (Arculus, 1994; Kushiro et al., 1968; along with Tl and Cu (Noll et al., 1996). The behavior of highly
Stolper and Newman, 1994). This occurs either through direct infiltration siderophile elements (HSE) such as Au and PGE is less well known under
metasomatism by slab-derived fluids percolating into the hot inner zone these conditions, but studies of metasomatized mantle xenoliths from
of the mantle wedge (Bourdon et al., 2003; Grove et al., 2006; Kelley et al., island arc lavas suggest that Au, Re, and the Pd-group elements
2010; Peacock, 1993), or by convective corner-flow mixing of metaso- (including Pt) are mobilized into the mantle wedge during subduction
matized peridotite into these hotter central regions (Fig. 1; Schmidt and metasomatism (Dale et al., 2009; Kepezhinskas et al., 2002; McInnes et
Poli, 1998; Tatsumi, 1986; Wysoczanski et al., 2006). al., 1999; Sun et al., 2004a; Widom et al., 2003).
Partial melting of hydrated peridotite under these conditions in The volumetric extent and efficiency of mobilization of metals into
the mantle wedge generates high-Mg basalts (Greene et al., 2006; the mantle wedge by this process are unknown, but fluid metasomatism
Pichavant et al., 2002; Smith et al., 2010). Such arc basalts are clearly represents one viable mechanism for metal transfer into arc
distinguished from MORB by higher contents of incompatible magma sources.
elements (as noted above) and water (up to 6 wt.% H2O; Cervantes The behavior of chalcophile and siderophile metals during subse-
and Wallace, 2003; Grove et al., 2003; Pichavant et al., 2002; Sobolev quent partial melting of the metasomatized mantle wedge depends
and Chaussidon, 1996). Critically, Hamada and Fujii (2008) and critically on oxidation state (fO2) and sulfur fugacity (fS2), because these
Zimmer et al. (2010) report that a water content of 2 wt.% separates parameters control the stability and abundance of sulfide phases. Gold
“dry” tholeiitic (olivine + plagioclase/orthopyroxene) from “wet” and PGE partition strongly into sulfide phases relative to silicate melts,
calc-alkaline (clinopyroxene + magnetite) magmatic fractionation but Cu to a somewhat lesser degree (Campbell and Naldrett, 1979;
trends. Peach et al., 1990), so if sulfides are abundant in the magma source
Arc basalts are also characterized by distinctly higher oxidation region, partial melts will be depleted in Au and PGE relative to Cu
states than MORB (up to 2 log units above the fayalite–magnetite– (Fig. 2). Under the high fO2 and fS2 conditions of the supra-subduction
quartz buffer: ΔFMQ + 2; Ballhaus, 1993; Brandon and Draper, 1996; zone mantle wedge, the bulk of the sulfur flux will likely consist of SO2 or
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 5

sulfate, dissolved first in slab fluids and then magma (Carroll and Nevertheless, Aizawa et al. (1999); Dreyer et al. (2010); and Duggen
Rutherford, 1985; Jenner et al., 2010; Jugo et al., 2005a). However, et al. (2007) have suggested that dehydration is the principal process
because of the equilibria between various sulfur species, at high fS2 some affecting sediments down to depths of ~100 km (i.e., to below the
condensed sulfide phases will likely also be present (McInnes et al., volcanic arc), with melting only occurring significantly at greater
2001). Consequently, Richards (2009) suggested that normal arc depths when temperatures exceed ~800 °C (possibly reflected in the
magmas will be minimally depleted in Cu (due to its higher abundance geochemistry of some back-arc magmas).
and moderate chalcophile nature) relative to sparse, highly siderophile Sediment contributions to the source of arc magmas have been the
elements such as Au and PGE, which will be strongly retained in residual subject of numerous studies, with the least ambiguous evidence
sulfide phases in the source region (e.g., Hamlyn et al., 1985; Mitchell coming from island arcs (e.g., MacDonald et al., 2000; Thirlwall et al.,
and Keays, 1981; Peach et al., 1990). This may explain the Cu-rich nature 1996; Wysoczanski et al., 2006). In continental arcs, it can be difficult
of typical arc-related porphyry deposits (relative to Au and PGE; to distinguish between chemical and isotopic signatures from
Richards, 2005). In contrast, Au-rich porphyry deposits may require subducted continent-derived sediment versus crustal contamination
atypical subduction-related or collisional tectonic settings and petro- during magma ascent (e.g., Hildreth and Moorbath, 1988; Kemp et al.,
genetic processes, which act to destabilize residual sulfide phases and 2007): both sources will contribute incompatible elements and
render Au incompatible (e.g., Jégo et al., 2010; Richards, 1995, 2009; crustal isotopic values to primary mantle-derived arc magmas
Sillitoe, 2000; Solomon, 1990; Wyborn and Sun, 1994; see Section 2.5). (Breeding et al., 2004).
The low abundances of PGE in many arc-related ore deposits suggest a Trace elements commonly used as indicators of sediment contribu-
further separation of these elements from Au and Cu, perhaps through tions to arc magmas are Ba, B, Be, Th, and Pb (Dreyer et al., 2010; Johnson
the formation of residual platinoid alloy phases (e.g., Barnes et al., 1985; and Plank, 1999), and Ba/La and Th/La ratios can be used as a measure of
Borisov and Palme, 1997; Kepezhinskas et al., 2002; Peach et al., 1990) sediment versus mantle source components (Plank, 2005; Walker et al.,
or Cr-spinels (into which Ir-group PGE strongly partition; Hattori et al., 2001). In particular, the cosmogenic radioisotope 10Be can be used as a
2010; Righter et al., 2004). tracer of recent (b10 Ma) introduction of sediments into arc magma
sources (Dreyer et al., 2010; Morris et al., 1990). However, although a
2.2. Sediment dehydration and/or melting in subduction zones clear sediment-derived isotopic signature can be observed in many
island arc systems, the volumetric contribution of sediments to island
Seafloor sediments on the surface of the downgoing slab are arc magmas seems to be relatively minor (Hawkesworth et al., 1994;
another potential source of metasomatic contributions to the mantle Kilian and Behrmann, 2003; Poli and Schmidt, 2002; Stern et al., 2006).
wedge. Much of this sedimentary material will be scraped off at the One additional element that may be added to the mantle wedge
trench to form an accretionary prism, but varying amounts may also from subducted sediments is sulfur, as suggested by the positive δ 34S
be subducted, depending on the degree of coupling between the compositions of arc magmas (de Hoog et al., 2001), which are similar
upper and lower plates, and also the sediment input load (Fig. 1). Such to those of seafloor sediments (Alt et al., 1993). Analyses of glass
sediments will be water-rich and pelitic in bulk composition, and thus inclusions in olivine from primitive arc magmas reveal concentrations
are more likely to undergo partial melting under subduction zone of up to 2900 ppm S (de Hoog et al., 2001), and Jugo et al. (2005b)
conditions than basaltic oceanic crust (Hermann and Spandler, 2008). measured experimental concentrations of up to 1.5 wt.% S in oxidized
arc basalts. These high sulfur contents have great significance for the
behavior of chalcophile and siderophile metals (see Sections 2.1.1,
2.5.1, and 3).

Porphyry Cu Porphyry Cu-Au 2.2.1. Behavior of metals


potential arc potential
6 magmas magmas Lead, which is significantly enriched in the continental crust (and
10
d in crustally-derived sediments) relative to the mantle, is the only metal for
che ide
105 e nri l sulf
Au idua which a clear sedimentary source can be inferred in some island arc
res
104 magmas, because it can be readily identified by its radiogenic isotopic
de composition compared with depleted mantle sources. However, in
103 ulfi
in s Cu maximized in
Cu
lfid
e magma (R ≥ 10 3) continental arcs, distinguishing a subducted sediment source of
102 ns
u
i 50 ppm Cu radiogenic Pb from crustal contamination during magma ascent is
Cu (ppm)

Au
Au (ppb)

10
ag
m a
5 ppb Au very difficult (e.g., Barreiro, 1984; Chiaradia et al., 2004; Kontak et al.,
m
1 Cu
in
gma Au maximized in 1990). This has led to diverging opinions: for example, Aitcheson et al.
ma magma (R ≥ 106)
ed
in (1995); Hildreth and Moorbath (1988); James (1982); Kay et al. (1999);
0.1 plet
de Sulfide/silicate melt partition and Tilton et al. (1981) concluded that the bulk of the radiogenic Pb in
10–2 Au coefficients:
a
D(Cu) = 103 central Andean magmas comes from upper plate crustal sources,
gm D(Au) = 105
10–3 ma whereas Macfarlane (1999); McNutt et al. (1979); Mukasa et al.
n Metal concentrations in
ui
10–4
A magma in absence of sulfide: (1990); and Sillitoe and Hart (1984) preferred a subducted sediment
Cu = 50 ppm
Au = 5 ppb source for Pb in some Andean igneous rocks and ore deposits.
10–5
Inferring a seafloor sediment source for other metallic components
1 10 102 103 104 105 106 107 108
R = (mass of silicate melt)/(mass of sulfide) in arc magmas (and related ore deposits) is much more speculative,
and generally involves the subduction of metal-rich manganese
nodules or even massive sulfide deposits. The latter, however, likely
Fig. 2. Concentrations of Cu and Au in silicate magma and coexisting sulfide liquid as a
oxidize and disperse geologically rapidly after formation on the
function 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, magmas seafloor (e.g., Edwards, 2004; Herzig et al., 1991) unless quickly
will be depleted in both Cu and Au. At R-factors between ~ 102–105, magmas will be buried by lava; they might thus only be expected to be subducted with
depleted in Au but essentially undepleted in Cu (porphyry Cu-potential magmas). At R- very young oceanic crust. Few studies have specifically invoked a
factors N105, magmas will be undepleted in Au and Cu (porphyry Cu–Au–potential subducted sediment source for ore metals other than a component of
magmas). A corollary of this diagram is that arc magmatism will leave small amounts of
relatively Au-rich sulfide in the mantle source or lithosphere during fractionation,
Pb, and the majority of authors have concluded that such a source is
which can be remelted during post-subduction tectonomagmatic processes, to form either unnecessary or unproven (e.g., Burnham, 1981; Chiaradia et al.,
small-volume, alkaline, porphyry Au-potential magmas. 2004; de Hoog et al., 2001; Fontboté et al., 1990).
6 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

2.3. Oceanic slab melting in subduction zones The chemical difference between slab dehydration and slab
melting would seem to be rather small, given that both media
The question of whether the downgoing oceanic slab, or more would be enriched in volatiles, incompatible elements, and silica.
specifically the basaltic oceanic crust, melts during subduction has Indeed, as noted in Section 2.1, there may well be a continuum
prompted lively debate recently, not only in the petrology literature between silica-rich aqueous fluids and aqueous silicate melts at
(e.g., Conrey, 2002; Defant and Kepezhinskas, 2001, 2002; Garrison and greater depths in subduction zones (Kawamoto, 2006; Kessel et al.,
Davidson, 2003), but also amongst economic geologists because of the 2005a,b; Manning, 2004; Portnyagin et al., 2007). This likely explains
suggestion that “slab melts” might in some way be uniquely fertile for later why the debate between slab melting and dehydration is somewhat
porphyry ore formation (e.g., Mungall, 2002; Oyarzun et al., 2001, 2002; intractable, and mostly hinges on subtle trace element characteristics.
Sajona and Maury, 1998; Thiéblemont et al., 1997; for contrary opinions,
see: Rabbia et al., 2002; Richards, 2002; Richards and Kerrich, 2007). 2.3.1. Behavior of metals
The idea that the hydrated basaltic ocean crust might melt during Slab melts are predicted to be volatile-rich (including H2O, S, and
subduction was an early assumption of the plate tectonic revolution, Cl) and oxidized, and thus, like hydrous slab fluids, would be expected
because it seemed conveniently to explain the relatively felsic to be able to transport base and precious metals at least to some
composition of arc magmas (as opposed to basalts formed by melting degree. However, analyses of such metals (except iron) are not
of peridotitic mantle). The theory was given credence by the reported in most melt inclusion studies (e.g., Kilian and Stern, 2002;
experiments of Rapp et al. (1991) and Rapp and Watson (1995), who McInnes and Cameron, 1994; Schiano et al., 1995), so there are no
showed that melting of amphibolite under upper mantle conditions direct constraints on the capacity of such melts to transfer chalcophile
(1025 °C and 0.8–1.6 GPa) could generate an intermediate composition and siderophile metals from the slab to the mantle wedge.
tonalite–trondhjemite melt, not dissimilar to an arc andesite. Defant and In a study of metasomatized mantle xenoliths from a submarine
Drummond (1990) termed the products of subducted slab melting volcano near Lihir Island, Papua New Guinea, McInnes et al. (1999)
“adakites”, after a single anomalous lava flow on Adak Island in the concluded that enrichments in Cu, Au, and PGE were caused by slab fluid
Aleutians described by Kay (1978). Because garnet should be present in metasomatism, rather than melts. In contrast, Kepezhinskas et al. (2002)
the eclogitic source of these magmas, Defant and Drummond (1990) measured the concentrations of Au and PGEs in mantle xenoliths from
argued that such melts could be distinguished by anomalously low the Kamchatka arc, and suggested that a fluid-transported component
concentrations of heavy rare earth elements (HREE) and Y (which are could be distinguished from a slab melt component by co-enrichments in
compatible with garnet) relative to light rare earth elements (LREE), and PGE and high field strength elements (HFSE) in the latter, because of the
high concentrations of Sr (because of the absence of plagioclase at such low capacity of aqueous fluids to carry HFSE. Intriguingly, they noted no
depths). Thus, slab melts, or adakites, could be distinguished on Sr/Y such correlation between HFSE-enrichments and Au, and concluded that,
versus Y, or La/Yb versus Yb diagrams from normal arc magmas formed whereas PGE might be transported into the mantle wedge by both fluids
in the absence of garnet. and melts, Au was likely only carried by hydrous fluids.
However, despite the theoretical possibility of melting subducted Two theoretical studies have proposed that slab melts should be
oceanic crust, most thermal models of subduction zones indicate that unusually effective as metal-transporting and ore-forming agents.
temperatures in the slab do not normally reach melting conditions Oyarzun et al. (2001) argued that slab melts should be unusually
(N800 °C) prior to dehydration and eclogitization (Fig. 1), which would oxidized and rich in H2O and SO2 (relative to normal arc magmas
render the slab infusible (e.g., Davies and Stevenson, 1992; Peacock, derived by asthenospheric partial melting), although no evidence was
1996; Poli and Schmidt, 2002). Thus, Defant and Drummond (1990) given for this assertion. Such magmas, they argued, should be
proposed that slab melting might be restricted to the subduction of particularly suitable for the formation of magmatic–hydrothermal
young (≤25 m.y. old) and therefore still hot oceanic crust, and Peacock porphyry copper deposits upon emplacement in the upper crust.
et al. (1994) were even more restrictive (b5 m.y. old). Other relatively Mungall (2002) presented a theoretical model for oxidation of the
uncommon scenarios that might result in slab melting include shallow mantle wedge by Fe 3+-rich slab melts to the point of complete sulfide
or stalled subduction (whereby the slab has more time to heat up at destruction, thereby rendering chalcophile and siderophile elements
shallow depths; Gutscher et al., 2000; Peacock et al., 1994), ridge incompatible in mantle phases, and free to partition into silicate
subduction (Guivel et al., 2003; Kay et al., 1993), or edge-melting of melts. He argued that ferric iron is a much stronger oxidant than slab-
detached slabs or slab windows (Haschke and Ben-Avraham, 2005; derived water, and that slab melts should be rich in Fe 3+ generated by
Thorkelson and Breitsprecher, 2005; Yogodzinski et al., 2001). oxidative seafloor alteration. Thus, slab melts might be uniquely
A further complication is added by the fact that most adakites favorable for the subsequent generation of metal-rich, and particu-
described in the petrology literature are not in fact primary slab melts, larly Au-rich, magmas derived from the mantle wedge.
but are substantially evolved, having reacted or hybridized with the Mungall's (2002) model may have applicability for less common Au-
asthenosphere during ascent (and likely also the upper plate rich porphyry deposits formed in atypical subduction zone settings that
lithosphere). This modification to the adakite slab-melting model is might cause slab melting, but does not seem well suited to explain
required to explain the high contents of MgO, Ni, and Cr present in regular arc porphyry Cu deposits. In either case, metals are envisaged to
some adakites relative to expected levels for hydrated basalt partial be sourced from the mantle wedge, not the slab. In contrast, Oyarzun et
melts (Defant and Kepezhinskas, 2001; Drummond et al., 1996; al.'s (2001) model does not address the source of metals, and is at root
Martin, 1999; Martin et al., 2005; Yogodzinski et al., 1995). based on the assumption that slab melts are uniquely more H2O- and
Direct evidence for slab melting is lacking, but supra-subduction zone SO2-rich, and more oxidized than normal arc magmas, leading to
xenoliths from the Tabar-Lihir-Tanga-Feni (Papua New Guinea), Philip- specific ore depositional processes rather than source processes. It is not
pine, and Patagonian arcs preserve hydrous, silica-rich glass inclusions clear that these assumptions are justified, and some have argued that
that are thought to represent migrating slab melts (respectively: Kilian slab melts might in fact be relatively reducing because of the additional
and Stern, 2002; McInnes and Cameron, 1994; Schiano et al., 1995). The presence of organic-rich sediment melts (Wang et al., 2007a).
glass inclusions characterized by Schiano et al. (1995) were calc-alkaline
in composition, with high incompatible element and low Ti, Nb, and Y 2.4. Supra-subduction zone lithospheric melting: the MASH process
contents, high LREE/HREE ratios, and homogenization temperatures of
~920 °C. They thus compositionally resemble melts that would be Hydrous basaltic magmas generated in the mantle wedge will have
predicted to form from slab melting, and appear to provide evidence that temperatures in excess of 1000 °C (Eggins, 1993; Grove et al., 2006;
this process occurs at least locally where conditions permit. MacDonald et al., 2000), and perhaps as high as 1350 °C (Schmidt and
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 7

Poli, 1998; Tatsumi, 2003). Because their densities will be lower than 2006; Richards and Kerrich, 2007; Tulloch and Kimbrough, 2003).
the mantle but not the crust (Herzberg et al., 1983), they will tend to Such common processes, affecting batches of magma crystallizing and
rise from their asthenospheric source region and penetrate the mantle fractionating at different crustal depths (e.g., Annen et al., 2006) are
lithosphere, but pool at the crust/mantle density barrier (“level of entirely consistent with petrological observations in arc volcanic
neutral buoyancy”: Fig. 3; Fyfe, 1992; Hildreth, 1981). Here, if the flux systems where “adakite-like” (i.e., high-Sr/Y) andesitic lavas may be
of magma and heat is maintained and supplemented by the latent interlayered with “normal” andesites in a single volcano, and do not
heat of crystallization as the magma begins to crystallize, high require a fundamental change in magma source 100 km below the
temperatures can be brought to bear on lower crustal rocks that will volcano (e.g., Feeley and Davidson, 1994; Grunder et al., 2008;
cause partial melting (Annen et al., 2006; Bergantz and Dawes, 1994; Richards et al., 2006a).
Huppert and Sparks, 1988; Klepeis et al., 2003; Petford and Gallagher, Once these hybrid magmas reach basaltic andesitic to andesitic
2001; Rushmer, 1993). Hildreth and Moorbath (1988) suggested that compositions, their densities are low enough to allow them to rise
it is the interaction between this hot, hydrous basalt flux from the through the lower continental crust (Herzberg et al., 1983), but they
subduction zone and felsic crustal partial melts that gives rise to the tend to stall again at a second density barrier in the middle to upper
uniform composition of andesites in continental volcanic arcs, by a crust below light supracrustal rocks. This is the level (5–10 km) at which
process they dubbed melting–assimilation–storage–homogenization large arc batholiths will form if the flux of magma is sustained, and is
(MASH). In a refinement of this model, Annen et al. (2006) referred to also the level at which evolved felsic melts and volatiles are accumulated
the region of magma–crust interaction as a “hot zone” (Fig. 3). (Fig. 3; see Richards, 2003, and references therein). These volatiles drive
Because garnet is a product of such lower crustal fractionation and buoyant, bubbly, evolved magma upwards into the cover rocks to form
partial melting processes (Alonso-Perez et al., 2009; Berger et al., subvolcanic stocks and dikes, or explosive volcanic eruptions if they
2009; Dufek and Bergantz, 2005; Garrido et al., 2006; Hansen et al., reach surface. The volatiles may also separate from the magma flux to
2002; Klepeis et al., 2003; Rushmer, 1993; Wolf and Wyllie, 1994), form a separate fluid plume, which ultimately vents at the surface
derivative calc-alkaline magmas may display trace element compo- (fumaroles) but may also form porphyry- and epithermal-type deposits
sitions that resemble adakites (see Section 2.3) but which are in the hypabyssal and near-surface environment (see discussion of
unrelated to slab melting (Klepeis et al., 2003; Macpherson et al., these processes in Sections 3 and 4).

Volcanic arc

Hydrothermal Epithermal deposits


alteration Porphyry deposits
Sea level
0 km
Mid- to upper crustal
batholithic complex

Feeder dikes
Continental crust

Lower crustal
MASH or “hot zone”
600°C
50 km

1000
°C
Subcontinental
De mantle lithosphere
hy Man
dr tle c
at orne
in r flow
g
oc
O ea Partial melting
lith macea ni
c
140
0°C
os nt nic cr Metasomatized
ph le us
60

10

er t asthenosphere
00

e
0°C

14 Asthenosphere
°C

00
°C 100 km

Fig. 3. Schematic section through a continental arc, showing the development of a MASH or “hot zone” at the base of the crust where basaltic arc magmas pool at their level of neutral
buoyancy, differentiate, and interact with crustal rocks and melts. Evolved, less dense, andesitic magmas rise into the mid-to-upper crust where they pool at their new level of neutral
buoyancy to form batholithic complexes. Along with volcanic structures, porphyry and epithermal deposits may form at shallower levels above these batholithic complexes where
exsolved magmatic fluids ascend, cool, and interact with near-surface upper crustal rocks. Modified from Richards (2003, 2005); sources: Hildreth and Moorbath (1988), Winter (2001),
Annen et al. (2006), and Sillitoe (2010).
8 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

2.4.1. Behavior of metals derived from continental crustal sources (Farmer and DePaolo, 1984;
Some of the clearest evidence for the involvement of crustal rocks in Stein, 1988; Klemm et al., 2008; White et al., 1981). On the other hand,
continental arc magmas comes from Pb isotopic data (e.g., Wörner et al., minor amounts of Mo do occur in some island arc-related porphyry
1992), although as noted in Section 2.2, there may be ambiguity deposits where no continental crustal sources are inferred (Westra and
between lower crustal melting and the melting of subducted continent- Keith, 1981), so a mantle (subduction zone) source for at least some Mo
derived sediments. Lead and uranium are much more abundant in the cannot be excluded. Moreover, Blevin and Chappell (1992) and Blevin
bulk continental crust (11 ppm Pb, 1.3 ppm U; Rudnick and Gao, 2003) et al. (1996) have demonstrated a continuum from Cu–Au deposits
or lower continental crust (4 ppm Pb, 0.2 ppm U; Rudnick and Gao, associated with unevolved, mafic I-type granitoids to W–Mo deposits
2003) than the primitive mantle (b0.2 ppm Pb, b0.02 ppm U; Taylor associated with cogenetic, evolved granites in eastern Australia,
and McLennan, 1985; Sun and McDonough, 1989), MORB (0.3 ppm Pb, suggesting a common, magmatic source for all of these elements.
0.05 ppm U; Sun and McDonough, 1989), or typical low-K mafic arc A complication is introduced in the Climax-type deposits, because
andesites (b1.8 ppm Pb, b0.2 ppm U; Gill, 1981), so it only requires although the immediate source of the Mo-bearing fluids is felsic
small amounts of contamination by radiogenic crustal lead to magma of clear crustal origin, many deposits also show a close genetic
significantly modify a magma's Pb isotopic composition. Thus, it is not association with mafic alkaline magmas, which may have introduced
clear that a particularly large amount of Pb in arc magmas is sourced volatiles, S, and possibly Mo into the evolved felsic magma chamber
from crustal rocks versus the subduction zone, and Macfarlane et al. (Audétat, 2010; Carten et al., 1993; Keith et al., 1986, 1998). Keith et
(1990) have argued that the crustal contribution is minimal in Central al. (1997), Hattori and Keith (2001), and Maughan et al. (2002) have
Andean magmas and ores. Moreover, porphyry-type systems are also suggested that injections of mafic alkaline magmas into the
typically not Pb-rich, except in late-stage skarns and distal veins evolving Bingham Canyon magmatic system may have given rise to
where some of the Pb may have been derived from local host rocks (e.g., the unusually large size and high grades of this porphyry Cu–Mo–Au
Mukasa et al., 1990). deposit. Along the same lines, Pettke et al. (2010) have proposed that
Most researchers have assumed that, because of the higher the unusual Cu–Mo–Au endowment of the southwestern USA (e.g.,
concentrations of Cu in primitive andesites (145 ppm; Gill, 1981) the giant Bingham, Butte, Climax, Henderson, and Questa porphyry
compared with the bulk continental crust (27 ppm; Rudnick and Gao, Cu–Mo–Au and porphyry–Mo deposits) reflects Cenozoic remobiliza-
2003), the bulk of Cu in porphyry-type deposits is mantle-derived. A tion of Proterozoic subduction-metasomatized subcontinental mantle
similar assumption is made for Au, although the primitive mantle and lithosphere (see Section 2.5).
continental crust actually have comparable concentrations (1–3 ppb Thus, at this time there is no consensus regarding the crustal
Au; Rudnick and Gao, 2003; Taylor and McLennan, 1985). Moreover, versus mantle origin of molybdenum in porphyry deposits, although it
porphyry Cu–(Au) deposits are found in association with mantle- is clear that the highest grade porphyry Mo deposits are formed in
derived arc magmas worldwide, regardless of crustal type (oceanic or intra-plate continental settings, and if a mantle source is important in
continental) or thickness (Kesler, 1973), so a crustal heritage for these these cases, it is not directly related to subduction activity but rather
metals does not appear to be critical. Nevertheless, a lower crustal to rifting or reactivation of previously subduction-enriched litho-
source, perhaps hybridized with mantle-derived magmas, has been spheric sources.
proposed by Bouse et al. (1999) for both magmas and metals in the
Laramide porphyry systems of Arizona. Moreover, Titley (1987, 2001)
has specifically suggested that Au and Ag are crustally derived in a 2.5. Lithospheric melting during post-subduction events
range of ore deposits including porphyries in southwestern USA,
because the ratios of these elements correlate closely with two A number of mineral deposits with broad similarities to those
distinct basement domains in this region. Given that Au is not formed by subduction-related processes are also found in post-
especially enriched in the mantle (see above), and that Ag is in fact subduction tectonic settings, such as subduction reversal or migration,
more abundant in the crust than the mantle (80 ppb in the bulk arc collision, continent–continent collision, and post-collisional rifting.
continental crust, versus b19 ppb in the primitive mantle; Taylor and They include porphyry Sn–W, Mo, Cu–Mo, and Cu–Au deposits and
McLennan, 1985), a crustal source for at least some proportion of epithermal Au deposits, and in many cases are only known not to be
these minor metals, and especially Ag, in arc magma-related systems directly related to subduction because of precise geochronology and
may be reasonable. However, it seems unlikely that this argument can plate tectonic reconstructions that place their formation after subduc-
be extended to copper, except perhaps on the margins. tion has demonstrably ceased. Associated magmas are typically calc-
alkaline, but tend towards somewhat more alkaline compositions
2.4.2. Sources of Mo compared with normal arc magmas (Richards, 2009).
Molybdenum occurs in varying amounts in porphyry-type deposits, In complex accretionary arcs, it can be very difficult to ascribe any
ranging from trace levels (b0.01 wt.% Mo) in porphyry Cu–(Mo) deposits, given pluton (and any associated mineral deposits) to a particular
where it may not even be recovered as a byproduct, to being the main ore subduction or collisional event, because subduction commonly con-
component (up to 0.3 wt.% Mo) in porphyry Mo deposits (Seedorff et al., tinues after collision, albeit normally with a shift in the locus of
2005; Westra and Keith, 1981). At the Mo-rich end of the spectrum, there magmatism. However, in continent–continent collision zones or where
are two clearly different tectonomagmatic associations, only one of which arc collision terminates subduction, there can be greater certainty about
is directly related to subduction: calc-alkaline porphyry Mo deposits are the timing of cessation of subduction magmatism. Consequently, it is in
generally relatively low grade (0.1–0.02 wt.% Mo; Carten et al., 1993), collisional orogens such as the Neo-Tethyan belt of southeastern Europe
whereas intra-cratonic rift-related deposits associated with high-silica, and southern Asia that some of the clearest examples of post-subduction
fluorine-rich, peraluminous granitoids are relatively high grade (0.1– magmatism and mineralization are found. These include, from east to
0.3 wt.% Mo; e.g., “Climax-type” deposits; Carten et al., 1993; Kirkham and west, the Miocene Gangdese porphyry Cu–Mo belt in the Tibetan orogen
Sinclair, 1996; Sinclair, 2007; Stein, 1988; White et al., 1981). (Hou et al., 2006, 2009; Yang et al., 2009), the Miocene Kerman porphyry
Kesler (1973) noted a general association (with exceptions) of Cu–Mo belt in southeastern Iran (Shafiei et al., 2009), the Miocene Sari
porphyry Cu–Au deposits in island arcs, and porphyry Cu–Mo deposits Gunay epithermal Au deposit in northwestern Iran (Richards et al.,
in continental arcs, and it is clear that the peraluminous felsic rocks 2006b), the Eocene Çöpler epithermal Au deposit in southeastern Turkey
associated with rift-related Climax-type porphyry Mo deposits are (Keskin et al., 2008; Kuscu et al., 2010), the Pliocene Kisladag porphyry
primarily of continental crustal origin (Farmer and DePaolo, 1984; Au deposit in western Turkey, the Miocene Skouries porphyry Cu–Au–
Stein, 1988). This has led to one view that Mo might be predominantly PGE deposit in Greece (Economou-Eliopoulos and Eliopoulos, 2000), and
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 9

the Roşia Montană epithermal Au deposit in Romania (Manske et al., predominantly present in the melt as sulfate or SO2 (Carroll and
2006; Neubauer et al., 2005). Rutherford, 1985). Xenoliths from supra-subduction zone mantle
Similarly, in the southwest Pacific ocean, accurate plate tectonic (McInnes et al., 1999) and samples of mafic cumulates from lower
reconstructions permit the identification of a number of post- crustal arc roots (Fig. 5; Canil et al., 2010; Greene et al., 2006; Jagoutz
subduction porphyry and epithermal deposits, such as the Grasberg et al., 2007) reveal the common presence of small amounts of sulfide,
porphyry Cu–Au deposit in Papua, Indonesia (Cloos et al., 2005; typically trapped as inclusions in silicate phases (suggesting a primary
Paterson and Cloos, 2005), the Ok Tedi porphyry Cu–Au deposit (van magmatic rather than secondary hydrothermal origin).
Dongen et al., 2010) and the Porgera alkalic-type epithermal Au deposit Hamlyn et al. (1985), Richards (1995, 2009), Solomon (1990), and
in mainland Papua New Guinea (Richards et al., 1990; Richards and Wyborn and Sun (1994) have explored the role of residual sulfide
Kerrich, 1993), the Lihir alkalic-type epithermal Au deposit on Lihir phases on the metal content of fractionating magmas, and also of
Island, Papua New Guinea (Carman, 2003; Kennedy et al., 1990), and the partial melts formed during later, post-subduction melting events.
Emperor alkalic-type epithermal Au deposit in Fiji (Gill and Whelan, The high partition coefficients for chalcophile and highly siderophile
1989; Setterfield et al., 1992). (For reviews of alkalic-type epithermal elements (HSE) between sulfide phases and silicate melt mean that
deposits, see Jensen and Barton, 2000 and Richards, 1995). such metals should be strongly partitioned into any coexisting sulfide
Because subduction has ceased in these regions, a fresh supply of phases (Campbell and Naldrett, 1979; Peach et al., 1990). As shown in
fluids, volatiles, and other slab-derived components to the mantle Fig. 2, at high abundances of sulfide relative to silicate melt (low R-
wedge no longer exists. Nevertheless, the broad geochemical similarity factor; Campbell and Naldrett, 1979), the melts will be depleted in all
of many of these magmas to normal arc magmas, including their of these chalcophile and siderophile elements. In contrast, at
hydrous and generally oxidized nature, suggests some link to intermediate abundances of sulfide (intermediate R-factor), only
subduction metasomatism. Consequently, many researchers have originally sparse HSE will show significant depletions. This led
implicated upper-plate lithospheric sources, modified by earlier Richards (2005, 2009) to propose that small amounts of sulfide left
subduction-related fluids and/or hydrous melts (e.g., Clemens et al., behind as residual phases from fractionation of arc magmas in the
2009; Cloos et al., 2005; Guo et al., 2007; Harris et al., 1986; Johnson deep lithosphere (or asthenosphere) will not significantly deplete
et al., 1978; Pearce et al., 1990; Pettke et al., 2010; Richards, 2009). those magmas in relatively abundant chalcophile elements such as Cu
Previously subduction-modified asthenosphere is unlikely to be a viable and Mo, but might significantly deplete them in highly siderophile
source except for a short period after subduction has ceased (e.g., elements such as Au. This would give rise to magmas with relatively
Richards et al., 1990; Solomon, 1990), because such material will be high Cu/Au ratios (which might form Cu-rich porphyry deposits), but
quickly dispersed by mantle convection. would leave a residue of potentially HSE-rich sulfides in the mantle
The key to all of these models is subduction-derived water, which is and/or lower crustal amphibolitic cumulate arc roots.
most likely stored in amphibolitic cumulates, residual from the earlier As noted in Section 2.5, subduction-modified asthenospheric sources
arc magma flux and located in the deep crust or mantle lithosphere (e.g., will be rapidly convected away when subduction ceases, and so could
Claeson and Meurer, 2004; Davidson et al., 2007; DeBari and Coleman, only contribute to immediately post-subduction magmatism. In
1989; Jagoutz et al., 2009; Larocque and Canil, 2010; Müntener and contrast, deep crustal amphibolites are preserved in the lithosphere,
Ulmer, 2006; Tiepolo and Tribuzio, 2008). Water lowers the solidus of and will be susceptible to partial melting at any later time due to thermal
silicate assemblages, and will lead to the formation of hydrous partial rebound or reheating by invading asthenospheric melts. Under lower fS2
melts during pro-grade metamorphism or mafic melt invasion (Beard post-subduction conditions (a flux of S from the subduction zone is no
and Lofgren, 1991; Rushmer, 1991; Wolf and Wyllie, 1994). longer present), any residual sulfide phases would likely dissolve into
Thermal rebound in thickened orogenic crust, delamination of the S-undersaturated silicate melt, carrying their metal loads with them
sub-continental mantle lithosphere, or post-collisional rifting (with (e.g., Ackerman et al., 2009). Richards (2009) proposed that this might
ingress of asthenospheric melts into the lower crust in the last two explain the occurrence of Au-rich porphyry and related epithermal
cases) can all cause small-volume partial melting of arc-metasoma- systems in some post-subduction settings, such as the alkalic-type
tized lithosphere and/or hydrous lower crustal cumulates (Fig. 4; epithermal Au deposits of the SW Pacific, and various post-collisional Au
Brown, 2010; Clemens et al., 2009; Harris et al., 1986; Richards, 2009). deposits in the Balkans–Turkey–Iran Neotethyan belt. Pettke et al.
Such melts, being derived from subduction-modified sources, will (2010) have proposed a similar model for giant porphyry Cu–Mo–Au
share many of the characteristic geochemical features of arc magmas, deposits in the southwestern USA.
including their relatively high water contents and oxidation states, This model can also explain the occurrence of Au–poor porphyry
and potentially metal contents (see Section 2.5.1). The smaller volume Cu–(Mo) deposits in post-subduction settings, such as those in Tibet
of partial melting to be expected in such tectonic settings will give and Iran (Hou et al., 2009; Shafiei et al., 2009; Wang et al., 2007b), the
these magmas a somewhat more alkaline composition than arc only difference being that in this case larger proportions of sulfide
magmas (e.g., Clemens et al., 2009), and will also mean that large may have fractionated out from the original arc magmas in the deep
batholithic complexes are unlikely to be formed, consistent with the crust. Such sulfides would have retained significant amounts of the
generally smaller and more isolated occurrence of such post- subduction flux of Cu and Mo, but HSE would be diluted to low
subduction magmatic systems (compared with arc-related Cordille- concentrations by the greater volume of sulfide (low R-factor; Fig. 2).
ran batholiths, or collisional S-type batholiths; Pitcher, 1997). Second-stage melts from such cumulate sources would therefore be
Because these post-subduction magmas are derived from amphibo- Cu–(Mo)-rich, but not necessarily Au-rich.
litic sources in which garnet (±titanite) is likely also present, and A control on these two scenarios (abundant Cu-rich residual sulfide
because their hydrous nature will suppress plagioclase fractionation versus sparse but HSE-rich residual sulfide, or low versus high R-factor)
(similar to other hydrous arc magmas), they may be characterized by might be the average oxidation state and sulfur fugacity of the generative
elevated Sr/Y and La/Yb ratios; that is, they may display adakite-like subduction system. In more oxidized or S-poor systems, smaller volumes
trace element characteristics. of HSE-rich sulfide would exsolve from the silicate melt (high R-factor;
Campbell and Naldrett, 1979), whereas in less oxidized or S-rich systems,
2.5.1. Behavior of metals in subduction-modified sources larger volumes of Cu-rich but HSE-poor sulfide would exsolve (low R-
As discussed in Section 2.1, arc magmas are characterized by high factor; Fig. 2). In particular, the proportion of sulfides exsolving from arc
fO2 and fS2 relative to normal melts from MORB-depleted astheno- magmas may be very sensitive to small changes in their oxidation state,
sphere. Consequently, such magmas may be sulfide-saturated (at because of the rapid change from sulfide to sulfate dominance in
oxidation states up to ΔFMQ + 2.3; Jugo, 2009) despite sulfur being magmatic systems between ΔFMQ +1 and +2 (Jugo et al., 2010). The
10 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

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 in mature island arcs. (b–d)
During post-subduction tectonic processes, previously subduction-modified sub-continental lithospheric mantle (SCLM) or lower crustal hydrous cumulate zones residual from
previous 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 arc
magmatism, 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 due
to the presence of hornblende (± garnet, titanite) in the amphibolitic lower crustal source rocks. See text for discussion.

oxidation state of the mantle wedge will depend on the character of the characteristic of such tectonic settings, resulting from partial melting of
flux from the subducting slab (e.g., a higher proportion of subducted pelitic protoliths in the deep crust triggered by the heat from upwelling
organic-rich sediment would lead to lower oxidation states; Wang et al., asthenospheric melts (Hildreth, 1981). Peraluminous S-type granites
2007a); this property is therefore likely to have a characteristic average (Chappell and White, 1974) are subsequently emplaced as large
value along any given arc at any particular period of time. batholith complexes in the mid- to upper orogenic crust (e.g., the
Variations in oxidation state over typical ranges for arc magmas Hercynian peraluminous granites of Europe; Barbarin, 1996; Clemens,
(ΔFMQ= 0 to +2; Ballhaus, 1993; Brandon and Draper, 1996; Blatter 2003; Darbyshire and Shepherd, 1994; Harris et al., 1986; Wyllie et al.,
and Carmichael, 1998; Malaspina et al., 2009; Parkinson and Arculus, 1976). These granites tend to be enriched in lithophile rather than
1999; Rowe et al., 2009) will not greatly affect the potential to form syn- chalcophile elements, reflecting their crustal origins, and may generate
subduction porphyry Cu–(Mo) deposits, but might control the Cu/HSE magmatic–hydrothermal deposits containing Sn, W, U, Mo, REE, Li, Be, B,
ratio in later magmas formed by post-subduction melting of these and F.
sulfide-bearing residues. Specifically, Au-rich (low Cu/Au) post-subduc-
tion porphyries might form in settings where previous arc magmatism 2.6.1. Sources of metals
was relatively oxidized (sparse but HSE-rich sulfide residue), whereas Tin and especially tungsten commonly accompany molybdenum in
Au-poor (high Cu/Au) post-subduction porphyries might form where porphyry deposits as trace metals and byproducts, but they also form a
previous arc magmatism was relatively reduced (more abundant Cu-rich class of porphyries on their own, associated with S-type granites in
sulfide residue). Such a mechanism might also explain why coeval belts continental orogens (Hart et al., 2005; Ishihara, 1981; Ishihara and
of porphyry deposits tend to have characteristic Cu/Au ratios. Murakami, 2006; Kerrich and Beckinsale, 1988; Kirkham and Sinclair,
Finally, partial melting of predominantly reduced, sulfide-rich crustal 1996; Lehmann, 1982). The Hercynian tin granites of Europe, and the
rocks in orogenic settings may lead to chalcophile and siderophile Bolivian and SE Asian tin belts are examples of such deposits, with
element-depleted, but potentially lithophile element-rich S-type mineralization occurring in skarns and greisens around the granite
magmas (see Section 2.6). intrusions, and to a lesser extent as internal stockworks and
disseminations (e.g., Černý et al., 2005; Meinert et al., 2005). As with
2.6. Crustal melting during post-collisional stress relaxation the source magmas, metals in these deposits appear to be predomi-
nantly of crustal origin (Hedenquist and Lowenstern, 1994). For
Collisional orogens commonly undergo crustal thickening followed example, in a recent assessment of the source of Sn in the Cornubian
by extensional or transpressional collapse. Bimodal magmatism is batholith of SW England, Williamson et al. (2010) concluded that all of
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 11

(a) strongly into sulfide phases exsolving or crystallizing from silicate


melts. Thus, if extensive fractionation and removal of magmatic
sulfide phases were to occur, the remaining silicate melt would be
strongly depleted in these elements (Jugo et al., 1999; Lynton et al.,
1993). This is in essence the model for formation of orthomagmatic
sulfide deposits from relatively reduced mafic magmas (e.g., Naldrett,
1989), and perhaps explains the deficiency of chalcophile elements
such as Cu and Au in more reduced, evolved, Sn–W-bearing S-type
Chalcopyrite magmas (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 their
evolution (as indicated by the common presence of sparse sulfide
inclusions in phenocrysts in arc volcanic rocks as well as lower crustal
arc 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 the
extent that sulfides physically separate from the magma, or at least
not in large amounts. Instead, small sulfide droplets or crystals may be
(b) entrained in magma ascending buoyantly through the crust (e.g.,
Bockrath et al., 2004; Tomkins and Mavrogenes, 2003) or as inclusions
in silicate phenocrysts, and will not be substantially lost to the overall
magma flux. Indeed, several authors have argued that pre-concen-
Pyrite tration of ore metals in magmatic sulfide phases may be an important
step in porphyry metallogenesis (e.g., Jenner et al., 2010). In these
models, sulfide phases subsequently break down due to changes in
oxidation state and sulfur fugacity in response to volatile exsolution
and magnetite crystallization upon emplacement in the upper crust,
thereby rendering metals available for redissolution in the volatile
Chalcopyrite phase (e.g., Cygan and Candela, 1995; Halter et al., 2002, 2005; Jugo et
al., 1999; Keith et al., 1997; Stavast et al., 2006). Other authors have
Pyrrhotite argued that this process, while it may occur, is not critical to metal
behavior in magmatic–hydrothermal systems, and that direct parti-
tioning from the silicate melt to the hydrothermal fluid phase is the
dominant 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 sulfide
from arc magmas at depth in lower crustal MASH cumulate zones may
Fig. 5. Reflected light photomicrographs of sulfide inclusions in amphibole-rich lower provide a source of metals (especially HSE) for later post-subduction
crustal arc cumulates from: (a) the Talkeetna arc, Alaska; (b) the Bonanza arc,
magmas, but that this process may not substantially affect the Cu
Vancouver Island, Canada (samples courtesy A. Greene and D. Canil, respectively).
content of the original arc magmas.
Regardless of the exact role of magmatic sulfides, the ultimate
the Sn could have been extracted from the crustally-derived granites. relationship is that of partitioning of metals between silicate melts and
Uranium is also significantly enriched in crustal rocks versus the mantle exsolving hydrothermal fluids, with sulfides as a possible intermediary
(0.91 ppm versus ~0.02 ppm, respectively; Taylor and McLennan, step. Our current understanding of these partitioning processes is now
1985), and so is unlikely to have a mantle source in such deposits. quite advanced following several decades of hydrothermal experiments
However, Dietrich et al. (1999) have suggested a possible role for (e.g., Candela and Piccoli, 1995) and more recently the advance of
mantle-derived magmas in triggering volatile (and metal) release from quantitative single fluid and melt inclusion analysis (e.g., Heinrich et al.,
evolved, felsic magmas in the Bolivian tin belt, and Walshe et al. (2011) 2003a,b), which has enabled direct measurement of metal contents in
have identified a mantle Nd isotopic signature in tin granites from ore-forming fluids and melts. In the following sections, I review some of
eastern Australia. the key factors in metal solvation and transport in magmatic–
In contrast, in the case of W skarns associated with Mo mineraliza- hydrothermal fluids.
tion in calc-alkaline I-type magmas, a shared mantle origin with Mo
might be indicated (e.g., Newberry and Swanson, 1986), consistent with 3.1. Partitioning of metals from magma into exsolving hydrothermal
the similar siderophile tendencies of these two elements, and their fluid
position in the periodic table (group VIB).
Subduction-related magmas commonly contain at least 4 wt.% H2O
3. Behavior of metals during magma fractionation and fluid during crustal ascent, as evidenced by the presence of hornblende and
exsolution in the upper crust biotite phenocrysts in many andesitic volcanic rocks and arc plutons
(Burnham, 1979; Naney, 1983; Rutherford and Devine, 1988). In
Key to the formation of magmatic–hydrothermal deposits of response to the decreasing solubility of water in silicate melts as
chalcophile and siderophile elements in the upper crust is the lack of pressure decreases, such hydrous magmas inevitably exsolve an
significant saturation with and loss of sulfide phases prior to aqueous aqueous volatile phase upon emplacement at shallow crustal levels or
volatile exsolution from a cooling magma (Candela, 1989b, 1992; on eruption (Burnham, 1979, 1997; Eichelberger, 1995). This process
Candela and Holland, 1986; Candela and Piccoli, 2005; Richards, 1995; has in the past rather confusingly been called first and second boiling
Richards and Kerrich, 1993; Spooner, 1993). As discussed in Sections (although neither process is technically “boiling”), the first event
2.1.1 and 2.5.1, chalcophile and siderophile elements partition occurring during ascent and depressurization of the magma, and the
12 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

second occurring after emplacement as crystallization progressively (a)


increases the concentration of volatiles in the residual melt (Candela, Deep
V-L isotherm (vapour)
supercritical
1989a). In reality, volatile exsolution is probably a more-or-less 2000 magmatic V-L isotherm (liquid)
1 fluid exsolution Critical / boiling curve
continuous process during arc magma ascent and cooling, starting at (10 wt.% NaCl) H2 O critical point
depth with the exsolution of relatively insoluble CO2 (Blundy et al., Halite saturation
2010; Holloway, 1976; Lowenstern, 2001; Wallace, 2005). However, Salinity isopleth
Supercritical Supercritical fluid
the bulk of the magmatic water content likely exsolves relatively 1500 fluid intersects Liquid path
rapidly as the magma approaches its solvus, at depths of ~ 5–10 km, 2-phase surface

P (bars)
Vapour path

depending on magma composition and water content (Burnham, Coexisting vapour+liquid

1979). V+L Liquid phase


progressively
The composition of this exsolved magmatic volatile phase is 1000 separates from

ite
al
dominantly aqueous, containing sulfur species (predominantly SO2 in vapour

H
L+
oxidized systems, but also some reduced S species), CO2, NaCl, KCl, HCl, Shallow
magmatic fluid
and metal chlorides. The exact composition depends on many variables, exsolution 2
including the depth of exsolution and the magma composition 500

0
ry

Cl 10
(especially the initial magmatic Cl/H2O ratio and alkali content; Candela, hy

Na 0 80
rp

ite
1989c; Candela and Piccoli, 2005; Cline and Bodnar, 1991; Webster, Po

al

t.% 6
+H
1992), but typical estimates for a single-phase supercritical fluid 0

W 40
0
90
exsolved at depths below the H2O–NaCl solvus are ~2–13 wt.% NaCl

0
80

0
70

20
0
0
60

50
equivalent (average 5 wt.% NaCl equivalent) with minor CO2 (Audétat

0
40
T (°C)

0
30

0
20

0
0
and Pettke, 2003; Audétat et al., 2008; Burnham, 1979; Candela, 1989c;

0
10
Hedenquist et al., 1998; John, 1991; Redmond et al., 2004), and up to
1.3 wt.% Cu and 0.3 wt.% Fe (Klemm et al., 2007; Rusk et al., 2004, 2008; (b)
Deep
Sawkins and Scherkenbach, 1981). These high observed metal solubil- supercritical V-L isotherm (vapour)
2000 magmatic V-L isotherm (liquid)
ities are consistent with or exceed experimental observations and fluid exsolution Critical / boiling curve
(10 wt.% NaCl)
theoretical predictions based on chloride complexing alone (e.g., H2 O critical point
Halite saturation
Candela and Holland, 1984, 1986), and suggest that other volatile
Supercritical Salinity isopleth
ligands such as sulfide species may enhance the solubility of chalcophile fluid never Supercritical fluid
elements such as Cu and Au in high temperature aqueous fluids (e.g., 1500 intersects Liquid path
V+L 3 2-phase
P (bars)

4 Vapour path
Heinrich et al., 1992; Pokrovski et al., 2005, 2008; Seo et al., 2009; Simon surface
Contraction path
et al., 2006; Zajacz et al., 2008, 2011). Supercritical Coexisting vapour+liquid

Shallowly emplaced magmas will exsolve fluids under pressure– fluid intersects
2-phase surface Liquid phase
1000 progressively
temperature (P–T) conditions that lie within the two-phase liquid–

ite
separates from

al
H
vapor field for the bulk fluid composition, resulting in immediate vapour

L+
formation of an immiscible low salinity vapor and high salinity brine
Vapour phase
(Fig. 6a). 500
departs from
2-phase surface Low to
The critical point in the H2O–NaCl system (which is commonly

0
moderate

Cl 10
salinity
used as a proxy for magmatic fluids) occurs between ~1.0 and 1.4 kb

Na 0 80
ite

liquids
al

for fluid temperatures between 600° and 800 °C (the typical al

t.% 6
+H

temperature range for fluids exsolved from intermediate to felsic


0
rm
V

W 40
e
0

ith
90

0
80

magmas) (Pitzer and Pabalan, 1986; Sourirajan and Kennedy, 1962), Ep


70

20
0
0
60

50

0
40

which is equivalent to depths of between ~ 3 and 5 km at lithostatic


30

T (°C)
0
20

0
0
10

pressures. Most porphyry deposits and their host plutons are


emplaced at depths between 1 and 6 km (Seedorff et al., 2005), so
fluids exsolving directly from magmas at these depths will typically Fig. 6. P–T–XNaCl phase diagram, modified from Driesner and Heinrich (2007), illustrating
fluid pathways for a magmatic–hydrothermal fluid exsolved with an initial bulk salinity of
form immiscible liquid and vapor plumes (e.g., Henley and McNabb,
10 wt.% NaCl. (a) Early, high thermal gradient fluids exsolved from deeply (path 1) and
1978; Nash, 1976). However, the bulk of the fluids and metals in shallowly (path 2) emplaced magmas (porphyry environment). (b) Late, low thermal
porphyry deposits are likely initially sourced at deeper levels (5– gradient fluids exsolved from deeply emplaced magma (high-sulfidation epithermal
10 km, as noted above) from larger volumes of magma in mid- to environment). Path 3 illustrates the supercritical fluid contraction path proposed by
upper crustal batholithic complexes (Candela and Piccoli, 2005; Cloos, Hedenquist et al. (1998), whereas path 4 illustrates a slightly steeper thermal gradient
with brief intersection of the 2-phase (L–V) field, as proposed by Heinrich et al. (2004).
2001; Damon, 1986; John, 1991; Richards, 2003, 2005; Shinohara and Note that in the two-phase field, the dense saline liquid phase progressively separates from
Hedenquist, 1997), at which depths the fluids will be supercritical. As the vapour phase, and the two-phase pathways shown do not represent a closed system.
these deep fluids rise into shallow cupola zones extending above the See text for discussion.
main 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). the brine phase (assuming an initial bulk salinity below ~20 wt.% NaCl),
Supercritical fluids are highly mobile (e.g., Coumou et al., 2008; Dunn and so has much greater potential as an efficient transporting medium for
and Hardee, 1981; Norton and Dutrow, 2001) and behave differently in ore components. However, until fairly recently, it was assumed that low
terms of magma–fluid partitioning compared to fluids exsolved at salinity vapors would not have the capacity to dissolve large quantities of
shallower depths in the two-phase field. In particular, Henley and base metals as chloride complexes, and the brine phase was therefore the
McNabb (1978) suggested that the higher density and viscosity of saline favored ore-forming medium (e.g., Bodnar and Beane, 1980; Cline and
brine condensates might restrict their flow, leaving them as a dense Bodnar, 1991; Eastoe, 1982; Hedenquist and Richards, 1998; Moore and
residual liquid in the deeper parts of evolving magmatic hydrothermal Nash, 1974; Nash, 1976; Shinohara, 1994; Williams et al., 1995).
systems (see also: Lewis and Lowell, 2009; White et al., 1971). The lower Observations of chalcopyrite crystals trapped in some vapor-rich fluid
density vapor or supercritical fluid would be expected to be highly inclusions (e.g., Bodnar and Beane, 1980) were explained by some as
upwardly mobile, as well as larger in terms of both volume and mass than products of heterogeneous trapping (see discussion in Mavrogenes and
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 13

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 epithermal
deposits. Also shown are the thermal structure, fluid flow pathways and characteristics during the main stage of hydrothermal activity, and overlapping hydrothermal alteration
zones. 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 et
al. (1998), and Fournier (1999).

Bodnar, 1994). The advent of quantitative analysis of single fluid may explain precipitation of the bulk of Cu, Mo, and some Au over a
inclusions by synchrotron X-ray microprobe, proton-induced X-ray relatively narrow temperature interval between 425°–320 °C (Hemley et
emission spectroscopy (PIXE), and laser ablation-inductively coupled al., 1992; Klemm et al., 2007; Landtwing et al., 2005). At shallower depths
plasma-mass spectrometry (LA-ICP-MS) revealed that, contrary to these and lower temperatures, the vapor phase may become too dilute to
earlier assumptions, considerable amounts of metal, including Cu and Au, transport significant amounts of base metals as chloride complexes, but
were indeed consistently present in some vapor-rich inclusions (Audétat may continue to carry some metals such as Au, Cu, As, and Sb as sulfide
et al., 2000; Harris et al., 2003; Heinrich et al., 1992, 1999; Klemm et al., complexes (Deditius et al., 2009; Simon et al., 2006, 2007), eventually
2007; Lowenstern et al., 1991; Simon et al., 2005, 2006; Ulrich et al., either venting them to the surface in high-temperature fumaroles (e.g.,
2001), and subsequent work has shown that much of this metal content Chaplygin et al., 2007; Hedenquist et al., 1994a; Symonds et al., 1987;
is transported as sulfide complexes rather than as (or in addition to) Tarana et al., 1995; Tessalina et al., 2008) or precipitating them in the
chloride complexes (Cauzid et al., 2007; Heinrich et al., 1992, 1999; near-surface high-sulfidation epithermal environment (see Section 4.2.1;
Pokrovski et al., 2005, 2008; Seo et al., 2009; Simon et al., 2006; Zajacz and Deditius et al., 2009; Hedenquist et al., 1993, 1994b; Heinrich et al., 1999,
Halter, 2009; Zajacz et al., 2010). 2004; Larocque et al., 2008; Murakami et al., 2010; Pudack et al., 2009;
Early resistance to these ideas was, I believe, at least in part due to Williams-Jones and Heinrich, 2005).
confusions of terminology: most papers dealing with this subject have
referred to metal solubility and transport in a vapor phase, suggesting to 4. Magmatic–hydrothermal ore formation
the unwary a low density, low salinity gas, whereas for the most part the
fluids in question were either single-phase supercritical fluids, or The focus of this paper is on the flux of metals in subduction-
relatively high density vapors just below their critical points. Under such related magmatic systems, but this would be of little practical interest
high P–T conditions, vapors are almost as saline as the initial single- if that flux did not ultimately lead to ore formation. Thus far, we have
phase fluid from which they evolved, and therefore they still contain focused on the importance of firstly not losing significant amounts of
plenty of chloride for base metal complexing. But perhaps more metal to a fractionating or residual sulfide phase, and then efficiently
importantly, as noted above, these vapors will also contain a high partitioning those metals into a highly mobile aqueous fluid phase.
concentration of volatile sulfur species, which now appear to be What subsequently happens to that fluid phase dictates whether
essential for the efficient solvation of chalcophile elements under high economic concentrations of metals are precipitated (grade), whereas
P–T conditions. When combined with the higher mass proportion of the the scale of the magmatic and derivative hydrothermal system
vapor phase (versus brine condensate) and its high mobility, a model for controls the total amount of metals precipitated (tonnage).
transport of the bulk of the metal flux in porphyries by relatively dense
vapors or supercritical fluids is now well established (Klemm et al., 4.1. Porphyry Cu ore formation
2007; Landtwing et al., 2010; Williams-Jones and Heinrich, 2005).
The rapid reduction in the efficiency of transport of metals as the In a landmark paper, Cline and Bodnar (1991) presented a model
vapor plume rises, cools, and becomes less dense by brine condensation, for the evolution of magmatic-hydrothermal systems from initial
14 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

aqueous fluid exsolution to cooling and mineral precipitation. A key efficiency in porphyry systems is likely achieved where volatile
finding of this work was that, although there are many variables that saturation occurs in large (≥ 100 km 3) mid- to upper crustal magma
can affect the specific evolutionary path of a given system, an chambers at depths ≥6 km, containing moderately hydrous (N4 wt.%
economic porphyry Cu deposit can potentially be formed from quite H2O) and Cl-rich magmas. These fluids either rise as bubbly magma or
small volumes of typical andesitic arc magma. For example, the as a separate volatile plume into the apical parts of the system
authors showed that 15–90 km 3 of andesitic magma initially contain- where decreasing pressure and temperature cause deposition of Cu
ing 50 ppm Cu, 2.5 wt.% H2O, and with a Cl/H2O ratio = 1, is sufficient and Mo ± Au (Candela, 1989b; Shinohara et al., 1995).
to generate a 250 Mt deposit with a grade of 0.75 wt.% Cu. The range of Focusing of magma ascent and fluid flow into narrow apical
required magma volumes reflects variables such as the depth of regions, or cupolas, is likely to be a function of structure in the brittle
magma emplacement and the compatibility of Cu with fractionating rocks overlying the batholithic system (Tosdal and Richards, 2001,
mineral phases (larger volumes are required if Cu is compatible with and references therein). Shallow crustal magma emplacement will
early fractionating silicate, oxide, or sulfide minerals). Details of the cause extensional doming in the cover rocks, with dilational fault
model, later updated in Cline (1995), show that magmas emplaced at zones providing high-permeability pathways for fluid and magma
moderate depths (1.0 to 2.0 kb, equivalent to depths of ~ 4–8 km) ascent (Burnham, 1979). Evidence from the dike emplacement
most efficiently partition Cu into early saturating saline fluids, literature suggests that such fractures may first be opened and
whereas Cu and Cl are only released from shallowly emplaced propagated by volatile pressure, and only later filled by more viscous
(0.5 kb, or ~2 km) magmas during late stages of crystallization (see magma (Burnham, 1979; Carrigan et al., 1992; Rubin, 1995). This
also Candela, 1989b). (Note that these depths reflect the locus of fluid raises the intriguing possibility that the cylindrical shapes of many
and metal exsolution from the magma under lithostatic pressure porphyry stocks may have arisen first as breccia pipes or diatremes
conditions, as opposed to the depth of subsequent hydrothermal bored out by rapidly escaping volatiles, only later to be back-filled
metal deposition, which will be at shallower levels and likely under with porphyritic magma (Fig. 8; Norton and Cathles, 1973; see also
hydrostatic pressure conditions; Fournier, 1999.) In large measure, Fig. 2 in Anderson et al., 2009, and Fig. 8 in Sillitoe, 2010, and Fig. 17 in
these differences reflect the change in properties of the magmatic– Vry et al., 2010).
hydrothermal fluid, which will separate initially as a moderately To some extent, focusing of magma and fluid flow along narrow
saline supercritical fluid in deeper systems (Fig. 6a, path 1), but as conduits may be self-organizational, because once initial channel
immiscible vapor and brine at shallower levels (Fig. 6a, path 2). Cline ways have developed, they will represent high-permeability path-
(1995) concluded that optimum conditions for porphyry Cu ore ways and are likely to thermally weaken the wall rocks and promote
formation are obtained where magma containing ~ 4 wt.% H2O and further fracturing and channeling.
with a high initial Cl/H2O (typical of most arc magmas) is emplaced at Narrow focusing of apical fracturing and subsequent fluid flow are
moderate crustal depths, thereby maximizing the efficiency of Cu likely to be critical to the formation of high grade porphyry deposits
partitioning into an early saline fluid phase. (e.g., El Teniente, Chile; Vry et al., 2010), while multiple breccia/
The results of these studies indicate that no special conditions or intrusive events can potentially increase tonnage (provided that later
magmatic metal enrichment are required to form even large porphyry events do not destroy earlier mineralization).
Cu deposits. Instead, the question is rather one of process efficiency: Given suitable fluid flow focusing, three further factors combine to
where (what depth) and how are metalliferous fluids released and cause maximum efficiency of metal deposition within relatively small
channeled? This finding is of fundamental importance from an volumes in porphyry Cu ± Mo ± Au deposits. All three effects are
exploration perspective, because it shifts the focus from seeking related to the steep temperature gradient in the cupola zone (Fig. 7),
anomalously metal-rich magmas (which have proven elusive; Jenner and they therefore control the vertical range of ore deposition. On the
et al., 2010) to searching for optimal ore depositional settings in other hand, fluid focusing controls the lateral extent of mineralization.
otherwise normal tectonomagmatic environments (e.g., Tosdal and In combination, the highest grades will occur where ore deposition is
Richards, 2001). both focused laterally and restricted vertically.
Cloos (2001), Shinohara et al. (1995), and Shinohara and Hedenquist The first factor is that Cu solubility (as chloride species) decreases
(1997) considered the question of process efficiency from the dramatically as fluids cool through the temperature interval ~ 400° to
perspective of physical separation and focusing of volatile release 300 °C (Crerar and Barnes, 1976; Hemley et al., 1992; Klemm et al.,
from the magma. Cloos (2001) suggested that the classic cupola shape 2007; Landtwing et al., 2005; Xiao et al., 1998). Given the very sharp
(Norton, 1982) of porphyry systems above larger batholithic magma temperature gradient implied by near-surface emplacement of
chambers reflects convective circulation of bubbly magma into the magma (Fig. 7), this temperature interval will correspond to a narrow
shallow apical parts of these systems (at 1–3 km depth), where volatiles depth range, likely with 1 or 2 km of the surface (although it may
physically separate from the melt and coalesce to form a discrete fluid- extend to greater depths with time as the magmatic–hydrothermal
filled cupola. The now-dense, volatile-depleted magma forms a system begins to cool). This depth range is typical for ore formation in
downward return flow to complete the convective cycle. In contrast, many porphyry systems.
Shinohara and Hedenquist (1997) envisaged fluids physically separat- The second factor is that SO2 dissolved in the magmatic–
ing from the magma at greater depth within the underlying magma hydrothermal fluid phase progressively disproportionates to H2S
chamber, and rising as a discrete plume up apical channelways formed and H2SO4 as the fluid cools below ~ 400 °C (Holland, 1965; Kusakabe
by fractures and dikes in the brittle carapace (Fig. 7). et al., 2000; Reeves et al., 2010; Sakai and Matsubaya, 1977):
A key aspect of both of these models is that volatile separation, and
Cu partitioning, occurs from a much larger volume of magma 4SO2 + 4H2 O ⇔ H2 S + 3H2 SO4 : ð1Þ
(emplaced at deeper levels) than that preserved and commonly
visible within the shallow-level ore body. In Cloos's (2001) model, This reaction generates both hydrogen sulfide, which initiates
volatile-rich, bubbly magma rising from the underlying batholith abundant precipitation of sulfide minerals (i.e., chalcopyrite, pyrite,
convects through the cupola zone where it releases its fluids, whereas molybdenite), and also sulfuric acid, which causes early deposition of
in Shinohara and Hedenquist's (1997) model, vesiculation and large volumes of anhydrite in the potassic alteration zone, and
convective circulation occur in the underlying magma chamber itself, progressively increasing degrees of hydrolytic alteration (an initial
and fluids are released as a plume into the base of the apical dike shift from feldspar-stable potassic alteration, to muscovite/sericite-
system. Combining these models with (Cline, 1995; Cline and stable phyllic alteration). Consequently, the bulk of Cu-sulfide miner-
Bodnar's, 1991) calculations suggests that maximum ore-forming alization occurs at the low-temperature, late-stage end of the potassic
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 15

(a) (b)

Mixed magmatic-
hydrothermal
breccia

Porphyry
intrusion
Porphyry
intruding
comagmatic
Cuspate breccia
intrusive
border Porphyry
dike

Hydrothermal
quartz filling
cavity
~20 cm

(c)

Limestone
wallrock
clast Cuspate
porphyry Mixed magmatic-
clast hydrothermal
breccia

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. The
breccia 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 the
magma 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 the
role of fluids in breccia formation. Scale in (c) is in centimeters.

alteration phase, just prior to the onset of phyllic alteration (i.e., large pressure drop), serve to narrowly focus Cu-sulfide mineralization
corresponding to the “ore shell” of the classic Lowell and Guilbert, 1970, both laterally and vertically within cupola zones above large mid- to
porphyry Cu model). upper crustal batholithic complexes. The most likely reasons for
The third factor is the increased permeability over this tempera- otherwise prospective porphyry systems to be unproductive will be
ture range caused by a combination of the transition in silicate rocks either a failure to focus fluid flow, or simply insufficient fluid supply
from ductile to brittle behavior at temperatures (between ~400°– (likely due to an insufficiently large underlying magmatic system).
350 °C; i.e., the brittle–ductile transition; Fig. 7; Cathles, 1991;
Fournier, 1999; Landtwing et al., 2005), and a window of retrograde 4.2. Epithermal Cu–Au ore formation
silica solubility (between ~550–350 °C; Fournier, 1985). Not only do
these processes result in the formation of open-standing brittle veins 4.2.1. High-sulfidation epithermal Cu–Au deposits
and porosity, thereby facilitating rapid upward fluid flow and As noted in Section 3.1, although the bulk of Cu (and Mo) appears to
wallrock permeation (e.g., the crackle breccias, stockworks, and be precipitated over the temperature interval 425°–320 °C, some Cu and
disseminated mineralization textures so characteristic of porphyry Cu other metals such as Au, Sb, and As may remain in solution as sulfide
deposits), but this transition also represents the boundary between complexes, to be carried into the shallow epithermal regime. Observa-
lithostatic and hydrostatic fluid pressures (a pressure differential of tions from rare fluid inclusions in high-sulfidation epithermal Cu–Au
~ 3×). Sudden depressurization of fluids across this boundary can be deposits suggest that the ore-forming fluid was a low- to moderate-
expected to have major effects on fluid properties, including phase salinity liquid (0.2 to 4.5 wt.% NaCl equivalent; Hedenquist et al., 1994b;
separation (Fig. 6) and consequent changes in metal solubility (e.g., Mancano and Campbell, 1995), which paragenetically post-dates
Landtwing et al., 2005, 2010; Murakami et al., 2010). advanced argillic alteration formed by highly acidic magmatic gasses
In combination, these four factors, (1) spatial focusing of fluid flow in (Arribas, 1995; Hedenquist et al., 1994b; Stoffregen, 1987). This
narrow cupolas, (2) reduction of metal solubility, (3) increased apparent inconsistency has been explained by Hedenquist et al.
dissolved sulfide activity, and (4) permeability increase due to transition (1998), Heinrich et al. (2004), and Heinrich (2005) in terms of
from ductile to brittle fracturing and retrograde silica solubility (with a contraction of a moderate salinity supercritical magmatic fluid or
16 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

vapor by rapid cooling during ascent such that it does not touch, or A more detailed analysis of the phase diagram shown in Fig. 6
barely touches, the two-phase solvus (Figs. 6b and 9, paths 3 and 4, reveals that a typical magmatic fluid of 2–13 wt.% NaCl would have to
respectively). Because of the curvature of the solvus crest (critical curve) first intersect the solvus at temperatures above ~ 400°–500 °C for this
to lower salinities at low temperatures and pressures, this moderately model to work optimally, otherwise it would fall on the liquid side of
saline fluid will lie on the liquid side of the solvus at shallow depths, the two-phase field, and would become more saline by boiling off a
although it may have contracted from what was originally a vapor or dilute vapor (Figs. 9 and 10). Assuming that this did indeed happen,
supercritical fluid phase at higher temperatures and depths. then the smaller (by mass) vapor component would have to leave the
Heinrich et al. (2004) specifically suggested that brief intersection two-phase surface again at pressure before cooling below ~400 °C,
with the solvus at high pressures and temperatures (Figs. 6b and 9, path otherwise its salinity rapidly falls to sub-weight percent levels at
4) might be an important way to shed chloride-complexed components lower temperatures and lower pressures (Figs. 6 and 10), which are
such as Fe from the vapor in a brine condensate, leaving bisulfide ligands inconsistent with fluid inclusion evidence for low to moderately saline
free to bond with Cu and Au in the residual vapor and transport it to liquids in high-sulfidation epithermal ore formation (Hedenquist et
shallower (epithermal) levels. This model requires that the vapor leaves al., 1994b; Mancano and Campbell, 1995). Thus, the P–T trajectory of a
the two-phase surface again by cooling at pressure, thereby passing over fluid that satisfies Heinrich et al.'s (2004) model for high-sulfidation
the crest of the solvus (the critical curve; (Figs. 6b and 9, path 4); upon epithermal Cu–Au mineralization (Figs. 6b and 9, path 4) is somewhat
subsequent ascent and depressurization, this fluid will be a liquid, as unique, and may occur only rarely or fleetingly during the waning
described above. Heinrich et al. (2004) argued that the Fe-rich brine stages of a cooling magmatic–hydrothermal system.
condensation step is essential for retention of Au (and Cu) in the vapor A more normal, or early ascent pathway for a magmatic–
phase, because otherwise Fe will tend to precipitate as Cu–Fe-sulfide hydrothermal fluid would be for it to rise more-or-less is enthalpically
minerals and strip the fluid of bisulfide ligands, thus causing Au to co- or quasi-adiabatically along a steep P–T gradient (e.g., Hemley and
precipitate at depth (a possible mechanism for the formation of Hunt, 1992; Henley and Hughes, 2000; Wood and Spera, 1984), and
porphyry Cu–Au deposits). However, this condensation step must therefore to dive deeply into the two-phase field and separate into an
then be followed by cooling while still at depth, in order to “lift” the increasingly dilute vapor phase and a saline brine (Figs. 6 and 9, path
vapor phase off the two-phase surface, and allow it to contract to a liquid 1), or even to boil dry to halite plus vapor (Figs. 6 and 9, path 2). Such
(Figs. 6b and 9, path 4). fluid pathways might be consistent with the widespread and intense

0 100 200 300 400 500 600 700 800 900 1000
Wet granite solidus

Wet granodiorite solid

V-L solvus (vapour)


8 (lith)
V-L solvus (liquid)
2000

.%
Deep
magmatic wt
7 (lith)
20

fluid
exsolution
t.%
us

w
(hydrostatic)

10
(lithostatic)
Ductile
Brittle

6 (lith)
1500 15 (hyd)
.%
5 wt 14 (hyd)
5 (lith)
13 (hyd)
Deep
Depth (km)

single- 12 (hyd)
P (bars)

phase
fluid
11 (hyd)
4 (lith)
1000 3 10 (hyd)

4 9 (hyd)
1 1 wt.%
km 8 (hyd) 3 (lith)
°C/
Early 300otherm
erm

porphyry Shallow
magmatic g e 7 (hyd)
oth

veins
(2-phase fluid
ge

fluid) exsolution 6 (hyd)


/km

2 (lith)
500
°C

2 5 (hyd)
30

4 (hyd)
3 (hyd)
Main-stage 1 (lith)
brittle
2 (hyd)
V/

porphyry
L+

veins V+Halite
Ha

1 (hyd)
li

Epithermal
te

0 0
0 100 200 300 400 500 600 700 800 900 1000
T (°C)

Fig. 9. Pressure (depth)–temperature section through the H2O–NaCl phase diagram, with vapour–liquid (V–L) solvi drawn for 1, 5, 10, and 20 wt.% NaCl (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; the transition point corresponds to the critical point for that
composition. Also shown are the wet granite and granodiorite solidi (Burnham, 1979), and average crustal (30 °C/km) and high (300 °C/km, near active volcanism) geothermal
gradients (Barbier, 2002; Goff et al., 1992; Noorollahi et al., 2007). Depths are indicated for hydrostatic (hyd) and lithostatic (lith) pressure conditions. Typical temperature–depth
ranges for supercritical magmatic fluid exsolution, early high-temperature porphyry veins, later main-stage brittle porphyry veins, and epithermal mineralization are indicated. Four
fluid P–T paths are shown corresponding to: (1) a typical porphyry-forming fluid path; (2) a shallow high-temperature path boiling to dryness (V + Halite field); (3) the deep
contraction path of Hedenquist et al. (1998); and (4) the contraction with minor brine condensation path of Heinrich et al. (2004).
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 17

-8 -7 -6 -5 -4 -3 -2 -1 0 10 20 30 40 50 60 70 80 90 100
1300 Single
1300
V-L solvus (liquid) phase 700°C
V-L solvus (vapour) mag-

rve
1200 matic 1200
V-H solvus (vapour) fluid

Critical cu
1100 V+L
1100
(700°C)

1000 1000

600°C
900 900

800 V 800
V+L
(600°C)
700
P (bars)

700

600 500°C 600

L+H (600°C)
L (600°C)
500 500
V+L
(500°C)

L+H (500°C)
L (500°C)

L+H (700°C)
400 400

L (700°C)
V+H (600°C)

V+H (500°C)

L+H (400°C)
300 400°C 300

L (400°C)
V+H (700°C)
Critical P V+L
200 of H2O (400°C) 200
V+H (400°C)

100 100

0 0
-8 -7 -6 -5 -4 -3 -2 -1 0/1 10 20 30 40 50 60 70 80 90 100
log (wt.% NaCl) wt.% NaCl
scale
change

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 scale
change 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 condense
a 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.

acidic (advanced argillic) alteration commonly found at shallow levels et al., 1998; Müller and Groves, 1993; Mutschler et al., 1985; Richards,
above porphyry systems, which is caused by acidic gasses (H2SO4, 1995; Thompson et al., 1985).
HCl) condensing from a low-density vapor plume. As Heinrich et al. Fluids in these alkalic-type deposits are commonly low- to
(2004) noted, such acidic vapors do not appear to transport Au (or Cu) moderate salinity (0–10 wt.% NaCl equiv.), low temperature (typically
effectively because bisulfide (HS –) ligands are hydrolized to H2S. This ≤250 °C) liquids, with evidence for decompressional boiling or fluid
may explain why advanced argillic alteration caps are commonly mixing as the prime ore depositional mechanism in high grade
barren (in terms of Au and Cu) and merely generate permeability that breccias and veins (Jensen and Barton, 2000; Richards, 1995). Stable
potentially focuses later ore-forming fluid flow. Thus, mineralization isotopic compositions of these fluids are generally ambiguous, and
may only occur where later moderate salinity liquids have followed a permit interpretations of the involvement of either isotopically
higher pressure, rather specialized cooling path, as described above exchanged meteoric waters or magmatic fluids, or both (Ahmad et
(Heinrich et al., 2004). al., 1987; Carman, 2003; Richards, 1995; Richards and Kerrich, 1993;
Ronacher et al., 2004; Scherbarth and Spry, 2006; Zhang and Spry,
4.2.2. Low-sulfidation epithermal Au deposits (including alkalic-type 1994). Similar stable isotopic data from other low-sulfidation deposits
deposits) are commonly interpreted to reflect a meteoric fluid source because of
Although low-sulfidation epithermal Au deposits are commonly the absence of clearly coeval magmatism (as noted above). However,
found in volcanic terrains, their link to magmatism is more tenuous the close link with magmatism in alkalic-type systems suggests that a
than high-sulfidation deposits, and there is commonly evidence for a magmatic fluid source is more likely (e.g., Carman, 2003; Scherbarth
greater involvement of meteoric groundwater in their formation than and Spry, 2006; Simmons and Brown, 2007), and the vapor
magmatic fluids (e.g., Faure et al., 2002; Field and Fifarek, 1985; Heald contraction model described in Section 4.2.1 could explain the
et al., 1987). Nevertheless, alkalic-type epithermal Au deposits, which observed characteristics of these ore fluids (i.e., direct contraction to
are mineralogically similar to low-sulfidation deposits (adularia and a moderate salinity liquid from a high temperature magmatic fluid at
sericite — or roscoelite [vanadium mica] — are stable), do show a depth). Because of the association of these deposits with relatively
strong temporal and genetic relationship to alkalic magmas, typically small intrusive complexes, and therefore a smaller crustal thermal
in relatively small and isolated intrusive complexes located in back- anomaly, a shallower fluid P–T path with cooling at depth is more
arc or post-subduction settings (e.g., Jensen and Barton, 2000; Kelley likely, consistent with a model of vapor contraction.
18 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

5. Summary and conclusions porphyry Cu–Mo deposits can also occur). Such magmatic–metallogenic
systems are thought to form by remelting of previously subduction-
5.1. Sources of magmas and metals modified upper plate lithosphere, and in particular the lower crustal
amphibolitic cumulate roots of former arc magmatic complexes.
Magmatic–hydrothermal porphyry Cu ±Mo±Au, Au, Mo, and Sn–W Remelting can be triggered by crustal thickening and thermal rebound
deposits (and related epithermal Au deposits), derive their metals from following arc or continent collision, delamination of sub-continental
their associated magmas. With the exception of porphyry Sn–W deposits mantle lithosphere causing direct exposure of the lower crust to
that are associated with crustally derived S-type granites, most asthenospheric temperatures and melts, and asthenospheric upwelling
other deposits in this grouping are formed by calc-alkaline to mildly during rifting of former arc crust. Sparse sulfide phases in these arc
alkaline I-type granitoids directly or indirectly related to subduction. cumulates, residual from fractionation of previous arc magmas, will
Sources of these magmas include subduction-metasomatized astheno- likely be rich in chalcophile and highly siderophile elements. During
spheric mantle wedge, basaltic oceanic crust and/or seafloor sediments, low-volume melting under relatively low fS2 conditions (in the absence
and, in post-subduction settings, subduction-modified upper plate of a flux of S from active subduction), these sulfide phases will likely
lithosphere. The majority of normal arc porphyry systems are generated redissolve in the mildly alkaline partial melt, and may provide a source
from hydrous mantle wedge melts that have interacted to varying for Au-rich (±PGE) post-subduction porphyry Cu–Au and epithermal
degrees with the upper plate lithosphere during passage towards the Au systems: examples include the Roşia Montană, Skouries, Kisladag,
surface. Assimilation and fractional crystallization processes fundamen- Çöpler, and Sari Gunay porphyry Cu–Au and epithermal Au deposits in
tally change the composition of the primary basaltic arc magmas to the Neo-Tethyan belt of Romania, Greece, Turkey, and Iran, and the
intermediate calc-alkaline compositions, with fractionated trace element Grasberg, Ok Tedi, Porgera, Lihir, and Emperor porphyry Cu–Au and
patterns and evolved (crustal) isotopic signatures. epithermal Au deposits in the southwest Pacific. However, gold
Seafloor sediments and basaltic oceanic crust only melt under enrichments may not occur where more abundant sulfides were
unusually hot subduction zone conditions or locally at plate edges, and present in the former arc complex, leading to more “normal” porphyry
despite widespread claims of the identification of slab melts (adakites) Cu± Mo systems: examples include the Kerman porphyry Cu belt of
in the literature based on high Sr/Y and La/Yb ratios in some evolved central Iran, and the Gangdese porphyry Cu belt of Tibet.
granitoids, this process is unlikely to be a major contributor to arc Porphyry Mo and Sn–W deposits associated with felsic magmas in
magmatism and metallogeny. Rather, these subtle trace element ratios continental interiors are thought to form mainly by partial melting of
can be readily explained by fractionation of hornblende ± titanite and continental crust during rifting to form S-type, lithophile element-rich
residual garnet, and suppression of plagioclase fractionation from granitic magmas. A role for mafic, mantle-derived magmas is suggested
water-rich mantle wedge basaltic magmas. These trace element by the common association with such rocks, but their role may be
characteristics are therefore an indicator of high magmatic water predominantly as a heat source for crustal melting and a trigger for
content rather than being a source signature, and this likely explains the volatile saturation and eruption, rather than as a unique source of
common association of high-Sr/Y (i.e., hydrous) magmas with mag- metals.
matic–hydrothermal ore deposits — without magmatic water, such
deposits cannot form.
Potential sources of metals in arc magmas include the oceanic crust 5.2. Porphyry and epithermal ore formation
and sediments (via dehydration fluids or melting), the mantle wedge,
and the upper plate crust. Fluid–mobile elements such as K, Rb, Cs, Ca, Sr, In the porphyry and epithermal ore depositional environment, a
Ba, U, B, Pb, As, Sb, Tl, and possibly Cu, Au, Re, and the Pd-group critical role is played by aqueous fluids exsolving from hydrous
elements, along with large amounts of H2O, Cl, and S, are fluxed from the magmas emplaced in the mid- to upper crust. The P–T–X properties of
dehydrating subducting slab into the mantle wedge, causing metaso- magmatic hydrothermal fluids, approximated by the H2O–NaCl
matism and partial melting by lowering the peridotite solidus. Although system, combined with volatile solubility in intermediate composition
there is some evidence for slab-derived fluid contributions of chalco- magmas, suggest that fluid saturation occurs at depths of 5–10 km in
phile and highly siderophile elements (Cu, Au, PGE) to the mantle the batholithic roots of arc magmatic systems. Volatile exsolution may
wedge, it is not clear that this is a necessary metallogenic step, the upper lead to the upward propagation of buoyant bubbly magma as dikes
mantle already containing significant amounts of these elements. Likely, and stocks intruded into the overlying shallow crust (with or without
a more important control on the metal content of subsequent partial subsequent eruption at surface), and/or the rapid ascent of a separate
melts is the abundance and stability of residual sulfide phases in the volatile plume. Structural focusing and chanelling of these evolved
asthenospheric mantle source. Under the high fO2 and fS2 conditions of magmas and fluids creates a cupola zone characterized by high
arc magmatism, sulfur will be dominantly present as sulfate and sulfate, thermal gradients and fluid flux. Metals (Cu, Mo, Au), which partition
but saturation in small amounts of sulfide phases is also likely. These strongly into the saline (2–13 wt.% NaCl equivalent) and S-rich
sulfide phases will tend to deplete the magma in highly siderophile magmatic hydrothermal phase at high P and T in the underlying
elements (Au and PGE), but will not be present in sufficient volume to batholithic magma chamber, experience rapid reduction in solubility
significantly deplete the magma in more abundant chalcophile elements as these fluids ascend, depressurize, and cool, with the bulk of Cu and
such as Cu and Mo. Such magmas therefore have the potential Mo being precipitated over a temperature range of 425°–320 °C at 1–
subsequently to form porphyry Cu–Mo deposits. Thus, Cu (and perhaps 6 km (commonly ≤2 km) depth.
Mo) are thought to be predominantly derived from the mantle, plus or This depth–temperature interval is critical because it also repre-
minus contributions from the subducting slab. Gold and silver present in sents: (1) the upward transition from ductile to brittle behavior in the
minor amounts in such deposits may also be derived from subduction cover rocks (~400°–350 °C), which facilitates fracturing and rapid fluid
sources, although there is some evidence for additional contributions depressurization; (2) a window of retrograde silica solubility (~550°–
from the upper plate crust (especially for Ag, and also perhaps Mo). 350 °C), which enhances permeability and porosity for ore deposition;
Arc-like magmas and related porphyry and epithermal ore deposits and (3) the temperature range (b400 °C) over which SO2 in the fluid
also occur in post-subduction tectonic settings, such as subduction phase begins to disproportionate to H2S and H2SO4, which causes
reversal or migration, arc collision, continent–continent collision, and precipitation of sulfide minerals (chalcopyrite, pyrite, molybdenite, rare
post-collisional rifting. They are distinguished from normal subduction- bornite). This reaction also generates increasingly acidic fluids, leading
related suites by slightly higher magmatic alkali (K2O and Na2O) to the characteristic progression from feldspar-stable alteration assem-
contents, and by the occurrence of Au-rich deposits (although normal blages (potassic), through muscovite/sericite-stable assemblages
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 19

(phyllic), to clay- (argillic) and alunite-stable assemblages (advanced Alonso-Perez, R., Müntener, O., Ulmer, P., 2009. Igneous garnet and amphibole
fractionation in the roots of island arcs: experimental constraints on andesitic
argillic). liquids. Contributions to Mineralogy and Petrology 157, 541–558.
Depending on the depth of exsolution, the initial magmatic fluid will Alt, J.C., Shanks, W.C., Jackson, M.C., 1993. Cycling of sulfur in subduction zones: the
either be a supercritical fluid (below ~6 km depth) or will exist as a two- geochemistry of sulfur in the Mariana Island Arc and back-arc trough. Earth and
Planetary Science Letters 119, 477–494.
phase moderate salinity vapor and high density brine. As the plume Anderson, E.D., Atkinson Jr., W.W., Marsh, T., Iriondo, A., 2009. Geology and
ascends, it will intersect its solvus (in the case of a deeply exsolved geochemistry of the Mammoth breccia pipe, Copper Creek mining district,
supercritical fluid), and the vapor phase will become progressively less southeastern Arizona: evidence for a magmatic–hydrothermal origin. Mineralium
Deposita 44, 151–170.
saline through brine condensation. At the level of porphyry ore Annen, C., Blundy, J.D., Sparks, R.S.J., 2006. The genesis of intermediate and silicic
formation, both the brine and vapor phase may contribute to metal magmas in deep crustal hot zones. Journal of Petrology 47, 505–539.
transport and deposition, although there is increasing evidence for the 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
importance of the vapor phase as a large-volume, highly upwardly
relation to magmatic fluid. In: Thompson, J.F.H. (Ed.), Magmas, fluids, and ore
mobile transportation medium. However, as this vapor phase continues deposits: Mineralogical Association of Canada Short Course, 23, pp. 419–454.
to ascend, it will rapidly decrease in salinity, such that chloride- Audétat, A., 2010. Source and evolution of molybdenum in the porphyry Mo(–Nb)
complexed metals are unlikely to be transported by such fluids to deposit at Cave Peak, Texas. Journal of Petrology 51, 1739–1760.
Audétat, A., Keppler, H., 2005. Solubility of rutile in subduction zone fluids, as
shallow epithermal levels. It will also increase in acidity, thereby determined by experiments in the hydrothermal diamond anvil cell. Earth and
reducing the solubility of bisulfide-complexed metals such as Au (and Planetary Science Letters 232, 393–402.
perhaps also Cu) by protonation of HS– to H2S. This acidic vapor phase is Audétat, A., Pettke, T., 2003. The magmatic-hydrothermal evolution of two barren
granites: a melt and fluid inclusion study of the Rito del Medio and Canada Pinabete
responsible for the extreme acid leaching in advanced argillic lithocaps plutons in northern New Mexico (USA). Geochimica et Cosmochimica Acta 67,
above porphyry systems. 97–121.
High-sulfidation epithermal Cu–Au deposits are hosted by these Audétat, A., Pettke, T., 2006. Evolution of a porphyry-Cu mineralized magma system at
Santa Rita, New Mexico (USA). Journal of Petrology 47, 2021–2046.
highly permeable advanced argillic alteration zones, but appear to have Audétat, A., Günther, D., Heinrich, C.A., 2000. Magmatic–hydrothermal evolution in a
been formed by later, moderately saline (0.2 to 4.5 wt.% NaCl fractionating granite: a microchemical study of the Sn–W–F-mineralized Mole
equivalent), less acidic liquids. Two models have been proposed to Granite (Australia). Geochimica et Cosmochimica Acta 64, 3373–3393.
Audétat, A., Pettke, T., Heinrich, C.A., Bodnar, R.J., 2008. The composition of magmatic-
explain the origin of these paragenetically late mineralizing fluids in hydrothermal fluids in barren and mineralized intrusions. Economic Geology 103,
terms of contraction of a single phase (supercritical) magmatic fluid, 877–908.
with (Heinrich et al., 2004) or without (Hedenquist et al., 1998) brief Ballhaus, C., 1993. Redox states of lithospheric and asthenospheric upper mantle.
Contributions to Mineralogy and Petrology 114, 331–348.
intersection of the solvus. Because of the topology of the P–T–XNaCl
Barbarin, B., 1996. Genesis of the two main types of peraluminous granitoids. Geology
phase diagram, such fluids contract to a liquid phase upon cooling at 24 (4), 295–298.
pressure. Thus, these authors propose that high-sulfidation epithermal Barbier, E., 2002. Geothermal energy technology and current status: an overview.
Cu–Au deposits may be formed directly from late-stage magmatic Renewable and Sustainable Energy Reviews 6, 3–65.
Barnes, S.-J., Naldrett, A.J., Gorton, M.P., 1985. The origin of the fractionation of
hydrothermal fluids. platinum-group elements in terrestrial magmas. Chemical Geology 53, 303–323.
Although low sulfidation epithermal Au deposits have not been Barreiro, B.A., 1984. Lead isotopes and Andean magmagenesis. In: Harmon, R.S.,
discussed in detail in this paper, because most such deposits are not Barreiro, B.A. (Eds.), Andean Magmatism Chemical and Isotopic Constraints.
Cheshire, Shiva, Nantwich, pp. 21–30.
directly related to porphyry-type magmatic–hydrothermal systems, Beard, J.S., Lofgren, G.E., 1991. Dehydration melting and water-saturated melting of
the vapor contraction mechanism might have applicability to alkalic- basaltic and andesitic greenstones and amphibolites at 1, 3, and 6.9 kb. Journal of
type epithermal gold deposits, which do show a close genetic Petrology 32, 365–401.
Bergantz, G.W., Dawes, R., 1994. Aspects of magma generation and ascent in
relationship to post-subduction alkalic magmas. continental lithosphere. In: Ryan, M.P. (Ed.), Magmatic Systems. Academic
Press, San Diego, pp. 291–317.
Berger, J., Caby, R., Liégeois, J.-P., Mercier, J.-C.C., Demaiffe, D., 2009. Dehydration,
Acknowledgments melting and related garnet growth in the deep root of the Amalaoulaou
Neoproterozoic magmatic arc (Gourma, NE Mali). Geological Magazine 146,
173–186.
I would like to thank Nigel Cook and Timothy Horscroft for inviting Blatter, D.L., Carmichael, I.S.E., 1998. Hornblende peridotite xenoliths from central
me to submit this paper. Review papers, by their nature, draw heavily on Mexico reveal the highly oxidized nature of subarc upper mantle. Geology 26,
the work of others, and I would particularly like to acknowledge the 1035–1038.
Blevin, P.L., Chappell, B.W., 1992. The role of magma sources, oxidation states and
following people who have influenced my thinking on this subject: P.
fractionation in determining the granite metallogeny of eastern Australia. Trans-
Candela, J. Cline, J. Dilles, J. Hedenquist, C. Heinrich, R. Sillitoe, and R. actions of the Royal Society of Edinburgh, Earth Sciences 83, 305–316.
Tosdal. An anonymous reviewer is thanked for helpful and constructive Blevin, P.L., Chappell, B.W., Allen, C.M., 1996. Intrusive metallogenic provinces in
comments. This work was funded by a Discovery Grant from the Natural eastern Australia based on granite source and composition. Geological Society of
America, Special Paper 315, 281–290.
Sciences and Engineering Research Council of Canada. 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-group
References elements during mantle melting. Science 305, 1951–1953.
Bodnar, R.J., Beane, R.E., 1980. Temporal and spatial variations in hydrothermal fluid
Ackerman, L., Walker, R.J., Puchtel, I.S., Pitcher, L., Jelínek, E., Strnad, L., 2009. Effects of characteristics during vein filling in preore cover overlying deeply buried porphyry
melt percolation on highly siderophile elements and Os isotopes in subcontinental copper-type mineralization at Red Mountain, Arizona. Economic Geology 75,
lithospheric mantle: a study of the upper mantle profile beneath Central Europe. 876–893.
Geochimica et Cosmochimica Acta 73, 2400–2414. Borisov, A., Palme, H., 1997. Experimental determination of the solubility of platinum in
Aerts, M., Hack, A.C., Reusser, E., Ulmer, P., 2010. Assessment of the diamond-trap silicate melts. Geochimica et Cosmochimica Acta 61, 4349–4357.
method for studying high-pressure fluids and melts and an improved freezing Bourdon, B., Turner, S., Dosseto, A., 2003. Dehydration and partial melting in subduction
stage design for laser ablation ICP-MS analysis. American Mineralogist 95, zones: constraints from U-series disequilibria. Journal Geophysical Research 108,
1523–1526. B6. doi:10.1029/2002JB001839.
Ahmad, S.N., Rose, A.W., 1980. Fluid inclusions in porphyry and skarn ore at Santa Rita, Bouse, R.M., Ruiz, J., Titley, S.R., Tosdal, R.M., Wooden, J.L., 1999. Lead isotope
New Mexico. Economic Geology 75, 229–250. compositions of Late Cretaceous and Early Tertiary igneous rocks and sulfide
Ahmad, M., Solomon, M., Walshe, J.L., 1987. Mineralogical and geochemical studies of minerals in Arizona: implications for the sources of plutons and metals in porphyry
the Emperor gold telluride deposit, Fiji. Economic Geology 82, 345–370. copper deposits. Economic Geology 94, 211–244.
Aitcheson, S.J., Harmon, R.S., Moorbath, S., Schneider, A., Soler, P., Soria-Escalante, E., Brandon, A.D., Draper, D.S., 1996. Constraints on the origin of the oxidation state of
Steele, G., Swainbank, I., Wörner, G., 1995. Pb isotopes define basement domains of mantle overlying subduction zones: an example from Simcoe, Washington, USA.
the Altiplano, central Andes. Geology 23, 555–558. Geochimica et Cosmochimica Acta 60, 1739–1749.
Aizawa, Y., Tatsumi, Y., Yamada, H., 1999. Element transport by dehydration of Breeding, C.M., Ague, J.J., Bröcker, M., 2004. Fluid–metasedimentary rock interactions in
subducted sediments: implications for arc and ocean island magmatism. Island Arc subduction-zone mélange: implications for the chemical composition of arc
8, 38–46. magmas. Geology 32, 1041–1044.
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 fluid Cline, J.S., 1995. Genesis of porphyry copper deposits: the behavior of water, chloride, and
partitioning of Nb, Ta, Hf, Zr, U and Th: implications for high field strength element copper in crystallizing melts. In: Pierce, F.W., Bolm, J.G. (Eds.), Porphyry Copper
depletions in island arc basalts. Earth and Planetary Science Letters 128, 327–339. Deposits of the American Cordillera: Arizona Geological Society Digest, 20, pp. 69–82.
Brown, M., 2010. Melting of the continental crust during orogenesis: the thermal, Cline, J.S., Bodnar, R.J., 1991. Can economic porphyry copper mineralization be
rheological, and compositional consequences of melt transport from lower to generated by a typical calc-alkaline melt? Journal of Geophysical Research 96,
upper continental crust. Canadian Journal of Earth Sciences 47, 655–694. 8113–8126.
Burnham, C.W., 1979. Magmas and hydrothermal fluids, In: Barnes, H.L. (Ed.), Cloos, M., 2001. Bubbling magma chambers, cupolas, and porphyry copper deposits.
Geochemistry of Hydrothermal Ore Deposits, 2nd edition. John Wiley and Sons, International Geology Review 43, 285–311.
New York, pp. 71–136. Cloos, M., Sapiie, B., van Ufford, A.Q., Weiland, R.J., Warren, P.Q., McMahon, T.P., 2005.
Burnham, C.W., 1981. Convergence and mineralization — is there a relation? Geol. Soc. Collisional delamination in New Guinea: the geotectonics of subducting slab
America Memoir 154, 761–768. breakoff. Special Paper, 400. Geological Society of America. 51 pp.
Burnham, C.W., 1997. Magmas and hydrothermal fluids, In: Barnes, H.L. (Ed.), Comment on Defant, M.J., Kepezhinskas, P., 2001Conrey, R.M., 2002. Adakites: a review
Geochemistry of Hydrothermal Ore Deposits, 3rd edition. John Wiley and Sons, of slab melting over the past decade and the case for a slab-melt component in arcs
New York, pp. 63–123. [EOS, Trans. AGU 82, 65–69]. EOS, Trans. AGU 83, 256.
Campbell, I.H., Naldrett, A.J., 1979. The influence of silicate:sulfide ratios on the Coumou, D., Driesner, T., Heinrich, C.A., 2008. Heat transport at boiling, near-critical
geochemistry of magmatic sulfides. Economic Geology 74, 1503–1506. conditions. Geofluids 8, 208–215.
Candela, P.A., 1989a. Felsic magmas, volatiles, and metallogenesis. In: Whitney, J.A., Crerar, D.A., Barnes, H.L., 1976. Ore solution chemistry V. Solubilities of chalcopyrite and
Naldrett, A.J. (Eds.), Ore Deposition Associated with Magmas: Soc. Econ. Geol., chalcocite assemblages in hydrothermal solution at 200 °C to 350 °C. Economic
Reviews in Economic Geology, 4, pp. 223–233. Geology 71, 772–794.
Candela, P.A., 1989b. Calculation of magmatic fluid contributions to porphyry-type ore Cygan, G.L., Candela, P.A., 1995. Preliminary study of gold partitioning among pyrrhotite,
systems: predicting fluid inclusion chemistries. Geochemical Journal 23, 295–305. pyrite, magnetite, and chalcopyrite in gold-saturated chloride solutions at 600 to
Candela, P.A., 1989c. Magmatic ore-forming fluids: thermodynamic and mass transfer 700 °C, 140 MPa (1400 bars). In: Thompson, J.F.H. (Ed.), Magmas, Fluids, and Ore
calculations of metal concentrations. In: Whitney, J.A., Naldrett, A.J. (Eds.), Ore Deposition Deposits. : Short Course Series, 23. Mineralogical Association of Canada, pp. 129–137.
Associated with Magmas: Soc. Econ. Geol., Reviews in Economic Geology, 4, pp. 203–221. Dale, C.W., Burton, K.W., Pearson, D.G., Gannoun, A., Alard, O., Argles, T.W., Parkinson,
Candela, P.A., 1992. Controls on ore metal ratios in granite-related ore systems: an I.J., 2009. Highly siderophile element behaviour accompanying subduction of
experimental and computational approach. Transactions of the Royal Society of oceanic crust: whole rock and mineral-scale insights from a high-pressure terrain.
Edinburgh, Earth Sciences 83, 317–326. Geochimica et Cosmochimica Acta 73, 1394–1416.
Candela, P.A., Holland, H.D., 1984. The partitioning of copper and molybdenum Damon, P.E., 1986. Batholith-volcano coupling in the metallogeny of porphyry copper
between silicate melts and aqueous fluids. Geochimica et Cosmochimica Acta 48, deposits. In: Friedrich, G.H. (Ed.), Geology and Metallogeny of Copper Deposits.
373–380. Springer-Verlag, Berlin, pp. 216–234.
Candela, P.A., Holland, H.D., 1986. A mass transfer model for copper and molybdenum Darbyshire, D.P.F., Shepherd, T.J., 1994. Nd and Sr isotope constraints on the origin of
in magmatic hydrothermal systems: the origin of porphyry-type ore deposits. the Cornubian batholith, SW England. Journal of the Geological Society, London
Economic Geology 81, 1–19. 151, 795–802.
Candela, P.A., Piccoli, P.M., 1995. Model ore—metal partitioning from melts into vapor and Davidson, J.P., 1996. Deciphering mantle and crustal signatures in subduction zone
bapor/brine mixtures. In: Thompson, J.F.H. (Ed.), Magmas, Fluids, and Ore Deposits. : magmatism. In: Bebout, G.E., Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction: Top
Short Course Series, 23. Mineralogical Association of Canada, pp. 101–127. ch. 5. to bottom. : Geophysical Monograph, 96. American Geophysical Union, pp. 251–262.
Candela, P.A., Piccoli, P.M., 2005. Magmatic processes in the development of porphyry- Davidson, J., Turner, S., Handley, H., Macpherson, C., Dosseto, A., 2007. Amphibole
type ore systems. In: Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. “sponge” in arc crust? Geology 35, 787–790.
(Eds.), Economic Geology 100th Anniversary Volume. Society of Economic Davies, J.H., Stevenson, D., 1992. Physical model of source region of subduction zone
Geologists, Littleton, CO, pp. 25–37. volcanics. Journal of Geophysical Research 97, 2037–2070.
Canil, D., Styan, J., Larocque, J., Bonnet, E., Kyba, J., 2010. Thickness and composition of de Hoog, J.C.M., Mason, P.R.D., van Bergen, M.J., 2001. Sulfur and chalcophile elements
the Bonanza arc crustal section, Vancouver Island, Canada. Geological Society of in subduction zones: constraints from a laser ablation ICP-MS study of melt
America, Bulletin 122, 1094–1105. inclusions from Galunggung Volcano, Indonesia. Geochimica et Cosmochimica Acta
Carman, G.D., 2003. Geology, mineralization, and hydrothermal evolution of the 65, 3147–3164.
Ladolam gold deposit, Lihir Island, Papua New Guinea. In: Simmons, S.F., Graham, I. DeBari, S.M., Coleman, R.G., 1989. Examination of the deep levels of an island arc:
(Eds.), Volcanic, Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of evidence from the Tonsina ultramafic–mafic assemblage, Tonsina, Alaska. Journal
Processes Within the Earth. Special Publication, No. 10. Society of Economic of Geophysical Research 94, 4373–4391.
Geologists, pp. 247–284. Deditius, A.P., Utsunomiya, S., Ewing, R.C., Chryssoulis, S.L., Venter, D., Kesler, S.E., 2009.
Carrigan, C.R., Schubert, G., Eichelberger, J.C., 1992. Thermal and dynamical regimes of Decoupled geochemical behavior of As and Cu in hydrothermal systems. Geology
single- and two-phase magmatic flow in dikes. Journal of Geophysical Research 97 37, 707–710.
(17), 392 377-17. Defant, M.J., Drummond, M.S., 1990. Derivation of some modern arc magmas by
Carroll, M.R., Rutherford, M.J., 1985. Sulfide and sulfate saturation in hydrous silicate melting of young subducted lithosphere. Nature 347, 662–665.
melts. Proceedings of the Fifteenth Lunar and Planetary Science Conference, Part 2: Defant, M.J., Kepezhinskas, P., 2001. Adakites: a review of slab melting over the past decade
Journal of Geophysical Research, 90, pp. C601–C612. Supplement. and the case for a slab-melt component in arcs. EOS, Trans. AGU 82 (65), 68–69.
Carten, R.B., White, W.H., Stein, H.J., 1993. High-grade granite-related molybdenum Defant, M.J., Kepezhinskas, P., 2002. Reply to Comment by Conrey, R. [Adakites: a
systems. In: Kirham, R.V., Sinclair, W.D., Thorpe, R.I., Duke, J.M. (Eds.), Mineral Deposit review of slab melting over the past decade and the case for a slab-melt component
Modeling. Special Paper, 40. Geological Association of Canada, pp. 521–554. in arcs: EOS, Trans. AGU 82, 65–69]. EOS, Trans AGU 83, 256–257.
Cathles, L.M., 1991. The importance of the 350 °C isotherm in ore-forming hydrother- DePaolo, D.J., 1981. Trace element and isotopic effects of combined wallrock
mal systems. Geological Society of America, Annual Meeting, October 21–24, 1991. assimilation and fractional crystallization. Earth and Planetary Science Letters 53,
San Diego, CA, Abstracts with Programs 23, p. A21. 189–202.
Cauzid, J., Philippot, P., Martinez-Criado, G., Ménez, B., Labouré, S., 2007. Contrasting Cu- Dietrich, A., Lehmann, B., Wallianos, A., Traxel, K., Palacios, C., 1999. Magma mixing in
complexing behaviour in vapour and liquid fluid inclusions from the Yankee Lode Bolivian tin porphyries. Die Naturwissenschaften 86, 40–43.
tin deposit, Mole Granite, Australia. Chemical Geology 246, 39–54. Dilles, J.H., 1987. Petrology of the Yerington batholith, Nevada: evidence for evolution
Černý, P., Blevin, P.L., Cuney, M., London, D., 2005. Granite-related ore deposits. In: of porphyry copper ore fluids. Economic Geology 82, 1750–1789.
Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), Economic Dreyer, B.M., Morris, J.D., Gill, J.B., 2010. Incorporation of subducted slab-derived
Geology 100th Anniversary Volume. Society of Economic Geologists, Littleton, CO, sediment and fluid in arc magmas: B–Be–10Be–εNd systematics of the Kurile
pp. 337–370. convergent margin, Russia. Journal of Petrology 51, 1761–1782.
Cervantes, P., Wallace, P.J., 2003. Role of H2O in subduction-zone magmatism: new Driesner, T., 2007. The system H2O–NaCl. Part II: Correlations for molar volume,
insights from melt inclusions in high-Mg basalts from central Mexico. Geology 31, enthalpy, and isobaric heat capacity from 0 to 1000 °C, 1 to 5000 bar, and 0 to 1
235–238. XNaCl. Geochimica et Cosmochimica Acta 71, 4902–4919.
Chaplygin, I.V., Mozgova, N.N., Mokhov, A.V., Koporulina, E.V., Bernhardt, H.-J., Bryzgalov, Driesner, T., Heinrich, C.A., 2007. The system H2O–NaCl. Part I: correlation formulae for
I.A., 2007. Minerals of the system ZnS–CdS from fumaroles of the Kudriavy Volcano, phase relations in temperature–pressure–composition space from 0 to 1000 °C,
Iturup Island, Kuriles, Russia. Canadian Mineralogist 45, 709–722. 0 to 5000 bar, and 0 to 1 XNaCl. Geochimica et Cosmochimica Acta 71, 4880–4901.
Chappell, B.W., White, A.J.R., 1974. Two contrasting granite types. Pacific Geology 8, Drummond, M.S., Defant, M.J., Kepezhinskas, P.K., 1996. Petrogenesis of slab-derived
173–174. trondhjemite–tonalite–dacite/adakite magmas. Special Paper, 315. Geological
Chiaradia, M., Fontboté, L., Paladines, A., 2004. Metal sources in mineral deposits and Society of America, pp. 205–215.
crustal rocks of Ecuador (1°N–4°S): a lead isotope synthesis. Economic Geology 99, Dufek, J., Bergantz, G.W., 2005. Lower crustal magma genesis and preservation: a
1085–1106. stochastic framework for the evaluation of basalt–crust interaction. Journal of
Claeson, D.T., Meurer, W.P., 2004. Fractional crystallization of hydrous basaltic arc-type Petrology 46, 2167–2195.
magmas and the formation of amphibole-bearing gabbroic cumulates. Contributions Duggen, S., Portnyagin, M., Baker, J., Ulfbeck, D., Hoernle, K., Garbe-Schönberg, D.,
to Mineralogy and Petrology 147, 288–304. Grassineau, N., 2007. Drastic shift in lava geochemistry in the volcanic-front to rear-arc
Clemens, J.D., 2003. S-type granitic magmas — petrogenetic issues, models and region of the Southern Kamchatkan subduction zone: evidence for the transition from
evidence. Earth-Science Reviews 61, 1–18. slab surface dehydration to sediment melting. Geochimica et Cosmochimica Acta 71,
Clemens, J.D., Darbyshire, D.P.F., Flinders, J., 2009. Sources of post-orogenic calcalkaline 452–480.
magmas: the Arrochar and Garabal Hill–Glen Fyne complexes, Scotland. Lithos 112, Dunn, J.C., Hardee, H.C., 1981. Superconvecting geothermal zones. Journal of
524–542. Volcanology and Geothermal Research 11, 189–201.
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 21

Dvir, O., Pettke, T., Fumagalli, P., Kessel, R., 2011. Fluids in the peridotite–water system Halter, W.E., Pettke, T., Heinrich, C.A., 2002. The origin of Cu/Au ratios in porphyry-type
up to 6 GPa and 800 °C: new experimental constrains on dehydration reactions. ore deposits. Science 296, 1844–1846.
Contributions to Mineralogy and Petrology 161, 829–844. Halter, W.E., Heinrich, C.A., Pettke, T., 2005. Magma evolution and the formation of
Eastoe, C.J., 1982. Physics and chemistry of the hydrothermal system at the Panguna porphyry porphyry Cu–Au ore fluids: evidence from silicate and sulfide melt inclusions.
copper deposit, Bougainville, Papua New Guinea. Economic Geology 77, 127–153. Mineralium Deposita 39, 845–863.
Economou-Eliopoulos, M., Eliopoulos, D.G., 2000. Palladium, platinum and gold Hamada, M., Fujii, T., 2008. Experimental constraints on the effects of pressure and H2O
concentration in porphyry copper systems of Greece and their genetic significance. on the fractional crystallization of high-Mg island arc basalt. Contributions to
Ore Geology Reviews 16, 59–70. Mineralogy and Petrology 155, 767–790.
Edwards, K.J., 2004. Formation and degradation of seafloor hydrothermal sulfide Hamlyn, P.R., Keays, R.R., Cameron, W.E., Crawford, A.J., Waldron, H.M., 1985.
deposits. Geological Society of America Special Papers 379, 83–96. Precious metals in magnesian low-Ti lavas: implications for metallogenesis and
Eggins, S.M., 1993. Origin and differentiation of picritic arc magmas, Ambae (Aoba), sulfur saturation in primary magmas. Geochimica et Cosmochimica Acta 49,
Vanuatu. Contributions to Mineralogy and Petrology 114, 79–100. 1797–1811.
Eichelberger, J.C., 1995. Silicic volcanism: ascent of viscous magmas from crustal Hansen, J., Skjerlie, K.P., Pedersen, R.B., De La Rosa, J., 2002. Crustal melting in the lower
reservoirs. Annual Review Earth Planet Science 23, 41–63. parts of island arcs: an example from the Bremanger Granitoid Complex, west
Farmer, G.L., DePaolo, D.J., 1984. Origin of Mesozoic and Tertiary granite in the western Norwegian Caledonides. Contributions to Mineralogy and Petrology 143, 316–335.
United States and implications for pre-Mesozoic crustal structure, 2. Nd and Sr Harris, N.B.W., Pearce, J.A., Tindle, A.G., 1986. Geochemical characteristics of collision
isotopic studies of unmineralized and Cu- and Mo-mineralized granite in the zone magmatism. Special Publication, 19. Geological Society, London, pp. 67–81.
Precambrian craton. Journal of Geophysical Research 89, 10141–10160. Harris, A.C., Kamenetsky, V.S., White, N.C., van Achterbergh, E., Ryan, C.G., 2003. Melt
Faure, K., Matsuhisa, Y., Metsugi, H., Mizota, C., Hayashi, S., 2002. The Hishikari Au–Ag inclusions in veins: linking magmas and porphyry Cu deposits. Science 302,
epithermal deposit, Japan: oxygen and hydrogen isotope evidence in determining 2109–2111.
the source of paleohydrothermal fluids. Economic Geology 97, 481–498. Hart, C.J.R., Mair, J.L., Goldfarb, R.J., Groves, D.I., 2005. Source and redox controls on
Feeley, T.C., Davidson, J.P., 1994. Petrology of calc-alkaline lavas at Volcán Ollagüe and metallogenic variations in intrusion-related ore systems, Tombstone-Tungsten
the origin of compositional diversity at Central Andean stratovolcanoes. Journal of Belt, Yukon Territory. In: Ishihara, S., Stephens, W.E., Harley, S.L., Arima, M.,
Petrology 35, 1295–1340. Nakajima, T. (Eds.), Fifth Hutton Symposium; the Origin of Granites and Related
Field, C.W., Fifarek, R.H., 1985. Light stable-isotope systematics in the epithermal Rocks. Special Paper, 389. Geological Society of America, pp. 339–356.
environment. In: Berger, B.R., Bethke, P.M. (Eds.), Geology and Geochemistry of Haschke, M., Ben-Avraham, Z., 2005. Adakites from collision-modified lithosphere.
Epithermal Systems. Reviews in Economic Geology, 2. Society of Economic Geophysical Research Letters 32 (15), L15302. doi:10.1029/2005GL023468.
Geologists, pp. 99–128. Hattori, K.H., 1997. Occurrence and origin of sulfide and sulfate in the 1991 Mount
Foley, S.F., Barth, M.G., Jenner, G.A., 2000. Rutile/melt partition coefficients for trace Pinatubo eruption products. In: Newhall, C.G., Punongbayan, R.S. (Eds.), Fire and
elements and an assessment of the influence of rutile on the trace element Mud: Eruptions and Lahars of Mount Pinatubo, Philippines. University of
characteristics of subduction zone magmas. Geochimica et Cosmochimica Acta 64, Washington Press, Seattle, pp. 807–824.
933–938. Hattori, K., Guillot, S., 2003. Volcanic fronts form as a consequence of serpentine
Fontboté, L., Gunnesch, K.A., Baumann, A., 1990. Metal sources in stratabound ore dehydration in the forearc mantle wedge. Geology 31, 525–528.
deposits in the Andes (Andean Cycle) — lead isotopic constraints. In: Fontboté, L., Hattori, K.H., Keith, J.D., 2001. Contribution of mafic melt to porphyry copper
Amstutz, G.C., Cardozo, M., Cedillo, E., Frutos, J. (Eds.), Stratabound Ore Deposits in mineralization: evidence from Mount Pinatubo, Philippines, and Bingham Canyon,
the Andes. Springer-Verlag, Berlin, pp. 759–773. Utah, USA. Mineralium Deposita 36, 799–806.
Forneris, J.F., Holloway, J.R., 2003. Phase equilibria in subducting basaltic crust: Hattori, K., Takahashi, Y., Guillot, S., Johanson, B., 2005. Occurrence of arsenic (V) in
implications for H2O release from the slab. Earth and Planetary Science Letters 214, forearc mantle serpentinites based on X-ray absorption spectroscopy study.
187–201. Geochimica et Cosmochimica Acta 69, 5585–5596.
Fournier, R.O., 1985. The behavior of silica in hydrothermal solutions. In: Berger, B.R., Hattori, K., Wallis, S., Enami, M., Mizukami, T., 2010. Subduction of mantle wedge
Bethke, P.M. (Eds.), Reviews in Economic Geology, 2. Society of Economic peridotites: evidence from the Higashi-akaishi ultramafic body in the Sanbagawa
Geologists, pp. 45–61. metamorphic belt. Island Arc 19, 192–207.
Fournier, R.O., 1999. Hydrothermal processes related to movement of fluid from plastic Hawkesworth, C.J., Gallagher, K., Hergt, J.M., McDermott, F., 1994. Destructive plate
into brittle rock in the magmatic–epithermal environment. Economic Geology 94, margin magmatism: geochemistry and melt generation. Lithos 33, 169–188.
1193–1212. Heald, P., Foley, N.K., Hayba, D.O., 1987. Comparative anatomy of volcanic-hosted
Fumagalli, P., Poli, S., 2005. Experimentally determined phase relations in hydrous epithermal deposits: acid-sulfate and adularia-sericite types. Economic Geology 82,
peridotites to 6.5 GPa and their consequences on the dynamics of subduction 1–26.
zones. Journal of Petrology 46, 555–578. Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of
Fyfe, W.S., 1992. Magma underplating of continental crust. Journal Volcanology hydrothermal ore deposits. Nature 370, 519–527.
Geothermal Research 50, 33–40. Hedenquist, J.W., Richards, J.P., 1998. The influence of geochemical techniques on the
Garrido, C.J., Bodinier, J.-L., Burg, J.-P., Zeilinger, G., Hussain, S.S., Dawood, H., Chaudhry, development of genetic models for porphyry copper deposits. In: Richards, J.P.,
M.N., Gervilla, F., 2006. Petrogenesis of mafic garnet granulite in the lower crust of Larson, P.B. (Eds.), Techniques in Hydrothermal Ore Deposits Geology: Reviews in
the Kohistan paleo-arc complex (northern Pakistan): implications for intra-crustal Economic Geology, 10, pp. 235–256. ch. 10.
differentiation of island arcs and generation of continental crust. Journal of Hedenquist, J.W., Simmons, S.F., Giggenbach, W.F., Eldridge, C.S., 1993. White Island,
Petrology 47, 1873–1914. New Zealand, volcanic–hydrothermal system represents the geochemical environ-
Garrison, J.M., Davidson, J.P., 2003. Dubious case for slab melting in the Northern ment of high-sulfidation Cu and Au ore deposition. Geology 21, 731–734.
volcanic zone of the Andes. Geology 31, 565–568. Hedenquist, J.W., Aoki, M., Shinohara, H., 1994a. Flux of volatiles and ore-forming
Gill, J.B., 1981. Orogenic Andesites and Plate Tectonics. Springer-Verlag, New York. metals from the magmatic–hydrothermal system of Satsuma Iwojima volcano.
390 pp. Geology 22, 585–588.
Gill, J., Whelan, P., 1989. Postsubduction ocean island alkali basalts in Fiji. Journal of Hedenquist, J.W., Matsuhisa, Y., Izawa, E., White, N.C., Giggenbach, W.F., Aoki, M.,
Geophysical Research 94, 4579–4588. 1994b. Geology, geochemistry, and origin of high sulfidation Cu–Au mineralization
Goff, S.J., Goff, F., Janik, C.J., 1992. Tecuamburro volcano, Guatemala: exploration in the Nansatsu district, Japan. Economic Geology 89, 1–30.
geothermal gradient drilling and results. Geothermics 21, 483–502. Hedenquist, J.W., Arribas Jr., A., Reynolds, J.R., 1998. Evolution of an intrusion-centered
Green, T.H., Adam, J., 2003. Experimentally-determined trace element characteristics of hydrothermal system: Far Southeast–Lepanto porphyry and epithermal Cu–Au
aqueous fluid from partially dehydrated mafic oceanic crust at 3.0 GPa, 650–700 °C. deposits, Philippines. Economic Geology 93, 373–404.
European Journal of Mineralogy 15, 815–830. Heinrich, C.A., 2005. The physical and chemical evolution of low-salinity magmatic
Greene, A.R., Debari, S.M., Kelemen, P.B., Blusztajn, J., Clift, P.D., 2006. A detailed fluids at the porphyry to epithermal transition: a thermodynamic study.
geochemical study of island arc crust: the Talkeetna Arc section, south-central Mineralium Deposita 39, 864–889.
Alaska. Journal of Petrology 47, 1051–1093. Heinrich, C.A., Ryan, G.G., Mernagh, T.P., Eadington, P.J., 1992. Segregation of ore metals
Grove, T.L., Elkins-Tanton, L.T., Parman, S.W., Chatterjee, N., Müntener, O., Gaetani, G.A., between magmatic brine and vapor: a fluid inclusion study using PIXE
2003. Fractional crystallization and mantle-melting controls on calc-alkaline microanalysis. Economic Geology 87, 1566–1583.
differentiation trends. Contributions to Mineralogy and Petrology 145, 515–533. Heinrich, C.A., Günther, D., Audétat, A., Ulrich, T., Frischknecht, R., 1999. Metal
Grove, T.L., Chatterjee, N., Parman, S.W., Médard, E., 2006. The influence of H2O on fractionation between magmatic brine and vapor, determined by microanalysis of
mantle wedge melting. Earth and Planetary Science Letters 249, 74–89. fluid inclusions. Geology 27, 755–758.
Grunder, A.L., Klemetti, E.W., Feeley, T.C., McKee, C.M., 2008. Eleven million years of arc Heinrich, C.A., Halter, W., Klemm, L., Landtwing, M.R., Pettke, T., 2003a. Laser ablation
volcanism at the Aucanquilcha Volcanic Cluster, northern Chilean Andes: micro-analysis of fluid and melt inclusions: towards understanding the process of
implications for the life span and emplacement of plutons. Transactions: Earth porphyry-style ore formation. Applied Earth Science (Transactions, Institutions of
Sciences, 97. Royal Society of Edinburgh, pp. 415–436. Mining and Metallurgy, Section B) 112, 185–186.
Guivel, C., Lagabrielle, Y., Bourgois, J., Martin, H., Arnaud, N., Fourcade, S., Cotten, J., Heinrich, C.A., Pettke, T., Halter, W.E., Aigner-Torres, M., Audétat, A., Günther, D.,
Maury, R.C., 2003. Very shallow melting of oceanic crust during spreading ridge Hattendorf, B., Bleiner, D., Guillong, M., Horn, I., 2003b. Quantitative multi-element
subduction: origin of near-trench Quaternary volcanism at the Chile Triple Junction. analysis of minerals, fluid and melt inclusions by laser-ablation inductively-
Journal of Geophysical Research 108 (B7), 2345. doi:10.1029/2002JB002119. coupled-plasma mass-spectrometry. Geochimica et Cosmochimica Acta 67,
Guo, F., Wilson, M., Li, C., 2007. Post-collisional adakites in south Tibet: products of 3473–3497.
partial melting of subduction-modified lower crust. Lithos 96, 205–224. Heinrich, C.A., Dreisner, T., Steffánson, A., Seward, T.M., 2004. Magmatic vapor
Gutscher, M.-A., Maury, R., Eissen, J.-P., Bourdon, E., 2000. Can slab melting be caused by contraction and the transport of gold from the porphyry environment to
flat subduction? Geology 28, 535–538. epithermal ore deposits. Geology 32, 761–764.
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 studies Kay, S.M., Mpodozis, C., Coira, B., 1999. Neogene magmatism, tectonism, and mineral
in rock-buffered systems: II. Some general geologic applications. Economic Geology deposits of the Central Andes (22° to 33°S latitude). In: Skinner, B.J. (Ed.), Geology
87, 23–43. and Ore Deposits of the Central Andes. : Special Publication, No. 7. Society of
Hemley, J.J., Cygan, G.L., Fein, J.B., Robinson, G.R., d'Angelo, W.M., 1992. Hydrothermal Economic Geologists, pp. 27–59.
ore-forming processes in the light of studies in rock-buffered systems: I. Iron– Keith, J.D., Shanks III, W.C., Archibald, D.A., Farrar, E., 1986. Volcanic and intrusive
copper–zinc–lead sulfide solubility relations. Economic Geology 87, 1–22. history of the Pine Grove porphyry molybdenum system, southwestern Utah.
Henley, R.W., Hughes, G.O., 2000. Underground fumaroles: “excess heat” effects in vein Economic Geology 81, 553–577.
formation. Economic Geology 95, 453–466. Keith, J.D., Whitney, J.A., Hattori, K., Ballantyne, G.H., Christiansen, E.H., Barr, D.L.,
Henley, R.W., McNabb, A., 1978. Magmatic vapor plumes and ground–water interaction Cannan, T.M., Hook, C.J., 1997. The role of magmatic sulfides and mafic alkaline
in porphyry copper emplacement. Economic Geology 73, 1–20. magmas in the Bingham and Tintic mining districts, Utah. Journal of Petrology 38,
Hermann, J., Spandler, C.J., 2008. Sediment melts at sub-arc depths: an experimental 1679–1690.
study. Journal of Petrology 49, 717–740. Keith, J.D., Christiansen, E.H., Maughan, D.T., Waite, K.A., 1998. The role of mafic alkaline
Herzberg, C.T., Fyfe, W.S., Carr, M.J., 1983. Density constraints on the formation of the magmas in felsic porphyry-Cu and Mo systems. In: Lentz, D.R. (Ed.), Mineralized
continental Moho and crust. Contributions to Mineralogy and Petrology 84, 1–5. Intrusion-Related Skarn Systems: Mineralogical Association of Canada Short Course
Herzig, P.M., Hannington, M.D., Scott, S.D., Maliotis, G., Rona, P.A., Thompson, G., 1991. Series, 26, pp. 211–243.
Gold-rich sea-floor gossans in the Troodos Ophiolite and on the Mid-Atlantic Ridge. Kelley, K.A., Cottrell, E., 2009. Water and the oxidation state of subduction zone
Economic Geology 86, 1747–1755. magmas. Science 325, 605–607.
Hildreth, W., 1981. Gradients in silicic magma chambers: implications for lithospheric Kelley, K.D., Romberger, S.B., Beaty, D.W., Pontius, J.A., Snee, L.W., Stein, H.J., Thompson,
magmatism. Journal of Geophysical Research 86 (10), 192 153-10. T.B., 1998. Geochemical and geochronological constraints on the genesis of Au–Te
Hildreth, W., Moorbath, S., 1988. Crustal contributions to arc magmatism in the Andes deposits at Cripple Creek, Colorado. Economic Geology 93, 981–1012.
of central Chile. Contributions to Mineralogy and Petrology 98, 455–489. Kelley, K.A., Plank, T., Newman, S., Stolper, E.M., Grove, T.L., Parman, S., Hauri, E., 2010.
Holland, H.D., 1965. Some applications of thermochemical data to problems of ore Mantle melting as a function of water content beneath the Mariana Arc. Journal of
deposits II. Mineral assemblages and the composition of ore forming fluids. Petrology 51, 1711–1738.
Economic Geology 60, 1101–1166. Kemp, A.I.S., Hawkesworth, C.J., Foster, G.L., Paterson, B.A., Woodhead, J.D., Hergt, J.M.,
Holloway, J.R., 1976. Fluids in the evolution of granitic magmas: consequences of finite Gray, C.M., Whitehouse, M.J., 2007. Magmatic and crustal differentiation history of
CO2 solubility. Geological Society of America Bulletin 87, 1513–1518. granitic rocks from Hf–O isotopes in zircon. Science 315, 980–983.
Hou, Z.Q., Zeng, P.S., Gao, Y.F., Dong, F.L., 2006. Himalayan Cu–Mo–Au mineralization in Kennedy, A.K., Hart, S.R., Frey, F.A., 1990. Composition and isotopic constraints on the
the eastern Indo-Asian collision zone: constraints from Re–Os dating of petrogenesis of alkaline arc lavas: Lihir Island, Papua New Guinea. Journal of
molybdenite. Mineralium Deposita 41, 33–45. Geophysical Research 95, 6929–6942.
Hou, Z., Yang, Z., Qu, X., Meng, X., Li, Z., Beaudoin, G., Rui, Z., Gao, Y., Zaw, K., 2009. The Kent, A.J.R., Peate, D.W., Newman, S., Stolper, E.M., Pearce, J.A., 2002. Chlorine in
Miocene Gangdese porphyry copper belt generated during post-collisional submarine glasses from the Lau Basin: seawater contamination and constraints on
extension in the Tibetan Orogen. Ore Geology Reviews 36, 25–51. the composition of slab-derived fluids. Earth and Planetary Science Letters 202,
Huppert, H.E., Sparks, R.S.J., 1988. The generation of granitic magmas by intrusion of 361–377.
basalt into continental crust. Journal of Petrology 29, 599–624. Kepezhinskas, P., Defant, M.J., Widom, E., 2002. Abundance and distribution of PGE and
Ishihara, S., 1981. The granitoid series and mineralization. Economic Geology 75th Au in the island-arc mantle: implications for sub-arc metasomatism. Lithos 60,
Anniversary Volume, pp. 458–484. 113–128.
Ishihara, S., Murakami, H., 2006. Fractionated ilmenite-series granites in southwest Japan: Kerrich, R., Beckinsale, R.D., 1988. Oxygen and strontium isotopic evidence for the
source magma for REE–Sn–W mineralizations. Resource Geology 56, 245–256. origin of granites in the tin belt of southeast Asia. In: Taylor, R.P., Strong, D.F. (Eds.),
Jagoutz, O., Muntener, O., Ulmer, P., Pettke, T., Burg, J.-P., Dawood, H., Hussain, S., 2007. Recent Advances in the Geology of Granite-Related Mineral Deposits: Canadian
Petrology and mineral chemistry of lower crustal intrusions: the Chilas Complex, Institute of Mining and Metallurgy, Special Volume 39, pp. 115–123.
Kohistan (NW Pakistan). Journal of Petrology 48, 1895–1953. Keskin, M., Genç, Ş.C., Tüysüz, O., 2008. Petrology and geochemistry of post-collisional
Jagoutz, O.E., Burg, J.-P., Hussain, S., Dawood, H., Pettke, T., Iizuka, T., Maruyama, S., Middle Eocene volcanic units in North-Central Turkey: evidence for magma
2009. Construction of the granitoid crust of an island arc part I: geochronological generation by slab breakoff following the closure of the Northern Neotethys Ocean.
and geochemical constraints from the plutonic Kohistan (NW Pakistan). Contri- Lithos 104, 267–305.
butions to Mineralogy and Petrology 158, 739–755. Kesler, S.E., 1973. Copper, molybdenum and gold abundances in porphyry copper
James, D.E., 1982. A combined O, Sr, Nd, and Pb isotopic and trace element study of deposits. Economic Geology 68, 106–112.
crustal contamination in central Andean lavas, I. Local geochemical variations. Kessel, R., Schmidt, M.W., Ulmer, P., Pettke, P., 2005a. Trace element signature of
Earth and Planetary Science Letters 57, 47–62. subduction-zone fluids, melts and supercritical liquids at 120–180 km depth.
Jégo, S., Pichavant, M., Mavrogenes, J.A., 2010. Controls on gold solubility in arc Nature 437, 724–727.
magmas: an experimental study at 1000 °C and 4 kbar. Geochimica et Cosmochi- Kessel, R., Ulmer, P., Pettke, P., Schmidt, M.W., Thompson, A.B., 2005b. The water–basalt
mica Acta 74, 2165–2189. system at 4 to 6 GPa: phase relations and second critical endpoint in a K-free
Jenner, F.E., O'Neill, H., St, C., Arculus, R.J., Mavrogenes, J.A., 2010. The magnetite crisis in eclogite at 700 to 1400 °C. Earth and Planetary Science Letters 237, 873–892.
the evolution of arc-related magmas and the initial concentration of Au, Ag and Cu. Kilian, R., Behrmann, J.H., 2003. Geochemical constraints on the sources of Southern
Journal of Petrology 51, 2445–2464. Chile Trench sediments and their recycling in arc magmas of the Southern Andes.
Jensen, E.P., Barton, M.D., 2000. Gold deposits related to alkaline magmatism. In: Journal of the Geological Society 160, 57–70.
Hagemann, S.G., Brown, P.E. (Eds.), Gold in 2000: Reviews in Economic Geology, 13, Kilian, R., Stern, C.R., 2002. Constraints on the interaction between slab melts and the
pp. 279–314. mantle wedge from adakitic glass in peridotite xenoliths. European Journal of
John, D.A., 1991. Evolution of hydrothermal fluids in the Alta Stock, Central Wasatch Mineralogy 14, 25–36.
Mountains, Utah. U.S. Geological Survey, Bulletin 1977. 51 pp. Kirkham, R.V., Sinclair, W.D., 1996. Porphyry copper, gold, molybdenum, tungsten, tin,
Johnson, M.C., Plank, T., 1999. Dehydration and melting experiments constrain the fate of silver. In: Eckstrand, O.R., Sinclair, W.D., Thorpe, R.I. (Eds.), Geology of Canadian
subducted sediments. Geochemistry Geophysics Geosystems 1, 1007. doi:10.1029/ mineral deposits. : Geology of Canada, 8. Geological Survey of Canada, pp. 421–446.
1999GC000014. Klemm, L.M., Pettke, T., Heinrich, C.A., Campos, E., 2007. Hydrothermal evolution of the
Johnson, R.W., Mackenzie, D.E., Smith, I.E.M., 1978. Delayed partial melting of El Teniente deposit, Chile: porphyry Cu–Mo ore deposit from low-salinity
subduction-modified mantle in Papua New Guinea. Tectonophysics 46, 197–216. magmatic fluids. Economic Geology 102, 1021–1045.
Jugo, P.J., 2009. Sulfur content at sulfide saturation in oxidized magmas. Geology 37, Klemm, L.M., Pettke, T., Heinrich, C.A., 2008. Fluid and source magma evolution of the
415–418. Questa porphyry Mo deposit, New Mexico, USA. Mineralium Deposita 43, 533–552.
Jugo, P.J., Candela, P.A., Piccoli, P.M., 1999. Magmatic sulfides and Au:Cu ratios in Klemme, S., Prowatke, S., Hametner, K., Gunther, D., 2005. Partitioning of trace elements
porphyry deposits: an experimental study of copper and gold partitioning at between rutile and silicate melts: implications for subduction zones. Geochimica et
850 °C, 100 MPa in a haplogranitic melt–pyrrhotite–intermediate solid solution– Cosmochimica Acta 69, 2361–2371.
gold metal assemblage, at gas saturation. Lithos 46, 573–589. Klepeis, K.A., Clarke, G.L., Rushmer, T., 2003. Magma transport and coupling between
Jugo, P.J., Luth, R.W., Richards, J.P., 2005a. Experimental data on the speciation of sulfur as a deformation and magmatism in the continental lithosphere. GSA Today 13, 4–11.
function of oxygen fugacity in basaltic melts. Geochimica et Cosmochimica Acta 69, Kogiso, T., Tatsumi, Y., Nakano, S., 1997. Trace element transport during dehydration
497–503. processes in the subducted oceanic crust: 1. Experiments and implications for the
Jugo, P.J., Luth, R.W., Richards, J.P., 2005b. An experimental study of the sulfur content in origin of ocean island basalts. Earth and Planetary Science Letters 148, 193–205.
basaltic melts saturated with immiscible sulfide or sulfate liquids at 1300 °C and Kontak, D.J., Cumming, G.L., Krstic, D., Clark, A.H., Farrar, E., 1990. Isotopic composition
1.0 GPa. Journal of Petrology 46, 783–798. of lead in ore deposits of the Cordillera Oriental, southeastern Peru. Economic
Jugo, P.J., Wilke, M., Botcharnikov, R.E., 2010. Sulfur K-edge XANES analysis of natural Geology 85, 1584–1603.
and synthetic basaltic glasses: implications for S speciation and S content as Kusakabe, M., Komoda, Y., Takano, B., Abiko, T., 2000. Sulfur isotopic effects in the
function of oxygen fugacity. Geochimica et Cosmochimica Acta 74, 5926–5938. disproportionation reaction of sulfur dioxide in hydrothermal fluids: implications
Kawamoto, T., 2006. Hydrous phases and water transport in the subducting slab. for the ™34S variations of dissolved bisulfate and elemental sulfur from active
Reviews in Mineralogy and Geochemistry 62, 273–289. crater lakes. Journal of Volcanology and Geothermal Research 97, 287–307.
Kay, R.W., 1978. Aleutian magnesian andesites: melts from subducted Pacific ocean Kuscu, I., Kuscu, G.G., Tosdal, R.M., Ulrich, T.D., Friedman, R., 2010. Magmatism in the
crust. Journal of Volcanology and Geothermal Research 4, 117–132. southeastern Anatolian orogenic belt: transition from arc to post-collisional setting in
Kay, S.M., Ramos, V.A., Marquez, M., 1993. Evidence in Cerro Pampa volcanic rocks for an evolving orogen. Special Publications, 340. Geological Society, London, pp. 437–460.
slab-melting prior to ridge–trench collision in southern South America. Journal of Kushiro, I., Syono, Y., Akimoto, S., 1968. Melting of a peridotite nodule at high pressures
Geology 101, 703–714. and high water pressures. Journal of Geophysical Research 73, 6023–6029.
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 23

Landtwing, M.R., Pettke, T., Halter, W.E., Heinrich, C.A., Redmond, P.B., Einaudi, M.T., Mukasa, S.B., Vidal, C.E., Injoque-Espinoza, J., 1990. Pb isotope bearing on the
Kunze, K., 2005. Copper deposition during quartz dissolution by cooling magmatic– metallogenesis of sulfide ore deposits in central and southern Peru. Economic
hydrothermal fluids: the Bingham porphyry. Earth and Planetary Science Letters Geology 85, 1438–1446.
235, 229–243. Müller, D., Groves, D.I., 1993. Direct and indirect associations between potassic igneous
Landtwing, M.R., Furrer, C., Redmond, P.B., Pettke, T., Guillong, M., Heinrich, C.A., 2010. rocks, shoshonites and gold–copper deposits. Ore Geology Reviews 8, 383–406.
The Bingham Canyon porphyry Cu–Mo–Au deposit III. Zoned copper–gold ore Mungall, J.E., 2002. Roasting the mantle: slab melting and the genesis of major Au and
deposition by magmatic vapor expansion. Economic Geology 105, 91–118. Au-rich Cu deposits. Geology 30, 915–918.
Larocque, J., Canil, D., 2010. The role of amphibole in the evolution of arc magmas and Müntener, O., Ulmer, P., 2006. Experimentally derived high-pressure cumulates from
crust: the case from the Jurassic Bonanza arc section, Vancouver Island, Canada. hydrous arc magmas and consequences for the seismic velocity structure of lower
Contributions to Mineralogy and Petrology 159, 475–492. arc crust. Geophysical Research Letters 33, L21308. doi:10.1029/2006GL027629.
Larocque, A.C.L., Stimac, J.A., Siebe, C., Greengrass, K., Chapman, R., Mejia, S.R., 2008. Murakami, H., Seo, J.H., Heinrich, C.A., 2010. The relation between Cu/Au ratio and formation
Deposition of a high-sulfidation Au assemblage from a magmatic volatile phase, Volcán depth of porphyry-style Cu–Au ±Mo deposits. Mineralium Deposita 45, 11–21.
Popocatépetl, Mexico. Journal of Volcanology and Geothermal Research 170, 51–60. Mutschler, F.E., Griffin, M.E., Stevens, D.S., Shannon, S.S., 1985. Precious metal deposits
Leeman, W.P., 1983. The influence of crustal structure on compositions of subduction- related to alkaline rocks in the North American Cordillera — an interpretive review.
related magmas. Journal Volcanology and Geothermal Research 18, 561–588. Transactions, 88. Geological Society of South Africa, pp. 355–377.
Lehmann, B., 1982. Metallogeny of tin: magmatic diferentiation versus geological Naldrett, A.J., 1989. Sulfide melts—crystallization temperatures, solubilities in silicate
heritage. Economic Geology 77, 50–59. melts, and Fe, Ni, and Cu partitioning between basaltic magmas, and olivine. In:
Lewis, K.C., Lowell, R.P., 2009. Numerical modeling of two-phase flow in the NaCl–H2O Whitney, J.A., Naldrett, A.J. (Eds.), Ore Deposition Associated with Magmas.
system: 2. Examples. Journal of Geophysical Research 114. doi:10.1029/2008JB006030 Reviews in Economic Geology, 4. Society of Economic Geologists, pp. 5–20. ch. 2,.
16 pp., B08204. Naney, M.T., 1983. Phase equilibria of rock-forming ferromagnesian silicates in granitic
Lowell, J.D., Guilbert, J.M., 1970. Lateral and vertical alteration-mineralization zoning in systems. American Journal of Science 283, 993–1033.
porphyry copper ore deposits. Economic Geology 65, 373–408. Nash, J.T., 1976. Fluid-inclusion petrology — data from porphyry copper deposits and
Lowenstern, J.B., 2001. Carbon dioxide in magmas and implications for hydrothermal applications to exploration. Professional Paper, 907-D. U.S. Geological Survey. 16 pp.
systems. Mineralium Deposita 36, 490–502. Neubauer, F., Lips, A., Kouzmanov, K., Lexa, J., Ivascanu, P., 2005. Subduction, slab
Lowenstern, J.B., Mahood, G.A., Rivers, M.L., Sutton, S.R., 1991. Evidence for extreme detachment and mineralization: the Neogene in the Apuseni Mountains and
partitioning of copper into a magmatic vapor phase. Science 252, 1405–1409. Carpathians. Ore Geology Reviews 27, 13–44.
Lynton, S.J., Candela, P.A., Piccoli, P.M., 1993. An experimental study of the partitioning Neuendorf, K.K.E., Mehl Jr., J.P., Jackson, J.A. (Eds.), 2005. Glossary of Geology, 5th
of copper between pyrrhotite and a high silica rhyolitic melt. Economic Geology 88, edition. American Geological Institute, Alexandria, Virginia. 800 pp.
901–915. Newberry, R.J., Swanson, S.E., 1986. Scheelite skarn granitoids: an evaluation of the
MacDonald, R., Hawkesworth, C.J., Heath, E., 2000. The Lesser Antilles volcanic chain: a roles of magmatic source and process. Ore Geology Reviews 1, 57–81.
study in arc magmatism. Earth-Science Reviews 49, 1–76. Noll, P.D., Newsom, H.E., Leeman, W.P., Ryan, J.G., 1996. The role of hydrothermal fluids
Macfarlane, A.W., 1999. Isotopic studies of northern Andean crustal evolution and ore in the production of subduction zone magmas: evidence from siderophile and
metal sources. In: Skinner, B.J. (Ed.), Geology and Ore Deposits of the Central Andes. : chalcophile trace elements and boron. Geochim. Cosmochim. Acta 60, 587–611.
Special Publication, No. 7. Society of Economic Geologists, pp. 195–217. Noorollahi, Y., Itoi, R., Fujii, H., Tanaka, T., 2007. GIS model for geothermal resource
Macfarlane, A.W., Marcet, P., LeHuray, A.P., Petersen, U., 1990. Lead isotope provinces of exploration in Akita and Iwate prefectures, northern Japan. Computers and
the Central Andes inferred from ores and crustal rocks. Economic Geology 85, Geosciences 33, 1008–1021.
1857–1880. Norton, D.L., 1982. Fluid and heat transport phenomena typical of copper-bearing
Macpherson, C.G., Dreher, S.T., Thirlwall, M.F., 2006. Adakites without slab melting: pluton environments. In: Titley, S.R. (Ed.), Advances in Geology of Porphyry Copper
high pressure differentiation of island arc magma, Mindanao, the Philippines. Earth Deposits of Southwestern North America. Tuscon, Univ, Arizona Press, pp. 59–72.
and Planetary Science Letters 243, 581–593. Norton, D.L., Cathles, L.M., 1973. Breccia pipes, products of exsolved vapor from
Malaspina, N., Poli, S., Fumagalli, P., 2009. The oxidation state of metasomatized mantle magmas. Economic Geology 68, 540–546.
wedge: insights from C–O–H-bearing garnet peridotite. Journal of Petrology 50, Norton, D.L., Dutrow, B.L., 2001. Complex behavior of magma–hydrothermal processes:
1533–1552. role of supercritical fluid. Geochimica et Cosmochimica Acta 65, 4009–4017.
Mancano, D.P., Campbell, A.R., 1995. Microthermometry of enargite-hosted fluid in- Oyarzun, R., Márquez, A., Lillo, J., López, I., Rivera, S., 2001. Giant versus small porphyry
clusions from the Lepanto, Philippines, high-sulfidation CuAu deposit. Geochimica et copper deposits of Cenozoic age in northern Chile: adakitic versus normal calc-
Cosmochimica Acta 59, 3909–3916. alkaline magmatism. Mineralium Deposita 36, 794–798.
Manning, C.E., 2004. The chemistry of subduction-zone fluids. Earth and Planetary Oyarzun, R., Márquez, A., Lillo, J., López, I., Rivera, S., 2002. Reply to discussion on “Giant
Science Letters 223, 1–16. versus small porphyry copper deposits of Cenozoic age in northern Chile: adakitic
Manske, S.L., Hedenquist, J.W., O'Connor, G., Tămaş, C., Cauuet, B., Leary, S., Minut, A., versus normal calc-alkaline magmatism” by Oyarzun et al. (Mineralium Deposita
2006. Roşia Montană, Romania: Europe's largest gold deposit January Society of 36: 794–798, 2001). Mineralium Deposita 37, 791–794.
Economic Geologists Newsletter 64 (1), 9–15. Parkinson, I.J., Arculus, R.J., 1999. The redox state of subduction zones: insights from
Martin, H., 1999. Adakitic magmas: modern analogues of Archaean granitoids. Lithos arc-peridotites. Chemical Geology 160, 409–423.
46, 411–429. Paterson, J.T., Cloos, M., 2005. Grasberg porphyry Cu–Au deposit, Papua, Indonesia: 1.
Martin, H., Smithies, R.H., Rapp, R., Moyen, J.-F., Champion, D., 2005. An overview of Magmatic history. In: Porter, T.M. (Ed.), Super Porphyry Copper and Gold Deposits:
adakite, tonalite–trondhjemite–granodiorite (TTG), and sanukitoid: relationships A Global Perspective, 2. Porter Geoscience Consulting Publishing, perspectiveLin-
and some implications for crustal evolution. Lithos 79, 1–24. den Park, South Australia, pp. 313–329.
Maughan, D.T., Keith, J.D., Christiansen, E.H., Pulsipher, T., Hattori, K., Evans, N.J., 2002. Peach, C.L., Mathez, E.A., Keays, R.R., 1990. Sulfide melt–silicate melt distribution
Contributions from mafic alkaline magmas to the Bingham porphyry Cu–Au–Mo coefficients for noble metals and other chalcophile elements as deduced from
deposit, Utah, USA. Mineralium Deposita 37, 14–37. MORB: implications for partial melting. Geochimica et Cosmochimica Acta 54,
Mavrogenes, J.A., Bodnar, R.J., 1994. Hydrogen movement into and out of fluid 3379–3389.
inclusions in quartz; experimental evidence and geologic implications. Geochimica Peacock, S.M., 1993. Large-scale hydration of the lithosphere above subducting slabs.
et Cosmochimica Acta 58, 141–148. Chemical Geology 108, 49–59.
McInnes, B.I.A., Cameron, E.M., 1994. Carbonated, alkaline hybridizing melts from a sub- Peacock, S.M., 1996. Thermal and petrologic structure of subduction zones. In: Bebout, G.E.,
arc environment: mantle wedge samples from the Tabar-Lihir-Tanga-Feni arc, Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction: Top to Bottom. : Geophysical
Papua New Guinea. Earth and Planetary Science Letters 122, 125–141. Monograph, 96. American Geophysical Union, Washington, DC, pp. 119–133.
McInnes, B.I.A., McBride, J.S., Evans, N.J., Lambert, D.D., Andrew, A.A., 1999. Osmium isotope Peacock, S.M., Rushmer, T., Thompson, A.B., 1994. Partial melting of subducting oceanic
constraints on ore metal recycling in subduction zones. Science 286, 512–516. crust. Earth and Planetary Science Letters 121, 227–244.
McInnes, B.I.A., Gregoire, M., Binns, R.A., Herzig, P.M., Hannington, M.D., 2001. Hydrous Pearce, J.A., Bender, J.F., De Long, S.E., Kidd, W.S.F., Low, P.J., Güner, Y., Saroglu, F., Yilmaz, Y.,
metasomatism of oceanic sub-arc mantle, Lihir, Papua New Guinea: petrology and Moorbath, S., Mitchell, J.G., 1990. Genesis of collision volcanism in eastern Anatolia,
geochemistry of fluid-metasomatised mantle wedge xenoliths. Earth and Planetary Turkey. Journal of Volcanology and Geothermal Research 44, 189–229.
Science Letters 188, 169–183. Petford, N., Gallagher, K., 2001. Partial melting of mafic (amphibolitic) lower crust by
McNutt, R.H., Clark, A.H., Zentilli, M., 1979. Lead isotopic compositions of Andean periodic influx of basaltic magma. Earth and Planetary Science Letters 193,
igneous rocks, Latitudes 26° to 29° S: petrologic and metallogenic implications. 483–499.
Economic Geology 74, 827–837. Pettke, T., Oberli, F., Heinrich, C.A., 2010. The magma and metal source of giant
Meinert, L.D., Dipple, G.M., Nicolescu, S., 2005. World skarn deposits. In: porphyry-type ore deposits, based on lead isotope microanalysis of individual fluid
Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), Economic inclusions. Earth and Planetary Science Letters 296, 267–277.
Geology 100th Anniversary Volume. Society of Economic Geologists, Littleton, Pichavant, M., Mysen, B.O., Macdonald, R., 2002. Source and H2O content of high-MgO
CO, pp. 299–336. magmas in island arc settings: an experimental study of a primitive calc-alkaline
Mitchell, R.H., Keays, R.R., 1981. Abundance and distribution of gold, palladium and basalt from St. Vincent, Lesser Antilles arc. Geochimica et Cosmochimica Acta 66,
iridium in some spinel and garnet lherzolites: implications for the nature and origin 2193–2209.
of precious metal-rich intergranular components in the upper mantle. Geochimica Pitcher, W.S., 1997. The Nature and Origin of Granite, 2nd edition. Chapman and Hall,
et Cosmochimica Acta 45, 2425–2442. London. 387 pp.
Moore, W.J., Nash, J.T., 1974. Alteration and fluid inclusion studies of the porphyry Pitzer, K.S., Pabalan, R.T., 1986. Thermodynamics of NaCl in steam. Geochimica et
copper ore body at Bingham, Utah. Economic Geology 69, 631–645. Cosmochimica Acta 50, 1445–1454.
Morris, J.D., Leeman, W.P., Tera, F., 1990. The subducted component in island arc lavas: Plank, T., 2005. Constraints from thorium/lanthanum on sediment recycling at subduction
constraints from Be isotopes and B–Be systematics. Nature 344, 31–36. zones and the evolution of the continents. Journal of Petrology 46, 921–944.
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–liquid Rutherford, M.J., Devine, J.D., 1988. The May 18, 1980, eruption of Mount St. Helens. 3.
partitioning of metals in hydrothermal systems. Geology 33, 657–660. Stability and chemistry of amphibole in the magma chamber. Journal of
Pokrovski, G.S., Borisova, A.Y., Harrichoury, J.-C., 2008. The effect of sulfur on vapor– Geophysical Research 93 (11), 959 949-11.
liquid fractionation of metals in hydrothermal systems. Earth and Planetary Science Ryerson, F.J., Watson, E.B., 1987. Rutile saturation in magmas: implications for Ti–Nb–
Letters 266, 345–362. Ta depletion in island-arc basalts. Earth and Planetary Science Letters 86, 225–239.
Poli, S., Schmidt, M.W., 2002. Petrology of subducted slabs. Annual Reviews of Earth and Sajona, F.G., Maury, R.C., 1998. Association of adakites with gold and copper
Planetary Sciences 30, 207–235. mineralization in the Philippines. Comptes rendus de l'Académie des sciences,
Portnyagin, M., Hoernle, K., Plechov, P., Mironov, N., Khubunaya, S., 2007. Constraints on Série II, Sciences de la terre et des planètes 326, 27–34.
mantle melting and composition and nature of slab components in volcanic arcs Sakai, H., Matsubaya, O., 1977. Stable isotopic studies of Japanese geothermal systems.
from volatiles (H2O, S, Cl, F) and trace elements in melt inclusions from the Geothermics 5, 97–124.
Kamchatka Arc. Earth and Planetary Science Letters 255, 53–69. Sawkins, F.J., Scherkenbach, D.A., 1981. High copper content of fluid inclusions in quartz
Pudack, C., Halter, W.E., Heinrich, C.A., Pettke, T., 2009. Evolution of magmatic vapor to from northern Sonora: implications for ore-genesis theory. Geology 9, 37–40.
gold-rich epithermal liquid: the porphyry to epithermal transition at Nevados de Scambelluri, M., Müntener, O., Ottolini, L., Pettke, T.P., Vannuccie, R., 2004. The fate of B,
Famatina, Northwest Argentina. Economic Geology 104, 449–477. Cl and Li in the subducted oceanic mantle and in the antigorite breakdown fluids.
Rabbia, O.M., Hernández, L.B., King, R.W., López-Escobar, L., 2002. Discussion on “Giant Earth and Planetary Science Letters 222, 217–234.
versus small porphyry copper deposits of Cenozoic age in northern Chile: adakitic Scherbarth, N.L., Spry, P.G., 2006. Mineralogical, petrological, stable isotope, and fluid
versus normal calc-alkaline magmatism” by Oyarzun et al. (Mineralium Deposita inclusion characteristics of the Tuvatu Gold–Silver Telluride Deposit, Fiji:
36: 794–798, 2001). Mineralium Deposita 37, 791–794. comparisons with the Emperor Deposit. Economic Geology 101, 135–158.
Rapp, R.P., Watson, E.B., 1995. Dehydration melting of metabasalt at 8–32 kbar: Schiano, P., Clocchiatti, R., Shimizu, N., Maury, R.C., Jochum, K.P., Hofmann, A.W., 1995.
implications for continental growth and crust–mantle recycling. Journal of Hydrous, silica-rich melts in the sub-arc mantle and their relationship with erupted
Petrology 36, 891–931. arc lavas. Nature 277, 595–600.
Rapp, R.P., Watson, E.B., Miller, C.F., 1991. Partial melting of amphibolite/eclogite and Schmidt, M.W., Poli, S., 1998. Experimentally based water budgets for dehydrating slabs
the origin of Archean trondhjemites and tholeiites. Precambrian Research 51, and consequences for arc magma generation. Earth and Planetary Science Letters
1–25. 163, 361–379.
Redmond, P.B., Einaudi, M.T., Inan, E.E., Landtwing, M.R., Heinrich, C.A., 2004. Copper Schmidt, M.W., Dardon, A., Chazot, G., Vannucci, R., 2004. The dependence of Nb and Ta
deposition by fluid cooling in intrusion-centered systems: new insights from the rutile-melt partitioning on melt composition and Nb/Ta fractionation during
Bingham porphyry ore deposit, Utah. Geology 32, 217–220. subduction processes. Earth and Planetary Science Letters 226, 415–432.
Reeves, E.P., Seewald, J.S., Saccocia, P., Walsh, E., Bach, W., Craddock, P.R., Shanks, W.C., Schmidt, A., Weyer, S., John, T., Brey, G.P., 2009. HFSE systematics of rutile-bearing
Sylva, S.P., Pichler, T., Rosner, M., 2010. Geochemistry of hydrothermal fluids from eclogites: new insights into subduction zone processes and implications for the
the PACMANUS, Northeast Pual and Vienna Woods hydrothermal fields, Manus earth's HFSE budget. Geochimica et Cosmochimica Acta 73, 455–468.
Basin, Papua New Guinea. Geochimica et Cosmochimica Acta 75, 1088–1123. Seedorff, E., Dilles, J.H., Proffett Jr., J.M., Einaudi, M.T., Zurcher, L., Stavast, W.J.A.,
Richards, J.P., 1995. Alkalic-type epithermal gold deposits — a review. In: Thompson, J.F.H. Johnson, D.A., Barton, M.D., 2005. Porphyry deposits: characteristics and origin of
(Ed.), Magmas, Fluids, and Ore Deposits. Short Course Series, 23. Mineralogical hypogene features. Economic Geology 100th Anniversary Volume, pp. 251–298.
Association of Canada, pp. 367–400. ch. 17. Seo, J.H., Guillong, M., Heinrich, C.A., 2009. The role of sulfur in the formation of
Richards, J.P., 2002. Discussion of “Giant versus small porphyry copper deposits of magmatic–hydrothermal copper–gold deposits. Earth and Planetary Science
Cenozoic age in northern Chile: adakitic versus normal calc-alkaline magmatism” Letters 282, 323–328.
by Oyarzun et al. (Mineralium Deposita 36: 794–798, 2001). Mineralium Deposita Setterfield, T.N., Mussett, A.E., Oglethorpe, R.D.J., 1992. Magmatism and associated
37, 788–790. hydrothermal activity during the evolution of the Tavua caldera: 40Ar/39Ar dating
Richards, J.P., 2003. Tectono-magmatic precursors for porphyry Cu–(Mo–Au) deposit of volcanic, intrusive, and hydrothermal events. Economic Geology 87, 1130–1140.
formation. Economic Geology 96, 1515–1533. Shafiei, B., Haschke, M., Shahabpour, J., 2009. Recycling of orogenic arc crust triggers
Richards, J.P., 2005. Cumulative factors in the generation of giant calc-alkaline porphyry porphyry Cu mineralization in Kerman Cenozoic arc rocks, southeastern Iran.
Cu deposits. In: Porter, T.M. (Ed.), Super Porphyry Copper and Gold Deposits: A Mineralium Deposita 44, 265–283.
Global Perspective, 1. Porter Geoscience Consulting Publishing, Linden Park, South Shinohara, H., 1994. Exsolution of immiscible vapor and liquid phases from a
Australia, pp. 7–25. crystallizing silicate melt: implications for chlorine and metal transport. Geochi-
Richards, J.P., 2009. Postsubduction porphyry Cu–Au and epithermal Au deposits: mica et Cosmochimica Acta 58, 5215–5221.
products of remelting of subduction-modified lithosphere. Geology 37, 247–250. Shinohara, H., Hedenquist, J.W., 1997. Constraints on magma degassing beneath the Far
Richards, J.P., Kerrich, R., 1993. The Porgera gold mine, Papua New Guinea: magmatic– Southeast porphyry Cu–Au deposit, Philippines. Journal of Petrology 38, 1741–1752.
hydrothermal to epithermal evolution of an alkalic-type precious metal deposit. Shinohara, H., Kazahaya, K., Lowenstern, J.B., 1995. Volatile transport in a convecting
Economic Geology 88, 1017–1052. magma column: implications for porphyry Mo mineralization. Geology 23,
Richards, J.P., Kerrich, R., 2007. Adakite-like rocks: their diverse origins and 1091–1094.
questionable role in metallogenesis. Economic Geology 102, 537–576. Sillitoe, R.H., 1973. The tops and bottoms of porphyry copper deposits. Economic
Richards, J.P., Chappell, B.W., McCulloch, M.T., 1990. Intraplate-type magmatism in a Geology 68, 799–815.
continent–island-arc collision zone: Porgera intrusive complex, Papua New Guinea. Sillitoe, R.H., 2000. Gold-rich porphyry deposits: descriptive and genetic models and
Geology 18, 958–961. their role in exploration and discovery. Reviews in Economic Geology 13,
Richards, J.P., Ullrich, T., Kerrich, R., 2006a. The Late Miocene–Quaternary Antofalla 315–345.
volcanic complex, southern Puna, NW Argentina: protracted history, diverse Sillitoe, R.H., 2010. Porphyry copper systems. Economic Geology 105, 3–41.
petrology, and economic potential. Journal of Volcanology and Geothermal Sillitoe, R.H., Hart, S.R., 1984. Lead-isotope signatures of porphyry copper deposits in
Research 152, 197–239. oceanic and continental settings, Colombian Andes. Geochimica et Cosmochimica
Richards, J.P., Wilkinson, D., Ullrich, T., 2006b. Geology of the Sari Gunay epithermal Acta 48, 2135–2142.
gold deposit, northwest Iran. Economic Geology 101, 1455–1496. Simmons, S.F., Brown, K.L., 2007. The flux of gold and related metals through a volcanic
Righter, K., Campbell, A.J., Humayun, M., Hervig, R.L., 2004. Partitioning of Ru, Rh, Pd, Re, arc, Taupo Volcanic Zone, New Zealand. Geology 35, 1099–1102.
Ir, and Au between Cr-bearing spinel, olivine, pyroxene and silicate melts. Simon, A.C., Frank, M.R., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2005. Gold
Geochimica et Cosmochimica Acta 68, 867–880. partitioning in melt–vapor–brine systems. Geochimica et Cosmochimica Acta 69,
Ronacher, E., Richards, J.P., Reed, M.H., Bray, C.J., Spooner, E.T.C., Adams, P.D., 2004. 3321–3335.
Characteristics and evolution of the hydrothermal fluid in the North Zone high- Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2006. Copper
grade area, Porgera gold deposit, Papua New Guinea. Economic Geology 99, partitioning in a melt–vapor–brine–magnetite–pyrrhotite assemblage. Geochimica
843–867. et Cosmochimica Acta 70, 5583–5600.
Rowe, M.C., Kent, A.J.R., Nielsen, R.L., 2009. Subduction influence on oxygen fugacity Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2007. The partitioning
and trace and volatile elements in basalts across the Cascade Volcanic Arc. Journal behavior of As and Au in S-free and S-bearing magmatic assemblages. Geochimica
of Petrology 50, 61–91. et Cosmochimica Acta 71, 1764–1782.
Rubin, A.M., 1995. Propagation of magma-filled cracks. Annual Review of Earth & Simon, A.C., Candela, P.A., Piccoli, P.M., Mengason, M., Englander, L., 2008. The effect of
Planetary Sciences 23, 287–336. crystal-melt partitioning on the budgets of Cu, Au, and Ag. American Mineralogist
Rudnick, R.L., Gao, S., 2003. Composition of the continental crust. In: Rudnick, R.L. (Ed.), 93, 1437–1448.
Treatise on Geochemistry 3: The Crust. Elsevier, Amsterdam, pp. 1–64. Sinclair, W.D., 2007. Porphyry deposits. In: Goodfellow, W.D. (Ed.), Mineral Deposits of
Rushmer, T., 1991. Partial melting of two amphibolites: contrasting experimental Canada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of
results under fluid-absent conditions. Contributions to Mineralogy and Petrology Geological Provinces, and Exploration Methods. Geological Association of Canada,
107, 41–59. St. John's, Newfoundland, pp. 223–243.
Rushmer, T., 1993. Experimental high-pressure granulites: some applications to natural Smith, I.E.M., Stewart, R.B., Price, R.C., Worthington, T.J., 2010. Are arc-type rocks the
mafic xenolith suites and Archean granulite terranes. Geology 21, 411–414. products of magma crystallisation? Observations from a simple oceanic arc
Rusk, B.G., Reed, M.H., Dilles, J.H., Klemm, L.M., Heinrich, C.A., 2004. Compositions of volcano: Raoul Island, Kermadec Arc, SW Pacific. Journal of Volcanology and
magmatic hydrothermal fluids determined by LA-ICP-MS of fluid inclusions from Geothermal Research 190, 219–234.
the porphyry copper–molybdenum deposit at Butte, MT. Chemical Geology 210, Sobolev, A., Chaussidon, M., 1996. H2O concentrations in primary melts from supra-
173–199. subduction zones and mid-ocean ridges: implications for H2O storage and recycling
Rusk, B.G., Reed, M.H., Dilles, J.H., 2008. Fluid inclusion evidence for magmatic– in the mantle. Earth and Planetary Science Letters 137, 45–55.
hydrothermal fluid evolution in the porphyry copper–molybdenum deposit at Solomon, M., 1990. Subduction, arc reversal, and the origin of porphyry copper–gold
Butte, Montana. Economic Geology 103, 307–334. deposits in island arcs. Geology 18, 630–633.
J.P. Richards / Ore Geology Reviews 40 (2011) 1–26 25

Sourirajan, S., Kennedy, G.C., 1962. The system H2O–NaCl at elevated temperatures and California and Median batholith of New Zealand. Special Paper, 374. Geological
pressures. American Journal of Science 260, 115–141. Society of America, pp. 1–21.
Spooner, E.T.C., 1993. Magmatic sulphide/volatile interaction as a mechanism for Ulmer, P., Trommsdorff, V., 1995. Serpentine stability to mantle depths and subduction-
producing chalcophile element enriched, Archean Au–quartz, epithermal Au–Ag related magmatism. Science 268, 858–861.
and Au skarn hydrothermal ore fluids. Ore Geology Reviews 7, 359–379. Ulrich, T., Günther, D., Heinrich, C.A., 2001. The evolution of a porphyry Cu–Au deposit,
Staudigel, H., Plank, T., White, B., Schmincke, H.-U., 1996. Geochemical fluxes during seafloor based on LA-ICP-MS analysis of fluid inclusions: Bajo de la Alumbrera, Argentina.
alteration of the basaltic upper oceanic crust: DSDP sites 417 and 418. In: Bebout, G.E., Economic Geology 96, 1743–1774.
Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction Top to Bottom. Geophysical van Dongen, M., Weinberg, R.F., Tomkins, A.G., Armstrong, R.A., Woodhead, J.D.,
Monograph, 96. American Geophysical Union, Washington, DC, pp. 19–38. 2010. Recycling of Proterozoic crust in Pleistocene juvenile magma and rapid
Stavast, W.J.A., Keith, J.D., Christiansen, E.H., Dorais, M.J., Tingey, D., 2006. The fate of formation of the Ok Tedi porphyry Cu–Au deposit, Papua New Guinea. Lithos
magmatic sulfides during intrusion or eruption, Bingham and Tintic districts, Utah. 114, 282–292.
Economic Geology 101, 329–345. Vry, V.H., Wilkinson, J.J., Seguel, J., Millan, J., 2010. Multistage intrusion, brecciation, and
Stein, H.J., 1988. Genetic traits of climax-type granites and molybdenum mineraliza- veining at El Teniente, Chile: evolution of a nested porphyry system. Economic
tion, Colorado Mineral Belt. In: Taylor, R.P., Strong, D.F. (Eds.), Recent Advances in Geology 105, 119–153.
the Geology of Granite-Related Mineral Deposits, Special Volume 39. Canadian Walker, J.A., Patino, L.C., Carr, M.J., Feigenson, M.D., 2001. Slab control over HFSE
Institute of Mining and Metallurgy, pp. 394–401. depletions in central Nicaragua. Earth and Planetary Science Letters 192, 533–543.
Stern, R.J., Kohut, E., Bloomer, S.H., Leybourne, M., Fouch, M., Vervoort, J., 2006. Wallace, P.J., 2005. Volatiles in subduction zone magmas: concentrations and fluxes
Subduction factory processes beneath the Guguan cross-chain, Mariana Arc: no role based on melt inclusion and volcanic gas data. Journal of Volcanology and
for sediments, are serpentinites important? Contributions to Mineralogy and Geothermal Research 140, 217–240.
Petrology 151, 202–221. Walshe, J.L., Solomon, M., Whitford, D.J., Sun, S.-S., Foden, J.D., 2011. The role of the
Stoffregen, R., 1987. Genesis of acid-sulfate alteration and Au–Cu–Ag mineralization at mantle in the genesis of tin deposits and tin provinces of eastern Australia.
Summitville, Colorado. Economic Geology 82, 1575–1591. Economic Geology 106, 297–305.
Stolper, E., Newman, S., 1994. The role of water in the petrogenesis of Mariana trough Wang, J., Hattori, K.H., Kilian, R., Stern, C.R., 2007a. Metasomatism of sub-arc mantle
magmas. Earth and Planetary Science Letters 121, 293–325. peridotites below southernmost South America: reduction of fO2 by slab-melt.
Sun, S.-s., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts: Contributions to Mineralogy and Petrology 153, 607–624.
implications for mantle composition and processes. In: Saunders, A.D., Norry, M.J. Wang, Q., Wyman, D.A., Xu, J.-F., Zhao, Z.-H., Jian, P., Zi, F., 2007b. Partial melting of
(Eds.), Magmatism in the Ocean Basins: Geological Society of London Special thickened or delaminated lower crust in the middle of Eastern China: implications
Publication, No. 42, pp. 313–345. for Cu–Au mineralization. Journal of Geology 115, 149–161.
Sun, W., Bennett, V.C., Kamenetsky, V.S., 2004a. The mechanism of Re enrichment in arc Webster, J.D., 1992. Water solubility and chlorine partitioning in Cl-rich granitic
magmas: evidence from Lau Basin basaltic glasses and primitive melt inclusions. systems: effects of melt composition at 2 kbar and 800 °C. Geochimica et
Earth and Planetary Science Letters 222, 101–114. Cosmochimica Acta 56, 679–687.
Sun, W.D., Arculus, R.J., Kamenetsky, V.S., Binns, R.A., 2004b. Release of gold-bearing Westra, G., Keith, S.B., 1981. Classification and genesis of stockwork molybdenum
fluids in convergent margin magmas prompted by magnetite crystallization. deposits. Economic Geology 76, 844–873.
Nature 431, 975–978. White, D.E., Muffler, L.J.P., Truesdell, A.H., 1971. Vapor-dominated hydrothermal
Symonds, R.B., Rose, W.I., Reed, M.H., Lichte, F.E., Finnegan, D.L., 1987. Volatilization, systems compared with hot-water systems. Economic Geology 66, 75–97.
transport and sublimation of metallic and non-metallic elements in high White, W.H., Bookstrom, A.A., Kamilli, R.J., Gangster, M.W., Smith, R.P., Ranta, D.E.,
temperature gases at Merapi Volcano, Indonesia. Geochimica et Cosmochimica Steininger, R.C., 1981. Character and origin of climax-type molybdenum deposits.
Acta 51, 2083–2101. Economic Geology 75th Anniversary volume, pp. 270–316.
Tarana, Y.A., Hedenquist, J.W., Korzhinsky, M.A., Tkachenko, S.I., Shmulovich, K.I., 1995. Widom, E., Kepezhinskas, P., Defant, M., 2003. The nature of metasomatism in the sub-
Geochemistry of magmatic gases from Kudryavy volcano, Iturup, Kuril Islands. arc mantle wedge: evidence from Re–Os isotopes in Kamchatka peridotite
Geochimica et Cosmochimica Acta 59, 1749–1761. xenoliths. Chemical Geology 196, 283–306.
Tatsumi, Y., 1986. Formation of the volcanic front in subduction zones. Geophysical Williams, T.J., Candela, P.A., Piccoli, P.M., 1995. The partitioning of copper between
Research Letters 17, 717–720. silicate melts and two-phase aqueous fluids: an experimental investigation at
Tatsumi, Y., 2003. Some constraints on arc magma genesis. In: Eiler, J. (Ed.), Inside the 1 kbar, 800 °C and 0.5 kbar, 850 °C. Contributions to Mineralogy and Petrology 121,
Subduction Factory. : Geophysical Monograph, 138. American Geophysical Union, 388–399.
Washington, DC, pp. 277–292. Williams-Jones, A.E., Heinrich, C.A., 2005. Vapor transport of metals and the formation
Tatsumi, Y., Eggins, S., 1995. Subduction Zone Magmatism. Blackwell, Oxford. 213 pp. of magmatic–hydrothermal ore deposits. Economic Geology 100, 1287–1312.
Tatsumi, Y., Hamilton, D.L., Nesbitt, R.W., 1986. Chemical characteristics of fluid phase Williamson, B.J., Müller, A., Shail, R.K., 2010. Source and partitioning of B and Sn in the
released from a subducted lithosphere and the origin of arc magmas: evidence from Cornubian batholith of southwest England. Ore Geology Reviews 38, 1–8.
high pressure experiments and natural rocks. Journal of Volcanology and Winter, J.D., 2001. An introduction to igneous and metamorphic petrology: Upper
Geothermal Research 29, 293–309. Saddle River. Prentice-Hall Inc., New Jersey. 697 p.
Taylor, S.R., McLennan, S.M., 1985. The continental crust: its composition and evolution: Wolf, M.B., Wyllie, P.J., 1994. Dehydration-melting of amphibolite at 10 kbar — the
an examination of the geochemical record preserved in sedimentary rocks. effects of temperature and time. Contributions to Mineralogy and Petrology 115,
Blackwell Scientific 312p. 369–383.
Tessalina, S.G., Yudovskaya, M.A., Chaplygin, I.V., Birck, J.-L., Capmas, F., 2008. Sources of Wood, S.A., Spera, F.J., 1984. Adiabatic decompression of aqueous solutions:
unique rhenium enrichment in fumaroles and sulphides at Kudryavy volcano. applications to hydrothermal fluid migration in the crust. Geology 12, 707–710.
Geochimica et Cosmochimica Acta 72, 889–909. Wörner, G., Moorbath, S., Harmon, R.S., 1992. Andean Cenozoic volcanic centers reflect
Thiéblemont, D., Stein, G., Lescuyer, J.-L., 1997. Gisements épithermaux et porphyr- basement isotopic domains. Geology 20, 1103–1106.
iques: la connexion adakite. C.R. Acad. Sci. Paris, Sciences de la terre et des planètes/ Wyborn, D., Sun, S.-s., 1994. Sulphur-undersaturated magmatism — a key factor for
Earth and Planetary Sciences 325, 103–109. generating magma-related copper–gold deposits. AGSO Research Newsletter 21,
Thirlwall, M.F., Graham, A.M., Arculus, R.J., Harmon, R.S., Macpherson, C.G., 1996. 7–8.
Resolution of the effects of crustal assimilation, sediment subduction, and fluid Wyllie, P.J., Huang, W.-L., Stern, C.R., Maaløe, S., 1976. Granitic magmas: possible and
transport in island arc magmas: Pb–Sr–Nd–O isotope geochemistry of Grenada, impossible sources, water contents, and crystallization sequences. Canadian
Lesser Antilles. Geochimica et Cosmochimica Acta 60, 4785–4810. Journal of Earth Sciences 13, 1007–1019.
Thompson, T.B., Trippel, A.D., Dwelley, P.C., 1985. Mineralized veins and breccias of the Wysoczanski, R.J., Wright, I.C., Gamble, J.A., Hauri, E.H., Luhr, J.F., Eggins, S.M., Handler,
Cripple Creek District, Colorado. Economic Geology 80, 1669–1688. M.R., 2006. Volatile contents of Kermadec Arc–Havre Trough pillow glasses:
Thorkelson, D.J., Breitsprecher, K., 2005. Partial melting of slab window margins: fingerprinting slab-derived aqueous fluids in the mantle sources of arc and back-
genesis of adakitic and non-adakitic magmas. Lithos 79, 25–41. arc lavas. Journal of Volcanology and Geothermal Research 152, 51–73.
Tiepolo, M., Tribuzio, R., 2008. Petrology and U–Pb zircon geochronology of amphibole- Xiao, Z., Gammons, C.H., Williams-Jones, A.E., 1998. Experimental study of copper(I)
rich cumulates with sanukitic affinity from Husky Ridge (Northern Victoria Land, chloride complexing in hydrothermal solutions at 40 to 300 °C and saturated water
Antarctica): insights into the role of amphibole in the petrogenesis of subduction- vapor pressure. Geochimica et Cosmochimica Acta 62, 2949–2964.
related magmas. Journal of Petrology 49, 937–970. Yang, Z., Hou, Z., White, N.C., Chang, Z., Li, Z., Song, Y., 2009. Geology of the post-
Tilton, G.R., Pollak, R.J., Clark, A.H., Robertson, R.C.R., 1981. Isotopic composition of Pb in collisional porphyry copper–molybdenum deposit at Qulong, Tibet. Ore Geology
central Andean ore deposits. Geological Society of America, Memoir 154, 791–816. Reviews 36, 133–159.
Titley, S.R., 1987. The crustal heritage of silver and gold ratios in Arizona ores. Yogodzinski, G.M., Kay, R.W., Volynets, O.N., Koloskov, A.V., Kay, S.M., 1995. Magnesian
Geological Society of America, Bulletin 99, 814–826. andesite in the western Aleutian Komandorsky region: implications for slab
Titley, S.R., 2001. Crustal affinities of metallogenesis in the American Southwest. melting and processes in the mantle wedge. Geological Society of America Bulletin
Economic Geology 96, 1323–1342. 107, 505–519.
Tomkins, A.G., Mavrogenes, J.A., 2003. Generation of metal-rich felsic magmas during Yogodzinski, G.M., Lees, J.M., Churikova, T.G., Dorendorf, F., Wöerner, G., Volynets, O.N.,
crustal anatexis. Geology 31, 765–768. 2001. Geochemical evidence for the melting of subducting oceanic lithosphere at
Tosdal, R.M., Richards, J.P., 2001. Magmatic and structural controls on the development plate edges. Nature 409, 500–504.
of porphyry Cu ± Mo ± Au deposits. In: Richards, J.P., Tosdal, R.M. (Eds.), Structural Zajacz, Z., Halter, W., 2009. Copper transport by high temperature, sulfur-rich magmatic
Controls on Ore Genesis. : Reviews in Economic Geology, 14. Society of Economic vapor: evidence from silicate melt and vapor inclusions in a basaltic andesite from the
Geologists, pp. 157–181. Villarrica volcano (Chile). Earth and Planetary Science Letters 282, 115–121.
Tulloch, A.J., Kimbrough, D.L., 2003. Paired plutonic belts in convergent margins and the Zajacz, Z., Halter, W.E., Pettke, T., Guillong, M., 2008. Determination of fluid/melt
development of high Sr/Y magmatism: Peninsular Ranges batholith of Baja– partition coefficients by LA-ICPMS analysis of co-existing fluid and silicate melt
26 J.P. Richards / Ore Geology Reviews 40 (2011) 1–26

inclusions: controls on element partitioning. Geochimica et Cosmochimica Acta 72, Zhang, X., Spry, P.G., 1994. Petrological, mineralogical, fluid inclusion, and stable
2169–2197. isotope studies of the Gies gold–silver telluride deposit, Judith Mountains,
Zajacz, Z., Seo, Z.H., Candela, P.A., Piccoli, P.M., Heinrich, C.A., Guillong, M., 2010. Alkali Montana. Economic Geology 89, 602–627.
metals control the release of gold from volatile-rich magmas. Earth and Planetary Zimmer, M.M., Plank, T., Hauri, E.H., Yogodzinski, G.M., Stelling, P., Larsen, J., Singer, B.,
Science Letters 297, 50–56. Jicha, B., Mandeville, C., Nye, C.J., 2010. The role of water in generating the calc-
Zajacz, Z., Seo, J.H., Candela, P.A., Piccoli, P.M., Tossell, J.A., 2011. The solubility of copper in alkaline trend: new volatile data for Aleutian magmas and a new Tholeiitic Index.
high-temperature magmatic vapors: a quest for the significance of various chloride Journal of Petrology 51, 2411–2444.
and sulfide complexes. Geochimica et Cosmochimica Acta 75, 2811–2827.

You might also like