Sei sulla pagina 1di 26

Review

Magmatic to hydrothermal metal uxes in convergent and collided margins


Jeremy P. Richards
Dept. of Earth and Atmospheric Sciences, University of Alberta, Edmonton, Alberta, Canada T6G 2E3
a b s t r a c t a r t i c l e i n f o
Article history:
Received 1 December 2010
Received in revised form 18 May 2011
Accepted 19 May 2011
Available online 27 May 2011
Keywords:
Porphyry deposit
Epithermal deposit
Subduction
Post-subduction
Magmatichydrothermal uid
Ore formation
Metals such as Cu, Mo, Au, Sn, and W in porphyry and related epithermal mineral deposits are derived
predominantly fromthe associated magmas, via magmatichydrothermal uids exsolved upon emplacement into
the mid- toupper crust. Four mainsources exist for magmas, andtherefore metals, inconvergent andcollidedplate
margins: the subducting oceanic plate basaltic crust, subducted seaoor sediments, the asthenospheric mantle
wedge between the subducting and overriding plates, and the upper plate lithosphere. This paper rstly examines
the source of normal arc magmas, and concludes that they are predominantly derived from partial melting of the
metasomatized mantle wedge, with possible minor contributions from subducted sediments. Although some
metals may be transferred from the subducting slab via dehydration uids, the bulk of the metals in the resultant
magmas are probably derived from the asthenospheric mantle. The most important contributions from the slab
fromametallogenic perspectiveareH
2
O, S, andCl, as well as oxidants. Partial meltingof thesubductedoceanic crust
and/or sediments may occur under some restricted conditions, but is unlikely to be a widespread process (in
Phanerozoic arcs), and does not signicantly differ metallogenically from slab-dehydration processes.
Primary, mantle-derived arc magmas are basaltic, but differ from mid-ocean ridge basalt in having higher water
contents (~10 higher), oxidation states (~2 log f
O2
units higher), and concentrations of incompatible elements
and other volatiles (e.g., S and Cl). Concentrations of chalcophile and siderophile metals in these partial melts
depend critically on the presence and abundance of residual sulde phases in the mantle source. At relatively
high abundances of suldes thought to be typical of active arcs where f
S2
and f
O2
are high (magma/sulde
ratio=10
2
10
5
), sparse, highly siderophile elements such as Au and PGE will be retained in the source, but
magmas may be relatively undepleted in abundant, moderately chalcophile elements such as Cu (and perhaps
Mo). Such magmas have the potential to formporphyry CuMo deposits upon emplacement inthe upper crust.
Gold-rich porphyry deposits would only form where residual sulde abundance was very low (magma/sulde
ratio N10
5
), perhaps due to unusually high mantle wedge oxidation states.
In contrast, some porphyry Mo and all porphyry SnW deposits are associated with felsic granitoids, derived
primarily from melting of continental crust during intra-plate rifting events. Nevertheless, mantle-derived
magmas may have a role to play as a heat source for anatexis and possibly as a source of volatiles and metals.
In post-subduction tectonic settings Tulloch and Kimbrough, 2003, such as subduction reversal or migration, arc
collision, continentcontinent collision, and post-collisional rifting, a subducting slab source no longer exists,
and magmas are predominantly derived frompartial melting of the upper plate lithosphere. This lithosphere will
have undergone signicant modication during the previous subduction cycle, most importantly with the
introduction of large volumes of hydrous, mac (amphibolitic) cumulates residual from lower crustal
differentiation of arc basalts. Small amounts of chalcophile and siderophile element-rich suldes may also be left
in these cumulates. Partial melting of these subduction-modied sources due to post-subduction thermal
readjustments or asthenospheric melt invasion will generate small volumes of calc-alkaline to mildly alkaline
magmas, which may redissolve residual suldes. Such magmas have the potential to form Au-rich as well as
normal CuMo porphyry and epithermal Au systems, depending on the amounts of sulde present inthe lower
crustal source. Alkalic-type epithermal Au deposits are an extreme end-member of this range of post-subduction
deposits, formed from subduction-modied mantle sources in extensional or transtensional environments.
Ore formation in porphyry and related epithermal environments is critically dependent on the partitioning of
metals from the magma into an exsolving magmatichydrothermal uid phase. This process occurs most
efciently at depths greater than ~6 km, within large mid- to upper crustal batholithic complexes fed by arc or
post-subduction magmas. Under such conditions, metals will partition efciently into a single-phase,
supercritical aqueous uid (~213 wt.% NaCl equivalent), which may exist as a separate volatile plume or as
bubbles entrained in buoyant magma. Focusing of upward ow of bubbly magma and/or uid into the apical
regions of the batholithic complex forms cupolas, which represent high mass- and heat-ux channelways
Ore Geology Reviews 40 (2011) 126
Tel.: +1 780 492 3430; fax: +1 780 492 2030.
E-mail address: Jeremy.Richards@ualberta.ca.
0169-1368/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.oregeorev.2011.05.006
Contents lists available at ScienceDirect
Ore Geology Reviews
j our nal homepage: www. el sevi er. com/ l ocat e/ or egeor ev
towards the surface. Cupolas may be self-organizing to the extent that once formed, further magma and uid
ow 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 extendto surface, where magmas and uids vent as volcanic products and fumaroles. Between
the surface and the underlying magma chamber, a very steep thermal gradient exists (700800 C over b5 km),
which is the primary cause of vertical focusing of ore mineral deposition. The bulk of metals (CuMoAu) that
forms porphyry ore bodies are precipitatedover a narrowtemperature interval between~425 and320 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) SO
2
dissolved in the magmatichydrothermal uid
phase disproportionates to H
2
S and H
2
SO
4
, leading to sulde and sulfate mineral deposition and the onset of
increasingly acidic alteration.
The bulk of the metal ux into the porphyry environment may be carried by moderately saline supercritical
uids 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 signicant concentrations of sulfur species, however, and bisulde complexing
of Cu and Au may enable their continued transport into the epithermal regime. In the high-suldation
epithermal environment, intense acidic (advanced-argillic) alteration is caused by the ux of highly acidic
magmatic volatiles (H
2
SO
4
, 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 uids, where observed, are low- to
moderate-salinity liquids, and are thought to represent later-stage magmatichydrothermal uids that have
ascended along shallower (cooler) geothermal gradients that either do not, or barely, intersect the liquidvapor
solvus. Such uids contract from the original supercritical uid or vapor to the liquid phase. Brief intersection
of the liquidvapor solvus may be important in shedding excess chloride and chloride-complexed metals (such
as Fe), so that bisulde-complexed metals remain in solution. Such a restrictive pressuretemperature path is
likely to occur only transiently during the evolutionof a magmatichydrothermal system, which may explain the
rarity of high-suldation CuAu 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-modied 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 uid exsolution in the upper crust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
3.1. Partitioning of metals from magma into exsolving hydrothermal uid. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
4. Magmatichydrothermal ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.1. Porphyry Cu ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
4.2. Epithermal CuAu ore formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2.1. High-suldation epithermal CuAu deposits. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4.2.2. Low-suldation 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
2 J.P. Richards / Ore Geology Reviews 40 (2011) 126
1. Introduction
The question of the source of various elements in convergent and
collided margin magmas has challenged geologists for decades. Igneous
petrologists seek to understand the petrogenesis of such magmas
through geochemical and isotopic tracing, whereas economic geologists
are generally more interested in the source of potentially valuable
elements suchas Cu, Mo, Sn, W, Au, andplatinumgroupelements (PGE),
which may ultimately be found in intrusion-related hydrothermal
deposits.
Igneous petrologists are broadly in agreement that arc magmas are
primarily derived from hydrous melting of the asthenospheric mantle
wedge above subducting plates, but melts fromthe subducted oceanic
crust (including sediments) and the upper plate lithosphere may also
be involved to varying degrees.
Economic geologists are also broadly in agreement that ore-forming
elements are partitioned from such magmas into an exsolving volatile
phase upon emplacement in the upper crust, and may then be
precipitated fromthese uids during cooling, uid mixing, and wallrock
reaction processes in porphyry-type and related epithermal mineral
deposits. However, these process theories do not address where the
metals originally came from, nor why porphyry deposits vary so widely
in their metal contents (from Au-rich, through CuMoAu, to Mo-
only deposits, with SnW deposits forming a distinct variant).
In addition to subduction-related calc-alkaline magmas, a diverse
suite of calc-alkaline to alkaline magmas is generated in post-
subduction and collisional tectonic settings, and these magmatic
systems may also generate porphyry and epithermal ore deposits.
Such systems raise an additional set of petrogenetic and metallogenic
questions.
It is the intent of this paper to merge these different geological
perspectives on magmagenesis and metallogeny in order to discuss
primary metal uxes in convergent and collisional margins in terms of
igneous petrogenetic and magmatichydrothermal processes. The
ultimate metal inventory and metal ratios in any given porphyry or
related deposit is secondarily controlled by late-stage magmatic and
shallowcrustal processes. These processes are examined, closing with a
review of uid and metal sources and behavior in related epithermal
environments.
2. Magma generation in convergent and collided margins:
geochemical characteristics and partitioning of metals
Most magmas erupted through or emplaced within the Earth's crust
are not primary magmas (in the sense of being chemically unmodied
since extraction fromtheir source), and most are not even primitive (in
the sense of being relatively unevolved; Hildreth and Moorbath, 1988;
Leeman, 1983; Neuendorf et al., 2005; Smith et al., 2010; Thirlwall et al.,
1996). Except for magmas producedanderuptedinextensional tectonic
regimes (where rapid ascent to the surface is facilitated by normal
faulting), most deeply-derived magmas undergo some degree of
fractionation and crustal contamination during their passage towards
the surface. It is therefore challenging to isolate geochemical and
isotopic characteristics of magma source regions fromthe effects of later
processes (Davidson, 1996). Magmas erupted through mature conti-
nental crust are the most difcult to ngerprint uniquely in terms of
source characteristics because wallrock assimilation and fractional
crystallization (AFC; DePaolo, 1981) are ubiquitous and commonly
extensive processes that will signicantly modifybulk rockgeochemical
and isotopic compositions; and yet, these are also the magmas that are
most commonly associated with porphyry- and epithermal-type
mineral deposits. The difculty in constraining source characteristics
in such magmas is perhaps responsible for the plethora of theories that
have beenproposedfor the originof ore-formingmagmas inconvergent
margin settings, ranging from the melting of subducting oceanic crust
and/or seaoor sediments, through melting of subduction metasoma-
tized asthenospheric or lithospheric mantle, to melting of underplated
or primitive lower crustal rocks, and even melting of evolved crustal
rocks in the case of some felsic porphyry Mo and SnW magmas.
Therefore, rather than start by trying to identify a unique source
for the typical intermediate-to-felsic calc-alkaline magmas that are
associated with ore deposits in mature convergent margins, I begin
this review by focusing on the much better constrained primitive
island arc environment, where the effects of fractionation and crustal
contamination, particularly by continentally derived materials, are
minimized, and processes in mantle source regions can be more
clearly dened.
2.1. Slab dehydration and asthenospheric melting in subduction zones
There is a general consensus that, with the exception of young
oceanic lithosphere (b25 m.y.-old; Defant and Drummond, 1990) or
plate edges (Yogodzinski et al., 2001), basaltic oceanic crust undergoes
low-temperature, high pressure metamorphism upon subduction,
whichreleases uids througha series of prograde dehydrationreactions
to form anhydrous eclogite (Fig. 1). Water and other volatile
components and solutes (including S and Cl) were originally incorpo-
rated into oceanic crustal and upper mantle rocks during oxidizing
seaoor alteration, generating hydrous minerals suchas serpentine, talc,
amphibole, micas, chlorite, zoisite, chloritoid, and lawsonite. Various
experimental studies have shown that these minerals undergo
dehydration reactions over a depth range extending to ~100 km,
corresponding to the blueschisteclogite transition in crustal rocks;
serpentine (antigorite) and the 10 equivalent of chlorite may extend
this range to200 km(e.g., Dvir et al., 2011; Forneris andHolloway, 2003;
Fumagalli and Poli, 2005; Poli and Schmidt, 2002; Schmidt and Poli,
1998; Ulmer and Trommsdorff, 1995). Below these depths, the
anhydrous eclogitic crust is essentially infusible, and the dense slab
continues its descent into the mantle without melting. See reviews of
this subject by Richards (2003, 2005, and references therein).
Numerous studies have explored the character of the uids that are
releasedfromthe dehydrating slab, because they are thought to account
for the unique geochemical character of subduction-related magmas
during later partial melting in the metasomatized asthenospheric
mantle wedge (located between the downgoing slab and the upper
plate). Slab-derived uids are thought to be water-rich at these depths,
and to carry signicant amounts of other volatile components suchas Cl
and S. For example, salinities in the range of 410 wt.% NaCl have been
inferred from primitive basalt or melt inclusion studies (de Hoog et al.,
2001; Kent et al., 2002; Portnyagin et al., 2007; Wallace, 2005;
Wysoczanski et al., 2006), and salinities of 0.42 wt.% NaCl equivalent
were measured in uid inclusions from high-pressure rocks thought to
represent subductedoceanic mantle (Scambelluri et al., 2004). At higher
pressures (~6 GPa) and greater depths (~175 km) there may no longer
be a physical distinction between solute-rich aqueous liquids and
hydrous silicate melts, and the uid may be supercritical in nature (e.g.,
Kessel et al., 2005a,b), but the role of such deeply released uids in
subduction zone magmatism is unclear (see discussion in Richards and
Kerrich, 2007).
In addition to volatiles, water-soluble large ion lithophile elements
(LILE: K, Rb, Cs, Ca, Sr, andBa, andU), andB, Pb, As, andSb(Breeding et al.,
2004; Hattori et al., 2005; Hattori and Guillot, 2003; Kogiso et al., 1997;
Manning, 2004; Tatsumi et al., 1986) are thought to be mobilizedinto the
forearc mantle wedge by dehydration uids, which may then be
convected by corner-ow into the sub-arc melting zone (Fig. 1). These
uidmobile components are also characteristically enriched in arc
magmas (e.g., typical ranges of: 13 wt.% H
2
O, 5002000 ppm Cl, 900
2500 ppm S; Davidson, 1996; Gill, 1981; Noll et al., 1996; Sobolev and
Chaussidon, 1996; Wallace, 2005; Portnyaginet al., 2007), whichis taken
as evidence of aqueous uid metasomatismof the mantle wedge magma
source. Silica may also be signicantly mobilized in these slab uids
(Aerts et al., 2010; Manning, 2004), as well as Tl and Cu (Noll et al., 1996;
3 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Stolper and Newman, 1994). The normally relatively incompatible high
eld strength elements (HFSE) Ti, Nb, and Ta are not signicantly uid
soluble under subduction zone conditions, and are retained in minerals
such as rutile either in the slab or the mantle wedge (Audtat and
Keppler, 2005; Brenan et al., 1994; Green and Adam, 2003). Arc magmas
derived from these sources therefore show characteristic negative
anomalies for these three elements on mantle-normalized spider
diagrams, but display enrichments in most other incompatible elements
(Foley et al., 2000; Gill, 1981; Klemme et al., 2005; Ryerson and Watson,
1987; Schmidt et al., 2004; Schmidt et al., 2009).
Hydrous metasomatism of the mid-ocean-ridge basalt (MORB)-
depleted asthenospheric mantle wedge causes partial melting by
lowering the solidus of peridotite (Arculus, 1994; Kushiro et al., 1968;
Stolper andNewman, 1994). This occurs either throughdirect inltration
metasomatismby slab-derived uids percolating into the hot inner zone
of themantle wedge (Bourdonet al., 2003; Groveet al., 2006; Kelleyet al.,
2010; Peacock, 1993), or by convective corner-ow mixing of metaso-
matized peridotite into these hotter central regions (Fig. 1; Schmidt and
Poli, 1998; Tatsumi, 1986; Wysoczanski et al., 2006).
Partial melting of hydrated peridotite under these conditions in
the mantle wedge generates high-Mg basalts (Greene et al., 2006;
Pichavant et al., 2002; Smith et al., 2010). Such arc basalts are
distinguished from MORB by higher contents of incompatible
elements (as noted above) and water (up to 6 wt.% H
2
O; Cervantes
and Wallace, 2003; Grove et al., 2003; Pichavant et al., 2002; Sobolev
and Chaussidon, 1996). Critically, Hamada and Fujii (2008) and
Zimmer et al. (2010) report that a water content of 2 wt.% separates
dry tholeiitic (olivine+plagioclase/orthopyroxene) from wet
calc-alkaline (clinopyroxene+magnetite) magmatic fractionation
trends.
Arc basalts are also characterized by distinctly higher oxidation
states than MORB (up to 2 log units above the fayalitemagnetite
quartz buffer: FMQ+2; Ballhaus, 1993; Brandon and Draper, 1996;
Parkinson and Arculus, 1999; Rowe et al., 2009). The relatively high
oxidation state of arc magmas is a critical factor in their subsequent
metallogeny, and originates from oxidative seaoor alteration of the
oceanic plate (Staudigel et al., 1996), transmittedinto the mantle wedge
by the metasomatic uid ux (Brandon and Draper, 1996; Kelley and
Cottrell, 2009; Malaspina et al., 2009).
2.1.1. Behavior of metals
Most base and precious metals would be expected to have at least
moderate solubilities in the hot, relatively oxidized, saline aqueous
uids exsolved from the downgoing slab. In particular, as noted in
Section 2.1, Pb, As, and Sb are strongly mobilized by such uids, possibly
along with Tl and Cu (Noll et al., 1996). The behavior of highly
siderophile elements (HSE) suchas AuandPGEis less well knownunder
these conditions, but studies of metasomatized mantle xenoliths from
island arc lavas suggest that Au, Re, and the Pd-group elements
(including Pt) are mobilized into the mantle wedge during subduction
metasomatism (Dale et al., 2009; Kepezhinskas et al., 2002; McInnes et
al., 1999; Sun et al., 2004a; Widom et al., 2003).
The volumetric extent and efciency of mobilization of metals into
the mantle wedge bythis process are unknown, but uidmetasomatism
clearly represents one viable mechanism for metal transfer into arc
magma sources.
The behavior of chalcophile and siderophile metals during subse-
quent partial melting of the metasomatized mantle wedge depends
critically on oxidation state (f
O2
) and sulfur fugacity (f
S2
), because these
parameters control the stability and abundance of sulde phases. Gold
and PGE partition strongly into sulde phases relative to silicate melts,
but Cu to a somewhat lesser degree (Campbell and Naldrett, 1979;
Peach et al., 1990), so if suldes are abundant in the magma source
region, partial melts will be depleted in Au and PGE relative to Cu
(Fig. 2). Under the high f
O2
and f
S2
conditions of the supra-subduction
zonemantle wedge, thebulkof thesulfur uxwill likely consist of SO
2
or
M
a
n
tle
c
o
rn
e
r flo
w
Sea level
Asthenosphere
Volcanic arc
Talc
S
e
r
p
e
n
t
i
n
e
Chlorite +
serpentine
Chlorite
1
0
0
0

C
6
0
0

C
O
c
e
a
n
i
c

m
a
n
t
l
e
l
i
t
h
o
s
p
h
e
r
e
Oceanic crust
D
e
h
y
d
r
a
t
i
o
n

o
f

o
c
e
a
n
i
c
l
i
t
h
o
s
p
h
e
r
e
1
0
0
0

C
A
m
p
h
i
b
o
l
e
Z
o
i
s
i
t
e
C
t
d
Oceanic crust
Partial melting
Mantle
lithosphere
C
h
l
o
r
i
t
e

1
0


p
h
a
s
e
+

s
e
r
p
e
n
t
i
n
e
0 km
100 km
200 km
1
4
0
0

C
1
4
0
0

C
600C
1
0
0
0
C
E
c
l
o
g
i
t
e
Sediment
Metasomatized
asthenosphere
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.
4 J.P. Richards / Ore Geology Reviews 40 (2011) 126
sulfate, dissolved rst in slab uids and then magma (Carroll and
Rutherford, 1985; Jenner et al., 2010; Jugo et al., 2005a). However,
because of the equilibria betweenvarious sulfur species, at highf
S2
some
condensed sulde phases will likely also be present (McInnes et al.,
2001). Consequently, Richards (2009) suggested that normal arc
magmas will be minimally depleted in Cu (due to its higher abundance
and moderate chalcophile nature) relative to sparse, highly siderophile
elements suchas AuandPGE, whichwill be strongly retainedinresidual
sulde phases in the source region (e.g., Hamlyn et al., 1985; Mitchell
andKeays, 1981; Peachet al., 1990). This may explainthe Cu-richnature
of typical arc-related porphyry deposits (relative to Au and PGE;
Richards, 2005). In contrast, Au-rich porphyry deposits may require
atypical subduction-related or collisional tectonic settings and petro-
genetic processes, which act to destabilize residual sulde phases and
render Au incompatible (e.g., Jgo et al., 2010; Richards, 1995, 2009;
Sillitoe, 2000; Solomon, 1990; Wyborn and Sun, 1994; see Section 2.5).
The low abundances of PGE in many arc-related ore deposits suggest a
further separation of these elements from Au and Cu, perhaps through
the formation of residual platinoid alloy phases (e.g., Barnes et al., 1985;
Borisov and Palme, 1997; Kepezhinskas et al., 2002; Peach et al., 1990)
or Cr-spinels (into which Ir-group PGE strongly partition; Hattori et al.,
2010; Righter et al., 2004).
2.2. Sediment dehydration and/or melting in subduction zones
Seaoor sediments on the surface of the downgoing slab are
another potential source of metasomatic contributions to the mantle
wedge. Much of this sedimentary material will be scraped off at the
trench to form an accretionary prism, but varying amounts may also
be subducted, depending on the degree of coupling between the
upper and lower plates, and also the sediment input load (Fig. 1). Such
sediments will be water-rich and pelitic in bulk composition, and thus
are more likely to undergo partial melting under subduction zone
conditions than basaltic oceanic crust (Hermann and Spandler, 2008).
Nevertheless, Aizawa et al. (1999); Dreyer et al. (2010); and Duggen
et al. (2007) have suggested that dehydration is the principal process
affecting sediments down to depths of ~100 km (i.e., to below the
volcanic arc), with melting only occurring signicantly at greater
depths when temperatures exceed ~800 C (possibly reected in the
geochemistry of some back-arc magmas).
Sediment contributions to the source of arc magmas have been the
subject of numerous studies, with the least ambiguous evidence
coming from island arcs (e.g., MacDonald et al., 2000; Thirlwall et al.,
1996; Wysoczanski et al., 2006). In continental arcs, it can be difcult
to distinguish between chemical and isotopic signatures from
subducted continent-derived sediment versus crustal contamination
during magma ascent (e.g., Hildreth and Moorbath, 1988; Kemp et al.,
2007): both sources will contribute incompatible elements and
crustal isotopic values to primary mantle-derived arc magmas
(Breeding et al., 2004).
Trace elements commonly used as indicators of sediment contribu-
tions to arc magmas are Ba, B, Be, Th, andPb(Dreyer et al., 2010; Johnson
and Plank, 1999), and Ba/La and Th/La ratios canbe used as a measure of
sediment versus mantle source components (Plank, 2005; Walker et al.,
2001). In particular, the cosmogenic radioisotope
10
Be can be used as a
tracer of recent (b10 Ma) introduction of sediments into arc magma
sources (Dreyer et al., 2010; Morris et al., 1990). However, although a
clear sediment-derived isotopic signature can be observed in many
island arc systems, the volumetric contribution of sediments to island
arc magmas seems to be relatively minor (Hawkesworth et al., 1994;
Kilian and Behrmann, 2003; Poli and Schmidt, 2002; Stern et al., 2006).
One additional element that may be added to the mantle wedge
from subducted sediments is sulfur, as suggested by the positive
34
S
compositions of arc magmas (de Hoog et al., 2001), which are similar
to those of seaoor sediments (Alt et al., 1993). Analyses of glass
inclusions in olivine fromprimitive arc magmas reveal concentrations
of up to 2900 ppm S (de Hoog et al., 2001), and Jugo et al. (2005b)
measured experimental concentrations of up to 1.5 wt.% S in oxidized
arc basalts. These high sulfur contents have great signicance for the
behavior of chalcophile and siderophile metals (see Sections 2.1.1,
2.5.1, and 3).
2.2.1. Behavior of metals
Lead, which is signicantly enriched in the continental crust (and
crustally-derived sediments) relative to the mantle, is the only metal for
which a clear sedimentary source can be inferred in some island arc
magmas, because it can be readily identied by its radiogenic isotopic
composition compared with depleted mantle sources. However, in
continental arcs, distinguishing a subducted sediment source of
radiogenic Pb from crustal contamination during magma ascent is
very difcult (e.g., Barreiro, 1984; Chiaradia et al., 2004; Kontak et al.,
1990). This has led to diverging opinions: for example, Aitcheson et al.
(1995); HildrethandMoorbath(1988); James (1982); Kay et al. (1999);
and Tilton et al. (1981) concluded that the bulk of the radiogenic Pb in
central Andean magmas comes from upper plate crustal sources,
whereas Macfarlane (1999); McNutt et al. (1979); Mukasa et al.
(1990); and Sillitoe and Hart (1984) preferred a subducted sediment
source for Pb in some Andean igneous rocks and ore deposits.
Inferring a seaoor sediment source for other metallic components
in arc magmas (and related ore deposits) is much more speculative,
and generally involves the subduction of metal-rich manganese
nodules or even massive sulde deposits. The latter, however, likely
oxidize and disperse geologically rapidly after formation on the
seaoor (e.g., Edwards, 2004; Herzig et al., 1991) unless quickly
buried by lava; they might thus only be expected to be subducted with
very young oceanic crust. Few studies have specically invoked a
subducted sediment source for ore metals other than a component of
Pb, and the majority of authors have concluded that such a source is
either unnecessary or unproven (e.g., Burnham, 1981; Chiaradia et al.,
2004; de Hoog et al., 2001; Fontbot et al., 1990).
A
u

(
p
p
b
)
C
u

(
p
p
m
)
R = (mass of silicate melt)/(mass of sulfide)
Cu maximized in
magma (R 10
3
)
Au maximized in
magma (R 10
6
)
0.1
1
10
10
4
10
5
10
6
10
5
1 10 10
4
10
5
10
6
10
7
10
8
C
u
in
s
u
lfid
e
C
u
in
m
a
g
m
a
A
u
in
s
u
lfid
e
A
u
in
m
a
g
m
a
Sulfide/silicate melt partition
coefficients:
D
(Cu)
= 10
3
D
(Au)
= 10
5
Metal concentrations in
magma in absence of sulfide:
Cu = 50 ppm
Au = 5 ppb
A
u
d
e
p
le
te
d
in
m
a
g
m
a
A
u
e
n
ric
h
e
d
in
re
s
id
u
a
l s
u
lfid
e
Porphyry Cu
potential arc
magmas
Porphyry Cu-Au
potential
magmas
10
2
10
3
10
2
10
3
10
4
10
3
10
2
5 ppb Au
50 ppm Cu
Fig. 2. Concentrations of Cu and Au in silicate magma and coexisting sulde liquid as a
function of R=[mass of silicate melt]/[mass of sulde melt] (Campbell and Naldrett,
1979; diagram modied from Richards, 2005, 2009). At R-factors below ~10
2
, magmas
will be depleted in both Cu and Au. At R-factors between ~10
2
10
5
, magmas will be
depleted in Au but essentially undepleted in Cu (porphyry Cu-potential magmas). At R-
factors N10
5
, magmas will be undepleted in Au and Cu (porphyry CuAupotential
magmas). A corollary of this diagram is that arc magmatism will leave small amounts of
relatively Au-rich sulde in the mantle source or lithosphere during fractionation,
which can be remelted during post-subduction tectonomagmatic processes, to form
small-volume, alkaline, porphyry Au-potential magmas.
5 J.P. Richards / Ore Geology Reviews 40 (2011) 126
2.3. Oceanic slab melting in subduction zones
The question of whether the downgoing oceanic slab, or more
specically the basaltic oceanic crust, melts during subduction has
prompted lively debate recently, not only in the petrology literature
(e.g., Conrey, 2002; Defant and Kepezhinskas, 2001, 2002; Garrison and
Davidson, 2003), but also amongst economic geologists because of the
suggestionthat slabmelts might insomewaybeuniquelyfertilefor later
porphyry ore formation (e.g., Mungall, 2002; Oyarzun et al., 2001, 2002;
Sajona and Maury, 1998; Thiblemont et al., 1997; for contrary opinions,
see: Rabbia et al., 2002; Richards, 2002; Richards and Kerrich, 2007).
The idea that the hydrated basaltic ocean crust might melt during
subduction was an early assumption of the plate tectonic revolution,
because it seemed conveniently to explain the relatively felsic
composition of arc magmas (as opposed to basalts formed by melting
of peridotitic mantle). The theory was given credence by the
experiments of Rapp et al. (1991) and Rapp and Watson (1995), who
showed that melting of amphibolite under upper mantle conditions
(1025 C and 0.81.6 GPa) could generate an intermediate composition
tonalitetrondhjemite melt, not dissimilar toanarc andesite. Defant and
Drummond (1990) termed the products of subducted slab melting
adakites, after a single anomalous lava ow on Adak Island in the
Aleutians described by Kay (1978). Because garnet should be present in
the eclogitic source of these magmas, Defant and Drummond (1990)
argued that such melts could be distinguished by anomalously low
concentrations of heavy rare earth elements (HREE) and Y (which are
compatible withgarnet) relativetolight rareearthelements (LREE), and
high concentrations of Sr (because of the absence of plagioclase at such
depths). Thus, slab melts, or adakites, could be distinguished on Sr/Y
versus Y, or La/Yb versus Yb diagrams fromnormal arc magmas formed
in the absence of garnet.
However, despite the theoretical possibility of melting subducted
oceanic crust, most thermal models of subduction zones indicate that
temperatures in the slab do not normally reach melting conditions
(N800 C) prior to dehydration and eclogitization (Fig. 1), which would
render the slab infusible (e.g., Davies and Stevenson, 1992; Peacock,
1996; Poli and Schmidt, 2002). Thus, Defant and Drummond (1990)
proposed that slab melting might be restricted to the subduction of
young (25 m.y. old) and therefore still hot oceanic crust, and Peacock
et al. (1994) were even more restrictive (b5 m.y. old). Other relatively
uncommon scenarios that might result in slab melting include shallow
or stalled subduction (whereby the slab has more time to heat up at
shallow depths; Gutscher et al., 2000; Peacock et al., 1994), ridge
subduction (Guivel et al., 2003; Kay et al., 1993), or edge-melting of
detached slabs or slab windows (Haschke and Ben-Avraham, 2005;
Thorkelson and Breitsprecher, 2005; Yogodzinski et al., 2001).
A further complication is added by the fact that most adakites
described in the petrology literature are not in fact primary slab melts,
but are substantially evolved, having reacted or hybridized with the
asthenosphere during ascent (and likely also the upper plate
lithosphere). This modication to the adakite slab-melting model is
required to explain the high contents of MgO, Ni, and Cr present in
some adakites relative to expected levels for hydrated basalt partial
melts (Defant and Kepezhinskas, 2001; Drummond et al., 1996;
Martin, 1999; Martin et al., 2005; Yogodzinski et al., 1995).
Direct evidence for slab melting is lacking, but supra-subductionzone
xenoliths from the Tabar-Lihir-Tanga-Feni (Papua New Guinea), Philip-
pine, and Patagonian arcs preserve hydrous, silica-rich glass inclusions
that are thought to represent migrating slab melts (respectively: Kilian
and Stern, 2002; McInnes and Cameron, 1994; Schiano et al., 1995). The
glass inclusions characterized by Schiano et al. (1995) were calc-alkaline
in composition, with high incompatible element and low Ti, Nb, and Y
contents, high LREE/HREE ratios, and homogenization temperatures of
~920 C. They thus compositionally resemble melts that would be
predicted to formfromslab melting, and appear to provide evidence that
this process occurs at least locally where conditions permit.
The chemical difference between slab dehydration and slab
melting would seem to be rather small, given that both media
would be enriched in volatiles, incompatible elements, and silica.
Indeed, as noted in Section 2.1, there may well be a continuum
between silica-rich aqueous uids and aqueous silicate melts at
greater depths in subduction zones (Kawamoto, 2006; Kessel et al.,
2005a,b; Manning, 2004; Portnyagin et al., 2007). This likely explains
why the debate between slab melting and dehydration is somewhat
intractable, and mostly hinges on subtle trace element characteristics.
2.3.1. Behavior of metals
Slab melts are predicted to be volatile-rich (including H
2
O, S, and
Cl) and oxidized, and thus, like hydrous slab uids, would be expected
to be able to transport base and precious metals at least to some
degree. However, analyses of such metals (except iron) are not
reported in most melt inclusion studies (e.g., Kilian and Stern, 2002;
McInnes and Cameron, 1994; Schiano et al., 1995), so there are no
direct constraints on the capacity of such melts to transfer chalcophile
and siderophile metals from the slab to the mantle wedge.
In a study of metasomatized mantle xenoliths from a submarine
volcano near Lihir Island, Papua New Guinea, McInnes et al. (1999)
concluded that enrichments in Cu, Au, and PGE were caused by slab uid
metasomatism, rather than melts. In contrast, Kepezhinskas et al. (2002)
measured the concentrations of Au and PGEs in mantle xenoliths from
the Kamchatka arc, and suggested that a uid-transported component
couldbe distinguishedfroma slab melt component by co-enrichments in
PGE and high eld strength elements (HFSE) in the latter, because of the
low capacity of aqueous uids to carry HFSE. Intriguingly, they noted no
suchcorrelationbetweenHFSE-enrichments andAu, andconcludedthat,
whereas PGE might be transported into the mantle wedge by both uids
and melts, Au was likely only carried by hydrous uids.
Two theoretical studies have proposed that slab melts should be
unusually effective as metal-transporting and ore-forming agents.
Oyarzun et al. (2001) argued that slab melts should be unusually
oxidized and rich in H
2
O and SO
2
(relative to normal arc magmas
derived by asthenospheric partial melting), although no evidence was
given for this assertion. Such magmas, they argued, should be
particularly suitable for the formation of magmatichydrothermal
porphyry copper deposits upon emplacement in the upper crust.
Mungall (2002) presented a theoretical model for oxidation of the
mantle wedge by Fe
3+
-rich slab melts to the point of complete sulde
destruction, thereby rendering chalcophile and siderophile elements
incompatible in mantle phases, and free to partition into silicate
melts. He argued that ferric iron is a much stronger oxidant than slab-
derived water, and that slab melts should be rich in Fe
3+
generated by
oxidative seaoor alteration. Thus, slab melts might be uniquely
favorable for the subsequent generation of metal-rich, and particu-
larly Au-rich, magmas derived from the mantle wedge.
Mungall's (2002) model may have applicability for less common Au-
rich porphyry deposits formed in atypical subduction zone settings that
might cause slab melting, but does not seem well suited to explain
regular arc porphyry Cu deposits. In either case, metals are envisaged to
be sourced fromthe mantle wedge, not the slab. In contrast, Oyarzun et
al.'s (2001) model does not address the source of metals, and is at root
based on the assumption that slab melts are uniquely more H
2
O- and
SO
2
-rich, and more oxidized than normal arc magmas, leading to
specic ore depositional processes rather thansource processes. It is not
clear that these assumptions are justied, and some have argued that
slab melts might in fact be relatively reducing because of the additional
presence of organic-rich sediment melts (Wang et al., 2007a).
2.4. Supra-subduction zone lithospheric melting: the MASH process
Hydrous basaltic magmas generated in the mantle wedge will have
temperatures in excess of 1000 C (Eggins, 1993; Grove et al., 2006;
MacDonald et al., 2000), and perhaps as high as 1350 C (Schmidt and
6 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Poli, 1998; Tatsumi, 2003). Because their densities will be lower than
the mantle but not the crust (Herzberg et al., 1983), they will tend to
rise fromtheir asthenospheric source region and penetrate the mantle
lithosphere, but pool at the crust/mantle density barrier (level of
neutral buoyancy: Fig. 3; Fyfe, 1992; Hildreth, 1981). Here, if the ux
of magma and heat is maintained and supplemented by the latent
heat of crystallization as the magma begins to crystallize, high
temperatures can be brought to bear on lower crustal rocks that will
cause partial melting (Annen et al., 2006; Bergantz and Dawes, 1994;
Huppert and Sparks, 1988; Klepeis et al., 2003; Petford and Gallagher,
2001; Rushmer, 1993). Hildreth and Moorbath (1988) suggested that
it is the interaction between this hot, hydrous basalt ux from the
subduction zone and felsic crustal partial melts that gives rise to the
uniform composition of andesites in continental volcanic arcs, by a
process they dubbed meltingassimilationstoragehomogenization
(MASH). In a renement of this model, Annen et al. (2006) referred to
the region of magmacrust interaction as a hot zone (Fig. 3).
Because garnet is a product of such lower crustal fractionation and
partial melting processes (Alonso-Perez et al., 2009; Berger et al.,
2009; Dufek and Bergantz, 2005; Garrido et al., 2006; Hansen et al.,
2002; Klepeis et al., 2003; Rushmer, 1993; Wolf and Wyllie, 1994),
derivative calc-alkaline magmas may display trace element compo-
sitions that resemble adakites (see Section 2.3) but which are
unrelated to slab melting (Klepeis et al., 2003; Macpherson et al.,
2006; Richards and Kerrich, 2007; Tulloch and Kimbrough, 2003).
Such common processes, affecting batches of magma crystallizing and
fractionating at different crustal depths (e.g., Annen et al., 2006) are
entirely consistent with petrological observations in arc volcanic
systems where adakite-like (i.e., high-Sr/Y) andesitic lavas may be
interlayered with normal andesites in a single volcano, and do not
require a fundamental change in magma source 100 km below the
volcano (e.g., Feeley and Davidson, 1994; Grunder et al., 2008;
Richards et al., 2006a).
Once these hybrid magmas reach basaltic andesitic to andesitic
compositions, their densities are low enough to allow them to rise
through the lower continental crust (Herzberg et al., 1983), but they
tend to stall again at a second density barrier in the middle to upper
crust belowlight supracrustal rocks. This is the level (510 km) at which
large arc batholiths will form if the ux of magma is sustained, and is
alsothe level at whichevolvedfelsic melts andvolatiles are accumulated
(Fig. 3; see Richards, 2003, and references therein). These volatiles drive
buoyant, bubbly, evolved magma upwards into the cover rocks to form
subvolcanic stocks and dikes, or explosive volcanic eruptions if they
reach surface. The volatiles may also separate from the magma ux to
form a separate uid plume, which ultimately vents at the surface
(fumaroles) but may also formporphyry- and epithermal-type deposits
in the hypabyssal and near-surface environment (see discussion of
these processes in Sections 3 and 4).
Volcanic arc
1
4
0
0

C
Sea level
600C
1
0
0
0
C
Continental crust
Metasomatized
asthenosphere
1
4
0
0

C
1
0
0
0

C
Subcontinental
mantle lithosphere
O
c
e
a
n
i
c
m
a
n
t
l
e
l
i
t
h
o
s
p
h
e
r
e
M
a
n
tle
c
o
rn
e
r flo
w
Asthenosphere
0 km
50 km
100 km
Mid- to upper crustal
batholithic complex
Lower crustal
MASH or hot zone
Feeder dikes
6
0
0

C
D
e
h
y
d
r
a
t
i
n
g

o
c
e
a
n
i
c

c
r
u
s
t
Epithermal deposits
Hydrothermal
alteration
Porphyry deposits
Partial melting
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 uids ascend, cool, and interact with near-surface upper crustal rocks. Modied fromRichards (2003, 2005); sources: Hildreth and Moorbath (1988), Winter (2001),
Annen et al. (2006), and Sillitoe (2010).
7 J.P. Richards / Ore Geology Reviews 40 (2011) 126
2.4.1. Behavior of metals
Some of the clearest evidence for the involvement of crustal rocks in
continental arc magmas comes fromPb isotopic data (e.g., Wrner et al.,
1992), although as noted in Section 2.2, there may be ambiguity
between lower crustal melting and the melting of subducted continent-
derived sediments. Lead and uranium are much more abundant in the
bulk continental crust (11 ppm Pb, 1.3 ppm U; Rudnick and Gao, 2003)
or lower continental crust (4 ppm Pb, 0.2 ppm U; Rudnick and Gao,
2003) than the primitive mantle (b0.2 ppm Pb, b0.02 ppm U; Taylor
and McLennan, 1985; Sun and McDonough, 1989), MORB (0.3 ppm Pb,
0.05 ppm U; Sun and McDonough, 1989), or typical low-K mac arc
andesites (b1.8 ppm Pb, b0.2 ppm U; Gill, 1981), so it only requires
small amounts of contamination by radiogenic crustal lead to
signicantly modify a magma's Pb isotopic composition. Thus, it is not
clear that a particularly large amount of Pb in arc magmas is sourced
from crustal rocks versus the subduction zone, and Macfarlane et al.
(1990) have argued that the crustal contribution is minimal in Central
Andean magmas and ores. Moreover, porphyry-type systems are
typically not Pb-rich, except in late-stage skarns and distal veins
where some of the Pb may have been derived fromlocal host rocks (e.g.,
Mukasa et al., 1990).
Most researchers have assumed that, because of the higher
concentrations of Cu in primitive andesites (145 ppm; Gill, 1981)
compared with the bulk continental crust (27 ppm; Rudnick and Gao,
2003), the bulk of Cu in porphyry-type deposits is mantle-derived. A
similar assumption is made for Au, although the primitive mantle and
continental crust actually have comparable concentrations (13 ppb
Au; Rudnick and Gao, 2003; Taylor and McLennan, 1985). Moreover,
porphyry Cu(Au) deposits are found in association with mantle-
derived arc magmas worldwide, regardless of crustal type (oceanic or
continental) or thickness (Kesler, 1973), so a crustal heritage for these
metals does not appear to be critical. Nevertheless, a lower crustal
source, perhaps hybridized with mantle-derived magmas, has been
proposed by Bouse et al. (1999) for both magmas and metals in the
Laramide porphyry systems of Arizona. Moreover, Titley (1987, 2001)
has specically suggested that Au and Ag are crustally derived in a
range of ore deposits including porphyries in southwestern USA,
because the ratios of these elements correlate closely with two
distinct basement domains in this region. Given that Au is not
especially enriched in the mantle (see above), and that Ag is in fact
more abundant in the crust than the mantle (80 ppb in the bulk
continental crust, versus b19 ppb in the primitive mantle; Taylor and
McLennan, 1985), a crustal source for at least some proportion of
these minor metals, and especially Ag, in arc magma-related systems
may be reasonable. However, it seems unlikely that this argument can
be extended to copper, except perhaps on the margins.
2.4.2. Sources of Mo
Molybdenum occurs in varying amounts in porphyry-type deposits,
ranging fromtrace levels (b0.01 wt.%Mo) inporphyry Cu(Mo) deposits,
where it may not evenbe recoveredas a byproduct, to being the mainore
component (up to 0.3 wt.% Mo) in porphyry Mo deposits (Seedorff et al.,
2005; Westra and Keith, 1981). At the Mo-rich end of the spectrum, there
are twoclearly different tectonomagmatic associations, only one of which
is directly related to subduction: calc-alkaline porphyry Mo deposits are
generally relatively low grade (0.10.02 wt.% Mo; Carten et al., 1993),
whereas intra-cratonic rift-related deposits associated with high-silica,
uorine-rich, peraluminous granitoids are relatively high grade (0.1
0.3 wt.%Mo; e.g., Climax-type deposits; Cartenet al., 1993; Kirkhamand
Sinclair, 1996; Sinclair, 2007; Stein, 1988; White et al., 1981).
Kesler (1973) noted a general association (with exceptions) of
porphyry CuAu deposits in island arcs, and porphyry CuMo deposits
in continental arcs, and it is clear that the peraluminous felsic rocks
associated with rift-related Climax-type porphyry Mo deposits are
primarily of continental crustal origin (Farmer and DePaolo, 1984;
Stein, 1988). This has led to one view that Mo might be predominantly
derived from continental crustal sources (Farmer and DePaolo, 1984;
Stein, 1988; Klemmet al., 2008; White et al., 1981). On the other hand,
minor amounts of Mo do occur in some island arc-related porphyry
deposits where no continental crustal sources are inferred (Westra and
Keith, 1981), so a mantle (subduction zone) source for at least some Mo
cannot be excluded. Moreover, Blevin and Chappell (1992) and Blevin
et al. (1996) have demonstrated a continuum from CuAu deposits
associated with unevolved, mac I-type granitoids to WMo deposits
associated with cogenetic, evolved granites in eastern Australia,
suggesting a common, magmatic source for all of these elements.
A complication is introduced in the Climax-type deposits, because
although the immediate source of the Mo-bearing uids is felsic
magma of clear crustal origin, many deposits also showa close genetic
association with mac alkaline magmas, which may have introduced
volatiles, S, and possibly Mo into the evolved felsic magma chamber
(Audtat, 2010; Carten et al., 1993; Keith et al., 1986, 1998). Keith et
al. (1997), Hattori and Keith (2001), and Maughan et al. (2002) have
also suggested that injections of mac alkaline magmas into the
evolving Bingham Canyon magmatic system may have given rise to
the unusually large size and high grades of this porphyry CuMoAu
deposit. Along the same lines, Pettke et al. (2010) have proposed that
the unusual CuMoAu endowment of the southwestern USA (e.g.,
the giant Bingham, Butte, Climax, Henderson, and Questa porphyry
CuMoAu and porphyryMo deposits) reects Cenozoic remobiliza-
tion of Proterozoic subduction-metasomatized subcontinental mantle
lithosphere (see Section 2.5).
Thus, at this time there is no consensus regarding the crustal
versus mantle origin of molybdenumin porphyry deposits, although it
is clear that the highest grade porphyry Mo deposits are formed in
intra-plate continental settings, and if a mantle source is important in
these cases, it is not directly related to subduction activity but rather
to rifting or reactivation of previously subduction-enriched litho-
spheric sources.
2.5. Lithospheric melting during post-subduction events
A number of mineral deposits with broad similarities to those
formed by subduction-related processes are also found in post-
subduction tectonic settings, such as subduction reversal or migration,
arc collision, continentcontinent collision, and post-collisional rifting.
They include porphyry SnW, Mo, CuMo, and CuAu deposits and
epithermal Au deposits, and in many cases are only known not to be
directly related to subduction because of precise geochronology and
plate tectonic reconstructions that place their formation after subduc-
tion has demonstrably ceased. Associated magmas are typically calc-
alkaline, but tend towards somewhat more alkaline compositions
compared with normal arc magmas (Richards, 2009).
In complex accretionary arcs, it can be very difcult to ascribe any
given pluton (and any associated mineral deposits) to a particular
subduction or collisional event, because subduction commonly con-
tinues after collision, albeit normally with a shift in the locus of
magmatism. However, in continentcontinent collision zones or where
arc collision terminates subduction, there can be greater certainty about
the timing of cessation of subduction magmatism. Consequently, it is in
collisional orogens such as the Neo-Tethyan belt of southeastern Europe
andsouthernAsia that some of the clearest examples of post-subduction
magmatism and mineralization are found. These include, from east to
west, the Miocene Gangdese porphyry CuMo belt inthe Tibetanorogen
(Houet al., 2006, 2009; Yanget al., 2009), the MioceneKermanporphyry
CuMo belt in southeastern Iran (Shaei et al., 2009), the Miocene Sari
Gunay epithermal Au deposit in northwestern Iran (Richards et al.,
2006b), theEocene pler epithermal Audeposit insoutheasternTurkey
(Keskin et al., 2008; Kuscu et al., 2010), the Pliocene Kisladag porphyry
Au deposit in western Turkey, the Miocene Skouries porphyry CuAu
PGEdeposit inGreece(Economou-Eliopoulos andEliopoulos, 2000), and
8 J.P. Richards / Ore Geology Reviews 40 (2011) 126
the Roia Montan epithermal Au deposit in Romania (Manske et al.,
2006; Neubauer et al., 2005).
Similarly, in the southwest Pacic ocean, accurate plate tectonic
reconstructions permit the identication of a number of post-
subduction porphyry and epithermal deposits, such as the Grasberg
porphyry CuAu deposit in Papua, Indonesia (Cloos et al., 2005;
Paterson and Cloos, 2005), the Ok Tedi porphyry CuAu deposit (van
Dongen et al., 2010) and the Porgera alkalic-type epithermal Au deposit
in mainland Papua New Guinea (Richards et al., 1990; Richards and
Kerrich, 1993), the Lihir alkalic-type epithermal Au deposit on Lihir
Island, Papua NewGuinea (Carman, 2003; Kennedyet al., 1990), andthe
Emperor alkalic-type epithermal Au deposit in Fiji (Gill and Whelan,
1989; Settereld et al., 1992). (For reviews of alkalic-type epithermal
deposits, see Jensen and Barton, 2000 and Richards, 1995).
Because subduction has ceased in these regions, a fresh supply of
uids, volatiles, and other slab-derived components to the mantle
wedge no longer exists. Nevertheless, the broad geochemical similarity
of many of these magmas to normal arc magmas, including their
hydrous and generally oxidized nature, suggests some link to
subduction metasomatism. Consequently, many researchers have
implicated upper-plate lithospheric sources, modied by earlier
subduction-related uids and/or hydrous melts (e.g., Clemens et al.,
2009; Cloos et al., 2005; Guo et al., 2007; Harris et al., 1986; Johnson
et al., 1978; Pearce et al., 1990; Pettke et al., 2010; Richards, 2009).
Previously subduction-modied asthenosphere is unlikely to be a viable
source except for a short period after subduction has ceased (e.g.,
Richards et al., 1990; Solomon, 1990), because such material will be
quickly dispersed by mantle convection.
The key to all of these models is subduction-derived water, which is
most likely stored in amphibolitic cumulates, residual from the earlier
arc magma uxandlocatedinthe deepcrust or mantle lithosphere (e.g.,
Claeson and Meurer, 2004; Davidson et al., 2007; DeBari and Coleman,
1989; Jagoutz et al., 2009; Larocque and Canil, 2010; Mntener and
Ulmer, 2006; Tiepolo and Tribuzio, 2008). Water lowers the solidus of
silicate assemblages, and will lead to the formation of hydrous partial
melts during pro-grade metamorphism or mac melt invasion (Beard
and Lofgren, 1991; Rushmer, 1991; Wolf and Wyllie, 1994).
Thermal rebound in thickened orogenic crust, delamination of
sub-continental mantle lithosphere, or post-collisional rifting (with
ingress of asthenospheric melts into the lower crust in the last two
cases) can all cause small-volume partial melting of arc-metasoma-
tized lithosphere and/or hydrous lower crustal cumulates (Fig. 4;
Brown, 2010; Clemens et al., 2009; Harris et al., 1986; Richards, 2009).
Such melts, being derived from subduction-modied sources, will
share many of the characteristic geochemical features of arc magmas,
including their relatively high water contents and oxidation states,
and potentially metal contents (see Section 2.5.1). The smaller volume
of partial melting to be expected in such tectonic settings will give
these magmas a somewhat more alkaline composition than arc
magmas (e.g., Clemens et al., 2009), and will also mean that large
batholithic complexes are unlikely to be formed, consistent with the
generally smaller and more isolated occurrence of such post-
subduction magmatic systems (compared with arc-related Cordille-
ran batholiths, or collisional S-type batholiths; Pitcher, 1997).
Because these post-subduction magmas are derived from amphibo-
litic sources in which garnet (titanite) is likely also present, and
because their hydrous nature will suppress plagioclase fractionation
(similar to other hydrous arc magmas), they may be characterized by
elevated Sr/Y and La/Yb ratios; that is, they may display adakite-like
trace element characteristics.
2.5.1. Behavior of metals in subduction-modied sources
As discussed in Section 2.1, arc magmas are characterized by high
f
O2
and f
S2
relative to normal melts from MORB-depleted astheno-
sphere. Consequently, such magmas may be sulde-saturated (at
oxidation states up to FMQ+2.3; Jugo, 2009) despite sulfur being
predominantly present in the melt as sulfate or SO
2
(Carroll and
Rutherford, 1985). Xenoliths from supra-subduction zone mantle
(McInnes et al., 1999) and samples of mac cumulates from lower
crustal arc roots (Fig. 5; Canil et al., 2010; Greene et al., 2006; Jagoutz
et al., 2007) reveal the common presence of small amounts of sulde,
typically trapped as inclusions in silicate phases (suggesting a primary
magmatic rather than secondary hydrothermal origin).
Hamlyn et al. (1985), Richards (1995, 2009), Solomon (1990), and
Wyborn and Sun (1994) have explored the role of residual sulde
phases on the metal content of fractionating magmas, and also of
partial melts formed during later, post-subduction melting events.
The high partition coefcients for chalcophile and highly siderophile
elements (HSE) between sulde phases and silicate melt mean that
such metals should be strongly partitioned into any coexisting sulde
phases (Campbell and Naldrett, 1979; Peach et al., 1990). As shown in
Fig. 2, at high abundances of sulde relative to silicate melt (low R-
factor; Campbell and Naldrett, 1979), the melts will be depleted in all
of these chalcophile and siderophile elements. In contrast, at
intermediate abundances of sulde (intermediate R-factor), only
originally sparse HSE will show signicant depletions. This led
Richards (2005, 2009) to propose that small amounts of sulde left
behind as residual phases from fractionation of arc magmas in the
deep lithosphere (or asthenosphere) will not signicantly deplete
those magmas in relatively abundant chalcophile elements such as Cu
and Mo, but might signicantly deplete them in highly siderophile
elements such as Au. This would give rise to magmas with relatively
high Cu/Au ratios (which might form Cu-rich porphyry deposits), but
would leave a residue of potentially HSE-rich suldes in the mantle
and/or lower crustal amphibolitic cumulate arc roots.
As notedinSection2.5, subduction-modiedasthenospheric sources
will be rapidly convected away when subduction ceases, and so could
only contribute to immediately post-subduction magmatism. In
contrast, deep crustal amphibolites are preserved in the lithosphere,
andwill besusceptible topartial meltingat anylater timeduetothermal
reboundor reheating byinvading asthenospheric melts. Under lower f
S2
post-subduction conditions (a ux of S from the subduction zone is no
longer present), any residual sulde phases would likely dissolve into
the S-undersaturated silicate melt, carrying their metal loads with them
(e.g., Ackerman et al., 2009). Richards (2009) proposed that this might
explain the occurrence of Au-rich porphyry and related epithermal
systems in some post-subduction settings, such as the alkalic-type
epithermal Audeposits of the SWPacic, andvarious post-collisional Au
deposits in the BalkansTurkeyIran Neotethyan belt. Pettke et al.
(2010) have proposed a similar model for giant porphyry CuMoAu
deposits in the southwestern USA.
This model can also explain the occurrence of Aupoor porphyry
Cu(Mo) deposits in post-subduction settings, such as those in Tibet
and Iran (Hou et al., 2009; Shaei et al., 2009; Wang et al., 2007b), the
only difference being that in this case larger proportions of sulde
may have fractionated out from the original arc magmas in the deep
crust. Such suldes would have retained signicant amounts of the
subduction ux of Cu and Mo, but HSE would be diluted to low
concentrations by the greater volume of sulde (low R-factor; Fig. 2).
Second-stage melts from such cumulate sources would therefore be
Cu(Mo)-rich, but not necessarily Au-rich.
A control on these two scenarios (abundant Cu-rich residual sulde
versus sparse but HSE-rich residual sulde, or low versus high R-factor)
might be the average oxidationstate andsulfur fugacity of the generative
subduction system. Inmore oxidized or S-poor systems, smaller volumes
of HSE-rich sulde would exsolve from the silicate melt (high R-factor;
Campbell andNaldrett, 1979), whereas inless oxidizedor S-richsystems,
larger volumes of Cu-rich but HSE-poor sulde would exsolve (low R-
factor; Fig. 2). In particular, the proportion of suldes exsolving fromarc
magmas may be very sensitive to small changes in their oxidation state,
because of the rapid change from sulde to sulfate dominance in
magmatic systems between FMQ+1 and +2 (Jugo et al., 2010). The
9 J.P. Richards / Ore Geology Reviews 40 (2011) 126
oxidation state of the mantle wedge will depend on the character of the
ux from the subducting slab (e.g., a higher proportion of subducted
organic-rich sediment would lead to lower oxidation states; Wang et al.,
2007a); this property is therefore likely to have a characteristic average
value along any given arc at any particular period of time.
Variations in oxidation state over typical ranges for arc magmas
(FMQ=0 to +2; Ballhaus, 1993; Brandon and Draper, 1996; Blatter
and Carmichael, 1998; Malaspina et al., 2009; Parkinson and Arculus,
1999; Rowe et al., 2009) will not greatly affect the potential to formsyn-
subduction porphyry Cu(Mo) deposits, but might control the Cu/HSE
ratio in later magmas formed by post-subduction melting of these
sulde-bearing residues. Specically, Au-rich (low Cu/Au) post-subduc-
tion porphyries might form in settings where previous arc magmatism
was relatively oxidized (sparse but HSE-rich sulde residue), whereas
Au-poor (high Cu/Au) post-subduction porphyries might form where
previous arc magmatismwas relatively reduced (more abundant Cu-rich
sulde residue). Such a mechanism might also explain why coeval belts
of porphyry deposits tend to have characteristic Cu/Au ratios.
Finally, partial melting of predominantly reduced, sulde-rich crustal
rocks in orogenic settings may lead to chalcophile and siderophile
element-depleted, but potentially lithophile element-rich S-type
magmas (see Section 2.6).
2.6. Crustal melting during post-collisional stress relaxation
Collisional orogens commonly undergo crustal thickening followed
by extensional or transpressional collapse. Bimodal magmatism is
characteristic of such tectonic settings, resulting frompartial melting of
pelitic protoliths in the deep crust triggered by the heat fromupwelling
asthenospheric melts (Hildreth, 1981). Peraluminous S-type granites
(Chappell and White, 1974) are subsequently emplaced as large
batholith complexes in the mid- to upper orogenic crust (e.g., the
Hercynian peraluminous granites of Europe; Barbarin, 1996; Clemens,
2003; Darbyshire and Shepherd, 1994; Harris et al., 1986; Wyllie et al.,
1976). These granites tend to be enriched in lithophile rather than
chalcophile elements, reecting their crustal origins, and may generate
magmatichydrothermal deposits containingSn, W, U, Mo, REE, Li, Be, B,
and F.
2.6.1. Sources of metals
Tin and especially tungsten commonly accompany molybdenum in
porphyry deposits as trace metals and byproducts, but they also form a
class of porphyries on their own, associated with S-type granites in
continental orogens (Hart et al., 2005; Ishihara, 1981; Ishihara and
Murakami, 2006; Kerrich and Beckinsale, 1988; Kirkham and Sinclair,
1996; Lehmann, 1982). The Hercynian tin granites of Europe, and the
Bolivian and SE Asian tin belts are examples of such deposits, with
mineralization occurring in skarns and greisens around the granite
intrusions, and to a lesser extent as internal stockworks and
disseminations (e.g., ern et al., 2005; Meinert et al., 2005). As with
the source magmas, metals in these deposits appear to be predomi-
nantly of crustal origin (Hedenquist and Lowenstern, 1994). For
example, in a recent assessment of the source of Sn in the Cornubian
batholith of SW England, Williamson et al. (2010) concluded that all of
Fig. 4. Post-subduction tectonic environments conducive to the formation of porphyry and epithermal deposits by remobilization of previously subduction-modied lithosphere
(modied fromRichards, 2009). (a) Porphyry CuMo deposits formed in normal arc settings; a continental arc is shown, but similar processes can occur in mature island arcs. (bd)
During post-subduction tectonic processes, previously subduction-modied 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 CuMo left behind in residual sulde phases by arc
magmatism, leading to the potential formation of porphyry CuAuMo 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.
10 J.P. Richards / Ore Geology Reviews 40 (2011) 126
the Sn could have been extracted from the crustally-derived granites.
Uraniumis also signicantly enriched in crustal rocks versus the mantle
(0.91 ppm versus ~0.02 ppm, respectively; Taylor and McLennan,
1985), and so is unlikely to have a mantle source in such deposits.
However, Dietrich et al. (1999) have suggested a possible role for
mantle-derived magmas in triggering volatile (and metal) release from
evolved, felsic magmas in the Bolivian tin belt, and Walshe et al. (2011)
have identied a mantle Nd isotopic signature in tin granites from
eastern Australia.
In contrast, in the case of W skarns associated with Mo mineraliza-
tion in calc-alkaline I-type magmas, a shared mantle origin with Mo
might be indicated(e.g., Newberry andSwanson, 1986), consistent with
the similar siderophile tendencies of these two elements, and their
position in the periodic table (group VIB).
3. Behavior of metals during magma fractionation and uid
exsolution in the upper crust
Key to the formation of magmatichydrothermal deposits of
chalcophile and siderophile elements in the upper crust is the lack of
signicant saturation with and loss of sulde phases prior to aqueous
volatile exsolution from a cooling magma (Candela, 1989b, 1992;
Candela and Holland, 1986; Candela and Piccoli, 2005; Richards, 1995;
Richards and Kerrich, 1993; Spooner, 1993). As discussed in Sections
2.1.1 and 2.5.1, chalcophile and siderophile elements partition
strongly into sulde phases exsolving or crystallizing from silicate
melts. Thus, if extensive fractionation and removal of magmatic
sulde 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
sulde deposits from relatively reduced mac magmas (e.g., Naldrett,
1989), and perhaps explains the deciency of chalcophile elements
such as Cu and Au in more reduced, evolved, SnW-bearing S-type
magmas (Blevin and Chappell, 1992; Hedenquist and Lowenstern,
1994).
In more oxidized subduction-related or post-subduction magmas,
although sulde saturation likely occurs at some point during their
evolution (as indicated by the common presence of sparse sulde
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 suldes physically separate from the magma, or at least
not in large amounts. Instead, small sulde droplets or crystals may be
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 ux. Indeed, several authors have argued that pre-concen-
tration of ore metals in magmatic sulde phases may be an important
step in porphyry metallogenesis (e.g., Jenner et al., 2010). In these
models, sulde 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
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
argued that this process, while it may occur, is not critical to metal
behavior in magmatichydrothermal systems, and that direct parti-
tioning from the silicate melt to the hydrothermal uid phase is the
dominant mechanism (e.g., Audtat 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 sulde
fromarc magmas at depth in lower crustal MASHcumulate zones may
provide a source of metals (especially HSE) for later post-subduction
magmas, but that this process may not substantially affect the Cu
content of the original arc magmas.
Regardless of the exact role of magmatic suldes, the ultimate
relationship is that of partitioning of metals between silicate melts and
exsolving hydrothermal uids, with suldes as a possible intermediary
step. Our current understanding of these partitioning processes is now
quite advanced following several decades of hydrothermal experiments
(e.g., Candela and Piccoli, 1995) and more recently the advance of
quantitative single uid and melt inclusion analysis (e.g., Heinrich et al.,
2003a,b), which has enabled direct measurement of metal contents in
ore-forming uids and melts. In the following sections, I reviewsome of
the key factors in metal solvation and transport in magmatic
hydrothermal uids.
3.1. Partitioning of metals from magma into exsolving hydrothermal
uid
Subduction-related magmas commonly contain at least 4 wt.% H
2
O
during crustal ascent, as evidenced by the presence of hornblende and
biotite phenocrysts in many andesitic volcanic rocks and arc plutons
(Burnham, 1979; Naney, 1983; Rutherford and Devine, 1988). In
response to the decreasing solubility of water in silicate melts as
pressure decreases, such hydrous magmas inevitably exsolve an
aqueous volatile phase upon emplacement at shallowcrustal levels or
on eruption (Burnham, 1979, 1997; Eichelberger, 1995). This process
has in the past rather confusingly been called rst and second boiling
(although neither process is technically boiling), the rst event
occurring during ascent and depressurization of the magma, and the
Chalcopyrite
Chalcopyrite
Pyrite
Pyrrhotite
(a)
(b)
Fig. 5. Reected light photomicrographs of sulde inclusions in amphibole-rich lower
crustal arc cumulates from: (a) the Talkeetna arc, Alaska; (b) the Bonanza arc,
Vancouver Island, Canada (samples courtesy A. Greene and D. Canil, respectively).
11 J.P. Richards / Ore Geology Reviews 40 (2011) 126
second occurring after emplacement as crystallization progressively
increases the concentration of volatiles in the residual melt (Candela,
1989a). In reality, volatile exsolution is probably a more-or-less
continuous process during arc magma ascent and cooling, starting at
depth with the exsolution of relatively insoluble CO
2
(Blundy et al.,
2010; Holloway, 1976; Lowenstern, 2001; Wallace, 2005). However,
the bulk of the magmatic water content likely exsolves relatively
rapidly as the magma approaches its solvus, at depths of ~510 km,
depending on magma composition and water content (Burnham,
1979).
The composition of this exsolved magmatic volatile phase is
dominantly aqueous, containing sulfur species (predominantly SO
2
in
oxidized systems, but also some reduced S species), CO
2
, NaCl, KCl, HCl,
and metal chlorides. The exact composition depends on many variables,
including the depth of exsolution and the magma composition
(especially the initial magmatic Cl/H
2
Oratio andalkali content; Candela,
1989c; Candela and Piccoli, 2005; Cline and Bodnar, 1991; Webster,
1992), but typical estimates for a single-phase supercritical uid
exsolved at depths below the H
2
ONaCl solvus are ~213 wt.% NaCl
equivalent (average 5 wt.% NaCl equivalent) with minor CO
2
(Audtat
and Pettke, 2003; Audtat et al., 2008; Burnham, 1979; Candela, 1989c;
Hedenquist et al., 1998; John, 1991; Redmond et al., 2004), and up to
1.3 wt.% Cu and 0.3 wt.% Fe (Klemmet al., 2007; Rusk et al., 2004, 2008;
Sawkins and Scherkenbach, 1981). These high observed metal solubil-
ities are consistent with or exceed experimental observations and
theoretical predictions based on chloride complexing alone (e.g.,
Candela and Holland, 1984, 1986), and suggest that other volatile
ligands such as sulde species may enhance the solubility of chalcophile
elements such as Cu and Au in high temperature aqueous uids (e.g.,
Heinrichet al., 1992; Pokrovski et al., 2005, 2008; Seo et al., 2009; Simon
et al., 2006; Zajacz et al., 2008, 2011).
Shallowly emplaced magmas will exsolve uids under pressure
temperature (PT) conditions that lie within the two-phase liquid
vapor eld for the bulk uid composition, resulting in immediate
formation of an immiscible low salinity vapor and high salinity brine
(Fig. 6a).
The critical point in the H
2
ONaCl system (which is commonly
used as a proxy for magmatic uids) occurs between ~1.0 and 1.4 kb
for uid temperatures between 600 and 800 C (the typical
temperature range for uids exsolved from intermediate to felsic
magmas) (Pitzer and Pabalan, 1986; Sourirajan and Kennedy, 1962),
which is equivalent to depths of between ~3 and 5 km at lithostatic
pressures. Most porphyry deposits and their host plutons are
emplaced at depths between 1 and 6 km (Seedorff et al., 2005), so
uids exsolving directly from magmas at these depths will typically
form immiscible liquid and vapor plumes (e.g., Henley and McNabb,
1978; Nash, 1976). However, the bulk of the uids and metals in
porphyry deposits are likely initially sourced at deeper levels (5
10 km, as noted above) from larger volumes of magma in mid- to
upper crustal batholithic complexes (Candela and Piccoli, 2005; Cloos,
2001; Damon, 1986; John, 1991; Richards, 2003, 2005; Shinohara and
Hedenquist, 1997), at which depths the uids will be supercritical. As
these deep uids rise into shallow cupola zones extending above the
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).
Supercritical uids are highly mobile (e.g., Coumou et al., 2008; Dunn
and Hardee, 1981; Norton and Dutrow, 2001) and behave differently in
terms of magmauid partitioning compared to uids exsolved at
shallower depths in the two-phase eld. In particular, Henley and
McNabb (1978) suggested that the higher density and viscosity of saline
brine condensates might restrict their ow, leaving them as a dense
residual liquid in the deeper parts of evolving magmatic hydrothermal
systems (see also: Lewis and Lowell, 2009; White et al., 1971). The lower
density vapor or supercritical uid would be expected to be highly
upwardly mobile, as well as larger interms of bothvolumeandmass than
the brine phase (assuming an initial bulk salinity below ~20 wt.% NaCl),
andsohas muchgreater potential as anefcient transportingmediumfor
ore components. However, until fairly recently, it was assumed that low
salinity vapors would not have the capacity to dissolve large quantities of
base metals as chloride complexes, andthe brine phase was therefore the
favored ore-forming medium (e.g., Bodnar and Beane, 1980; Cline and
Bodnar, 1991; Eastoe, 1982; Hedenquist and Richards, 1998; Moore and
Nash, 1974; Nash, 1976; Shinohara, 1994; Williams et al., 1995).
Observations of chalcopyrite crystals trapped in some vapor-rich uid
inclusions (e.g., Bodnar and Beane, 1980) were explained by some as
products of heterogeneous trapping (see discussion in Mavrogenes and
2000
1500
1
0
0
0
5
0
0
0
9
0
0
8
0
0
7
0
0
6
0
0
5
0
0
4
0
0
3
0
0
2
0
0
1
0
0 0
2
0
4
0
6
0
8
0
1
0
0
P

(
b
a
r
s
)
W
t
.
%

N
a
C
l
0
V
+
L
V-L isotherm (vapour)
V-L isotherm (liquid)
Critical / boiling curve
H
2
O critical point
Halite saturation
Salinity isopleth
Supercritical fluid
Liquid path
Vapour path
Coexisting vapour+liquid
Shallow
magmatic fluid
exsolution
L
+
H
a
l
i
t
e
V
+
H
a
l
i
t
e
Supercritical
fluid intersects
2-phase surface
Deep
supercritical
magmatic
fluid exsolution
(10 wt.% NaCl)
P
o
r
p
h
y
r
y
(a)
2
1
Liquid phase
progressively
separates from
vapour
2000
1500
1000
5
0
0
0
9
0
0
8
0
0
7
0
0
6
0
0
5
0
0
4
0
0
3
0
0
2
0
0
1
0
0 0
2
0
4
0
6
0
8
0
1
0
0
T
(C
)
T
(C
)
P

(
b
a
r
s
)
W
t
.
%

N
a
C
l
0
V
+
L
L
+
H
a
l
i
t
e
V
+
H
a
l
i
t
e
V-L isotherm (vapour)
V-L isotherm (liquid)
Critical / boiling curve
H
2
O critical point
Halite saturation
Salinity isopleth
Supercritical fluid
Liquid path
Vapour path
Contraction path
Coexisting vapour+liquid
Deep
supercritical
magmatic
fluid exsolution
(10 wt.% NaCl)
Vapour phase
departs from
2-phase surface
Supercritical
fluid never
intersects
2-phase
surface
Supercritical
fluid intersects
2-phase surface
Low to
moderate
salinity
liquids
4
3
E
p
i
t
h
e
r
m
a
l
(b)
Liquid phase
progressively
separates from
vapour
Fig. 6. PTX
NaCl
phase diagram, modied fromDriesner and Heinrich (2007), illustrating
uid pathways for a magmatichydrothermal uid exsolved with an initial bulk salinity of
10 wt.% NaCl. (a) Early, high thermal gradient uids exsolved from deeply (path 1) and
shallowly (path 2) emplaced magmas (porphyry environment). (b) Late, low thermal
gradient uids exsolved from deeply emplaced magma (high-suldation epithermal
environment). Path 3 illustrates the supercritical uid contraction path proposed by
Hedenquist et al. (1998), whereas path 4 illustrates a slightly steeper thermal gradient
with brief intersection of the 2-phase (LV) eld, as proposed by Heinrich et al. (2004).
Notethat inthetwo-phase eld, thedensesalineliquidphase progressivelyseparates from
the vapour phase, and the two-phase pathways shown do not represent a closed system.
See text for discussion.
12 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Bodnar, 1994). The advent of quantitative analysis of single uid
inclusions by synchrotron X-ray microprobe, proton-induced X-ray
emission spectroscopy (PIXE), and laser ablation-inductively coupled
plasma-mass spectrometry (LA-ICP-MS) revealed that, contrary to these
earlier assumptions, considerable amounts of metal, including CuandAu,
were indeed consistently present insome vapor-rich inclusions (Audtat
et al., 2000; Harris et al., 2003; Heinrich et al., 1992, 1999; Klemm et al.,
2007; Lowenstern et al., 1991; Simon et al., 2005, 2006; Ulrich et al.,
2001), and subsequent work has shown that much of this metal content
is transported as sulde complexes rather than as (or in addition to)
chloride complexes (Cauzid et al., 2007; Heinrich et al., 1992, 1999;
Pokrovski et al., 2005, 2008; Seoet al., 2009; Simonet al., 2006; Zajacz and
Halter, 2009; Zajacz et al., 2010).
Early resistance to these ideas was, I believe, at least in part due to
confusions of terminology: most papers dealing with this subject have
referred to metal solubility and transport ina vapor phase, suggesting to
the unwary a lowdensity, lowsalinitygas, whereas for the most part the
uids in question were either single-phase supercritical uids, or
relatively highdensityvapors just belowtheir critical points. Under such
high PT conditions, vapors are almost as saline as the initial single-
phase uid from which they evolved, and therefore they still contain
plenty of chloride for base metal complexing. But perhaps more
importantly, as noted above, these vapors will also contain a high
concentration of volatile sulfur species, which now appear to be
essential for the efcient solvation of chalcophile elements under high
PT conditions. When combined with the higher mass proportion of the
vapor phase (versus brine condensate) andits highmobility, a model for
transport of the bulk of the metal ux in porphyries by relatively dense
vapors or supercritical uids is now well established (Klemm et al.,
2007; Landtwing et al., 2010; Williams-Jones and Heinrich, 2005).
The rapid reduction in the efciency of transport of metals as the
vapor plume rises, cools, and becomes less dense by brine condensation,
may explain precipitation of the bulk of Cu, Mo, and some Au over a
relatively narrowtemperature interval between425320 C(Hemley et
al., 1992; Klemmet al., 2007; Landtwing et al., 2005). At shallower depths
and lower temperatures, the vapor phase may become too dilute to
transport signicant amounts of base metals as chloride complexes, but
may continue to carry some metals such as Au, Cu, As, and Sb as sulde
complexes (Deditius et al., 2009; Simon et al., 2006, 2007), eventually
either venting them to the surface in high-temperature fumaroles (e.g.,
Chaplygin et al., 2007; Hedenquist et al., 1994a; Symonds et al., 1987;
Tarana et al., 1995; Tessalina et al., 2008) or precipitating them in the
near-surfacehigh-suldationepithermal environment (seeSection4.2.1;
Deditius et al., 2009; Hedenquist et al., 1993, 1994b; Heinrich et al., 1999,
2004; Larocque et al., 2008; Murakami et al., 2010; Pudack et al., 2009;
Williams-Jones and Heinrich, 2005).
4. Magmatichydrothermal ore formation
The focus of this paper is on the ux of metals in subduction-
related magmatic systems, but this would be of little practical interest
if that ux did not ultimately lead to ore formation. Thus far, we have
focused on the importance of rstly not losing signicant amounts of
metal to a fractionating or residual sulde phase, and then efciently
partitioning those metals into a highly mobile aqueous uid phase.
What subsequently happens to that uid phase dictates whether
economic concentrations of metals are precipitated (grade), whereas
the scale of the magmatic and derivative hydrothermal system
controls the total amount of metals precipitated (tonnage).
4.1. Porphyry Cu ore formation
In a landmark paper, Cline and Bodnar (1991) presented a model
for the evolution of magmatic-hydrothermal systems from initial
Fig. 7. Schematic cross-section through a typical coupled arc batholithcupolavolcanic system, with associated porphyry CuAu and linked high suldation CuAu epithermal
deposits. Also shown are the thermal structure, uid ow 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 eld of view, with greatest intensity (epidote, actinolite)
close to the intrusions, fading to background distally. Modied from Richards (2005); sources: Sillitoe (1973, 2010), Dilles (1987), Shinohara and Hedenquist (1997), Hedenquist et
al. (1998), and Fournier (1999).
13 J.P. Richards / Ore Geology Reviews 40 (2011) 126
aqueous uid exsolution to cooling and mineral precipitation. A key
nding of this work was that, although there are many variables that
can affect the specic evolutionary path of a given system, an
economic porphyry Cu deposit can potentially be formed from quite
small volumes of typical andesitic arc magma. For example, the
authors showed that 1590 km
3
of andesitic magma initially contain-
ing 50 ppm Cu, 2.5 wt.% H
2
O, and with a Cl/H
2
O ratio=1, is sufcient
to generate a 250 Mt deposit with a grade of 0.75 wt.% Cu. The range of
required magma volumes reects variables such as the depth of
magma emplacement and the compatibility of Cu with fractionating
mineral phases (larger volumes are required if Cu is compatible with
early fractionating silicate, oxide, or sulde minerals). Details of the
model, later updated in Cline (1995), show that magmas emplaced at
moderate depths (1.0 to 2.0 kb, equivalent to depths of ~48 km)
most efciently partition Cu into early saturating saline uids,
whereas Cu and Cl are only released from shallowly emplaced
(0.5 kb, or ~2 km) magmas during late stages of crystallization (see
also Candela, 1989b). (Note that these depths reect the locus of uid
and metal exsolution from the magma under lithostatic pressure
conditions, as opposed to the depth of subsequent hydrothermal
metal deposition, which will be at shallower levels and likely under
hydrostatic pressure conditions; Fournier, 1999.) In large measure,
these differences reect the change in properties of the magmatic
hydrothermal uid, which will separate initially as a moderately
saline supercritical uid in deeper systems (Fig. 6a, path 1), but as
immiscible vapor and brine at shallower levels (Fig. 6a, path 2). Cline
(1995) concluded that optimum conditions for porphyry Cu ore
formation are obtained where magma containing ~4 wt.% H
2
O and
with a high initial Cl/H
2
O (typical of most arc magmas) is emplaced at
moderate crustal depths, thereby maximizing the efciency of Cu
partitioning into an early saline uid phase.
The results of these studies indicate that no special conditions or
magmatic metal enrichment are required to formeven large porphyry
Cu deposits. Instead, the question is rather one of process efciency:
where (what depth) and how are metalliferous uids released and
channeled? This nding is of fundamental importance from an
exploration perspective, because it shifts the focus from seeking
anomalously metal-rich magmas (which have proven elusive; Jenner
et al., 2010) to searching for optimal ore depositional settings in
otherwise normal tectonomagmatic environments (e.g., Tosdal and
Richards, 2001).
Cloos (2001), Shinohara et al. (1995), andShinohara andHedenquist
(1997) considered the question of process efciency from the
perspective of physical separation and focusing of volatile release
from the magma. Cloos (2001) suggested that the classic cupola shape
(Norton, 1982) of porphyry systems above larger batholithic magma
chambers reects convective circulation of bubbly magma into the
shallowapical parts of these systems (at 13 kmdepth), where volatiles
physically separate from the melt and coalesce to form a discrete uid-
lled cupola. The now-dense, volatile-depleted magma forms a
downward return ow to complete the convective cycle. In contrast,
Shinohara and Hedenquist (1997) envisaged uids physically separat-
ing from the magma at greater depth within the underlying magma
chamber, and rising as a discrete plume up apical channelways formed
by fractures and dikes in the brittle carapace (Fig. 7).
A key aspect of both of these models is that volatile separation, and
Cu partitioning, occurs from a much larger volume of magma
(emplaced at deeper levels) than that preserved and commonly
visible within the shallow-level ore body. In Cloos's (2001) model,
volatile-rich, bubbly magma rising from the underlying batholith
convects through the cupola zone where it releases its uids, whereas
in Shinohara and Hedenquist's (1997) model, vesiculation and
convective circulation occur in the underlying magma chamber itself,
and uids are released as a plume into the base of the apical dike
system. Combining these models with (Cline, 1995; Cline and
Bodnar's, 1991) calculations suggests that maximum ore-forming
efciency in porphyry systems is likely achieved where volatile
saturation occurs in large (100 km
3
) mid- to upper crustal magma
chambers at depths 6 km, containing moderately hydrous (N4 wt.%
H
2
O) and Cl-rich magmas. These uids either rise as bubbly magma or
as a separate volatile plume into the apical parts of the system
where decreasing pressure and temperature cause deposition of Cu
and MoAu (Candela, 1989b; Shinohara et al., 1995).
Focusing of magma ascent and uid ow into narrow apical
regions, or cupolas, is likely to be a function of structure in the brittle
rocks overlying the batholithic system (Tosdal and Richards, 2001,
and references therein). Shallow crustal magma emplacement will
cause extensional doming in the cover rocks, with dilational fault
zones providing high-permeability pathways for uid and magma
ascent (Burnham, 1979). Evidence from the dike emplacement
literature suggests that such fractures may rst be opened and
propagated by volatile pressure, and only later lled by more viscous
magma (Burnham, 1979; Carrigan et al., 1992; Rubin, 1995). This
raises the intriguing possibility that the cylindrical shapes of many
porphyry stocks may have arisen rst as breccia pipes or diatremes
bored out by rapidly escaping volatiles, only later to be back-lled
with porphyritic magma (Fig. 8; Norton and Cathles, 1973; see also
Fig. 2 in Anderson et al., 2009, and Fig. 8 in Sillitoe, 2010, and Fig. 17 in
Vry et al., 2010).
To some extent, focusing of magma and uid ow along narrow
conduits may be self-organizational, because once initial channel
ways have developed, they will represent high-permeability path-
ways and are likely to thermally weaken the wall rocks and promote
further fracturing and channeling.
Narrow focusing of apical fracturing and subsequent uid ow are
likely to be critical to the formation of high grade porphyry deposits
(e.g., El Teniente, Chile; Vry et al., 2010), while multiple breccia/
intrusive events can potentially increase tonnage (provided that later
events do not destroy earlier mineralization).
Given suitable uid ow focusing, three further factors combine to
cause maximum efciency of metal deposition within relatively small
volumes in porphyry CuMoAu deposits. All three effects are
related to the steep temperature gradient in the cupola zone (Fig. 7),
and they therefore control the vertical range of ore deposition. On the
other hand, uid focusing controls the lateral extent of mineralization.
In combination, the highest grades will occur where ore deposition is
both focused laterally and restricted vertically.
The rst factor is that Cu solubility (as chloride species) decreases
dramatically as uids cool through the temperature interval ~400 to
300 C (Crerar and Barnes, 1976; Hemley et al., 1992; Klemm et al.,
2007; Landtwing et al., 2005; Xiao et al., 1998). Given the very sharp
temperature gradient implied by near-surface emplacement of
magma (Fig. 7), this temperature interval will correspond to a narrow
depth range, likely with 1 or 2 km of the surface (although it may
extend to greater depths with time as the magmatichydrothermal
systembegins to cool). This depth range is typical for ore formation in
many porphyry systems.
The second factor is that SO
2
dissolved in the magmatic
hydrothermal uid phase progressively disproportionates to H
2
S
and H
2
SO
4
as the uid cools below ~400 C (Holland, 1965; Kusakabe
et al., 2000; Reeves et al., 2010; Sakai and Matsubaya, 1977):
4SO
2
+ 4H
2
O H
2
S + 3H
2
SO
4
: 1
This reaction generates both hydrogen sulde, which initiates
abundant precipitation of sulde minerals (i.e., chalcopyrite, pyrite,
molybdenite), and also sulfuric acid, which causes early deposition of
large volumes of anhydrite in the potassic alteration zone, and
progressively increasing degrees of hydrolytic alteration (an initial
shift from feldspar-stable potassic alteration, to muscovite/sericite-
stable phyllic alteration). Consequently, the bulk of Cu-sulde miner-
alization occurs at the low-temperature, late-stage end of the potassic
14 J.P. Richards / Ore Geology Reviews 40 (2011) 126
alteration phase, just prior to the onset of phyllic alteration (i.e.,
corresponding to the ore shell of the classic Lowell and Guilbert, 1970,
porphyry Cu model).
The third factor is the increased permeability over this tempera-
ture range caused by a combination of the transition in silicate rocks
from ductile to brittle behavior at temperatures (between ~400
350 C; i.e., the brittleductile transition; Fig. 7; Cathles, 1991;
Fournier, 1999; Landtwing et al., 2005), and a window of retrograde
silica solubility (between ~550350 C; Fournier, 1985). Not only do
these processes result in the formation of open-standing brittle veins
and porosity, thereby facilitating rapid upward uid ow and
wallrock permeation (e.g., the crackle breccias, stockworks, and
disseminated mineralization textures so characteristic of porphyry Cu
deposits), but this transition also represents the boundary between
lithostatic and hydrostatic uid pressures (a pressure differential of
~3). Sudden depressurization of uids across this boundary can be
expected to have major effects on uid properties, including phase
separation (Fig. 6) and consequent changes in metal solubility (e.g.,
Landtwing et al., 2005, 2010; Murakami et al., 2010).
Incombination, these four factors, (1) spatial focusingof uidowin
narrow cupolas, (2) reduction of metal solubility, (3) increased
dissolvedsuldeactivity, and(4) permeabilityincrease due totransition
from ductile to brittle fracturing and retrograde silica solubility (with a
large pressure drop), serve to narrowly focus Cu-sulde mineralization
both laterally and vertically within cupola zones above large mid- to
upper crustal batholithic complexes. The most likely reasons for
otherwise prospective porphyry systems to be unproductive will be
either a failure to focus uid ow, or simply insufcient uid supply
(likely due to an insufciently large underlying magmatic system).
4.2. Epithermal CuAu ore formation
4.2.1. High-suldation epithermal CuAu deposits
As noted in Section 3.1, although the bulk of Cu (and Mo) appears to
be precipitatedover the temperature interval 425320 C, some Cuand
other metals such as Au, Sb, and As may remain in solution as sulde
complexes, to be carried into the shallow epithermal regime. Observa-
tions from rare uid inclusions in high-suldation epithermal CuAu
deposits suggest that the ore-forming uid was a low- to moderate-
salinity liquid (0.2 to 4.5 wt.%NaCl equivalent; Hedenquist et al., 1994b;
Mancano and Campbell, 1995), which paragenetically post-dates
advanced argillic alteration formed by highly acidic magmatic gasses
(Arribas, 1995; Hedenquist et al., 1994b; Stoffregen, 1987). This
apparent inconsistency has been explained by Hedenquist et al.
(1998), Heinrich et al. (2004), and Heinrich (2005) in terms of
contraction of a moderate salinity supercritical magmatic uid or
Porphyry
intrusion
Mixed magmatic-
hydrothermal
breccia
Cuspate
intrusive
border
Porphyry
intruding
comagmatic
breccia
Cuspate
porphyry
clast
Mixed magmatic-
hydrothermal
breccia
Hydrothermal
quartz filling
cavity
Limestone
wallrock
clast
Porphyry
dike
(a) (b)
(c)
~20 cm
Fig. 8. Photographs of porphyritic magma invading co-magmatic hydrothermal breccia pipe (a, c) and breccia vein/dike (b) from the Pachapaqui AgCuPbZn deposit, Pru. 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 lled with hydrothermal quartz, reecting the
role of uids in breccia formation. Scale in (c) is in centimeters.
15 J.P. Richards / Ore Geology Reviews 40 (2011) 126
vapor by rapid cooling during ascent such that it does not touch, or
barely touches, the two-phase solvus (Figs. 6b and 9, paths 3 and 4,
respectively). Because of the curvature of the solvus crest (critical curve)
to lower salinities at low temperatures and pressures, this moderately
saline uid will lie on the liquid side of the solvus at shallow depths,
although it may have contracted from what was originally a vapor or
supercritical uid phase at higher temperatures and depths.
Heinrich et al. (2004) specically suggested that brief intersection
with the solvus at high pressures and temperatures (Figs. 6b and 9, path
4) might be animportant way to shed chloride-complexed components
suchas Fefromthe vapor ina brine condensate, leaving bisulde ligands
free to bond with Cu and Au in the residual vapor and transport it to
shallower (epithermal) levels. This model requires that the vapor leaves
the two-phase surfaceagainby coolingat pressure, therebypassingover
the crest of the solvus (the critical curve; (Figs. 6b and 9, path 4); upon
subsequent ascent and depressurization, this uid will be a liquid, as
described above. Heinrich et al. (2004) argued that the Fe-rich brine
condensation step is essential for retention of Au (and Cu) in the vapor
phase, because otherwise Fe will tend to precipitate as CuFe-sulde
minerals and strip the uid of bisulde ligands, thus causing Au to co-
precipitate at depth (a possible mechanism for the formation of
porphyry CuAu deposits). However, this condensation step must
then be followed by cooling while still at depth, in order to lift the
vapor phase off the two-phasesurface, andallowit tocontract toa liquid
(Figs. 6b and 9, path 4).
A more detailed analysis of the phase diagram shown in Fig. 6
reveals that a typical magmatic uid of 213 wt.% NaCl would have to
rst intersect the solvus at temperatures above ~400500 C for this
model to work optimally, otherwise it would fall on the liquid side of
the two-phase eld, and would become more saline by boiling off a
dilute vapor (Figs. 9 and 10). Assuming that this did indeed happen,
then the smaller (by mass) vapor component would have to leave the
two-phase surface again at pressure before cooling below ~400 C,
otherwise its salinity rapidly falls to sub-weight percent levels at
lower temperatures and lower pressures (Figs. 6 and 10), which are
inconsistent with uid inclusion evidence for lowto moderately saline
liquids in high-suldation epithermal ore formation (Hedenquist et
al., 1994b; Mancano and Campbell, 1995). Thus, the PT trajectory of a
uid that satises Heinrich et al.'s (2004) model for high-suldation
epithermal CuAu mineralization (Figs. 6b and 9, path 4) is somewhat
unique, and may occur only rarely or eetingly during the waning
stages of a cooling magmatichydrothermal system.
A more normal, or early ascent pathway for a magmatic
hydrothermal uid would be for it to rise more-or-less is enthalpically
or quasi-adiabatically along a steep PT gradient (e.g., Hemley and
Hunt, 1992; Henley and Hughes, 2000; Wood and Spera, 1984), and
therefore to dive deeply into the two-phase eld and separate into an
increasingly dilute vapor phase and a saline brine (Figs. 6 and 9, path
1), or even to boil dry to halite plus vapor (Figs. 6 and 9, path 2). Such
uid pathways might be consistent with the widespread and intense
s
u
d
i
l
o
s
e
t
i
n
a
r
g
t
e
W
s
u
d
i
l
o
s
e
t
i
r
o
i
d
o
n
a
r
g
t
e
W
Epithermal
Early
porphyry
veins
(2-phase
fluid)
Deep
magmatic
fluid
exsolution
3
4
1
V-L solvus (vapour)
V-L solvus (liquid)
B
r
i
t
t
l
e
(
h
y
d
r
o
s
t
a
t
i
c
)
D
u
c
t
i
l
e
(
l
i
t
h
o
s
t
a
t
i
c
)
1 wt.%
5
w
t.%
1
0

w
t
.
%
2
0

w
t
.
%
3
0

C
/
k
m

g
e
o
t
h
e
r
m
3
0
0

C
/k
m
g
e
o
th
e
rm
V+Halite
V
/
L
+
H
a
l
i
t
e
Main-stage
brittle
porphyry
veins
Shallow
magmatic
fluid
exsolution
2
Deep
single-
phase
fluid
D
e
p
t
h

(
k
m
)
1 (hyd)
2 (hyd)
1 (lith)
2 (lith)
8 (hyd) 3 (lith)
4 (lith)
5 (lith)
6 (lith)
7 (lith)
8 (lith)
3 (hyd)
4 (hyd)
5 (hyd)
6 (hyd)
7 (hyd)
9 (hyd)
10 (hyd)
0
T (C)
0 100 200 300 400 500 600 700 800 900 1000
P

(
b
a
r
s
)
500
0
1000
1500
2000
0 100 200 300 400 500 600 700 800 900 1000
11 (hyd)
12 (hyd)
13 (hyd)
14 (hyd)
15 (hyd)
Fig. 9. Pressure (depth)temperature section through the H
2
ONaCl phase diagram, with vapourliquid (VL) 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 temperaturedepth
ranges for supercritical magmatic uid exsolution, early high-temperature porphyry veins, later main-stage brittle porphyry veins, and epithermal mineralization are indicated. Four
uid PT paths are shown corresponding to: (1) a typical porphyry-forming uid path; (2) a shallow high-temperature path boiling to dryness (V+Halite eld); (3) the deep
contraction path of Hedenquist et al. (1998); and (4) the contraction with minor brine condensation path of Heinrich et al. (2004).
16 J.P. Richards / Ore Geology Reviews 40 (2011) 126
acidic (advanced argillic) alteration commonly found at shallowlevels
above porphyry systems, which is caused by acidic gasses (H
2
SO
4
,
HCl) condensing from a low-density vapor plume. As Heinrich et al.
(2004) noted, such acidic vapors do not appear to transport Au (or Cu)
effectively because bisulde (HS

) ligands are hydrolized to H


2
S. This
may explain why advanced argillic alteration caps are commonly
barren (in terms of Au and Cu) and merely generate permeability that
potentially focuses later ore-forming uid ow. Thus, mineralization
may only occur where later moderate salinity liquids have followed a
higher pressure, rather specialized cooling path, as described above
(Heinrich et al., 2004).
4.2.2. Low-suldation epithermal Au deposits (including alkalic-type
deposits)
Although low-suldation epithermal Au deposits are commonly
found in volcanic terrains, their link to magmatism is more tenuous
than high-suldation deposits, and there is commonly evidence for a
greater involvement of meteoric groundwater in their formation than
magmatic uids (e.g., Faure et al., 2002; Field and Fifarek, 1985; Heald
et al., 1987). Nevertheless, alkalic-type epithermal Au deposits, which
are mineralogically similar to low-suldation deposits (adularia and
sericite or roscoelite [vanadium mica] are stable), do show a
strong temporal and genetic relationship to alkalic magmas, typically
in relatively small and isolated intrusive complexes located in back-
arc or post-subduction settings (e.g., Jensen and Barton, 2000; Kelley
et al., 1998; Mller and Groves, 1993; Mutschler et al., 1985; Richards,
1995; Thompson et al., 1985).
Fluids in these alkalic-type deposits are commonly low- to
moderate salinity (010 wt.% NaCl equiv.), lowtemperature (typically
250 C) liquids, with evidence for decompressional boiling or uid
mixing as the prime ore depositional mechanism in high grade
breccias and veins (Jensen and Barton, 2000; Richards, 1995). Stable
isotopic compositions of these uids are generally ambiguous, and
permit interpretations of the involvement of either isotopically
exchanged meteoric waters or magmatic uids, or both (Ahmad et
al., 1987; Carman, 2003; Richards, 1995; Richards and Kerrich, 1993;
Ronacher et al., 2004; Scherbarth and Spry, 2006; Zhang and Spry,
1994). Similar stable isotopic data fromother low-suldation deposits
are commonly interpreted to reect a meteoric uid source because of
the absence of clearly coeval magmatism (as noted above). However,
the close link with magmatism in alkalic-type systems suggests that a
magmatic uid source is more likely (e.g., Carman, 2003; Scherbarth
and Spry, 2006; Simmons and Brown, 2007), and the vapor
contraction model described in Section 4.2.1 could explain the
observed characteristics of these ore uids (i.e., direct contraction to
a moderate salinity liquid from a high temperature magmatic uid at
depth). Because of the association of these deposits with relatively
small intrusive complexes, and therefore a smaller crustal thermal
anomaly, a shallower uid PT path with cooling at depth is more
likely, consistent with a model of vapor contraction.
C
r
i
t
i
c
a
l

c
u
r
v
e
L
+
H

(
4
0
0

C
)
V+H (400C)
V+L
(700C)
V
wt.% NaCl log (wt.% NaCl)
P

(
b
a
r
s
)
scale
change
100
0
200
300
400
500
600
700
800
900
1000
1100
1200
1300
700C
600C
500C
400C
Single
phase
mag-
matic
fluid
V+L
(600C)
V+L
(400C)
V+L
(500C)
V+H (700C)
V+H (600C)
V+H (500C)
L
+
H

(
5
0
0

C
)
L
+
H

(
7
0
0

C
)
L
+
H

(
6
0
0

C
)
Critical P
of H
2
O
L

(
4
0
0

C
)
L

(
5
0
0

C
)
L

(
7
0
0

C
)
L

(
6
0
0

C
)
-8 -7 -6 -5 -4 -3 -2 -1 0/1 10 20 30 40 50 60 70 80 90 100
100
0
200
300
400
500
600
700
800
900
1000
1100
1200
1300
-8 -7 -6 -5 -4 -3 -2 -1 0 10 20 30 40 50 60 70 80 90 100
V-L solvus (liquid)
V-L solvus (vapour)
V-H solvus (vapour)
Fig. 10. PressureX
NaCl
section through the H
2
ONaCl phase diagram, with vaporliquid (VL) solvi drawn at various temperatures (data fromDriesner, 2007; Driesner and Heinrich,
2007). Red curves indicate that the solvus phase is a vapor, blue curves that it is a liquid; below the vaporhalite (VH) 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 uids (213 wt.% NaCl) is shown in gray. Fluids that intersect the VL solvus above ~400500 C will be moderate-density vapors and will condense
a small amount of dense, saline liquid; uids that intersect the solvus below this temperature will be liquids and will boil off a dilute, low-density vapor phase.
17 J.P. Richards / Ore Geology Reviews 40 (2011) 126
5. Summary and conclusions
5.1. Sources of magmas and metals
Magmatichydrothermal porphyry CuMoAu, Au, Mo, and SnW
deposits (and related epithermal Au deposits), derive their metals from
their associated magmas. Withthe exceptionof porphyry SnWdeposits
that are associated with crustally derived S-type granites, most
other deposits in this grouping are formed by calc-alkaline to mildly
alkaline I-type granitoids directly or indirectly related to subduction.
Sources of these magmas include subduction-metasomatized astheno-
spheric mantle wedge, basaltic oceanic crust and/or seaoor sediments,
and, in post-subduction settings, subduction-modied upper plate
lithosphere. The majority of normal arc porphyry systems are generated
from hydrous mantle wedge melts that have interacted to varying
degrees with the upper plate lithosphere during passage towards the
surface. Assimilation and fractional crystallization processes fundamen-
tally change the composition of the primary basaltic arc magmas to
intermediate calc-alkaline compositions, withfractionated trace element
patterns and evolved (crustal) isotopic signatures.
Seaoor sediments and basaltic oceanic crust only melt under
unusually hot subduction zone conditions or locally at plate edges, and
despite widespread claims of the identication of slab melts (adakites)
in the literature based on high Sr/Y and La/Yb ratios in some evolved
granitoids, this process is unlikely to be a major contributor to arc
magmatism and metallogeny. Rather, these subtle trace element ratios
can be readily explained by fractionation of hornblendetitanite and
residual garnet, and suppression of plagioclase fractionation from
water-rich mantle wedge basaltic magmas. These trace element
characteristics are therefore an indicator of high magmatic water
content rather than being a source signature, and this likely explains the
common association of high-Sr/Y (i.e., hydrous) magmas with mag-
matichydrothermal ore deposits without magmatic water, such
deposits cannot form.
Potential sources of metals in arc magmas include the oceanic crust
and sediments (via dehydration uids or melting), the mantle wedge,
andthe upper platecrust. Fluidmobile elements suchas K, Rb, Cs, Ca, Sr,
Ba, U, B, Pb, As, Sb, Tl, and possibly Cu, Au, Re, and the Pd-group
elements, along withlarge amounts of H
2
O, Cl, andS, are uxedfromthe
dehydrating subducting slab into the mantle wedge, causing metaso-
matismand partial melting by lowering the peridotite solidus. Although
there is some evidence for slab-derived uid contributions of chalco-
phile and highly siderophile elements (Cu, Au, PGE) to the mantle
wedge, it is not clear that this is a necessary metallogenic step, the upper
mantle already containing signicant amounts of these elements. Likely,
a more important control on the metal content of subsequent partial
melts is the abundance and stability of residual sulde phases in the
asthenospheric mantle source. Under the high f
O2
and f
S2
conditions of
arc magmatism, sulfur will be dominantly present as sulfate and sulfate,
but saturation in small amounts of sulde phases is also likely. These
sulde phases will tend to deplete the magma in highly siderophile
elements (Au and PGE), but will not be present in sufcient volume to
signicantly depletethemagmainmore abundant chalcophileelements
such as Cu and Mo. Such magmas therefore have the potential
subsequently to formporphyry CuMo deposits. Thus, Cu (and perhaps
Mo) are thought to be predominantly derived from the mantle, plus or
minus contributions fromthe subductingslab. Goldandsilver present in
minor amounts in such deposits may also be derived from subduction
sources, although there is some evidence for additional contributions
from the upper plate crust (especially for Ag, and also perhaps Mo).
Arc-like magmas and related porphyry and epithermal ore deposits
also occur in post-subduction tectonic settings, such as subduction
reversal or migration, arc collision, continentcontinent collision, and
post-collisional rifting. They are distinguished fromnormal subduction-
related suites by slightly higher magmatic alkali (K
2
O and Na
2
O)
contents, and by the occurrence of Au-rich deposits (although normal
porphyryCuModeposits canalsooccur). Suchmagmaticmetallogenic
systems are thought to form by remelting of previously subduction-
modied upper plate lithosphere, and in particular the lower crustal
amphibolitic cumulate roots of former arc magmatic complexes.
Remelting can be triggered by crustal thickening and thermal rebound
following arc or continent collision, delamination of sub-continental
mantle lithosphere causing direct exposure of the lower crust to
asthenospheric temperatures and melts, and asthenospheric upwelling
during rifting of former arc crust. Sparse sulde phases in these arc
cumulates, residual from fractionation of previous arc magmas, will
likely be rich in chalcophile and highly siderophile elements. During
low-volume melting under relatively lowf
S2
conditions (in the absence
of a ux of S from active subduction), these sulde phases will likely
redissolve in the mildly alkaline partial melt, and may provide a source
for Au-rich (PGE) post-subduction porphyry CuAu and epithermal
Au systems: examples include the Roia Montan, Skouries, Kisladag,
pler, and Sari Gunay porphyry CuAu and epithermal Au deposits in
the Neo-Tethyan belt of Romania, Greece, Turkey, and Iran, and the
Grasberg, Ok Tedi, Porgera, Lihir, and Emperor porphyry CuAu and
epithermal Au deposits in the southwest Pacic. However, gold
enrichments may not occur where more abundant suldes were
present in the former arc complex, leading to more normal porphyry
CuMo systems: examples include the Kerman porphyry Cu belt of
central Iran, and the Gangdese porphyry Cu belt of Tibet.
Porphyry Mo and SnW deposits associated with felsic magmas in
continental interiors are thought to form mainly by partial melting of
continental crust during rifting to form S-type, lithophile element-rich
granitic magmas. A role for mac, mantle-derived magmas is suggested
by the common association with such rocks, but their role may be
predominantly as a heat source for crustal melting and a trigger for
volatile saturation and eruption, rather than as a unique source of
metals.
5.2. Porphyry and epithermal ore formation
In the porphyry and epithermal ore depositional environment, a
critical role is played by aqueous uids exsolving from hydrous
magmas emplaced in the mid- to upper crust. The PTX properties of
magmatic hydrothermal uids, approximated by the H
2
ONaCl
system, combined with volatile solubility in intermediate composition
magmas, suggest that uid saturation occurs at depths of 510 km in
the batholithic roots of arc magmatic systems. Volatile exsolution may
lead to the upward propagation of buoyant bubbly magma as dikes
and stocks intruded into the overlying shallow crust (with or without
subsequent eruption at surface), and/or the rapid ascent of a separate
volatile plume. Structural focusing and chanelling of these evolved
magmas and uids creates a cupola zone characterized by high
thermal gradients and uid ux. Metals (Cu, Mo, Au), which partition
strongly into the saline (213 wt.% NaCl equivalent) and S-rich
magmatic hydrothermal phase at high P and T in the underlying
batholithic magma chamber, experience rapid reduction in solubility
as these uids ascend, depressurize, and cool, with the bulk of Cu and
Mo being precipitated over a temperature range of 425320 C at 1
6 km (commonly 2 km) depth.
This depthtemperature interval is critical because it also repre-
sents: (1) the upward transition from ductile to brittle behavior in the
cover rocks (~400350 C), which facilitates fracturing and rapid uid
depressurization; (2) a window of retrograde silica solubility (~550
350 C), which enhances permeability and porosity for ore deposition;
and (3) the temperature range (b400 C) over which SO
2
in the uid
phase begins to disproportionate to H
2
S and H
2
SO
4
, which causes
precipitation of sulde minerals (chalcopyrite, pyrite, molybdenite, rare
bornite). This reaction also generates increasingly acidic uids, leading
to the characteristic progression from feldspar-stable alteration assem-
blages (potassic), through muscovite/sericite-stable assemblages
18 J.P. Richards / Ore Geology Reviews 40 (2011) 126
(phyllic), to clay- (argillic) and alunite-stable assemblages (advanced
argillic).
Depending on the depth of exsolution, the initial magmatic uid will
either be a supercritical uid(below~6 kmdepth) or will exist as a two-
phase moderate salinity vapor and high density brine. As the plume
ascends, it will intersect its solvus (in the case of a deeply exsolved
supercritical uid), and the vapor phase will become progressively less
saline through brine condensation. At the level of porphyry ore
formation, both the brine and vapor phase may contribute to metal
transport and deposition, although there is increasing evidence for the
importance of the vapor phase as a large-volume, highly upwardly
mobile transportation medium. However, as this vapor phase continues
to ascend, it will rapidly decrease in salinity, such that chloride-
complexed metals are unlikely to be transported by such uids to
shallow epithermal levels. It will also increase in acidity, thereby
reducing the solubility of bisulde-complexed metals such as Au (and
perhaps also Cu) by protonation of HS

to H
2
S. This acidic vapor phase is
responsible for the extreme acid leaching in advanced argillic lithocaps
above porphyry systems.
High-suldation epithermal CuAu deposits are hosted by these
highly permeable advanced argillic alteration zones, but appear to have
been formed by later, moderately saline (0.2 to 4.5 wt.% NaCl
equivalent), less acidic liquids. Two models have been proposed to
explain the origin of these paragenetically late mineralizing uids in
terms of contraction of a single phase (supercritical) magmatic uid,
with (Heinrich et al., 2004) or without (Hedenquist et al., 1998) brief
intersection of the solvus. Because of the topology of the PTX
NaCl
phase diagram, such uids contract to a liquid phase upon cooling at
pressure. Thus, these authors propose that high-suldation epithermal
CuAu deposits may be formed directly from late-stage magmatic
hydrothermal uids.
Although low suldation epithermal Au deposits have not been
discussed in detail in this paper, because most such deposits are not
directly related to porphyry-type magmatichydrothermal systems,
the vapor contraction mechanism might have applicability to alkalic-
type epithermal gold deposits, which do show a close genetic
relationship to post-subduction alkalic magmas.
Acknowledgments
I would like to thank Nigel Cook and Timothy Horscroft for inviting
me to submit this paper. Reviewpapers, by their nature, drawheavilyon
the work of others, and I would particularly like to acknowledge the
following people who have inuenced my thinking on this subject: P.
Candela, J. Cline, J. Dilles, J. Hedenquist, C. Heinrich, R. Sillitoe, and R.
Tosdal. An anonymous reviewer is thanked for helpful and constructive
comments. This work was funded by a Discovery Grant fromthe Natural
Sciences and Engineering Research Council of Canada.
References
Ackerman, L., Walker, R.J., Puchtel, I.S., Pitcher, L., Jelnek, E., Strnad, L., 2009. Effects of
melt percolation on highly siderophile elements and Os isotopes in subcontinental
lithospheric mantle: a study of the upper mantle prole beneath Central Europe.
Geochimica et Cosmochimica Acta 73, 24002414.
Aerts, M., Hack, A.C., Reusser, E., Ulmer, P., 2010. Assessment of the diamond-trap
method for studying high-pressure uids and melts and an improved freezing
stage design for laser ablation ICP-MS analysis. American Mineralogist 95,
15231526.
Ahmad, S.N., Rose, A.W., 1980. Fluid inclusions in porphyry and skarn ore at Santa Rita,
New Mexico. Economic Geology 75, 229250.
Ahmad, M., Solomon, M., Walshe, J.L., 1987. Mineralogical and geochemical studies of
the Emperor gold telluride deposit, Fiji. Economic Geology 82, 345370.
Aitcheson, S.J., Harmon, R.S., Moorbath, S., Schneider, A., Soler, P., Soria-Escalante, E.,
Steele, G., Swainbank, I., Wrner, G., 1995. Pb isotopes dene basement domains of
the Altiplano, central Andes. Geology 23, 555558.
Aizawa, Y., Tatsumi, Y., Yamada, H., 1999. Element transport by dehydration of
subducted sediments: implications for arc and ocean island magmatism. Island Arc
8, 3846.
Alonso-Perez, R., Mntener, O., Ulmer, P., 2009. Igneous garnet and amphibole
fractionation in the roots of island arcs: experimental constraints on andesitic
liquids. Contributions to Mineralogy and Petrology 157, 541558.
Alt, J.C., Shanks, W.C., Jackson, M.C., 1993. Cycling of sulfur in subduction zones: the
geochemistry of sulfur in the Mariana Island Arc and back-arc trough. Earth and
Planetary Science Letters 119, 477494.
Anderson, E.D., Atkinson Jr., W.W., Marsh, T., Iriondo, A., 2009. Geology and
geochemistry of the Mammoth breccia pipe, Copper Creek mining district,
southeastern Arizona: evidence for a magmatichydrothermal origin. Mineralium
Deposita 44, 151170.
Annen, C., Blundy, J.D., Sparks, R.S.J., 2006. The genesis of intermediate and silicic
magmas in deep crustal hot zones. Journal of Petrology 47, 505539.
Arculus, R.J., 1994. Aspects of magma genesis in arcs. Lithos 33, 189208.
Arribas Jr., A., 1995. Characteristics of high-suldation epithermal deposits, and their
relation to magmatic uid. In: Thompson, J.F.H. (Ed.), Magmas, uids, and ore
deposits: Mineralogical Association of Canada Short Course, 23, pp. 419454.
Audtat, A., 2010. Source and evolution of molybdenum in the porphyry Mo(Nb)
deposit at Cave Peak, Texas. Journal of Petrology 51, 17391760.
Audtat, A., Keppler, H., 2005. Solubility of rutile in subduction zone uids, as
determined by experiments in the hydrothermal diamond anvil cell. Earth and
Planetary Science Letters 232, 393402.
Audtat, A., Pettke, T., 2003. The magmatic-hydrothermal evolution of two barren
granites: a melt and uid inclusion study of the Rito del Medio and Canada Pinabete
plutons in northern New Mexico (USA). Geochimica et Cosmochimica Acta 67,
97121.
Audtat, A., Pettke, T., 2006. Evolution of a porphyry-Cu mineralized magma system at
Santa Rita, New Mexico (USA). Journal of Petrology 47, 20212046.
Audtat, A., Gnther, D., Heinrich, C.A., 2000. Magmatichydrothermal evolution in a
fractionating granite: a microchemical study of the SnWF-mineralized Mole
Granite (Australia). Geochimica et Cosmochimica Acta 64, 33733393.
Audtat, A., Pettke, T., Heinrich, C.A., Bodnar, R.J., 2008. The composition of magmatic-
hydrothermal uids in barren and mineralized intrusions. Economic Geology 103,
877908.
Ballhaus, C., 1993. Redox states of lithospheric and asthenospheric upper mantle.
Contributions to Mineralogy and Petrology 114, 331348.
Barbarin, B., 1996. Genesis of the two main types of peraluminous granitoids. Geology
24 (4), 295298.
Barbier, E., 2002. Geothermal energy technology and current status: an overview.
Renewable and Sustainable Energy Reviews 6, 365.
Barnes, S.-J., Naldrett, A.J., Gorton, M.P., 1985. The origin of the fractionation of
platinum-group elements in terrestrial magmas. Chemical Geology 53, 303323.
Barreiro, B.A., 1984. Lead isotopes and Andean magmagenesis. In: Harmon, R.S.,
Barreiro, B.A. (Eds.), Andean Magmatism Chemical and Isotopic Constraints.
Cheshire, Shiva, Nantwich, pp. 2130.
Beard, J.S., Lofgren, G.E., 1991. Dehydration melting and water-saturated melting of
basaltic and andesitic greenstones and amphibolites at 1, 3, and 6.9 kb. Journal of
Petrology 32, 365401.
Bergantz, G.W., Dawes, R., 1994. Aspects of magma generation and ascent in
continental lithosphere. In: Ryan, M.P. (Ed.), Magmatic Systems. Academic
Press, San Diego, pp. 291317.
Berger, J., Caby, R., Ligeois, J.-P., Mercier, J.-C.C., Demaiffe, D., 2009. Dehydration,
melting and related garnet growth in the deep root of the Amalaoulaou
Neoproterozoic magmatic arc (Gourma, NE Mali). Geological Magazine 146,
173186.
Blatter, D.L., Carmichael, I.S.E., 1998. Hornblende peridotite xenoliths from central
Mexico reveal the highly oxidized nature of subarc upper mantle. Geology 26,
10351038.
Blevin, P.L., Chappell, B.W., 1992. The role of magma sources, oxidation states and
fractionation in determining the granite metallogeny of eastern Australia. Trans-
actions of the Royal Society of Edinburgh, Earth Sciences 83, 305316.
Blevin, P.L., Chappell, B.W., Allen, C.M., 1996. Intrusive metallogenic provinces in
eastern Australia based on granite source and composition. Geological Society of
America, Special Paper 315, 281290.
Blundy, J., Cashman, K.V., Rust, A., Witham, F., 2010. A case for CO
2
-rich arc magmas.
Earth and Planetary Science Letters 290, 289301.
Bockrath, C., Ballhaus, C., Holzheid, A., 2004. Fractionation of the platinum-group
elements during mantle melting. Science 305, 19511953.
Bodnar, R.J., Beane, R.E., 1980. Temporal and spatial variations in hydrothermal uid
characteristics during vein lling in preore cover overlying deeply buried porphyry
copper-type mineralization at Red Mountain, Arizona. Economic Geology 75,
876893.
Borisov, A., Palme, H., 1997. Experimental determination of the solubility of platinumin
silicate melts. Geochimica et Cosmochimica Acta 61, 43494357.
Bourdon, B., Turner, S., Dosseto, A., 2003. Dehydration and partial melting in subduction
zones: constraints from U-series disequilibria. Journal Geophysical Research 108,
B6. doi:10.1029/2002JB001839.
Bouse, R.M., Ruiz, J., Titley, S.R., Tosdal, R.M., Wooden, J.L., 1999. Lead isotope
compositions of Late Cretaceous and Early Tertiary igneous rocks and sulde
minerals in Arizona: implications for the sources of plutons and metals in porphyry
copper deposits. Economic Geology 94, 211244.
Brandon, A.D., Draper, D.S., 1996. Constraints on the origin of the oxidation state of
mantle overlying subduction zones: an example from Simcoe, Washington, USA.
Geochimica et Cosmochimica Acta 60, 17391749.
Breeding, C.M., Ague, J.J., Brcker, M., 2004. Fluidmetasedimentary rock interactions in
subduction-zone mlange: implications for the chemical composition of arc
magmas. Geology 32, 10411044.
19 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Brenan, J.M., Shaw, H.F., Phinney, D.L., Ryerson, F.J., 1994. Rutileaqueous uid
partitioning of Nb, Ta, Hf, Zr, U and Th: implications for high eld strength element
depletions in island arc basalts. Earth and Planetary Science Letters 128, 327339.
Brown, M., 2010. Melting of the continental crust during orogenesis: the thermal,
rheological, and compositional consequences of melt transport from lower to
upper continental crust. Canadian Journal of Earth Sciences 47, 655694.
Burnham, C.W., 1979. Magmas and hydrothermal uids, In: Barnes, H.L. (Ed.),
Geochemistry of Hydrothermal Ore Deposits, 2nd edition. John Wiley and Sons,
New York, pp. 71136.
Burnham, C.W., 1981. Convergence and mineralization is there a relation? Geol. Soc.
America Memoir 154, 761768.
Burnham, C.W., 1997. Magmas and hydrothermal uids, In: Barnes, H.L. (Ed.),
Geochemistry of Hydrothermal Ore Deposits, 3rd edition. John Wiley and Sons,
New York, pp. 63123.
Campbell, I.H., Naldrett, A.J., 1979. The inuence of silicate:sulde ratios on the
geochemistry of magmatic suldes. Economic Geology 74, 15031506.
Candela, P.A., 1989a. Felsic magmas, volatiles, and metallogenesis. In: Whitney, J.A.,
Naldrett, A.J. (Eds.), Ore Deposition Associated with Magmas: Soc. Econ. Geol.,
Reviews in Economic Geology, 4, pp. 223233.
Candela, P.A., 1989b. Calculation of magmatic uid contributions to porphyry-type ore
systems: predicting uid inclusion chemistries. Geochemical Journal 23, 295305.
Candela, P.A., 1989c. Magmatic ore-forming uids: thermodynamic and mass transfer
calculations of metal concentrations. In: Whitney, J.A., Naldrett, A.J. (Eds.), OreDeposition
AssociatedwithMagmas: Soc. Econ. Geol., Reviews inEconomic Geology, 4, pp. 203221.
Candela, P.A., 1992. Controls on ore metal ratios in granite-related ore systems: an
experimental and computational approach. Transactions of the Royal Society of
Edinburgh, Earth Sciences 83, 317326.
Candela, P.A., Holland, H.D., 1984. The partitioning of copper and molybdenum
between silicate melts and aqueous uids. Geochimica et Cosmochimica Acta 48,
373380.
Candela, P.A., Holland, H.D., 1986. A mass transfer model for copper and molybdenum
in magmatic hydrothermal systems: the origin of porphyry-type ore deposits.
Economic Geology 81, 119.
Candela, P.A., Piccoli, P.M., 1995. Model oremetal partitioning frommelts into vapor and
bapor/brine mixtures. In: Thompson, J.F.H. (Ed.), Magmas, Fluids, and Ore Deposits. :
Short Course Series, 23. Mineralogical Association of Canada, pp. 101127. ch. 5.
Candela, P.A., Piccoli, P.M., 2005. Magmatic processes in the development of porphyry-
type ore systems. In: Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P.
(Eds.), Economic Geology 100th Anniversary Volume. Society of Economic
Geologists, Littleton, CO, pp. 2537.
Canil, D., Styan, J., Larocque, J., Bonnet, E., Kyba, J., 2010. Thickness and composition of
the Bonanza arc crustal section, Vancouver Island, Canada. Geological Society of
America, Bulletin 122, 10941105.
Carman, G.D., 2003. Geology, mineralization, and hydrothermal evolution of the
Ladolam gold deposit, Lihir Island, Papua New Guinea. In: Simmons, S.F., Graham, I.
(Eds.), Volcanic, Geothermal, and Ore-Forming Fluids: Rulers and Witnesses of
Processes Within the Earth. Special Publication, No. 10. Society of Economic
Geologists, pp. 247284.
Carrigan, C.R., Schubert, G., Eichelberger, J.C., 1992. Thermal and dynamical regimes of
single- and two-phase magmatic ow in dikes. Journal of Geophysical Research 97
(17), 392 377-17.
Carroll, M.R., Rutherford, M.J., 1985. Sulde and sulfate saturation in hydrous silicate
melts. Proceedings of the Fifteenth Lunar and Planetary Science Conference, Part 2:
Journal of Geophysical Research, 90, pp. C601C612. Supplement.
Carten, R.B., White, W.H., Stein, H.J., 1993. High-grade granite-related molybdenum
systems. In: Kirham, R.V., Sinclair, W.D., Thorpe, R.I., Duke, J.M. (Eds.), Mineral Deposit
Modeling. Special Paper, 40. Geological Association of Canada, pp. 521554.
Cathles, L.M., 1991. The importance of the 350 C isotherm in ore-forming hydrother-
mal systems. Geological Society of America, Annual Meeting, October 2124, 1991.
San Diego, CA, Abstracts with Programs 23, p. A21.
Cauzid, J., Philippot, P., Martinez-Criado, G., Mnez, B., Labour, S., 2007. Contrasting Cu-
complexing behaviour in vapour and liquid uid inclusions from the Yankee Lode
tin deposit, Mole Granite, Australia. Chemical Geology 246, 3954.
ern, P., Blevin, P.L., Cuney, M., London, D., 2005. Granite-related ore deposits. In:
Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), Economic
Geology 100th Anniversary Volume. Society of Economic Geologists, Littleton, CO,
pp. 337370.
Cervantes, P., Wallace, P.J., 2003. Role of H
2
O in subduction-zone magmatism: new
insights from melt inclusions in high-Mg basalts from central Mexico. Geology 31,
235238.
Chaplygin, I.V., Mozgova, N.N., Mokhov, A.V., Koporulina, E.V., Bernhardt, H.-J., Bryzgalov,
I.A., 2007. Minerals of the systemZnSCdS from fumaroles of the Kudriavy Volcano,
Iturup Island, Kuriles, Russia. Canadian Mineralogist 45, 709722.
Chappell, B.W., White, A.J.R., 1974. Two contrasting granite types. Pacic Geology 8,
173174.
Chiaradia, M., Fontbot, L., Paladines, A., 2004. Metal sources in mineral deposits and
crustal rocks of Ecuador (1N4S): a lead isotope synthesis. Economic Geology 99,
10851106.
Claeson, D.T., Meurer, W.P., 2004. Fractional crystallization of hydrous basaltic arc-type
magmas and the formation of amphibole-bearing gabbroic cumulates. Contributions
to Mineralogy and Petrology 147, 288304.
Clemens, J.D., 2003. S-type granitic magmas petrogenetic issues, models and
evidence. Earth-Science Reviews 61, 118.
Clemens, J.D., Darbyshire, D.P.F., Flinders, J., 2009. Sources of post-orogenic calcalkaline
magmas: the Arrochar and Garabal HillGlen Fyne complexes, Scotland. Lithos 112,
524542.
Cline, J.S., 1995. Genesis of porphyry copper deposits: the behavior of water, chloride, and
copper in crystallizing melts. In: Pierce, F.W., Bolm, J.G. (Eds.), Porphyry Copper
Deposits of the American Cordillera: Arizona Geological Society Digest, 20, pp. 6982.
Cline, J.S., Bodnar, R.J., 1991. Can economic porphyry copper mineralization be
generated by a typical calc-alkaline melt? Journal of Geophysical Research 96,
81138126.
Cloos, M., 2001. Bubbling magma chambers, cupolas, and porphyry copper deposits.
International Geology Review 43, 285311.
Cloos, M., Sapiie, B., van Ufford, A.Q., Weiland, R.J., Warren, P.Q., McMahon, T.P., 2005.
Collisional delamination in New Guinea: the geotectonics of subducting slab
breakoff. Special Paper, 400. Geological Society of America. 51 pp.
Comment on Defant, M.J., Kepezhinskas, P., 2001Conrey, R.M., 2002. Adakites: a review
of slab melting over the past decade and the case for a slab-melt component in arcs
[EOS, Trans. AGU 82, 6569]. EOS, Trans. AGU 83, 256.
Coumou, D., Driesner, T., Heinrich, C.A., 2008. Heat transport at boiling, near-critical
conditions. Geouids 8, 208215.
Crerar, D.A., Barnes, H.L., 1976. Ore solution chemistry V. Solubilities of chalcopyrite and
chalcocite assemblages in hydrothermal solution at 200 C to 350 C. Economic
Geology 71, 772794.
Cygan, G.L., Candela, P.A., 1995. Preliminary study of gold partitioning among pyrrhotite,
pyrite, magnetite, and chalcopyrite in gold-saturated chloride solutions at 600 to
700 C, 140 MPa (1400 bars). In: Thompson, J.F.H. (Ed.), Magmas, Fluids, and Ore
Deposits. : Short Course Series, 23. Mineralogical Association of Canada, pp. 129137.
Dale, C.W., Burton, K.W., Pearson, D.G., Gannoun, A., Alard, O., Argles, T.W., Parkinson,
I.J., 2009. Highly siderophile element behaviour accompanying subduction of
oceanic crust: whole rock and mineral-scale insights from a high-pressure terrain.
Geochimica et Cosmochimica Acta 73, 13941416.
Damon, P.E., 1986. Batholith-volcano coupling in the metallogeny of porphyry copper
deposits. In: Friedrich, G.H. (Ed.), Geology and Metallogeny of Copper Deposits.
Springer-Verlag, Berlin, pp. 216234.
Darbyshire, D.P.F., Shepherd, T.J., 1994. Nd and Sr isotope constraints on the origin of
the Cornubian batholith, SW England. Journal of the Geological Society, London
151, 795802.
Davidson, J.P., 1996. Deciphering mantle and crustal signatures in subduction zone
magmatism. In: Bebout, G.E., Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction: Top
to bottom. : Geophysical Monograph, 96. American Geophysical Union, pp. 251262.
Davidson, J., Turner, S., Handley, H., Macpherson, C., Dosseto, A., 2007. Amphibole
sponge in arc crust? Geology 35, 787790.
Davies, J.H., Stevenson, D., 1992. Physical model of source region of subduction zone
volcanics. Journal of Geophysical Research 97, 20372070.
de Hoog, J.C.M., Mason, P.R.D., van Bergen, M.J., 2001. Sulfur and chalcophile elements
in subduction zones: constraints from a laser ablation ICP-MS study of melt
inclusions fromGalunggung Volcano, Indonesia. Geochimica et Cosmochimica Acta
65, 31473164.
DeBari, S.M., Coleman, R.G., 1989. Examination of the deep levels of an island arc:
evidence from the Tonsina ultramacmac assemblage, Tonsina, Alaska. Journal
of Geophysical Research 94, 43734391.
Deditius, A.P., Utsunomiya, S., Ewing, R.C., Chryssoulis, S.L., Venter, D., Kesler, S.E., 2009.
Decoupled geochemical behavior of As and Cu in hydrothermal systems. Geology
37, 707710.
Defant, M.J., Drummond, M.S., 1990. Derivation of some modern arc magmas by
melting of young subducted lithosphere. Nature 347, 662665.
Defant, M.J., Kepezhinskas, P., 2001. Adakites: a reviewof slabmeltingover the past decade
and the case for a slab-melt component in arcs. EOS, Trans. AGU 82 (65), 6869.
Defant, M.J., Kepezhinskas, P., 2002. Reply to Comment by Conrey, R. [Adakites: a
reviewof slab melting over the past decade and the case for a slab-melt component
in arcs: EOS, Trans. AGU 82, 6569]. EOS, Trans AGU 83, 256257.
DePaolo, D.J., 1981. Trace element and isotopic effects of combined wallrock
assimilation and fractional crystallization. Earth and Planetary Science Letters 53,
189202.
Dietrich, A., Lehmann, B., Wallianos, A., Traxel, K., Palacios, C., 1999. Magma mixing in
Bolivian tin porphyries. Die Naturwissenschaften 86, 4043.
Dilles, J.H., 1987. Petrology of the Yerington batholith, Nevada: evidence for evolution
of porphyry copper ore uids. Economic Geology 82, 17501789.
Dreyer, B.M., Morris, J.D., Gill, J.B., 2010. Incorporation of subducted slab-derived
sediment and uid in arc magmas: BBe10BeNd systematics of the Kurile
convergent margin, Russia. Journal of Petrology 51, 17611782.
Driesner, T., 2007. The system H
2
ONaCl. Part II: Correlations for molar volume,
enthalpy, and isobaric heat capacity from 0 to 1000 C, 1 to 5000 bar, and 0 to 1
X
NaCl
. Geochimica et Cosmochimica Acta 71, 49024919.
Driesner, T., Heinrich, C.A., 2007. The system H
2
ONaCl. Part I: correlation formulae for
phase relations in temperaturepressurecomposition space from 0 to 1000 C,
0 to 5000 bar, and 0 to 1 X
NaCl
. Geochimica et Cosmochimica Acta 71, 48804901.
Drummond, M.S., Defant, M.J., Kepezhinskas, P.K., 1996. Petrogenesis of slab-derived
trondhjemitetonalitedacite/adakite magmas. Special Paper, 315. Geological
Society of America, pp. 205215.
Dufek, J., Bergantz, G.W., 2005. Lower crustal magma genesis and preservation: a
stochastic framework for the evaluation of basaltcrust interaction. Journal of
Petrology 46, 21672195.
Duggen, S., Portnyagin, M., Baker, J., Ulfbeck, D., Hoernle, K., Garbe-Schnberg, D.,
Grassineau, N., 2007. Drastic shift inlavageochemistryinthevolcanic-front torear-arc
region of the SouthernKamchatkan subduction zone: evidence for the transition from
slab surface dehydration to sediment melting. Geochimica et Cosmochimica Acta 71,
452480.
Dunn, J.C., Hardee, H.C., 1981. Superconvecting geothermal zones. Journal of
Volcanology and Geothermal Research 11, 189201.
20 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Dvir, O., Pettke, T., Fumagalli, P., Kessel, R., 2011. Fluids in the peridotitewater system
up to 6 GPa and 800 C: new experimental constrains on dehydration reactions.
Contributions to Mineralogy and Petrology 161, 829844.
Eastoe, C.J., 1982. Physics andchemistryof thehydrothermal systemat thePanguna porphyry
copper deposit, Bougainville, Papua New Guinea. Economic Geology 77, 127153.
Economou-Eliopoulos, M., Eliopoulos, D.G., 2000. Palladium, platinum and gold
concentration in porphyry copper systems of Greece and their genetic signicance.
Ore Geology Reviews 16, 5970.
Edwards, K.J., 2004. Formation and degradation of seaoor hydrothermal sulde
deposits. Geological Society of America Special Papers 379, 8396.
Eggins, S.M., 1993. Origin and differentiation of picritic arc magmas, Ambae (Aoba),
Vanuatu. Contributions to Mineralogy and Petrology 114, 79100.
Eichelberger, J.C., 1995. Silicic volcanism: ascent of viscous magmas from crustal
reservoirs. Annual Review Earth Planet Science 23, 4163.
Farmer, G.L., DePaolo, D.J., 1984. Origin of Mesozoic and Tertiary granite in the western
United States and implications for pre-Mesozoic crustal structure, 2. Nd and Sr
isotopic studies of unmineralized and Cu- and Mo-mineralized granite in the
Precambrian craton. Journal of Geophysical Research 89, 1014110160.
Faure, K., Matsuhisa, Y., Metsugi, H., Mizota, C., Hayashi, S., 2002. The Hishikari AuAg
epithermal deposit, Japan: oxygen and hydrogen isotope evidence in determining
the source of paleohydrothermal uids. Economic Geology 97, 481498.
Feeley, T.C., Davidson, J.P., 1994. Petrology of calc-alkaline lavas at Volcn Ollage and
the origin of compositional diversity at Central Andean stratovolcanoes. Journal of
Petrology 35, 12951340.
Field, C.W., Fifarek, R.H., 1985. Light stable-isotope systematics in the epithermal
environment. In: Berger, B.R., Bethke, P.M. (Eds.), Geology and Geochemistry of
Epithermal Systems. Reviews in Economic Geology, 2. Society of Economic
Geologists, pp. 99128.
Foley, S.F., Barth, M.G., Jenner, G.A., 2000. Rutile/melt partition coefcients for trace
elements and an assessment of the inuence of rutile on the trace element
characteristics of subduction zone magmas. Geochimica et Cosmochimica Acta 64,
933938.
Fontbot, L., Gunnesch, K.A., Baumann, A., 1990. Metal sources in stratabound ore
deposits in the Andes (Andean Cycle) lead isotopic constraints. In: Fontbot, L.,
Amstutz, G.C., Cardozo, M., Cedillo, E., Frutos, J. (Eds.), Stratabound Ore Deposits in
the Andes. Springer-Verlag, Berlin, pp. 759773.
Forneris, J.F., Holloway, J.R., 2003. Phase equilibria in subducting basaltic crust:
implications for H
2
O release from the slab. Earth and Planetary Science Letters 214,
187201.
Fournier, R.O., 1985. The behavior of silica in hydrothermal solutions. In: Berger, B.R.,
Bethke, P.M. (Eds.), Reviews in Economic Geology, 2. Society of Economic
Geologists, pp. 4561.
Fournier, R.O., 1999. Hydrothermal processes related to movement of uid from plastic
into brittle rock in the magmaticepithermal environment. Economic Geology 94,
11931212.
Fumagalli, P., Poli, S., 2005. Experimentally determined phase relations in hydrous
peridotites to 6.5 GPa and their consequences on the dynamics of subduction
zones. Journal of Petrology 46, 555578.
Fyfe, W.S., 1992. Magma underplating of continental crust. Journal Volcanology
Geothermal Research 50, 3340.
Garrido, C.J., Bodinier, J.-L., Burg, J.-P., Zeilinger, G., Hussain, S.S., Dawood, H., Chaudhry,
M.N., Gervilla, F., 2006. Petrogenesis of mac garnet granulite in the lower crust of
the Kohistan paleo-arc complex (northern Pakistan): implications for intra-crustal
differentiation of island arcs and generation of continental crust. Journal of
Petrology 47, 18731914.
Garrison, J.M., Davidson, J.P., 2003. Dubious case for slab melting in the Northern
volcanic zone of the Andes. Geology 31, 565568.
Gill, J.B., 1981. Orogenic Andesites and Plate Tectonics. Springer-Verlag, New York.
390 pp.
Gill, J., Whelan, P., 1989. Postsubduction ocean island alkali basalts in Fiji. Journal of
Geophysical Research 94, 45794588.
Goff, S.J., Goff, F., Janik, C.J., 1992. Tecuamburro volcano, Guatemala: exploration
geothermal gradient drilling and results. Geothermics 21, 483502.
Green, T.H., Adam, J., 2003. Experimentally-determined trace element characteristics of
aqueous uid frompartially dehydrated mac oceanic crust at 3.0 GPa, 650700 C.
European Journal of Mineralogy 15, 815830.
Greene, A.R., Debari, S.M., Kelemen, P.B., Blusztajn, J., Clift, P.D., 2006. A detailed
geochemical study of island arc crust: the Talkeetna Arc section, south-central
Alaska. Journal of Petrology 47, 10511093.
Grove, T.L., Elkins-Tanton, L.T., Parman, S.W., Chatterjee, N., Mntener, O., Gaetani, G.A.,
2003. Fractional crystallization and mantle-melting controls on calc-alkaline
differentiation trends. Contributions to Mineralogy and Petrology 145, 515533.
Grove, T.L., Chatterjee, N., Parman, S.W., Mdard, E., 2006. The inuence of H
2
O on
mantle wedge melting. Earth and Planetary Science Letters 249, 7489.
Grunder, A.L., Klemetti, E.W., Feeley, T.C., McKee, C.M., 2008. Eleven million years of arc
volcanism at the Aucanquilcha Volcanic Cluster, northern Chilean Andes:
implications for the life span and emplacement of plutons. Transactions: Earth
Sciences, 97. Royal Society of Edinburgh, pp. 415436.
Guivel, C., Lagabrielle, Y., Bourgois, J., Martin, H., Arnaud, N., Fourcade, S., Cotten, J.,
Maury, R.C., 2003. Very shallow melting of oceanic crust during spreading ridge
subduction: origin of near-trench Quaternary volcanismat the Chile Triple Junction.
Journal of Geophysical Research 108 (B7), 2345. doi:10.1029/2002JB002119.
Guo, F., Wilson, M., Li, C., 2007. Post-collisional adakites in south Tibet: products of
partial melting of subduction-modied lower crust. Lithos 96, 205224.
Gutscher, M.-A., Maury, R., Eissen, J.-P., Bourdon, E., 2000. Can slab melting be caused by
at subduction? Geology 28, 535538.
Halter, W.E., Pettke, T., Heinrich, C.A., 2002. The origin of Cu/Au ratios in porphyry-type
ore deposits. Science 296, 18441846.
Halter, W.E., Heinrich, C.A., Pettke, T., 2005. Magma evolution and the formation of
porphyry CuAu ore uids: evidence from silicate and sulde melt inclusions.
Mineralium Deposita 39, 845863.
Hamada, M., Fujii, T., 2008. Experimental constraints on the effects of pressure and H
2
O
on the fractional crystallization of high-Mg island arc basalt. Contributions to
Mineralogy and Petrology 155, 767790.
Hamlyn, P.R., Keays, R.R., Cameron, W.E., Crawford, A.J., Waldron, H.M., 1985.
Precious metals in magnesian low-Ti lavas: implications for metallogenesis and
sulfur saturation in primary magmas. Geochimica et Cosmochimica Acta 49,
17971811.
Hansen, J., Skjerlie, K.P., Pedersen, R.B., De La Rosa, J., 2002. Crustal melting in the lower
parts of island arcs: an example from the Bremanger Granitoid Complex, west
Norwegian Caledonides. Contributions to Mineralogy and Petrology 143, 316335.
Harris, N.B.W., Pearce, J.A., Tindle, A.G., 1986. Geochemical characteristics of collision
zone magmatism. Special Publication, 19. Geological Society, London, pp. 6781.
Harris, A.C., Kamenetsky, V.S., White, N.C., van Achterbergh, E., Ryan, C.G., 2003. Melt
inclusions in veins: linking magmas and porphyry Cu deposits. Science 302,
21092111.
Hart, C.J.R., Mair, J.L., Goldfarb, R.J., Groves, D.I., 2005. Source and redox controls on
metallogenic variations in intrusion-related ore systems, Tombstone-Tungsten
Belt, Yukon Territory. In: Ishihara, S., Stephens, W.E., Harley, S.L., Arima, M.,
Nakajima, T. (Eds.), Fifth Hutton Symposium; the Origin of Granites and Related
Rocks. Special Paper, 389. Geological Society of America, pp. 339356.
Haschke, M., Ben-Avraham, Z., 2005. Adakites from collision-modied lithosphere.
Geophysical Research Letters 32 (15), L15302. doi:10.1029/2005GL023468.
Hattori, K.H., 1997. Occurrence and origin of sulde and sulfate in the 1991 Mount
Pinatubo eruption products. In: Newhall, C.G., Punongbayan, R.S. (Eds.), Fire and
Mud: Eruptions and Lahars of Mount Pinatubo, Philippines. University of
Washington Press, Seattle, pp. 807824.
Hattori, K., Guillot, S., 2003. Volcanic fronts form as a consequence of serpentine
dehydration in the forearc mantle wedge. Geology 31, 525528.
Hattori, K.H., Keith, J.D., 2001. Contribution of mac melt to porphyry copper
mineralization: evidence from Mount Pinatubo, Philippines, and Bingham Canyon,
Utah, USA. Mineralium Deposita 36, 799806.
Hattori, K., Takahashi, Y., Guillot, S., Johanson, B., 2005. Occurrence of arsenic (V) in
forearc mantle serpentinites based on X-ray absorption spectroscopy study.
Geochimica et Cosmochimica Acta 69, 55855596.
Hattori, K., Wallis, S., Enami, M., Mizukami, T., 2010. Subduction of mantle wedge
peridotites: evidence from the Higashi-akaishi ultramac body in the Sanbagawa
metamorphic belt. Island Arc 19, 192207.
Hawkesworth, C.J., Gallagher, K., Hergt, J.M., McDermott, F., 1994. Destructive plate
margin magmatism: geochemistry and melt generation. Lithos 33, 169188.
Heald, P., Foley, N.K., Hayba, D.O., 1987. Comparative anatomy of volcanic-hosted
epithermal deposits: acid-sulfate and adularia-sericite types. Economic Geology 82,
126.
Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of
hydrothermal ore deposits. Nature 370, 519527.
Hedenquist, J.W., Richards, J.P., 1998. The inuence of geochemical techniques on the
development of genetic models for porphyry copper deposits. In: Richards, J.P.,
Larson, P.B. (Eds.), Techniques in Hydrothermal Ore Deposits Geology: Reviews in
Economic Geology, 10, pp. 235256. ch. 10.
Hedenquist, J.W., Simmons, S.F., Giggenbach, W.F., Eldridge, C.S., 1993. White Island,
New Zealand, volcanichydrothermal system represents the geochemical environ-
ment of high-suldation Cu and Au ore deposition. Geology 21, 731734.
Hedenquist, J.W., Aoki, M., Shinohara, H., 1994a. Flux of volatiles and ore-forming
metals from the magmatichydrothermal system of Satsuma Iwojima volcano.
Geology 22, 585588.
Hedenquist, J.W., Matsuhisa, Y., Izawa, E., White, N.C., Giggenbach, W.F., Aoki, M.,
1994b. Geology, geochemistry, and origin of high suldation CuAu mineralization
in the Nansatsu district, Japan. Economic Geology 89, 130.
Hedenquist, J.W., Arribas Jr., A., Reynolds, J.R., 1998. Evolution of an intrusion-centered
hydrothermal system: Far SoutheastLepanto porphyry and epithermal CuAu
deposits, Philippines. Economic Geology 93, 373404.
Heinrich, C.A., 2005. The physical and chemical evolution of low-salinity magmatic
uids at the porphyry to epithermal transition: a thermodynamic study.
Mineralium Deposita 39, 864889.
Heinrich, C.A., Ryan, G.G., Mernagh, T.P., Eadington, P.J., 1992. Segregation of ore metals
between magmatic brine and vapor: a uid inclusion study using PIXE
microanalysis. Economic Geology 87, 15661583.
Heinrich, C.A., Gnther, D., Audtat, A., Ulrich, T., Frischknecht, R., 1999. Metal
fractionation between magmatic brine and vapor, determined by microanalysis of
uid inclusions. Geology 27, 755758.
Heinrich, C.A., Halter, W., Klemm, L., Landtwing, M.R., Pettke, T., 2003a. Laser ablation
micro-analysis of uid and melt inclusions: towards understanding the process of
porphyry-style ore formation. Applied Earth Science (Transactions, Institutions of
Mining and Metallurgy, Section B) 112, 185186.
Heinrich, C.A., Pettke, T., Halter, W.E., Aigner-Torres, M., Audtat, A., Gnther, D.,
Hattendorf, B., Bleiner, D., Guillong, M., Horn, I., 2003b. Quantitative multi-element
analysis of minerals, uid and melt inclusions by laser-ablation inductively-
coupled-plasma mass-spectrometry. Geochimica et Cosmochimica Acta 67,
34733497.
Heinrich, C.A., Dreisner, T., Steffnson, A., Seward, T.M., 2004. Magmatic vapor
contraction and the transport of gold from the porphyry environment to
epithermal ore deposits. Geology 32, 761764.
21 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Hemley, J.J., Hunt, J.P., 1992. Hydrothermal ore-forming processes in the light of studies
in rock-buffered systems: II. Some general geologic applications. Economic Geology
87, 2343.
Hemley, J.J., Cygan, G.L., Fein, J.B., Robinson, G.R., d'Angelo, W.M., 1992. Hydrothermal
ore-forming processes in the light of studies in rock-buffered systems: I. Iron
copperzinclead sulde solubility relations. Economic Geology 87, 122.
Henley, R.W., Hughes, G.O., 2000. Underground fumaroles: excess heat effects in vein
formation. Economic Geology 95, 453466.
Henley, R.W., McNabb, A., 1978. Magmatic vapor plumes and groundwater interaction
in porphyry copper emplacement. Economic Geology 73, 120.
Hermann, J., Spandler, C.J., 2008. Sediment melts at sub-arc depths: an experimental
study. Journal of Petrology 49, 717740.
Herzberg, C.T., Fyfe, W.S., Carr, M.J., 1983. Density constraints on the formation of the
continental Moho and crust. Contributions to Mineralogy and Petrology 84, 15.
Herzig, P.M., Hannington, M.D., Scott, S.D., Maliotis, G., Rona, P.A., Thompson, G., 1991.
Gold-rich sea-oor gossans in the Troodos Ophiolite and on the Mid-Atlantic Ridge.
Economic Geology 86, 17471755.
Hildreth, W., 1981. Gradients in silicic magma chambers: implications for lithospheric
magmatism. Journal of Geophysical Research 86 (10), 192 153-10.
Hildreth, W., Moorbath, S., 1988. Crustal contributions to arc magmatism in the Andes
of central Chile. Contributions to Mineralogy and Petrology 98, 455489.
Holland, H.D., 1965. Some applications of thermochemical data to problems of ore
deposits II. Mineral assemblages and the composition of ore forming uids.
Economic Geology 60, 11011166.
Holloway, J.R., 1976. Fluids in the evolution of granitic magmas: consequences of nite
CO
2
solubility. Geological Society of America Bulletin 87, 15131518.
Hou, Z.Q., Zeng, P.S., Gao, Y.F., Dong, F.L., 2006. Himalayan CuMoAu mineralization in
the eastern Indo-Asian collision zone: constraints from ReOs dating of
molybdenite. Mineralium Deposita 41, 3345.
Hou, Z., Yang, Z., Qu, X., Meng, X., Li, Z., Beaudoin, G., Rui, Z., Gao, Y., Zaw, K., 2009. The
Miocene Gangdese porphyry copper belt generated during post-collisional
extension in the Tibetan Orogen. Ore Geology Reviews 36, 2551.
Huppert, H.E., Sparks, R.S.J., 1988. The generation of granitic magmas by intrusion of
basalt into continental crust. Journal of Petrology 29, 599624.
Ishihara, S., 1981. The granitoid series and mineralization. Economic Geology 75th
Anniversary Volume, pp. 458484.
Ishihara, S., Murakami, H., 2006. Fractionated ilmenite-series granites in southwest Japan:
source magma for REESnW mineralizations. Resource Geology 56, 245256.
Jagoutz, O., Muntener, O., Ulmer, P., Pettke, T., Burg, J.-P., Dawood, H., Hussain, S., 2007.
Petrology and mineral chemistry of lower crustal intrusions: the Chilas Complex,
Kohistan (NW Pakistan). Journal of Petrology 48, 18951953.
Jagoutz, O.E., Burg, J.-P., Hussain, S., Dawood, H., Pettke, T., Iizuka, T., Maruyama, S.,
2009. Construction of the granitoid crust of an island arc part I: geochronological
and geochemical constraints from the plutonic Kohistan (NW Pakistan). Contri-
butions to Mineralogy and Petrology 158, 739755.
James, D.E., 1982. A combined O, Sr, Nd, and Pb isotopic and trace element study of
crustal contamination in central Andean lavas, I. Local geochemical variations.
Earth and Planetary Science Letters 57, 4762.
Jgo, S., Pichavant, M., Mavrogenes, J.A., 2010. Controls on gold solubility in arc
magmas: an experimental study at 1000 C and 4 kbar. Geochimica et Cosmochi-
mica Acta 74, 21652189.
Jenner, F.E., O'Neill, H., St, C., Arculus, R.J., Mavrogenes, J.A., 2010. The magnetite crisis in
the evolution of arc-related magmas and the initial concentration of Au, Ag and Cu.
Journal of Petrology 51, 24452464.
Jensen, E.P., Barton, M.D., 2000. Gold deposits related to alkaline magmatism. In:
Hagemann, S.G., Brown, P.E. (Eds.), Gold in 2000: Reviews in Economic Geology, 13,
pp. 279314.
John, D.A., 1991. Evolution of hydrothermal uids in the Alta Stock, Central Wasatch
Mountains, Utah. U.S. Geological Survey, Bulletin 1977. 51 pp.
Johnson, M.C., Plank, T., 1999. Dehydration and melting experiments constrain the fate of
subducted sediments. Geochemistry Geophysics Geosystems 1, 1007. doi:10.1029/
1999GC000014.
Johnson, R.W., Mackenzie, D.E., Smith, I.E.M., 1978. Delayed partial melting of
subduction-modied mantle in Papua New Guinea. Tectonophysics 46, 197216.
Jugo, P.J., 2009. Sulfur content at sulde saturation in oxidized magmas. Geology 37,
415418.
Jugo, P.J., Candela, P.A., Piccoli, P.M., 1999. Magmatic suldes and Au:Cu ratios in
porphyry deposits: an experimental study of copper and gold partitioning at
850 C, 100 MPa in a haplogranitic meltpyrrhotiteintermediate solid solution
gold metal assemblage, at gas saturation. Lithos 46, 573589.
Jugo, P.J., Luth, R.W., Richards, J.P., 2005a. Experimental data onthe speciationof sulfur as a
function of oxygen fugacity in basaltic melts. Geochimica et Cosmochimica Acta 69,
497503.
Jugo, P.J., Luth, R.W., Richards, J.P., 2005b. An experimental study of the sulfur content in
basaltic melts saturated with immiscible sulde or sulfate liquids at 1300 C and
1.0 GPa. Journal of Petrology 46, 783798.
Jugo, P.J., Wilke, M., Botcharnikov, R.E., 2010. Sulfur K-edge XANES analysis of natural
and synthetic basaltic glasses: implications for S speciation and S content as
function of oxygen fugacity. Geochimica et Cosmochimica Acta 74, 59265938.
Kawamoto, T., 2006. Hydrous phases and water transport in the subducting slab.
Reviews in Mineralogy and Geochemistry 62, 273289.
Kay, R.W., 1978. Aleutian magnesian andesites: melts from subducted Pacic ocean
crust. Journal of Volcanology and Geothermal Research 4, 117132.
Kay, S.M., Ramos, V.A., Marquez, M., 1993. Evidence in Cerro Pampa volcanic rocks for
slab-melting prior to ridgetrench collision in southern South America. Journal of
Geology 101, 703714.
Kay, S.M., Mpodozis, C., Coira, B., 1999. Neogene magmatism, tectonism, and mineral
deposits of the Central Andes (22 to 33S latitude). In: Skinner, B.J. (Ed.), Geology
and Ore Deposits of the Central Andes. : Special Publication, No. 7. Society of
Economic Geologists, pp. 2759.
Keith, J.D., Shanks III, W.C., Archibald, D.A., Farrar, E., 1986. Volcanic and intrusive
history of the Pine Grove porphyry molybdenum system, southwestern Utah.
Economic Geology 81, 553577.
Keith, J.D., Whitney, J.A., Hattori, K., Ballantyne, G.H., Christiansen, E.H., Barr, D.L.,
Cannan, T.M., Hook, C.J., 1997. The role of magmatic suldes and mac alkaline
magmas in the Bingham and Tintic mining districts, Utah. Journal of Petrology 38,
16791690.
Keith, J.D., Christiansen, E.H., Maughan, D.T., Waite, K.A., 1998. The role of mac alkaline
magmas in felsic porphyry-Cu and Mo systems. In: Lentz, D.R. (Ed.), Mineralized
Intrusion-Related Skarn Systems: Mineralogical Association of Canada Short Course
Series, 26, pp. 211243.
Kelley, K.A., Cottrell, E., 2009. Water and the oxidation state of subduction zone
magmas. Science 325, 605607.
Kelley, K.D., Romberger, S.B., Beaty, D.W., Pontius, J.A., Snee, L.W., Stein, H.J., Thompson,
T.B., 1998. Geochemical and geochronological constraints on the genesis of AuTe
deposits at Cripple Creek, Colorado. Economic Geology 93, 9811012.
Kelley, K.A., Plank, T., Newman, S., Stolper, E.M., Grove, T.L., Parman, S., Hauri, E., 2010.
Mantle melting as a function of water content beneath the Mariana Arc. Journal of
Petrology 51, 17111738.
Kemp, A.I.S., Hawkesworth, C.J., Foster, G.L., Paterson, B.A., Woodhead, J.D., Hergt, J.M.,
Gray, C.M., Whitehouse, M.J., 2007. Magmatic and crustal differentiation history of
granitic rocks from HfO isotopes in zircon. Science 315, 980983.
Kennedy, A.K., Hart, S.R., Frey, F.A., 1990. Composition and isotopic constraints on the
petrogenesis of alkaline arc lavas: Lihir Island, Papua New Guinea. Journal of
Geophysical Research 95, 69296942.
Kent, A.J.R., Peate, D.W., Newman, S., Stolper, E.M., Pearce, J.A., 2002. Chlorine in
submarine glasses from the Lau Basin: seawater contamination and constraints on
the composition of slab-derived uids. Earth and Planetary Science Letters 202,
361377.
Kepezhinskas, P., Defant, M.J., Widom, E., 2002. Abundance and distribution of PGE and
Au in the island-arc mantle: implications for sub-arc metasomatism. Lithos 60,
113128.
Kerrich, R., Beckinsale, R.D., 1988. Oxygen and strontium isotopic evidence for the
origin of granites in the tin belt of southeast Asia. In: Taylor, R.P., Strong, D.F. (Eds.),
Recent Advances in the Geology of Granite-Related Mineral Deposits: Canadian
Institute of Mining and Metallurgy, Special Volume 39, pp. 115123.
Keskin, M., Gen, .C., Tysz, O., 2008. Petrology and geochemistry of post-collisional
Middle Eocene volcanic units in North-Central Turkey: evidence for magma
generation by slab breakoff following the closure of the Northern Neotethys Ocean.
Lithos 104, 267305.
Kesler, S.E., 1973. Copper, molybdenum and gold abundances in porphyry copper
deposits. Economic Geology 68, 106112.
Kessel, R., Schmidt, M.W., Ulmer, P., Pettke, P., 2005a. Trace element signature of
subduction-zone uids, melts and supercritical liquids at 120180 km depth.
Nature 437, 724727.
Kessel, R., Ulmer, P., Pettke, P., Schmidt, M.W., Thompson, A.B., 2005b. The waterbasalt
system at 4 to 6 GPa: phase relations and second critical endpoint in a K-free
eclogite at 700 to 1400 C. Earth and Planetary Science Letters 237, 873892.
Kilian, R., Behrmann, J.H., 2003. Geochemical constraints on the sources of Southern
Chile Trench sediments and their recycling in arc magmas of the Southern Andes.
Journal of the Geological Society 160, 5770.
Kilian, R., Stern, C.R., 2002. Constraints on the interaction between slab melts and the
mantle wedge from adakitic glass in peridotite xenoliths. European Journal of
Mineralogy 14, 2536.
Kirkham, R.V., Sinclair, W.D., 1996. Porphyry copper, gold, molybdenum, tungsten, tin,
silver. In: Eckstrand, O.R., Sinclair, W.D., Thorpe, R.I. (Eds.), Geology of Canadian
mineral deposits. : Geology of Canada, 8. Geological Survey of Canada, pp. 421446.
Klemm, L.M., Pettke, T., Heinrich, C.A., Campos, E., 2007. Hydrothermal evolution of the
El Teniente deposit, Chile: porphyry CuMo ore deposit from low-salinity
magmatic uids. Economic Geology 102, 10211045.
Klemm, L.M., Pettke, T., Heinrich, C.A., 2008. Fluid and source magma evolution of the
Questa porphyry Mo deposit, New Mexico, USA. Mineralium Deposita 43, 533552.
Klemme, S., Prowatke, S., Hametner, K., Gunther, D., 2005. Partitioning of trace elements
between rutile and silicate melts: implications for subduction zones. Geochimica et
Cosmochimica Acta 69, 23612371.
Klepeis, K.A., Clarke, G.L., Rushmer, T., 2003. Magma transport and coupling between
deformation and magmatism in the continental lithosphere. GSA Today 13, 411.
Kogiso, T., Tatsumi, Y., Nakano, S., 1997. Trace element transport during dehydration
processes in the subducted oceanic crust: 1. Experiments and implications for the
origin of ocean island basalts. Earth and Planetary Science Letters 148, 193205.
Kontak, D.J., Cumming, G.L., Krstic, D., Clark, A.H., Farrar, E., 1990. Isotopic composition
of lead in ore deposits of the Cordillera Oriental, southeastern Peru. Economic
Geology 85, 15841603.
Kusakabe, M., Komoda, Y., Takano, B., Abiko, T., 2000. Sulfur isotopic effects in the
disproportionation reaction of sulfur dioxide in hydrothermal uids: implications
for the
34
S variations of dissolved bisulfate and elemental sulfur from active
crater lakes. Journal of Volcanology and Geothermal Research 97, 287307.
Kuscu, I., Kuscu, G.G., Tosdal, R.M., Ulrich, T.D., Friedman, R., 2010. Magmatism in the
southeastern Anatolian orogenic belt: transition from arc to post-collisional setting in
anevolving orogen. Special Publications, 340. Geological Society, London, pp. 437460.
Kushiro, I., Syono, Y., Akimoto, S., 1968. Melting of a peridotite nodule at high pressures
and high water pressures. Journal of Geophysical Research 73, 60236029.
22 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Landtwing, M.R., Pettke, T., Halter, W.E., Heinrich, C.A., Redmond, P.B., Einaudi, M.T.,
Kunze, K., 2005. Copper deposition during quartz dissolution by cooling magmatic
hydrothermal uids: the Bingham porphyry. Earth and Planetary Science Letters
235, 229243.
Landtwing, M.R., Furrer, C., Redmond, P.B., Pettke, T., Guillong, M., Heinrich, C.A., 2010.
The Bingham Canyon porphyry CuMoAu deposit III. Zoned coppergold ore
deposition by magmatic vapor expansion. Economic Geology 105, 91118.
Larocque, J., Canil, D., 2010. The role of amphibole in the evolution of arc magmas and
crust: the case from the Jurassic Bonanza arc section, Vancouver Island, Canada.
Contributions to Mineralogy and Petrology 159, 475492.
Larocque, A.C.L., Stimac, J.A., Siebe, C., Greengrass, K., Chapman, R., Mejia, S.R., 2008.
Deposition of a high-suldation Au assemblage froma magmatic volatile phase, Volcn
Popocatpetl, Mexico. Journal of Volcanology and Geothermal Research 170, 5160.
Leeman, W.P., 1983. The inuence of crustal structure on compositions of subduction-
related magmas. Journal Volcanology and Geothermal Research 18, 561588.
Lehmann, B., 1982. Metallogeny of tin: magmatic diferentiation versus geological
heritage. Economic Geology 77, 5059.
Lewis, K.C., Lowell, R.P., 2009. Numerical modeling of two-phase ow in the NaClH
2
O
system: 2. Examples. Journal of Geophysical Research114. doi:10.1029/2008JB006030
16 pp., B08204.
Lowell, J.D., Guilbert, J.M., 1970. Lateral and vertical alteration-mineralization zoning in
porphyry copper ore deposits. Economic Geology 65, 373408.
Lowenstern, J.B., 2001. Carbon dioxide in magmas and implications for hydrothermal
systems. Mineralium Deposita 36, 490502.
Lowenstern, J.B., Mahood, G.A., Rivers, M.L., Sutton, S.R., 1991. Evidence for extreme
partitioning of copper into a magmatic vapor phase. Science 252, 14051409.
Lynton, S.J., Candela, P.A., Piccoli, P.M., 1993. An experimental study of the partitioning
of copper between pyrrhotite and a high silica rhyolitic melt. Economic Geology 88,
901915.
MacDonald, R., Hawkesworth, C.J., Heath, E., 2000. The Lesser Antilles volcanic chain: a
study in arc magmatism. Earth-Science Reviews 49, 176.
Macfarlane, A.W., 1999. Isotopic studies of northern Andean crustal evolution and ore
metal sources. In: Skinner, B.J. (Ed.), Geology and Ore Deposits of the Central Andes. :
Special Publication, No. 7. Society of Economic Geologists, pp. 195217.
Macfarlane, A.W., Marcet, P., LeHuray, A.P., Petersen, U., 1990. Lead isotope provinces of
the Central Andes inferred from ores and crustal rocks. Economic Geology 85,
18571880.
Macpherson, C.G., Dreher, S.T., Thirlwall, M.F., 2006. Adakites without slab melting:
high pressure differentiation of island arc magma, Mindanao, the Philippines. Earth
and Planetary Science Letters 243, 581593.
Malaspina, N., Poli, S., Fumagalli, P., 2009. The oxidation state of metasomatized mantle
wedge: insights from COH-bearing garnet peridotite. Journal of Petrology 50,
15331552.
Mancano, D.P., Campbell, A.R., 1995. Microthermometry of enargite-hosted uid in-
clusions from the Lepanto, Philippines, high-suldation CuAu deposit. Geochimica et
Cosmochimica Acta 59, 39093916.
Manning, C.E., 2004. The chemistry of subduction-zone uids. Earth and Planetary
Science Letters 223, 116.
Manske, S.L., Hedenquist, J.W., O'Connor, G., Tma, C., Cauuet, B., Leary, S., Minut, A.,
2006. Roia Montan, Romania: Europe's largest gold deposit January Society of
Economic Geologists Newsletter 64 (1), 915.
Martin, H., 1999. Adakitic magmas: modern analogues of Archaean granitoids. Lithos
46, 411429.
Martin, H., Smithies, R.H., Rapp, R., Moyen, J.-F., Champion, D., 2005. An overview of
adakite, tonalitetrondhjemitegranodiorite (TTG), and sanukitoid: relationships
and some implications for crustal evolution. Lithos 79, 124.
Maughan, D.T., Keith, J.D., Christiansen, E.H., Pulsipher, T., Hattori, K., Evans, N.J., 2002.
Contributions from mac alkaline magmas to the Bingham porphyry CuAuMo
deposit, Utah, USA. Mineralium Deposita 37, 1437.
Mavrogenes, J.A., Bodnar, R.J., 1994. Hydrogen movement into and out of uid
inclusions in quartz; experimental evidence and geologic implications. Geochimica
et Cosmochimica Acta 58, 141148.
McInnes, B.I.A., Cameron, E.M., 1994. Carbonated, alkaline hybridizing melts froma sub-
arc environment: mantle wedge samples from the Tabar-Lihir-Tanga-Feni arc,
Papua New Guinea. Earth and Planetary Science Letters 122, 125141.
McInnes, B.I.A., McBride, J.S., Evans, N.J., Lambert, D.D., Andrew, A.A., 1999. Osmiumisotope
constraints on ore metal recycling in subduction zones. Science 286, 512516.
McInnes, B.I.A., Gregoire, M., Binns, R.A., Herzig, P.M., Hannington, M.D., 2001. Hydrous
metasomatism of oceanic sub-arc mantle, Lihir, Papua New Guinea: petrology and
geochemistry of uid-metasomatised mantle wedge xenoliths. Earth and Planetary
Science Letters 188, 169183.
McNutt, R.H., Clark, A.H., Zentilli, M., 1979. Lead isotopic compositions of Andean
igneous rocks, Latitudes 26 to 29 S: petrologic and metallogenic implications.
Economic Geology 74, 827837.
Meinert, L.D., Dipple, G.M., Nicolescu, S., 2005. World skarn deposits. In:
Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), Economic
Geology 100th Anniversary Volume. Society of Economic Geologists, Littleton,
CO, pp. 299336.
Mitchell, R.H., Keays, R.R., 1981. Abundance and distribution of gold, palladium and
iridiumin some spinel and garnet lherzolites: implications for the nature and origin
of precious metal-rich intergranular components in the upper mantle. Geochimica
et Cosmochimica Acta 45, 24252442.
Moore, W.J., Nash, J.T., 1974. Alteration and uid inclusion studies of the porphyry
copper ore body at Bingham, Utah. Economic Geology 69, 631645.
Morris, J.D., Leeman, W.P., Tera, F., 1990. The subducted component in island arc lavas:
constraints from Be isotopes and BBe systematics. Nature 344, 3136.
Mukasa, S.B., Vidal, C.E., Injoque-Espinoza, J., 1990. Pb isotope bearing on the
metallogenesis of sulde ore deposits in central and southern Peru. Economic
Geology 85, 14381446.
Mller, D., Groves, D.I., 1993. Direct and indirect associations between potassic igneous
rocks, shoshonites and goldcopper deposits. Ore Geology Reviews 8, 383406.
Mungall, J.E., 2002. Roasting the mantle: slab melting and the genesis of major Au and
Au-rich Cu deposits. Geology 30, 915918.
Mntener, O., Ulmer, P., 2006. Experimentally derived high-pressure cumulates from
hydrous arc magmas and consequences for the seismic velocity structure of lower
arc crust. Geophysical Research Letters 33, L21308. doi:10.1029/2006GL027629.
Murakami, H., Seo, J.H., Heinrich, C.A., 2010. The relationbetweenCu/Auratioandformation
depth of porphyry-style CuAuMo deposits. Mineralium Deposita 45, 1121.
Mutschler, F.E., Grifn, M.E., Stevens, D.S., Shannon, S.S., 1985. Precious metal deposits
related to alkaline rocks in the North American Cordillera an interpretive review.
Transactions, 88. Geological Society of South Africa, pp. 355377.
Naldrett, A.J., 1989. Sulde meltscrystallization temperatures, solubilities in silicate
melts, and Fe, Ni, and Cu partitioning between basaltic magmas, and olivine. In:
Whitney, J.A., Naldrett, A.J. (Eds.), Ore Deposition Associated with Magmas.
Reviews in Economic Geology, 4. Society of Economic Geologists, pp. 520. ch. 2,.
Naney, M.T., 1983. Phase equilibria of rock-forming ferromagnesian silicates in granitic
systems. American Journal of Science 283, 9931033.
Nash, J.T., 1976. Fluid-inclusion petrology data from porphyry copper deposits and
applications to exploration. Professional Paper, 907-D. U.S. Geological Survey. 16 pp.
Neubauer, F., Lips, A., Kouzmanov, K., Lexa, J., Ivascanu, P., 2005. Subduction, slab
detachment and mineralization: the Neogene in the Apuseni Mountains and
Carpathians. Ore Geology Reviews 27, 1344.
Neuendorf, K.K.E., Mehl Jr., J.P., Jackson, J.A. (Eds.), 2005. Glossary of Geology, 5th
edition. American Geological Institute, Alexandria, Virginia. 800 pp.
Newberry, R.J., Swanson, S.E., 1986. Scheelite skarn granitoids: an evaluation of the
roles of magmatic source and process. Ore Geology Reviews 1, 5781.
Noll, P.D., Newsom, H.E., Leeman, W.P., Ryan, J.G., 1996. The role of hydrothermal uids
in the production of subduction zone magmas: evidence from siderophile and
chalcophile trace elements and boron. Geochim. Cosmochim. Acta 60, 587611.
Noorollahi, Y., Itoi, R., Fujii, H., Tanaka, T., 2007. GIS model for geothermal resource
exploration in Akita and Iwate prefectures, northern Japan. Computers and
Geosciences 33, 10081021.
Norton, D.L., 1982. Fluid and heat transport phenomena typical of copper-bearing
pluton environments. In: Titley, S.R. (Ed.), Advances in Geology of Porphyry Copper
Deposits of Southwestern North America. Tuscon, Univ, Arizona Press, pp. 5972.
Norton, D.L., Cathles, L.M., 1973. Breccia pipes, products of exsolved vapor from
magmas. Economic Geology 68, 540546.
Norton, D.L., Dutrow, B.L., 2001. Complex behavior of magmahydrothermal processes:
role of supercritical uid. Geochimica et Cosmochimica Acta 65, 40094017.
Oyarzun, R., Mrquez, A., Lillo, J., Lpez, I., Rivera, S., 2001. Giant versus small porphyry
copper deposits of Cenozoic age in northern Chile: adakitic versus normal calc-
alkaline magmatism. Mineralium Deposita 36, 794798.
Oyarzun, R., Mrquez, A., Lillo, J., Lpez, I., Rivera, S., 2002. Reply to discussion on Giant
versus small porphyry copper deposits of Cenozoic age in northern Chile: adakitic
versus normal calc-alkaline magmatism by Oyarzun et al. (Mineralium Deposita
36: 794798, 2001). Mineralium Deposita 37, 791794.
Parkinson, I.J., Arculus, R.J., 1999. The redox state of subduction zones: insights from
arc-peridotites. Chemical Geology 160, 409423.
Paterson, J.T., Cloos, M., 2005. Grasberg porphyry CuAu deposit, Papua, Indonesia: 1.
Magmatic history. In: Porter, T.M. (Ed.), Super Porphyry Copper and Gold Deposits:
A Global Perspective, 2. Porter Geoscience Consulting Publishing, perspectiveLin-
den Park, South Australia, pp. 313329.
Peach, C.L., Mathez, E.A., Keays, R.R., 1990. Sulde meltsilicate melt distribution
coefcients for noble metals and other chalcophile elements as deduced from
MORB: implications for partial melting. Geochimica et Cosmochimica Acta 54,
33793389.
Peacock, S.M., 1993. Large-scale hydration of the lithosphere above subducting slabs.
Chemical Geology 108, 4959.
Peacock, S.M., 1996. Thermal andpetrologic structure of subductionzones. In: Bebout, G.E.,
Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction: Top to Bottom. : Geophysical
Monograph, 96. American Geophysical Union, Washington, DC, pp. 119133.
Peacock, S.M., Rushmer, T., Thompson, A.B., 1994. Partial melting of subducting oceanic
crust. Earth and Planetary Science Letters 121, 227244.
Pearce, J.A., Bender, J.F., De Long, S.E., Kidd, W.S.F., Low, P.J., Gner, Y., Saroglu, F., Yilmaz, Y.,
Moorbath, S., Mitchell, J.G., 1990. Genesis of collision volcanism in eastern Anatolia,
Turkey. Journal of Volcanology and Geothermal Research 44, 189229.
Petford, N., Gallagher, K., 2001. Partial melting of mac (amphibolitic) lower crust by
periodic inux of basaltic magma. Earth and Planetary Science Letters 193,
483499.
Pettke, T., Oberli, F., Heinrich, C.A., 2010. The magma and metal source of giant
porphyry-type ore deposits, based on lead isotope microanalysis of individual uid
inclusions. Earth and Planetary Science Letters 296, 267277.
Pichavant, M., Mysen, B.O., Macdonald, R., 2002. Source and H
2
O content of high-MgO
magmas in island arc settings: an experimental study of a primitive calc-alkaline
basalt from St. Vincent, Lesser Antilles arc. Geochimica et Cosmochimica Acta 66,
21932209.
Pitcher, W.S., 1997. The Nature and Origin of Granite, 2nd edition. Chapman and Hall,
London. 387 pp.
Pitzer, K.S., Pabalan, R.T., 1986. Thermodynamics of NaCl in steam. Geochimica et
Cosmochimica Acta 50, 14451454.
Plank, T., 2005. Constraints fromthorium/lanthanumon sediment recycling at subduction
zones and the evolution of the continents. Journal of Petrology 46, 921944.
23 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Pokrovski, G.S., Roux, J., Harrichoury, J.-C., 2005. Fluid density control on vaporliquid
partitioning of metals in hydrothermal systems. Geology 33, 657660.
Pokrovski, G.S., Borisova, A.Y., Harrichoury, J.-C., 2008. The effect of sulfur on vapor
liquid fractionation of metals in hydrothermal systems. Earth and Planetary Science
Letters 266, 345362.
Poli, S., Schmidt, M.W., 2002. Petrology of subducted slabs. Annual Reviews of Earth and
Planetary Sciences 30, 207235.
Portnyagin, M., Hoernle, K., Plechov, P., Mironov, N., Khubunaya, S., 2007. Constraints on
mantle melting and composition and nature of slab components in volcanic arcs
from volatiles (H
2
O, S, Cl, F) and trace elements in melt inclusions from the
Kamchatka Arc. Earth and Planetary Science Letters 255, 5369.
Pudack, C., Halter, W.E., Heinrich, C.A., Pettke, T., 2009. Evolution of magmatic vapor to
gold-rich epithermal liquid: the porphyry to epithermal transition at Nevados de
Famatina, Northwest Argentina. Economic Geology 104, 449477.
Rabbia, O.M., Hernndez, L.B., King, R.W., Lpez-Escobar, L., 2002. Discussion on Giant
versus small porphyry copper deposits of Cenozoic age in northern Chile: adakitic
versus normal calc-alkaline magmatism by Oyarzun et al. (Mineralium Deposita
36: 794798, 2001). Mineralium Deposita 37, 791794.
Rapp, R.P., Watson, E.B., 1995. Dehydration melting of metabasalt at 832 kbar:
implications for continental growth and crustmantle recycling. Journal of
Petrology 36, 891931.
Rapp, R.P., Watson, E.B., Miller, C.F., 1991. Partial melting of amphibolite/eclogite and
the origin of Archean trondhjemites and tholeiites. Precambrian Research 51,
125.
Redmond, P.B., Einaudi, M.T., Inan, E.E., Landtwing, M.R., Heinrich, C.A., 2004. Copper
deposition by uid cooling in intrusion-centered systems: new insights from the
Bingham porphyry ore deposit, Utah. Geology 32, 217220.
Reeves, E.P., Seewald, J.S., Saccocia, P., Walsh, E., Bach, W., Craddock, P.R., Shanks, W.C.,
Sylva, S.P., Pichler, T., Rosner, M., 2010. Geochemistry of hydrothermal uids from
the PACMANUS, Northeast Pual and Vienna Woods hydrothermal elds, Manus
Basin, Papua New Guinea. Geochimica et Cosmochimica Acta 75, 10881123.
Richards, J.P., 1995. Alkalic-type epithermal golddeposits a review. In: Thompson, J.F.H.
(Ed.), Magmas, Fluids, and Ore Deposits. Short Course Series, 23. Mineralogical
Association of Canada, pp. 367400. ch. 17.
Richards, J.P., 2002. Discussion of Giant versus small porphyry copper deposits of
Cenozoic age in northern Chile: adakitic versus normal calc-alkaline magmatism
by Oyarzun et al. (Mineralium Deposita 36: 794798, 2001). Mineralium Deposita
37, 788790.
Richards, J.P., 2003. Tectono-magmatic precursors for porphyry Cu(MoAu) deposit
formation. Economic Geology 96, 15151533.
Richards, J.P., 2005. Cumulative factors in the generation of giant calc-alkaline porphyry
Cu deposits. In: Porter, T.M. (Ed.), Super Porphyry Copper and Gold Deposits: A
Global Perspective, 1. Porter Geoscience Consulting Publishing, Linden Park, South
Australia, pp. 725.
Richards, J.P., 2009. Postsubduction porphyry CuAu and epithermal Au deposits:
products of remelting of subduction-modied lithosphere. Geology 37, 247250.
Richards, J.P., Kerrich, R., 1993. The Porgera gold mine, Papua New Guinea: magmatic
hydrothermal to epithermal evolution of an alkalic-type precious metal deposit.
Economic Geology 88, 10171052.
Richards, J.P., Kerrich, R., 2007. Adakite-like rocks: their diverse origins and
questionable role in metallogenesis. Economic Geology 102, 537576.
Richards, J.P., Chappell, B.W., McCulloch, M.T., 1990. Intraplate-type magmatism in a
continentisland-arc collision zone: Porgera intrusive complex, Papua NewGuinea.
Geology 18, 958961.
Richards, J.P., Ullrich, T., Kerrich, R., 2006a. The Late MioceneQuaternary Antofalla
volcanic complex, southern Puna, NW Argentina: protracted history, diverse
petrology, and economic potential. Journal of Volcanology and Geothermal
Research 152, 197239.
Richards, J.P., Wilkinson, D., Ullrich, T., 2006b. Geology of the Sari Gunay epithermal
gold deposit, northwest Iran. Economic Geology 101, 14551496.
Righter, K., Campbell, A.J., Humayun, M., Hervig, R.L., 2004. Partitioning of Ru, Rh, Pd, Re,
Ir, and Au between Cr-bearing spinel, olivine, pyroxene and silicate melts.
Geochimica et Cosmochimica Acta 68, 867880.
Ronacher, E., Richards, J.P., Reed, M.H., Bray, C.J., Spooner, E.T.C., Adams, P.D., 2004.
Characteristics and evolution of the hydrothermal uid in the North Zone high-
grade area, Porgera gold deposit, Papua New Guinea. Economic Geology 99,
843867.
Rowe, M.C., Kent, A.J.R., Nielsen, R.L., 2009. Subduction inuence on oxygen fugacity
and trace and volatile elements in basalts across the Cascade Volcanic Arc. Journal
of Petrology 50, 6191.
Rubin, A.M., 1995. Propagation of magma-lled cracks. Annual Review of Earth &
Planetary Sciences 23, 287336.
Rudnick, R.L., Gao, S., 2003. Composition of the continental crust. In: Rudnick, R.L. (Ed.),
Treatise on Geochemistry 3: The Crust. Elsevier, Amsterdam, pp. 164.
Rushmer, T., 1991. Partial melting of two amphibolites: contrasting experimental
results under uid-absent conditions. Contributions to Mineralogy and Petrology
107, 4159.
Rushmer, T., 1993. Experimental high-pressure granulites: some applications to natural
mac xenolith suites and Archean granulite terranes. Geology 21, 411414.
Rusk, B.G., Reed, M.H., Dilles, J.H., Klemm, L.M., Heinrich, C.A., 2004. Compositions of
magmatic hydrothermal uids determined by LA-ICP-MS of uid inclusions from
the porphyry coppermolybdenum deposit at Butte, MT. Chemical Geology 210,
173199.
Rusk, B.G., Reed, M.H., Dilles, J.H., 2008. Fluid inclusion evidence for magmatic
hydrothermal uid evolution in the porphyry coppermolybdenum deposit at
Butte, Montana. Economic Geology 103, 307334.
Rutherford, M.J., Devine, J.D., 1988. The May 18, 1980, eruption of Mount St. Helens. 3.
Stability and chemistry of amphibole in the magma chamber. Journal of
Geophysical Research 93 (11), 959 949-11.
Ryerson, F.J., Watson, E.B., 1987. Rutile saturation in magmas: implications for TiNb
Ta depletion in island-arc basalts. Earth and Planetary Science Letters 86, 225239.
Sajona, F.G., Maury, R.C., 1998. Association of adakites with gold and copper
mineralization in the Philippines. Comptes rendus de l'Acadmie des sciences,
Srie II, Sciences de la terre et des plantes 326, 2734.
Sakai, H., Matsubaya, O., 1977. Stable isotopic studies of Japanese geothermal systems.
Geothermics 5, 97124.
Sawkins, F.J., Scherkenbach, D.A., 1981. High copper content of uid inclusions in quartz
from northern Sonora: implications for ore-genesis theory. Geology 9, 3740.
Scambelluri, M., Mntener, O., Ottolini, L., Pettke, T.P., Vannuccie, R., 2004. The fate of B,
Cl and Li in the subducted oceanic mantle and in the antigorite breakdown uids.
Earth and Planetary Science Letters 222, 217234.
Scherbarth, N.L., Spry, P.G., 2006. Mineralogical, petrological, stable isotope, and uid
inclusion characteristics of the Tuvatu GoldSilver Telluride Deposit, Fiji:
comparisons with the Emperor Deposit. Economic Geology 101, 135158.
Schiano, P., Clocchiatti, R., Shimizu, N., Maury, R.C., Jochum, K.P., Hofmann, A.W., 1995.
Hydrous, silica-rich melts in the sub-arc mantle and their relationship with erupted
arc lavas. Nature 277, 595600.
Schmidt, M.W., Poli, S., 1998. Experimentally based water budgets for dehydrating slabs
and consequences for arc magma generation. Earth and Planetary Science Letters
163, 361379.
Schmidt, M.W., Dardon, A., Chazot, G., Vannucci, R., 2004. The dependence of Nb and Ta
rutile-melt partitioning on melt composition and Nb/Ta fractionation during
subduction processes. Earth and Planetary Science Letters 226, 415432.
Schmidt, A., Weyer, S., John, T., Brey, G.P., 2009. HFSE systematics of rutile-bearing
eclogites: new insights into subduction zone processes and implications for the
earth's HFSE budget. Geochimica et Cosmochimica Acta 73, 455468.
Seedorff, E., Dilles, J.H., Proffett Jr., J.M., Einaudi, M.T., Zurcher, L., Stavast, W.J.A.,
Johnson, D.A., Barton, M.D., 2005. Porphyry deposits: characteristics and origin of
hypogene features. Economic Geology 100th Anniversary Volume, pp. 251298.
Seo, J.H., Guillong, M., Heinrich, C.A., 2009. The role of sulfur in the formation of
magmatichydrothermal coppergold deposits. Earth and Planetary Science
Letters 282, 323328.
Settereld, T.N., Mussett, A.E., Oglethorpe, R.D.J., 1992. Magmatism and associated
hydrothermal activity during the evolution of the Tavua caldera:
40
Ar/
39
Ar dating
of volcanic, intrusive, and hydrothermal events. Economic Geology 87, 11301140.
Shaei, B., Haschke, M., Shahabpour, J., 2009. Recycling of orogenic arc crust triggers
porphyry Cu mineralization in Kerman Cenozoic arc rocks, southeastern Iran.
Mineralium Deposita 44, 265283.
Shinohara, H., 1994. Exsolution of immiscible vapor and liquid phases from a
crystallizing silicate melt: implications for chlorine and metal transport. Geochi-
mica et Cosmochimica Acta 58, 52155221.
Shinohara, H., Hedenquist, J.W., 1997. Constraints on magma degassing beneath the Far
Southeast porphyry CuAu deposit, Philippines. Journal of Petrology 38, 17411752.
Shinohara, H., Kazahaya, K., Lowenstern, J.B., 1995. Volatile transport in a convecting
magma column: implications for porphyry Mo mineralization. Geology 23,
10911094.
Sillitoe, R.H., 1973. The tops and bottoms of porphyry copper deposits. Economic
Geology 68, 799815.
Sillitoe, R.H., 2000. Gold-rich porphyry deposits: descriptive and genetic models and
their role in exploration and discovery. Reviews in Economic Geology 13,
315345.
Sillitoe, R.H., 2010. Porphyry copper systems. Economic Geology 105, 341.
Sillitoe, R.H., Hart, S.R., 1984. Lead-isotope signatures of porphyry copper deposits in
oceanic and continental settings, Colombian Andes. Geochimica et Cosmochimica
Acta 48, 21352142.
Simmons, S.F., Brown, K.L., 2007. The ux of gold and related metals through a volcanic
arc, Taupo Volcanic Zone, New Zealand. Geology 35, 10991102.
Simon, A.C., Frank, M.R., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2005. Gold
partitioning in meltvaporbrine systems. Geochimica et Cosmochimica Acta 69,
33213335.
Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2006. Copper
partitioning in a meltvaporbrinemagnetitepyrrhotite assemblage. Geochimica
et Cosmochimica Acta 70, 55835600.
Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2007. The partitioning
behavior of As and Au in S-free and S-bearing magmatic assemblages. Geochimica
et Cosmochimica Acta 71, 17641782.
Simon, A.C., Candela, P.A., Piccoli, P.M., Mengason, M., Englander, L., 2008. The effect of
crystal-melt partitioning on the budgets of Cu, Au, and Ag. American Mineralogist
93, 14371448.
Sinclair, W.D., 2007. Porphyry deposits. In: Goodfellow, W.D. (Ed.), Mineral Deposits of
Canada: A Synthesis of Major Deposit-Types, District Metallogeny, the Evolution of
Geological Provinces, and Exploration Methods. Geological Association of Canada,
St. John's, Newfoundland, pp. 223243.
Smith, I.E.M., Stewart, R.B., Price, R.C., Worthington, T.J., 2010. Are arc-type rocks the
products of magma crystallisation? Observations from a simple oceanic arc
volcano: Raoul Island, Kermadec Arc, SW Pacic. Journal of Volcanology and
Geothermal Research 190, 219234.
Sobolev, A., Chaussidon, M., 1996. H
2
O concentrations in primary melts from supra-
subduction zones and mid-ocean ridges: implications for H
2
Ostorage and recycling
in the mantle. Earth and Planetary Science Letters 137, 4555.
Solomon, M., 1990. Subduction, arc reversal, and the origin of porphyry coppergold
deposits in island arcs. Geology 18, 630633.
24 J.P. Richards / Ore Geology Reviews 40 (2011) 126
Sourirajan, S., Kennedy, G.C., 1962. The system H
2
ONaCl at elevated temperatures and
pressures. American Journal of Science 260, 115141.
Spooner, E.T.C., 1993. Magmatic sulphide/volatile interaction as a mechanism for
producing chalcophile element enriched, Archean Auquartz, epithermal AuAg
and Au skarn hydrothermal ore uids. Ore Geology Reviews 7, 359379.
Staudigel, H., Plank, T., White, B., Schmincke, H.-U., 1996. Geochemical uxes during seaoor
alteration of the basaltic upper oceanic crust: DSDP sites 417 and 418. In: Bebout, G.E.,
Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction Top to Bottom. Geophysical
Monograph, 96. American Geophysical Union, Washington, DC, pp. 1938.
Stavast, W.J.A., Keith, J.D., Christiansen, E.H., Dorais, M.J., Tingey, D., 2006. The fate of
magmatic suldes during intrusion or eruption, Bingham and Tintic districts, Utah.
Economic Geology 101, 329345.
Stein, H.J., 1988. Genetic traits of climax-type granites and molybdenum mineraliza-
tion, Colorado Mineral Belt. In: Taylor, R.P., Strong, D.F. (Eds.), Recent Advances in
the Geology of Granite-Related Mineral Deposits, Special Volume 39. Canadian
Institute of Mining and Metallurgy, pp. 394401.
Stern, R.J., Kohut, E., Bloomer, S.H., Leybourne, M., Fouch, M., Vervoort, J., 2006.
Subduction factory processes beneath the Guguan cross-chain, Mariana Arc: no role
for sediments, are serpentinites important? Contributions to Mineralogy and
Petrology 151, 202221.
Stoffregen, R., 1987. Genesis of acid-sulfate alteration and AuCuAg mineralization at
Summitville, Colorado. Economic Geology 82, 15751591.
Stolper, E., Newman, S., 1994. The role of water in the petrogenesis of Mariana trough
magmas. Earth and Planetary Science Letters 121, 293325.
Sun, S.-s., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
implications for mantle composition and processes. In: Saunders, A.D., Norry, M.J.
(Eds.), Magmatism in the Ocean Basins: Geological Society of London Special
Publication, No. 42, pp. 313345.
Sun, W., Bennett, V.C., Kamenetsky, V.S., 2004a. The mechanism of Re enrichment in arc
magmas: evidence from Lau Basin basaltic glasses and primitive melt inclusions.
Earth and Planetary Science Letters 222, 101114.
Sun, W.D., Arculus, R.J., Kamenetsky, V.S., Binns, R.A., 2004b. Release of gold-bearing
uids in convergent margin magmas prompted by magnetite crystallization.
Nature 431, 975978.
Symonds, R.B., Rose, W.I., Reed, M.H., Lichte, F.E., Finnegan, D.L., 1987. Volatilization,
transport and sublimation of metallic and non-metallic elements in high
temperature gases at Merapi Volcano, Indonesia. Geochimica et Cosmochimica
Acta 51, 20832101.
Tarana, Y.A., Hedenquist, J.W., Korzhinsky, M.A., Tkachenko, S.I., Shmulovich, K.I., 1995.
Geochemistry of magmatic gases from Kudryavy volcano, Iturup, Kuril Islands.
Geochimica et Cosmochimica Acta 59, 17491761.
Tatsumi, Y., 1986. Formation of the volcanic front in subduction zones. Geophysical
Research Letters 17, 717720.
Tatsumi, Y., 2003. Some constraints on arc magma genesis. In: Eiler, J. (Ed.), Inside the
Subduction Factory. : Geophysical Monograph, 138. American Geophysical Union,
Washington, DC, pp. 277292.
Tatsumi, Y., Eggins, S., 1995. Subduction Zone Magmatism. Blackwell, Oxford. 213 pp.
Tatsumi, Y., Hamilton, D.L., Nesbitt, R.W., 1986. Chemical characteristics of uid phase
released froma subducted lithosphere and the origin of arc magmas: evidence from
high pressure experiments and natural rocks. Journal of Volcanology and
Geothermal Research 29, 293309.
Taylor, S.R., McLennan, S.M., 1985. The continental crust: its composition and evolution:
an examination of the geochemical record preserved in sedimentary rocks.
Blackwell Scientic 312p.
Tessalina, S.G., Yudovskaya, M.A., Chaplygin, I.V., Birck, J.-L., Capmas, F., 2008. Sources of
unique rhenium enrichment in fumaroles and sulphides at Kudryavy volcano.
Geochimica et Cosmochimica Acta 72, 889909.
Thiblemont, D., Stein, G., Lescuyer, J.-L., 1997. Gisements pithermaux et porphyr-
iques: la connexion adakite. C.R. Acad. Sci. Paris, Sciences de la terre et des plantes/
Earth and Planetary Sciences 325, 103109.
Thirlwall, M.F., Graham, A.M., Arculus, R.J., Harmon, R.S., Macpherson, C.G., 1996.
Resolution of the effects of crustal assimilation, sediment subduction, and uid
transport in island arc magmas: PbSrNdO isotope geochemistry of Grenada,
Lesser Antilles. Geochimica et Cosmochimica Acta 60, 47854810.
Thompson, T.B., Trippel, A.D., Dwelley, P.C., 1985. Mineralized veins and breccias of the
Cripple Creek District, Colorado. Economic Geology 80, 16691688.
Thorkelson, D.J., Breitsprecher, K., 2005. Partial melting of slab window margins:
genesis of adakitic and non-adakitic magmas. Lithos 79, 2541.
Tiepolo, M., Tribuzio, R., 2008. Petrology and UPb zircon geochronology of amphibole-
rich cumulates with sanukitic afnity from Husky Ridge (Northern Victoria Land,
Antarctica): insights into the role of amphibole in the petrogenesis of subduction-
related magmas. Journal of Petrology 49, 937970.
Tilton, G.R., Pollak, R.J., Clark, A.H., Robertson, R.C.R., 1981. Isotopic composition of Pb in
central Andean ore deposits. Geological Society of America, Memoir 154, 791816.
Titley, S.R., 1987. The crustal heritage of silver and gold ratios in Arizona ores.
Geological Society of America, Bulletin 99, 814826.
Titley, S.R., 2001. Crustal afnities of metallogenesis in the American Southwest.
Economic Geology 96, 13231342.
Tomkins, A.G., Mavrogenes, J.A., 2003. Generation of metal-rich felsic magmas during
crustal anatexis. Geology 31, 765768.
Tosdal, R.M., Richards, J.P., 2001. Magmatic and structural controls on the development
of porphyry CuMoAu deposits. In: Richards, J.P., Tosdal, R.M. (Eds.), Structural
Controls on Ore Genesis. : Reviews in Economic Geology, 14. Society of Economic
Geologists, pp. 157181.
Tulloch, A.J., Kimbrough, D.L., 2003. Paired plutonic belts in convergent margins and the
development of high Sr/Y magmatism: Peninsular Ranges batholith of Baja
California and Median batholith of New Zealand. Special Paper, 374. Geological
Society of America, pp. 121.
Ulmer, P., Trommsdorff, V., 1995. Serpentine stability to mantle depths and subduction-
related magmatism. Science 268, 858861.
Ulrich, T., Gnther, D., Heinrich, C.A., 2001. The evolution of a porphyry CuAu deposit,
based on LA-ICP-MS analysis of uid inclusions: Bajo de la Alumbrera, Argentina.
Economic Geology 96, 17431774.
van Dongen, M., Weinberg, R.F., Tomkins, A.G., Armstrong, R.A., Woodhead, J.D.,
2010. Recycling of Proterozoic crust in Pleistocene juvenile magma and rapid
formation of the Ok Tedi porphyry CuAu deposit, Papua New Guinea. Lithos
114, 282292.
Vry, V.H., Wilkinson, J.J., Seguel, J., Millan, J., 2010. Multistage intrusion, brecciation, and
veining at El Teniente, Chile: evolution of a nested porphyry system. Economic
Geology 105, 119153.
Walker, J.A., Patino, L.C., Carr, M.J., Feigenson, M.D., 2001. Slab control over HFSE
depletions in central Nicaragua. Earth and Planetary Science Letters 192, 533543.
Wallace, P.J., 2005. Volatiles in subduction zone magmas: concentrations and uxes
based on melt inclusion and volcanic gas data. Journal of Volcanology and
Geothermal Research 140, 217240.
Walshe, J.L., Solomon, M., Whitford, D.J., Sun, S.-S., Foden, J.D., 2011. The role of the
mantle in the genesis of tin deposits and tin provinces of eastern Australia.
Economic Geology 106, 297305.
Wang, J., Hattori, K.H., Kilian, R., Stern, C.R., 2007a. Metasomatism of sub-arc mantle
peridotites below southernmost South America: reduction of fO
2
by slab-melt.
Contributions to Mineralogy and Petrology 153, 607624.
Wang, Q., Wyman, D.A., Xu, J.-F., Zhao, Z.-H., Jian, P., Zi, F., 2007b. Partial melting of
thickened or delaminated lower crust in the middle of Eastern China: implications
for CuAu mineralization. Journal of Geology 115, 149161.
Webster, J.D., 1992. Water solubility and chlorine partitioning in Cl-rich granitic
systems: effects of melt composition at 2 kbar and 800 C. Geochimica et
Cosmochimica Acta 56, 679687.
Westra, G., Keith, S.B., 1981. Classication and genesis of stockwork molybdenum
deposits. Economic Geology 76, 844873.
White, D.E., Mufer, L.J.P., Truesdell, A.H., 1971. Vapor-dominated hydrothermal
systems compared with hot-water systems. Economic Geology 66, 7597.
White, W.H., Bookstrom, A.A., Kamilli, R.J., Gangster, M.W., Smith, R.P., Ranta, D.E.,
Steininger, R.C., 1981. Character and origin of climax-type molybdenum deposits.
Economic Geology 75th Anniversary volume, pp. 270316.
Widom, E., Kepezhinskas, P., Defant, M., 2003. The nature of metasomatism in the sub-
arc mantle wedge: evidence from ReOs isotopes in Kamchatka peridotite
xenoliths. Chemical Geology 196, 283306.
Williams, T.J., Candela, P.A., Piccoli, P.M., 1995. The partitioning of copper between
silicate melts and two-phase aqueous uids: an experimental investigation at
1 kbar, 800 C and 0.5 kbar, 850 C. Contributions to Mineralogy and Petrology 121,
388399.
Williams-Jones, A.E., Heinrich, C.A., 2005. Vapor transport of metals and the formation
of magmatichydrothermal ore deposits. Economic Geology 100, 12871312.
Williamson, B.J., Mller, A., Shail, R.K., 2010. Source and partitioning of B and Sn in the
Cornubian batholith of southwest England. Ore Geology Reviews 38, 18.
Winter, J.D., 2001. An introduction to igneous and metamorphic petrology: Upper
Saddle River. Prentice-Hall Inc., New Jersey. 697 p.
Wolf, M.B., Wyllie, P.J., 1994. Dehydration-melting of amphibolite at 10 kbar the
effects of temperature and time. Contributions to Mineralogy and Petrology 115,
369383.
Wood, S.A., Spera, F.J., 1984. Adiabatic decompression of aqueous solutions:
applications to hydrothermal uid migration in the crust. Geology 12, 707710.
Wrner, G., Moorbath, S., Harmon, R.S., 1992. Andean Cenozoic volcanic centers reect
basement isotopic domains. Geology 20, 11031106.
Wyborn, D., Sun, S.-s., 1994. Sulphur-undersaturated magmatism a key factor for
generating magma-related coppergold deposits. AGSO Research Newsletter 21,
78.
Wyllie, P.J., Huang, W.-L., Stern, C.R., Maale, S., 1976. Granitic magmas: possible and
impossible sources, water contents, and crystallization sequences. Canadian
Journal of Earth Sciences 13, 10071019.
Wysoczanski, R.J., Wright, I.C., Gamble, J.A., Hauri, E.H., Luhr, J.F., Eggins, S.M., Handler,
M.R., 2006. Volatile contents of Kermadec ArcHavre Trough pillow glasses:
ngerprinting slab-derived aqueous uids in the mantle sources of arc and back-
arc lavas. Journal of Volcanology and Geothermal Research 152, 5173.
Xiao, Z., Gammons, C.H., Williams-Jones, A.E., 1998. Experimental study of copper(I)
chloride complexing in hydrothermal solutions at 40 to 300 C and saturated water
vapor pressure. Geochimica et Cosmochimica Acta 62, 29492964.
Yang, Z., Hou, Z., White, N.C., Chang, Z., Li, Z., Song, Y., 2009. Geology of the post-
collisional porphyry coppermolybdenum deposit at Qulong, Tibet. Ore Geology
Reviews 36, 133159.
Yogodzinski, G.M., Kay, R.W., Volynets, O.N., Koloskov, A.V., Kay, S.M., 1995. Magnesian
andesite in the western Aleutian Komandorsky region: implications for slab
melting and processes in the mantle wedge. Geological Society of America Bulletin
107, 505519.
Yogodzinski, G.M., Lees, J.M., Churikova, T.G., Dorendorf, F., Werner, G., Volynets, O.N.,
2001. Geochemical evidence for the melting of subducting oceanic lithosphere at
plate edges. Nature 409, 500504.
Zajacz, Z., Halter, W., 2009. Copper transport by high temperature, sulfur-rich magmatic
vapor: evidence fromsilicate melt and vapor inclusions ina basaltic andesite fromthe
Villarrica volcano (Chile). Earth and Planetary Science Letters 282, 115121.
Zajacz, Z., Halter, W.E., Pettke, T., Guillong, M., 2008. Determination of uid/melt
partition coefcients by LA-ICPMS analysis of co-existing uid and silicate melt
25 J.P. Richards / Ore Geology Reviews 40 (2011) 126
inclusions: controls on element partitioning. Geochimica et Cosmochimica Acta 72,
21692197.
Zajacz, Z., Seo, Z.H., Candela, P.A., Piccoli, P.M., Heinrich, C.A., Guillong, M., 2010. Alkali
metals control the release of gold from volatile-rich magmas. Earth and Planetary
Science Letters 297, 5056.
Zajacz, Z., Seo, J.H., Candela, P.A., Piccoli, P.M., Tossell, J.A., 2011. The solubility of copper in
high-temperature magmatic vapors: a quest for the signicance of various chloride
and sulde complexes. Geochimica et Cosmochimica Acta 75, 28112827.
Zhang, X., Spry, P.G., 1994. Petrological, mineralogical, uid inclusion, and stable
isotope studies of the Gies goldsilver telluride deposit, Judith Mountains,
Montana. Economic Geology 89, 602627.
Zimmer, M.M., Plank, T., Hauri, E.H., Yogodzinski, G.M., Stelling, P., Larsen, J., Singer, B.,
Jicha, B., Mandeville, C., Nye, C.J., 2010. The role of water in generating the calc-
alkaline trend: new volatile data for Aleutian magmas and a new Tholeiitic Index.
Journal of Petrology 51, 24112444.
26 J.P. Richards / Ore Geology Reviews 40 (2011) 126

Potrebbero piacerti anche