Sunteți pe pagina 1din 19

Lithos 233 (2015) 27–45

Contents lists available at ScienceDirect

Lithos
journal homepage: www.elsevier.com/locate/lithos

Invited review article

The oxidation state, and sulfur and Cu contents of arc magmas:


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

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

Article history: Global data for measured Fe2O3/FeO ratios and Cu contents in unaltered volcanic and intrusive arc rocks indicate
Received 8 July 2014 that, on average, they are slightly more oxidized than other magmas derived from depleted upper mantle (such
Accepted 11 December 2014 as MORB), but contain similar Cu contents across their compositional ranges. Although Cu scatters to elevated
Available online 26 December 2014
values in some intermediate composition samples, the bulk of the data show a steady but gentle trend to
lower concentrations with differentiation, reaching modal values of ~ 50–100 ppm in andesitic rocks. These
Keywords:
Oxidation state
data suggest that Cu is mildly compatible during partial melting and fractionation processes, likely reflecting
Magma minor degrees of sulfide saturation throughout the magmatic cycle. However, the volume of sulfides must be
Sulfur small such that significant proportions of the metal content remain in the magma during fractionation to inter-
Copper mediate compositions. Previous studies have shown that andesitic magmas containing ~50 ppm Cu can readily
Gold form large porphyry-type Cu deposits upon emplacement in the upper crust.
Porphyry A review of the literature suggests that the elevated oxidation state in the asthenospheric mantle wedge source of
arc magmas (ΔFMQ ≈ +1 ± 1) derives from the subduction of seawater-altered and oxidized oceanic crust, and
is transmitted into the mantle wedge via prograde metamorphic dehydration fluids carrying sulfate and other
oxidizing components. Progressive hydration and oxidation of the mantle wedge may take up to ~10 m.y. to
reach a steady state from the onset of subduction, explaining the rarity of porphyry deposits in primitive island
arcs, and the late formation of porphyries in continental arc magmatic cycles. Magmas generated from this
metasomatized and moderately oxidized mantle source will be hydrous basalts containing high concentrations
of sulfur, mainly dissolved as sulfate or sulfite. Some condensed sulfides (melt or minerals) may be present
due to the high overall fS2, despite the moderately high oxidation state. These sulfides may retain some highly
siderophile elements in the source, but are unlikely to be sufficiently voluminous to significantly affect the budget
of more modestly sulfide-compatible and more abundant elements such as Cu and Mo. These primary magmas
can therefore be considered to be largely Cu-Mo-undepleted, although highly siderophile elements such as Au
and platinum group elements (PGE) may be depleted unless no sulfides remain in the source. The latter condition
seems unlikely during active subduction because of the continuous flux of fresh sulfur from the slab, but may
occur during post-subduction re-melting (leading to potentially Au-rich post-subduction porphyry and alkalic-
type epithermal systems).
Lower crustal differentiation of main-stage arc magmas results in some loss of Cu to residual or cumulate sulfides,
but again the amount appears to be minor, and does not drastically reduce the Cu content of derivative
intermediate-composition melts. Fractionation and devolatilization affect the oxidation state of the magma in
competing ways, but, while crystallization and segregation of Fe3+-rich magnetite can cause reduction in re-
duced to moderately oxidized evolved magmas, this effect appears to be outweighed by the oxidative effects of
degassing reduced or weakly oxidized gaseous species such as H2, H2S, and SIVO2, and preferential solvation
and removal of Fe2+ in saline hydrothermal fluids. Consequently, most arc magmatic suites show slight increases
in oxidation state during differentiation, reaching typical values of ΔFMQ = +1 to +2.
This oxidation state is significant, because it corresponds to the transition from dissolved sulfide to sulfate dom-
inance in magmas. It has been shown that Cu and Au solubilities in silicate magma increase up to this level
(ΔFMQ ≈ +1), but while Cu solubility continues to increase at higher oxidation states, Au shows a precipitous
drop as sulfide, which solvates Au in the melt, is converted to sulfate. This may explain the somewhat restricted
distribution of Au-rich porphyry Cu deposits, but the general association of porphyry Cu deposits with relatively
oxidized magmas.
Exsolution of a saline, high temperature aqueous fluid enables metals to partition from the magma into a highly
mobile volatile phase. Sulfur also partitions strongly into this fluid phase, predominantly as SO2 at ΔFMQ = +1

⁎ Tel.: +1 780 492 3430.


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

http://dx.doi.org/10.1016/j.lithos.2014.12.011
0024-4937/© 2014 Elsevier B.V. All rights reserved.
28 J.P. Richards / Lithos 233 (2015) 27–45

to +2. However, as the fluid cools below ~400 °C, SIVO2 disproportionates to form reduced H2S−II and oxidized
H2SVIO4. The H2S bonds with metals in solution to precipitate as Cu- and Mo-sulfides, while the H2SO4 (and HCl)
generates progressively acidic wallrock alteration (phyllic, argillic, advanced argillic). Gold may precipitate with
early Cu/Mo-sulfides, but some may also stay in solution as bisulfide complexes, eventually reaching the
epithermal environment.
Thus, three components, [S], [H2O], and fO2 work together throughout subduction and arc magmatic processes to
transport chalcophile and siderophile metals from the mantle into the upper crust, where they may be concen-
trated by hydrothermal processes to form ore deposits. These processes are far from 100% efficient, and metals
(especially highly siderophile elements such as Au and PGE) may be left behind at various stages of the passage
of arc magmas through the lithosphere, where they may form potentially metalliferous source rocks for partial
melts and subsequent magmatic-hydrothermal ore deposits generated during later tectonomagmatic events.
© 2014 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2. Evidence for, and causes and effects of oxidation in sub-arc mantle and arc magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1. Methods of estimating magmatic fO2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1.1. Whole-rock Fe3 +/Fe2 + ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1.2. Whole-rock chemical ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1.3. Rare earth elements in rocks and minerals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.1.4. Mineral equilibria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2. Oxidation state of sub-arc mantle xenoliths and orogenic peridotites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.3. Oxidation state of arc magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.4. Redox effects of subduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.5. Partial melting in the mantle wedge and the calc-alkaline trend . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.6. Redox effects on metallogeny during partial melting and magmatic differentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.7. The role of sulfur on metal solubility during partial melting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.8. The source of sulfur in arc magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
3. Lower crustal MASH processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4. Fractional crystallization and devolatilization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
5. The magmatic to hydrothermal transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

1. Introduction being formed from larger volumes of magma. A caveat to this model,
however, is that Cu must behave incompatibly in such magmas: in
A major proportion of the world’s Cu, Mo, and Au is mined from hy- other words, it must not be lost significantly from the melt phase
drothermal ore deposits (typically porphyry and epithermal deposits) through crystallization or early (deep) fluid exsolution prior to
associated either directly or indirectly with subduction-related shallow-level partitioning into a magmatic-hydrothermal system
magmas. While all volcanic arcs feature hydrothermal systems, large (Spooner, 1993). Compatible behavior rapidly strips Cu from the
mineral deposits tend to form in clusters or age-related belts within magma, and reduces the potential to form porphyry-type ore deposits
these arcs, or as isolated deposits in back-arc settings. Cline and (or requires larger volumes of magma to form an equivalent sized de-
Bodnar (1991; updated in Cline, 1995) convincingly demonstrated posit). Copper does not partition strongly into most igneous silicate
that an average arc andesite containing 50 ppm Cu has the potential and oxide minerals, but is readily sequestered by sulfide phases
to form an economic porphyry Cu deposit, but clearly not all do. A ques- (melts, or minerals such as pyrrhotite and intermediate solid solution,
tion therefore remains as to what conditions convert potential to actual ISS; Jugo et al., 1999). Consequently, the quickest way to reduce
ore formation? the economic potential of a magma is to saturate it with a sulfide
The formation of a large porphyry ore deposit requires the consecu- phase, and allow the dense sulfides to settle out at depth prior to hydro-
tive, cumulative, and optimized execution of numerous steps, starting thermal fluid exsolution (unless of course the objective is to form an
with the generation of sufficient volumes of fertile magma, and ending orthomagmatic sulfide deposit; e.g., Hamlyn and Keays, 1986; Hamlyn
with efficient hydrothermal transport and ore depositional processes. et al., 1985).
Malfunctioning or omission of any one of these steps can reduce or Arc magmatic systems are characteristically S-rich (de Hoog et al.,
eliminate the potential for subsequent ore formation within an individ- 2001; Wallace, 2005; Wallace and Edmonds, 2011), and therefore
ual deposit, but failure of an early step is most likely to lead to belt-scale might be expected to undergo early sulfide saturation were it not for an-
barren systems (Richards, 2013). other key characteristic: their relatively high (but variable) oxidation
Cline and Bodnar (1991) modeled the partitioning of Cu and Cl state, compared to most other mantle-derived magmas (Evans, 2012;
between hydrothermal fluids and andesitic melt with an initial compo- Evans et al., 2012; Gill, 1981; Kelley and Cottrell, 2009; Laubier et al.,
sition of 2.5 wt.% H2O, Cl/H2O = 0.1, and 50 ppm Cu, at various pres- 2014; Stamper et al., 2014a). Under relatively oxidizing conditions
sures as a function of the degree of magmatic crystallization. They (ΔFMQ ≥ +1, where ΔFMQ is the log fO2 difference between the sample
concluded that ≤30 km3 of magma with this composition could gener- value and the fayalite–magnetite-quartz mineral buffer), the bulk of the
ate an economic porphyry deposit (≥250 million tons of 0.75 wt.% Cu) sulfur will be present in the melt as SVIO2−4 or SIVO2 rather than S2−,
if emplaced at moderate depths (2 kb), with larger deposits potentially and there will therefore be less tendency to exsolve large volumes of
J.P. Richards / Lithos 233 (2015) 27–45 29

sulfide melt or precipitate sulfide minerals (Carroll and Rutherford, state, and to what degree they are independent of other processes
1985; Métrich and Clocchiatti, 1996; Métrich et al., 2002; Nilsson and such as original (pre-subduction) source composition and partitioning
Peach, 1993). Thus, relatively oxidized arc magmas, with average con- effects (Brounce et al., 2013).
centrations of H2O, Cl, Cu, and S, should meet the primary criteria
for porphyry ore formation of Cline and Bodnar (1991). However, 2.1.3. Rare earth elements in rocks and minerals
unlike the concentrations of these chemical components, which might The rare earth elements (REE) Eu and Ce also occur in two valence
be expected to control ore potential in a proportional fashion, magmatic states in magmas (Eu2+ and Eu3+, and Ce3+ and Ce4+) with different
redox state appears to have an almost stepwise effect, with porphyry Cu partition coefficients into minerals such as plagioclase (in which only
ore potential falling suddenly at oxidation states below ΔFMQ ≈ + 1 Eu2 + is compatible as a substitution for Ca2 +; Frey et al., 1978;
(Botcharnikov et al., 2011; Zajacz et al., 2012). Thus, it is particularly im- Hanson, 1980) and zircon (in which Eu2+ is excluded and Ce4+ is pref-
portant to understand what is controlling the overall redox state of arc erentially accommodated as substitutions for Zr4+; Ballard et al., 2002;
systems, and what conditions might cause changes around the critical Dilles et al., 2015). Consequently, relatively reduced magmas that frac-
value of ΔFMQ ≈ +1. tionate Eu2+-rich plagioclase will show negative Eu anomalies on nor-
This paper explores the controls on this parameter from oxidative malized rare-earth-element diagrams, but oxidized magmas will not.
seafloor alteration, through mantle wedge metasomatism and partial Rare earth element analyses of zircons will also show negative Eu anom-
melting, deep crustal magmatic processes, and finally to shallow-level alies in reduced rocks because Eu2+ is preferentially excluded from zir-
crystallization and devolatilization. Many of the arc-scale variations in con (and the magma may already have been depleted in Eu by
arc magma fertility can be understood in terms of the effects of fO2 on plagioclase fractionation). In contrast, Ce will show positive anomalies
these processes. in zircons from oxidized magmas because Ce4 + is preferentially
partitioned into this mineral. Ballard et al. (2002), Trail et al. (2011,
2. Evidence for, and causes and effects of oxidation in sub-arc mantle 2012) and Burnham and Berry (2012, 2014) have provided empirical
and arc magmas relationships to estimate magmatic oxidation state from the magnitude
of Ce and Eu anomalies in zircon, which offer insights into relatively
Subduction zones have long been thought to generate relatively ox- late stage conditions during crystallization of intermediate-to-felsic
idized magmas (ΔFMQ = +0.5 to + 2; Ballhaus et al., 1991; Brounce magmas.
et al., 2014; Gill, 1981; Rowe et al., 2009; Zimmer et al., 2010), compared
with other depleted mantle-derived magmas such as tholeiitic mid- 2.1.4. Mineral equilibria
ocean ridge basalts (e.g., MORB, ΔFMQ = −1 to 0; Bézos and Humler, Electron microprobe analyses of coexisting mineral pairs are widely
2005; Christie et al., 1986; Frost and McCammon, 2008; Laubier et al., used to estimate conditions of crystallization such as pressure, temper-
2014). However, estimating the oxidation state of magmas and source ature, and fO2, assuming chemical equilibrium and lack of recrystalliza-
rocks from chemical information stored in cooled rocks is complex, tion or exsolution. One of the most commonly used mineral pairs is
and is based on numerous assumptions of equilibrium and estimates magnetite–ilmenite, whose compositions, corrected stoichiometrically
of partitioning relationships. The methods most commonly used to esti- for Fe3+/Fe2+ distribution (Droop, 1987), can be used to estimate fO2
mate fO2 in magmas are reviewed below, together with the various lines from the equilibrium reaction:
of evidence for sub-arc mantle and magma oxidation state. These data
confirm the elevated but variable oxidation state of the sub-arc mantle SiO2 þ 2Fe2 TiO4½Usp ¼ 2FeTiO3½ilmenite þ Fe2 SiO4½Fay ð1Þ
and derived magmas.
where Usp is the ulvospinel component in magnetite, and Fay is the
2.1. Methods of estimating magmatic fO2 fayalite component in olivine, or equivalent in pyroxenes (Andersen
et al., 1993; Lepage, 2003). The great advantage of this method is that
2.1.1. Whole-rock Fe3+/Fe2+ ratios it uses cheap and quick analytical techniques with microbeam resolu-
Wet-chemical analysis of FeO concentrations in whole-rock samples tion. However, the relatively low precision and accuracy of electron mi-
can be used to calculate actual Fe2O3 concentrations from Fe2O3(total) croprobe analyses, combined with assumptions of chemical equilibrium
data commonly generated by X-ray fluorescence or inductively coupled and stoichiometry, mean that this method is likely only accurate to
plasma analyses. Whole-rock Fe3+/Fe2+ ratios can then be used to esti- ± 0.15 log fO2 units at best (Andersen et al., 1993; Sauerzapf et al.,
mate the oxidation state of rock samples (Blevin, 2004; Carmichael and 2008). Moreover, coexisting magnetite and ilmenite are not present in
Ghiorso, 1986). A problem with this method, however, is that Fe2O3/FeO all igneous rocks, and even where present they are commonly not clear-
ratios in igneous rocks are highly susceptible to weathering, alteration, ly in textural equilibrium (e.g., magnetite microphenocrysts, but ilmen-
and redox changes due to near-solidus degassing. Glasses tend to yield ite crystallites in the matrix), or have experienced exsolution or
the best estimates of original magmatic fO2, but even otherwise fresh oxidation reactions during cooling (e.g., Saito et al., 2004).
crystalline rocks may yield results 1–2 log units higher than comagmatic Amphibole compositions have also been used as an oxybarometer,
glasses due to processes such as degassing (Christie et al., 1986). Alter- based on an empirical relationship between magmatic fO2 and a Mg
native approaches to measuring whole-rock Fe3 +/Fe2 + ratios are index (Ridolfi et al., 2010). This method has the advantage that amphi-
Mössbauer spectroscopy (Mysen et al., 1985; Wood and Virgo, 1989), bole is commonly present in arc plutonic and volcanic rocks, but its de-
and micro X-ray absorption near-edge structure (μ-XANES; Cottrell pendence on multiple element electron microprobe analyses, and the
et al., 2009); the latter method permits direct microbeam-scale mea- commonly complex history of amphiboles in igneous rocks leads to a
surement on glasses, thereby avoiding alteration, crystals, etc. relatively high degree of uncertainty in the interpretation of the results.

2.1.2. Whole-rock chemical ratios 2.2. Oxidation state of sub-arc mantle xenoliths and orogenic peridotites
In addition to Fe, V and Cr occur in igneous rocks and magmas in
multiple valence states, with different partitioning behaviors between Direct estimates of oxidation state in the sub-arc mantle have been
melts and minerals (Canil, 1999). When normalized to a chemically obtained from ultramafic xenoliths in primitive arc magmas from a
similar, but non redox-dependent element such as Sc or Yb, empirical number of locations around the world, and from orogenic peridotites
indications of magmatic oxidation state can be inferred (Laubier et al., thought to represent obducted fragments of sub-arc mantle wedge. Rel-
2014; Lee et al., 2005; Li and Lee, 2004). However, it is unclear how ac- atively high oxidation states are indicated in the majority of cases (Arai
curately these trace element proxies reflect actual magmatic oxidation and Ishimaru, 2008; Frost and McCammon, 2008). Examples are
30 J.P. Richards / Lithos 233 (2015) 27–45

summarized in Fig. 1, and include: spinel lherzolite xenoliths from analyses of arc volcanic and plutonic rocks that include analytical data
Ichinomegata, Japan, with ΔFMQ = 0 to + 1.2 (Wood and Virgo, for both FeO and Fe2O3 have been selected from the GEOROC database
1989); spinel peridotite xenoliths from Simcoe volcano, Washington, (http://georoc.mpch-mainz.gwdg.de/georoc/), filtered for alteration to
USA, with ΔFMQ = + 0.3 to + 1.4 (Brandon and Draper, 1996); exclude any samples where alteration is noted, or that have LOI
hornblende-bearing spinel lherzolite and websterite xenoliths in andes- ≥1.0 wt.% (see caption to Fig. 2 for details on selection procedures; de-
ites from central Mexico with ΔFMQ = + 1.5 to + 2.4 (Blatter and spite these precautions, it is acknowledged that some altered samples
Carmichael, 1998a); spinel peridotite xenoliths from several arcs with are likely still present in the datasets). The arc samples are divided
ΔFMQ = + 0.3 to + 2.5 (Parkinson and Arculus, 1999; Parkinson into continental plus mature island arc (hereafter referred to as “mature
et al., 2003); metasomatized peridotite xenoliths from Lihir Island, arc”) and island arc suites. They are compared to a similarly selected
Papua New Guinea, with ΔFMQ = + 1.8 to + 2.0 (McInnes et al., suite of 1147 non-arc tholeiitic (MORB) volcanic and plutonic rocks,
2001); peridotite xenoliths from Shiveluch volcano, Kamchatka, with presumed to be derived from normal depleted upper mantle.
ΔFMQ = +1.4 to +2.6 (Bryant et al., 2007); and orogenic peridotites The median log(Fe2O3/FeO) value is highest in the mature arc suite
from the Sulu belt of eastern China and the Italian Alps with ΔFMQ = (− 0.22 ± 0.39, n = 7672), lower in the island arc suite (− 0.27 ±
0 to +2 (Malaspina et al., 2009). However, there are also a few exam- 0.33, n = 2432), and lowest in the tholeiitic suite (− 0.69 ± 0.30,
ples of lower oxidation states, such as spinel harzburgite and websterite n = 1147) (Fig. 2). When plotted against FeOTotal, the arc rocks predom-
xenoliths from Patagonia, with ΔFMQ = –1.1 to –0.2 (Wang et al., 2007; inantly plot in the “moderately oxidized” to “strongly oxidized” fields of
see also Malaspina et al., 2010). Blevin (2004), whereas the tholeiites are centered on the “moderately
These data suggest that the asthenospheric mantle wedge is com- reduced” field (Fig. 3). Blevin (2004) suggests that the boundary be-
monly up to 3 log units more oxidized than typical MORB-depleted tween the “moderately reduced” and “moderately oxidized” fields ap-
mantle (ΔFMQ = − 1 to 0), but that more reducing conditions may proximates the FMQ buffer curve, which is consistent with the
occur locally. Wang et al. (2007), Song et al. (2009) and Malaspina majority of tholeiites being at or below FMQ, and arc magmas lying
et al. (2010) have suggested that subduction of reduced, organic above this buffer.
carbon-bearing sediment may be responsible for these lower values, Mathez (1984), Mallmann and O'Neill (2009) and Lee et al. (2010)
but the data suggest that this is not a widespread condition in arcs. In- have suggested that whole-rock Fe3+/Fe2+ ratios may be affected by
stead, as reviewed below, subduction (at least since deep ocean oxida- late stage differentiation or degassing processes, but the clear relative
tion following the Neoproterozoic Oxygenation Event: Evans and difference between these two major magma suites, which is largely in-
Tomkins, 2011; Prouteau and Scaillet, 2013; Richards and Mumin, dependent of degree of differentiation (the weak inverse correlation
2013a; Shields-Zhou and Och, 2011; or earlier: Jagoutz, 2013) appears with [FeOTotal] is accounted for in the slope of the field boundary lines
to be associated with variable degrees of oxidation of the overlying in Fig. 3) indicates that this is a source feature and not a process over-
mantle. print (Carmichael, 1991; Crabtree and Lange, 2012; De Hoog et al.,
2004; Kelley and Cottrell, 2012). Kelley and Cottrell (2009) further
2.3. Oxidation state of arc magmas showed that Fe3+/ΣFe ratios in arc magmas increase towards the trench
and correlate with concentrations of H2O and fluid-mobile elements
As noted above, arc magmas have long been recognized as being (e.g., Ba/La), indicating that the oxidation effect is a function of the
more oxidized than typical tholeiitic magmas, typically in the range
ΔFMQ = +0.5 to +2, and locally up to or above +3 (Ballhaus, 1993;
Eggins, 1993; Luhr, 1992; Mullen and McCallum, 2013; Pichavant and
3500
Macdonald, 2007; Stamper et al., 2014a). Relatively few suites of geo-
Mature arc Median values
chemical data for arc rocks include direct estimations of oxidation Island arc
3000 Mature arc
state, but where separate chemical analyses of FeO have been made, Tholeiite

the Fe2O3/FeO ratio can be used as a proxy. A compilation of 10,104 Island arc
2500
Tholeiites
Frequency

2000

1500

1000

500

0
-2.0 -1.5 -1.0 -0.50 0.0 0.50 1.0 1.5 2.0
log10 (Fe2O3/FeO)

Fig. 2. Histogram of log(Fe2O3/FeO) values for mature arc, island arc, and non-arc tholeiitic
volcanic and plutonic rocks from the GEOROC database (http://georoc.mpch-mainz.gwdg.
de/georoc/). All available files from the pre-compiled “Convergent margins” dataset were
used for the arc group, and the “Basalt tholeiitic”, “Tholeiite 1”, and “Tholeiite 2” files from
the pre-compiled “Rocks” dataset were used for the tholeiite group. In the latter case,
analyses reported to be from convergent margins and Archean settings were excluded
([Tectonic setting] column heading in GEOROC database files). The datasets were trimmed
Fig. 1. Histogram of ΔFMQ values for mantle peridotite xenoliths in arc magmas. Data from by deleting analyses: (1) that were not of volcanic or plutonic rocks, identified in [Rock
Blatter and Carmichael (1998a; Central Mexico), Parkinson and Arculus (1999; various Type] column; (2) that were recorded as being altered to any degree, identified in
locations), Parkinson et al. (2003; Grenada), Bryant et al. (2007; Shiveluch volcano, [Alteration] column; (3) that did not have independently measured FeO, identified in
Kamchatka), Wang et al. (2007; Cerro del Fraile, Patagonia) and Malaspina et al. (2010; [FeO(wt%)] column; and (4) that had total volatile contents ≤0.0 or ≥1.0 wt.% (sum of col-
Bardane, Norway). umns [H2O(wt%) to [LOI(wt%)]).
J.P. Richards / Lithos 233 (2015) 27–45 31

3
wedge transfer process, the chief contenders are hydrothermal fluids
Mature arcs
Island arcs (Bali et al., 2011; Peacock, 1993; Perfit et al., 1980; Ringwood, 1977;
Tholeiites Tatsumi, 1989), partial melts of hydrated basalt (Defant and
2
Drummond, 1990; Kay, 1978; Rapp and Watson, 1995), partial melts
of sediments (Behn et al., 2011; Hermann and Spandler, 2008;
log10 (Fe2O3 / FeO)

1 Johnson and Plank, 1999; Labanieh et al., 2012), supercritical fusions


of these components (Bureau and Keppler, 1999; Kawamoto, 2006;
Very strongly oxi
dized Kessel et al., 2005; Portnyagin et al., 2007), or combinations of all of
0 the above at different depths in subduction zones. Mungall (2002) has
Strongly oxidized
argued that slab melts would carry more iron, and in particular Fe3+,
Moderately oxidiz
ed than hydrothermal fluids, and would therefore potentially be stronger
-1
Moderately red
uced oxidants than aqueous fluids. However, petrogenetic and geochemical
considerations suggest that slab melting only occurs in exceptional
Strongly reduce
d
-2 circumstances in Phanerozoic arcs, such as the subduction of
young, hot oceanic lithosphere, or at slab windows or tears where hot
asthenosphere directly contacts subducted oceanic crust (Davies and
-3 Stevenson, 1992; Defant and Drummond, 1990; Peacock et al., 1994;
0 5 10 15 20
Yogodzinski et al., 2001). Although some have argued that slab melting
FeOTotal (wt.%) might be uniquely linked to arc metallogeny because of this strong oxi-
dative effect (e.g., Mungall, 2002; Oyarzun et al., 2001; Sajona and
Fig. 3. Log(Fe2O3/FeO) versus FeOTotal for mature arc (red), island arc (blue), and non-arc Maury, 1998; Sun et al., 2011), others have questioned the evidence
tholeiitic (black) volcanic and plutonic rocks (sourced and selected as in Fig. 2). Relative
oxidation state boundaries are from Blevin (2004), who suggests that the thick black
for this relationship (Rabbia et al., 2002; Richards, 2002, 2012;
line between the “moderately reduced” and “moderately oxidized” fields approximates Richards and Kerrich, 2007).
the FMQ buffer curve. Regardless of the exact process, it is clear that both fluids and melts
are capable of transferring the oxidized signature of seafloor alteration
degree of subduction metasomatism (see also: Brounce et al., 2014; and sediments into the mantle wedge source of arc magmas, and that
Carmichael et al., 2006; Rowe et al., 2009; Zimmer et al., 2010). the effect does not need to be extreme to raise the fO2 of these magmas
Evidence from other redox-sensitive elements such as vanadium 1 or 2 log units above normal MORB-depleted mantle values to the
have yielded conflicting results. Lee et al. (2005) interpreted relatively range typical or arc magmas (Evans, 2012). Likely, this effect is also to
uniform V/Sc ratios in both MORB and arc basalts to indicate fO2 values some extent cumulative, resulting in a progressive increase in oxidation
of ΔFMQ = −1.25 to +0.5, and argued that the higher fO2 values calcu- state of the supra-subduction zone mantle wedge from typical mantle
lated for many arc volcanic rocks by other methods might reflect differ- values (ΔFMQ = − 1 to 0) at the onset of subduction, to ΔFMQ ≥ + 1
entiation processes rather than source characteristics. Similar after a few million years of static arc development (i.e., arc magmatism
arguments have been made by Mallmann and O'Neill (2009) and by focused along a fixed axis in the upper plate, implying steady-state sub-
Lee et al. (2010) on the basis of Zn/Fe ratios. On the other hand, duction conditions; Evans, 2012). This may explain why calculated oxi-
Laubier et al. (2014) used V/Yb ratios to argue for increasing magmatic dation states of primitive magmas from nascent oceanic arcs are
oxidation state from MORB (ΔFMQ ≈ 0), through back-arc basins, to commonly indistinguishable from MORB (e.g., Lee et al., 2005), but in-
arcs (~ ΔFMQ ≥ + 2). The reliability of these trace element ratios as crease with arc evolution (Evans et al., 2012).
proxies for magmatic oxidation states is thus unclear at the present Progressive mantle wedge oxidation may also be reflected in arc
time. metallogeny. Few large ore deposits occur in primitive island arcs, but
develop commonly after several million years of arc magmatism, sug-
2.4. Redox effects of subduction gesting that the arc magmatic system requires time to evolve to a
more “fertile” state (Richards, 2003). These processes are discussed fur-
The source of the oxidizing signature in arc magmas is thought to be ther below.
subducting oceanic crust. Oxidation of the deep oceans since the
Neoproterozoic Oxygenation Event (Canfield et al., 2006) has led to ox- 2.5. Partial melting in the mantle wedge and the calc-alkaline trend
idative alteration of oceanic crust at mid-ocean ridges, which generates
secondary hematite and anhydrite in altered basalts (Alt et al., 2013; The primary control on partial melting in the asthenospheric mantle
Bischoff and Dickson, 1975; Shanks et al., 1981; Staudigel et al., 1996), wedge is the influx of water from the dehydrating subducting slab,
and the deposition of relatively oxidized deep-ocean sediments (Chen which lowers the solidus temperature of peridotite (Arculus, 1994;
et al., 1996; Piper et al., 1984; Wilson et al., 1985; with the exceptions Grove et al., 2006, 2012; Kushiro et al., 1968; Ringwood, 1977; Stolper
of reduced sediment deposition as noted by Malaspina et al., 2010 and and Newman, 1994; Ulmer, 2001). The resulting primary magmas are
Wang et al., 2007). The main carriers for oxygen are Fe3+, Mn4+, S6+, hydrous, tholeiitic, high-Mg basalts (Greene et al., 2006; Pichavant
C4+, and to a lesser extent H2O, in the form of dissolved Fe and Mn ox- et al., 2002; Smith et al., 2010; Thirlwall et al., 1996), which evolve
ides, sulfates, and carbonates (Andreani et al., 2013; Brandon and to calc-alkaline compositions by early fractionation of olivine,
Draper, 1996; Evans, 2012; Evans and Tomkins, 2011; Kelley and clinopyroxene, and magnetite, and later crystallization of hornblende
Cottrell, 2009). Variability in the extent of oxidative alteration and the and plagioclase (Grove and Kinzler, 1986; Grove et al., 2003; Moore
amount of oxidized (or reduced) sediment may, therefore, be the pri- and Carmichael, 1998). Hamada and Fujii (2008) and Zimmer et al.
mary cause of the wide range of oxidation states found in arc magmas, (2010) suggest that primary magmatic water contents of 2 wt.% sepa-
from essentially unoxidized (MORB-like) to highly oxidized composi- rate the “dry” tholeiitic olivine-plagioclase-orthopyroxene fractionation
tions (Fig. 3). trend from the “wet” calc-alkaline trend, which is consistent with esti-
Subduction of this seafloor-altered material and subsequent remobi- mates of water contents in primary and primitive arc magmas of be-
lization during prograde high-pressure metamorphism introduces tween 1 and 8 wt.% H2O (typically 2–6 wt.%; Carmichael, 2002;
these redox components into the metasomatized mantle wedge (Alt Cervantes and Wallace, 2003; Edmonds et al., 2014; Grove et al., 2003;
et al., 2012, 2013; Andreani et al., 2013). Although there is considerable Kimura and Ariskin, 2014; Pichavant et al., 2002; Plank et al., 2013;
debate in the literature about the exact mechanism of the slab-to- Ruscitto et al., 2012; Sobolev and Chaussidon, 1996; Wallace, 2005).
32 J.P. Richards / Lithos 233 (2015) 27–45

Zimmer et al. (2010) also stress the additional effect of elevated fO2 in a 500
establishing the low-Fe calc-alkaline trend by early crystallization of Mature arc data plotted as a
Fe-rich spinels. function of log(Fe2O3/FeO)
These hydrous compositions predispose arc magmas to developing 400
<-0.5
“adakite-like” trace element signatures of Sr enrichment and heavy -0.5 to 0
rare earth element (HREE) depletion, through early and abundant crys- >0
Median (-0.5 to 0)

Cu (ppm)
tallization of pyroxene and hornblende (which preferentially partition 300
middle (MREE) and HREE), and late crystallization of plagioclase
(which would otherwise deplete the melt in Sr; Naney, 1983; Richards
and Kerrich, 2007; Ridolfi et al., 2010; Rutherford and Devine, 1988; 200
Stamper et al., 2014b).
In addition to elevated water contents and oxidation states (in many
100
cases, as noted above), primary arc magmas are also enriched in a suite
of volatile and fluid-mobile elements, including Cl (500–2000 ppm Cl;
Wallace, 2005), S (900–2500 ppm S; de Hoog et al., 2001; Wallace,
0
2005), large-ion lithophile elements (LILE: Rb, K, Cs, Sr, Ba, and U), 0 5 10 15 20
and Li, B, Pb, As, and Sb (Gill, 1981; Perfit et al., 1980; Wysoczanski MgO (wt.%)
et al., 2012). These magmas are also characteristically depleted (relative
to adjacent elements on primitive mantle- or MORB-normalized trace- b 500
element diagrams) in Ti, Ta, and Nb due to their low mobility in slab Island arc data plotted as a
fluids and/or retention in restite phases (Foley et al., 2000; Klemme function of log(Fe2O3/FeO)
et al., 2005; Ryerson and Watson, 1987; Schmidt et al., 2004, 2009). 400
<-0.5
Whether or not they are also enriched in metals is discussed below. -0.5 to 0
>0
Median (-0.5 to 0)

Cu (ppm)
2.6. Redox effects on metallogeny during partial melting and magmatic 300
differentiation

200
A perennial debate in economic geology revolves around whether
ore-generating magmas are unusually metal-rich at source, or whether
ore formation results from magmatic and/or hydrothermal processes
100
operating on otherwise “normal” magmas. Examples of such debates in-
clude the abundance of Au in lamprophyres (Rock and Groves, 1988;
Tilling et al, 1973; Wyman and Kerrich, 1989), and Cu and Au in arc
0
magmas (Cline and Bodnar, 1991; Core et al., 2006; Keith et al., 1997; 0 5 10 15 20
Richards, 2013; Timm et al., 2012; Wilkinson, 2013). To date, no con- MgO (wt.%)
vincing evidence has been found for the necessity of unusually metal-
rich magmas in most ore-forming environments, but it is also obvious c 500
Tholeiite data plotted as a
that unusually metal-poor magmas would not be optimal for ore forma- function of log(Fe2O3/FeO)
tion. Thus, processes that could deplete magmas in ore metals could sig-
400
nificantly reduce their fertility, and vice versa.
<-0.5
Several recent studies have compared the Cu contents of arc and -0.5 to 0
>0
mid-ocean-ridge basalts (MORB), and have generally concluded that
Cu (ppm)

300
there is little difference in the compositions of primitive magmas from
these distinct tectonic settings (Jenner et al., 2010; Lee et al., 2012), al-
though Timm et al. (2012) found that some mafic rocks from the 200
Kermadec arc extend to higher Cu contents than MORB. Using the
same dataset as for Figs. 2 and 3, Cu contents of least-altered mature
arc, island arc, and tholeiitic (MORB) igneous rocks are plotted against 100
MgO in Fig. 4. The data are subdivided by log(Fe2O3/FeO) range
(b −0.5, − 0.5 to 0, and N0, roughly corresponding to the moderately
reduced, moderately oxidized, and strongly oxidized ranges of Blevin, 0
0 5 10 15 20
2004) based on the distribution shown in Figs. 2 and 3. The distribution
of data in Fig. 4 is virtually indistinguishable for the three tectonic set- MgO (wt.%)
tings, and also appears to be independent of oxidation state, as shown
Fig. 4. Plots of Cu versus MgO content in (a) mature arc, (b) island arc, and (c) non-arc tho-
by the similar distribution of data for rocks with low (b−0.5), medium leiitic igneous rocks, subdivided by log(Fe2O3/FeO) b −0.5, −0.5 to 0, and N0 (roughly
(− 0.5 to 0), and high (N 0) log(Fe2O3/FeO) values. Primitive rocks corresponding to the moderately reduced, moderately oxidized, and strongly oxidized
(N10 wt.% MgO) from all settings have moderate Cu contents of ranges of Blevin, 2004; sources of data as in Fig. 2). The distribution of data are broadly
~ 75–100 ppm Cu: 78.0 ± 63.6 ppm Cu (n = 53) in mature arcs; similar in all three petrogenetic cases, and for all oxidation states. Median Cu contents
(binned over 1 wt.% MgO ranges up to 10 wt.% MgO, and then between 10–12.5 wt.%,
74.7 ± 33.9 ppm Cu (n = 41) in island arcs; and 97.2 ± 26.7 ppm Cu
and N12.5 wt.% MgO) are shown for the moderately oxidized arc suites in (a) and (b),
(n = 52) in primitive tholeiites. More evolved rocks (10–5 wt.% MgO) with a polynomial line fit: (a) y = − 3.50 + 16.68x − 1.42x2 + 0.04x3 (R = 0.93);
expand to lower and higher Cu concentrations (0 to N200 ppm Cu), be- (b) y = −21.70 + 39.39x − 4.14x2 + 0.13x3 (R = 0.97).
fore the higher values decrease again towards 0 ppm Cu at lower MgO
contents. (The tholeiite data are mainly for basaltic rocks, with few
data for more differentiated compositions.) Overall, the bulk of the calculated for moderately oxidized samples (log[Fe2O3/FeO] = −0.5 to
data from mature arcs show a steady decrease in Cu values from mafic 0) shown in Fig. 4a. Island arc data show a slight increase in median Cu
to felsic compositions, as illustrated by the distribution of median values contents from primitive to intermediate compositions (up from
J.P. Richards / Lithos 233 (2015) 27–45 33

~75 ppm Cu in primitive rocks to ~95 ppm Cu at ~7 wt.% MgO, although mean = 94.0 ± 58.9 ppm, median = 85.0 ppm, n = 372; tholeiite:
relatively few data constrain the primitive range; Fig. 4b), before de- mean = 117.5 ± 64.2 ppm, median = 110.0 ppm, n = 115).
creasing again towards 0 ppm Cu in felsic rocks. The distribution of data shown in Figs. 4 and 5 suggests that Cu con-
There is a broader scatter to higher Cu contents in tholeiitic and is- tents in mafic to intermediate magmas from both arc and ridge tectonic
land arc rocks in the range 3–7 wt.% MgO (which corresponds to the settings are similar and generally quite low (b200 ppm, similar to
peak in Cu concentration) compared with mature arcs (Fig. 5; mature values obtained by Lee et al., 2012, in mantle melting models). This sug-
arc: mean = 59.9 ± 42.6, median = 46.9 ppm, n = 1247; island arc: gests that magmatic Cu concentrations are moderated by the presence
of minor amounts of sulfides in the source, into which chalcophile and
siderophile metals (such as Cu, Au, and the platinum group elements;
a 400
Cu in mature arc rocks
PGE) will preferentially partition (Keays, 1987, 1995; Lee et al., 2012),
even for moderately to strongly oxidized rocks.
350 In somewhat more evolved rocks (3–7 wt.% MgO), the data scatter to
Mean = 59.9 ± 42.6 ppm Cu (n = 1247)
Median = 46.9 ppm Cu higher Cu contents in both arc and ridge settings, but the bulk of the
Skewness = 2.02
300 data remains below ~ 150 ppm Cu. This distribution is illustrated by
the skewness of the histograms in Fig. 5 (up to 2.02 in mature arc
250
Frequency

rocks; Fig. 5a). For mature and island arc suites, the overriding trend
200
of the data seems to be towards decreasing Cu contents with differenti-
ation (decreasing MgO content), and the scatter to higher values ap-
150 pears more random. It also does not correlate with Fe3+/Fe2+ ratios in
the rocks, and both relatively oxidized and relatively reduced rocks
100 can have high Cu contents, although this is not common in either case.
Copper therefore appears to behave mildly compatibly for the majority
50 of samples analyzed from arcs and ridges, indicating the continual pres-
ence of condensed sulfide phases that hold the Cu content of the melt to
0
0 100 200 300 400 500 relatively low and decreasing concentrations. The scatter to higher Cu
Cu (ppm) concentrations in some intermediate-composition samples might re-
flect local incompatible behavior (as described by Jenner et al., 2010;
b 60 Lee et al., 2012), or might be a product of sulfide accumulation or
Cu in island arc rocks subsolidus processes (despite efforts to screen for alteration).
50 Assuming that these high-Cu values do reflect magmatic composi-
Mean = 94.0 ± 58.9 ppm Cu (n = 372) tions, the reason for this incompatible behavior is not clear, but could
Median = 85.0 ppm Cu
Skewness = 1.37 be due either to low total sulfur fugacity (unlikely in most arc settings)
40
or unusually high oxidation states, such that sulfide solubility was not
Frequency

exceeded or sulfides were destabilized by oxidation (e.g., Ballhaus


30 et al., 1991; Jenner et al., 2010; Mungall, 2002). However, a clear corre-
lation with high magmatic oxidation state is not observed, and it cannot
therefore be concluded that oxidized magmas are inherently more fer-
20
tile (with respect to Cu content) than more reduced magmas. A similar
conclusion has been drawn for gold in arc magmas (Jégo and Pichavant,
10 2012). The data also do not support the notion that arc magmas in gen-
eral are fundamentally more fertile (Cu-rich) than other mantle-derived
0 magmas such as MORB tholeiites; if anything the opposite is true, which
0 100 200 300 400 500 must mean that other processes are involved in the relationship
Cu (ppm) between arc magmas and porphyry Cu deposits. On the other hand, it
is also clear that most arc magmas (whether relatively oxidized or
c 25 reduced) retain ~50 ppm Cu on average into the andesitic range of com-
Cu in MORB tholeiites positions, which is also the value used by Cline and Bodnar (1991) to
show that normal arc magmas are capable of generating large porphyry
20
Mean = 117.5 ± 64.2 ppm Cu (n = 115)
Cu deposits.
Median = 110.0 ppm Cu Contrary to the findings of Jenner et al. (2010), the data do not sup-
Skewness = 1.19
port a general and sudden switch from incompatible to compatible be-
Frequency

15
havior of Cu at intermediate compositions (suggested to be caused by
sulfide-saturation triggered by magnetite crystallization). While this
10
switch may occur in some cases, the broad arrays of data shown in
Fig. 4 are more consistent with mildly compatible behavior of Cu
throughout the differentiation range for the bulk of magmas, with in-
5 compatible behavior (as indicated by higher Cu contents) being the ex-
ception. It might be argued that it is these exceptions that give rise to ore
deposits, but the effect is not particularly large – rarely more than 2–3
0 times average Cu contents – and thus at most might be expected to af-
0 100 200 300 400 500 fect the size but not the existence of a deposit, and it occurs over a rela-
Cu (ppm) tively narrow range of intermediate magma compositions (3–7 wt.%
MgO). This narrow peak of Cu concentrations would not be expected
Fig. 5. Histograms of Cu concentrations in rocks between 3–7 wt.% MgO, which span the
peak of Cu contents, for (a) mature arcs, (b) island arcs, and (c) and non-arc tholeiites
to significantly enhance the net flux of Cu from the mantle source region
(MORB) (sources of data as in Fig. 2). MORB and island arc suites show higher mean because Cu contents of primitive magmas are uniformly moderate.
and median Cu contents than mature arcs. Moreover, these anomalously Cu-rich magmas (if these compositions
34 J.P. Richards / Lithos 233 (2015) 27–45

indeed reflect magmatic values) do not clearly correlate with known 1999; Park et al., 2013; Righter et al., 2008; Widom et al., 2003). These
porphyry Cu deposits. differences are explored further below.
In summary, Cu contents of arc magmas globally show a general Where condensed sulfide phases coexist with silicate magma, the
weak trend of decreasing concentration with differentiation, with a ratio of the volume of sulfide to silicate melt (inverse of the R-factor of
scatter to higher concentrations in some intermediate composition Campbell and Naldrett, 1979, where R = [mass of silicate melt]/[mass
rocks. These trends appear to be independent of oxidation state, and of sulfide]) can affect the ratios of metals left in the magma, as well as
there is no clear correlation between rocks with higher Cu concentra- their abundance. As discussed by Richards (2009), the presence of mod-
tions and porphyry ore deposits. Intermediate composition arc magmas erate amounts of residual sulfide relative to a large volume of silicate
reaching the upper crust typically contain 50–100 ppm Cu (Fig. 5), melt (e.g., ~ 0.001–1 wt.% sulfide; intermediate R-factor of 102–105;
which has been shown to be sufficient for the generation of large por- Fig. 6) will deplete the magma in low-abundance, high partition coeffi-
phyry deposits from reasonable volumes of magma (Cline and Bodnar, cient elements such as Au and PGE (highly siderophile elements) but
1991). may have little effect on the concentration of more abundant elements
with lower partition coefficient such as Cu, leading to high Cu/Au, Cu/
2.7. The role of sulfur on metal solubility during partial melting PGE magmas (e.g., Park et al., 2013; Fig. 6). Such magmas should be con-
sidered Cu-undepleted but Au-PGE-depleted, rather than Cu-rich, and
In combination with oxygen fugacity, the sulfur fugacity of magmat- may be typical of normal arc environments where several hundred to
ic systems plays a critical role in the behavior of chalcophile a few thousand ppm S may be present (de Hoog et al., 2001; Jugo,
and siderophile metals. For sulfur, an important transition occurs be- 2009; Wallace, 2005). These magmas have the potential to form Cu-
tween dissolved sulfide and sulfate dominance in silicate melts at rich porphyry-type deposits (Richards, 2009, 2011a).
ΔFMQ ≈ +1 (Carroll and Rutherford, 1985; Jugo et al., 2010; Nilsson If residual sulfide abundances are high relative to the volume of
and Peach, 1993), with high concentrations of S (dissolved as SO24 −) magma (R b 102), all chalcophile and siderophile elements will be de-
possible in more oxidized magmas (e.g., up to 1.5 wt.% S; Beermann pleted in the magma (Fig. 6; Keays, 1987). However, such high
et al., 2011; Jugo et al., 2005a). This means that oxidized magmas dom- sulfide/magma ratios are rarely achieved in nature (except during the
inated by sulfate (ΔFMQ ≥ +1) will be less likely to become saturated in formation of some orthomagmatic sulfide deposits by bulk sulfide con-
sulfide phases. But sulfide saturation is also a function of the overall sul- tamination of S-undersaturated mafic melts; e.g., Naldrett, 1989; Keays,
fur content of the system. Thus, sulfide minerals or melts may be stable 1995). In contrast, under high R-factor conditions (R N 105), where van-
even in quite oxidized melts if fS2 conditions are high (Jugo, 2009), ishingly small amounts of (or no) residual sulfides are present in the
which is a characteristic feature of arc systems. This may explain why mantle source and during differentiation, derivative magmas will be
high Cu concentrations, reflecting the “burn-out” of condensed sulfide undepleted even in highly siderophile elements, and can thus potential-
phases in highly oxidized mantle sources and magmas (as predicted ly lead to the formation of Au- or PGE-rich deposits (Fig. 6; Hamlyn
by Jugo et al., 2005a; Mungall, 2002; Parkinson and Arculus, 1999; et al., 1985; Richards, 2009). Such conditions appear to be relatively
Righter et al., 2008; Timm et al., 2012) is not clearly observed in the rare in normal arc settings, but may arise in alkaline back-arc magmas
datasets shown in Fig. 4. Instead, the data show that relatively reduced (Métrich and Clocchiatti, 1996; Richards et al., 1991), and/or during
magmas are just as likely to, but rarely do, contain anomalously high Cu second-stage S-undersaturated melting of subduction modified mantle,
contents as more oxidized magmas. Sparsity of data for elements such where a flux of fresh S from the subducting slab is no longer present
as Au and PGE means that it is not clear whether these metals behave (Dale et al., 2012; Richards, 1995, 2009; Wyborn and Sun, 1994).
in the same way (see discussion below). Finally, the nature of the restite or fractionating sulfide or alloy phase
The distribution of the bulk of the data suggests that, under most also has a significant effect on metal partitioning. In particular, the pres-
conditions, the mantle sources of arc and ridge magmas are sulfide- ence of small amounts of Pt-Fe alloy as restite in the mantle source re-
saturated (Keays, 1995), and that most of these magmas remain satu- gion may have a profound effect on PGE abundances and ratios in
rated or close to saturation throughout their differentiation history. partial melts (Kepezhinskas et al., 2002; Mungall and Brenan, 2014;
The net effect of sulfide saturation on metal concentration in the Park et al., 2013). In addition, Righter et al. (2004) have shown that,
coexisting melt depends on several factors, including: (1) the metal in
question; (2) the volume of the sulfide phase relative to the silicate Cu-undepleted Au-undepleted post-
arc magmas subduction magmas
melt; and (3) the nature of the condensed sulfide phase (sulfide melt, 106
d in
pyrrhotite, monosulfide solid solution, intermediate solid solution). An che ide
105 e nri l sulf
additional factor may be the presence of platinoid metal alloys in the Au idua
res
mantle source, which can dramatically fractionate the PGE (Mungall 104
de
and Brenan, 2014; Park et al., 2013). 103 ulfi
in s Cu maximized in
The metals of greatest economic interest in subduction zone Cu e magma (R ≥ 103)
102 u lfid
magmas are Cu, Au, and the PGE. Although these metals can all be char- ins
Au
Cu (ppm)
Au (ppb)

acterized as chalcophile and siderophile in character, Au and the PGE are 10 ma


ag
considered to be highly siderophile, and partition strongly into most i nm ma
1 Cu ag Au maximized in
sulfide phases relative to silicate melts (Campbell and Naldrett, 1979; inm magma (R ≥ 106)
Laurenz et al., 2013; Peach et al., 1990), while the PGE also partition dif- 0.1 ted
ple
de Sulfide/silicate melt partition
ferentially into metal alloy phases (Kepezhinskas et al., 2002; Mungall 10–2 Au coefficients:
D(Cu) = 103
and Brenan, 2014). Consequently, the chemical behavior of these ele- a D(Au) = 105
10–3 gm
ments in the presence of sulfides and alloys can be expected to be sub- n ma Metal concentrations in
Aui magma in absence of sulfide:
stantially different, and this may explain the widely variable Cu/Au and 10–4 Cu = 50 ppm
Cu/PGE ratios in resultant ore deposits. Unfortunately, because of the Au = 5 ppb
10–5
difficulty and expense of obtaining accurate trace-level analyses of Au
1 10 102 103 104 105 106 107 108
and PGE in igneous rocks, there is very little information on their abun-
R = (mass of silicate melt)/(mass of sulfide)
dance in arc rocks (certainly compared to the wealth of data for Cu).
Nevertheless, the differential behavior of these elements has been doc- Fig. 6. Variations of Cu and Au contents in silicate melts and coexisting sulfides as a func-
umented by several studies (e.g., Hamlyn et al., 1985; Jégo and tion of the volume ratio of these two phases (the R-factor of Campbell and Naldrett, 1979);
Pichavant, 2012; Jégo et al., 2010; Li and Audétat, 2012; McInnes et al., modified from Richards (2009).
J.P. Richards / Lithos 233 (2015) 27–45 35

while Re, Au and Pd are incompatible in Cr-spinel, Rh, Ru and Ir are volumes of magma crystallize as plutonic rocks, with the potential for
strongly compatible in this mineral, potentially leading to significant subsurface volatile release and hydrothermal ore formation. The time
fractionation between these elements. Copper, Au, and PGE also have required for this transition can be expected to vary depending on the
different partition coefficients between various sulfide phases (liquids subduction rate, angle of subduction, and other factors, but likely takes
and minerals) and silicate melt. In general, partition coefficients seem on the order of 10 m.y. (Deering et al., 2012; Evans, 2012; Ishizuka
to be highest for sulfide liquids (Laurenz et al., 2013; Li and Audétat, et al., 2011; Reagan et al., 2008). This timeframe is consistent with the
2012; Mungall and Brenan, 2014), followed by intermediate solid solu- relatively late timing of porphyry systems in island arc settings and in
tion (ISS) for Cu and Au (Jugo et al., 1999; Kesler et al., 2002), continental arc magmatic cycles (Kesler et al., 1977; Mitchell and Bell,
monosulfide solid solution (MSS; Li and Audétat, 2012), and pyrrhotite 1973; Richards, 2003, 2005; Rohrlach and Loucks, 2005).
(Jugo et al., 1999; Zajacz et al., 2013). Under mantle conditions, it would In addition to promoting plutonism, the thickening of arc crust has
seem likely that some sulfide liquid is present, and that it will have a multiple chemical effects on arc magmatism, including extending the
major control on the abundances and ratios of chalcophile and range of magmatic differentiation and interaction with older crustal ma-
siderophile elements in arc magmas, particularly if these sulfides are terials. These processes have been collectively termed MASH, standing
preferentially left in the restite rather than being entrained in the for crustal melting, assimilation, storage, and homogenization of arc
magma, due to their density and wetting properties (as suggested by magmas (Hildreth and Moorbath, 1988). Their effects on metallogenic
Mungall and Su, 2005). potential depend largely on whether they cause substantial loss of
metals from the magma to crystallizing phases, prior to ascent into the
2.8. The source of sulfur in arc magmas upper crust and partitioning of those metals into an exsolving
magmatic-hydrothermal fluid (Spooner, 1993). The same principles
In the preceding discussion, the high abundance of sulfur in arc mag- that were used to consider the loss of metals to restite in the mantle
matic systems has been noted. But because of its critical importance in wedge apply to fractionating minerals and melts. Although minor
both the fixing and solvating of metals, it is worth briefly reviewing amounts of Cu can be incorporated into igneous silicate and oxides
the evidence for the source of this sulfur, and how its nature may have phases, partition coefficients are typically less than 1 indicating that
changed over geological time. Cu behaves incompatibly with respect to these minerals (Cline and
Arc magmas are rich in sulfur relative to most other mantle-derived Bodnar, 1991; Core et al., 2005; Lee et al., 2012; Liu et al., 2014); conse-
magmas, and this has been referred to as the “excess sulfur” problem by quently, their fractional crystallization will not significantly affect the
some authors (e.g., Sharma et al., 2004; Wallace, 2005), reflecting the concentration of metals in the melt phase. On the other hand, segrega-
much larger volumes of SO2 emitted from some arc volcanoes than tion of sulfide melts and minerals will strongly partition chalcophile
would be predicted from the solubility of S in co-erupted felsic lavas. and siderophile elements, and if large volumes of sulfide are fractionat-
It is now clear that the excess S comes from magma chamber recharge ed they will strongly deplete the magma in these metals (potentially
by more mafic magmas with higher S solubility (De Hoog et al., 2004; forming orthomagmatic sulfide deposits within mafic cumulate rocks;
Hattori, 1993; Roberge et al., 2009; Van Hoose et al., 2013). Spandler et al., 2000; Fig. 6). Sato (2012) and Tomkins et al. (2012)
Sulfur isotopic compositions in arc magmas are typically higher than have proposed that such conditions might occur where reduced, gra-
depleted mantle values of δ34SVCDT ≈ 0‰ (up to and rarely above 13‰; phitic rocks are present in the lower crust. Assimilation of this material
Alt et al., 1993; Ishihara and Sasaki, 1989), and this has been interpreted by oxidized S-rich arc magmas would cause reduction of SO2− 4 to S2−,
to reflect a contribution from seawater sulfate, fixed in the subducted voluminous sulfide immiscibility, and depletion of metals in the melt,
oceanic crust during earlier seafloor hydrothermal alteration (Alt with subsequent decreased potential for magmatic-hydrothermal ore
et al., 2013; Debret et al., 2014; Evans et al., 2014; Sasaki and Ishihara, formation upon emplacement in the upper crust.
1979; Shanks et al., 1981; Wallace and Edmonds, 2011) or in sulfate- To date, no examples of large sulfide accumulations in lower crustal
bearing sediments (De Hoog et al., 2001). As a consequence of the arc magmatic systems have been reported, with the possible exception
Neoproterozoic Oxygenation Event, sulfate in oxidized deep ocean wa- of Alaskan-type PGE deposits (Batanova et al., 2005; Slansky et al., 1991;
ters reacts with Fe2+ in basaltic oceanic crust to form Fe3+-oxides, sul- Spandler et al., 2000; Thakurta et al., 2008, 2014). Nevertheless, small
fates, and some sulfides. Subduction of this material, along with sulfate- amounts of sulfide minerals are commonly present in lithospheric man-
bearing sediments, introduces a flux of oxidized sulfur (and other com- tle xenoliths (Szabó et al., 2004) and arc cumulate sequences (Lee et al.,
ponents; Debret et al., 2014) into the mantle wedge that easily swamps 2012; Li et al., 2013; Peng et al., 2013; Richards, 2011a; Zhang et al.,
background mantle concentrations, with up to ~ 40% of the subducted 2014), indicating, as noted previously, that arc magmas remain close
sulfur being transferred to the wedge via dehydration fluids (Jégo and to sulfide-saturation throughout much of their evolution. Core et al.
Dasgupta, 2013), and 15–30% of the flux reaching the surface in volcanic (2006), Hou et al. (2009, 2011), Richards (2009), Shafiei et al. (2009)
emissions (Wallace, 2005). This flux of sulfate may be the main vector and Pettke et al. (2010) have suggested that such minor accumulations
for oxidation of the mantle wedge (Alt et al., 2012; Evans and of sulfides might represent easily remobilized sources of metals during
Tomkins, 2011; Jégo and Dasgupta, 2014; Kelley and Cottrell, 2009). later, post-subduction magmatism and ore deposit formation, whereas
Evans and Tomkins (2011), Prouteau and Scaillet (2013) and Chiaradia et al. (2009), Jenner et al. (2010), Lee et al. (2012), Sillitoe
Richards and Mumin (2013a,b) have proposed that, prior to the (2012) and Chiaradia (2014) consider that these sulfides are an essen-
Neoproterozoic Oxidation Event, arc magmas were more reduced and tial precursor concentration step for the formation of normal arc por-
less S-rich, and thus less prospective for the formation of porphyry- phyry Cu systems. In the former case, numerous post-subduction
type magmatic-hydrothermal Cu ± Au deposits. In particular, scenarios, such as collision or post-collisional rifting, can be envisaged
Richards and Mumin (2013a,b) suggested that this characteristic that would re-melt sulfide-bearing lower crustal rocks under sulfur-
might explain the relative rarity of Precambrian porphyry deposits (in undersaturated conditions, leading to redissolution of sulfides and
addition to preservation considerations; Kesler and Wilkinson, 2006). their metals, and the potential formation of post-subduction porphyry
Cu ± Au deposits. In contrast, in the latter case (preconcentration of
3. Lower crustal MASH processes metals in lower crustal sulfide cumulates) it is not clear what would
cause a change from sulfide-saturated to sulfide-undersaturated condi-
Nascent island arcs are characterized by the eruption of primitive tions if nothing else changed during a normal cycle of sulfur-rich arc
tholeiitic basalts onto thin oceanic crust, and are associated with few magmatism. The logical result of sequestering metals in cumulates at
magmatic-hydrothermal ore deposits. Only after the arc crust has thick- the base of the crust would be to reduce the potential for the fractionat-
ened and differentiated to lower density compositions will significant ed magmas subsequently to form magmatic-hydrothermal ore deposits,
36 J.P. Richards / Lithos 233 (2015) 27–45

rather than being a key step in their formation. Chiaradia (2014, alkaline (low-Fe) geochemical character is imprinted on arc magmatic
p. 44–45) explains this apparent contradiction by appealing to “a special suites, caused by early crystallization of Fe-rich minerals (Grove et al.,
combination of regional to local scale conditions [to] allow the success- 2003; Hamada and Fujii, 2008; Hildreth and Moorbath, 1988;
ful transfer of the deep Cu accumulations to shallower levels and the Pichavant and Macdonald, 2007; Romick et al., 1992; Stamper et al.,
formation of large porphyry copper systems”, but calling upon unde- 2014b). Evidence from experimental petrology and deep arc crustal sec-
fined special processes does not seem a very satisfactory model to ex- tions confirm that olivine, pyroxene, amphibole, and magnetite are
plain the formation of a common and widespread ore deposit type. early crystallizing phases from hydrous arc basalts, with plagioclase
Jenner et al. (2010, 2012), Sun et al. (2011), Lee et al. (2012) and appearing relatively late on the cotectic (Blatter and Carmichael,
Chiaradia (2014) all note an apparent increase in Cu content with differ- 1998b; Larocque and Canil, 2010; Moore and Carmichael, 1998;
entiation in mafic arc magmas prior to a decrease once intermediate Müntener and Ulmer, 2006; Müntener et al., 2001; Nandedkar et al.,
compositions are reached (cf. Fig. 4). Jenner et al. (2010) attribute this 2014; Tollan et al., 2012). This crystallization sequence not only results
trend to initial incompatible behavior of Cu, followed by change to sul- in early Fe-depletion relative to alkalis, but can also generate high Sr/Y
fide saturation triggered by the onset of magnetite crystallization (the and HREE-depleted compositions (high La/Yb, low Yb and Y concentra-
“magnetite crisis”), or volatile exsolution (Sun et al., 2004). Lee et al. tions) that resemble adakites (Castillo et al., 1999; Klepeis et al., 2003;
(2012) and Chiaradia (2014) further suggest that such processes could Macpherson et al., 2006; Richards and Kerrich, 2007; Takahashi et al.,
lead to a significant accumulation of metals within the lower crust, 2005; Tiepolo and Tribuzio, 2008; Tulloch and Kimbrough, 2003). Frac-
such that 70–80% of the Cu in the magma flux might be sequestered tionation of mafic and largely anhydrous minerals also enriches the
in arc cumulate rocks (Lee et al., 2012). However, it is notable that this melt in incompatible elements and water (Rohrlach and Loucks, 2005)
proportion is similar to the volume of arc magma crystallized in plutons and Plank et al. (2013) have noted that erupted arc magmas have glob-
versus being erupted at surface (~ 80% vs. ~ 20%, respectively; Barker ally uniform and elevated water contents of ~4 wt.% H2O.
et al., 2013; Carmichael, 2002; Sadofsky et al., 2008). Consequently, Fractionation of Fe-bearing minerals can also affect the Fe3+/Fe2+
this Cu loss might simply reflect background-level disseminations of ratio of the melt. Most mafic silicates preferentially incorporate Fe2+
sulfides in cumulate and plutonic rocks, and might not indicate signifi- rather than Fe3 +, so their crystallization might be expected to
cant depletion of Cu in the residual magmas that reach the upper crust. increase the ferric content of the melt. But crystallization of magnetite
In summary, deep crustal MASH processes affecting arc magmas re- (FeIII2FeIIO4) would have the opposite effect for magmas with Fe3 +/
sult in variable degrees of Cu loss to cumulate sulfides, the amount de- Fe2+ ratios less than 2:
pending on the oxidation state of the local lower crust. Large amounts
II III III II
of sulfide (and contained metals) might be lost if the lower crust is re- Fe OðmeltÞ þ Fe 2 O3ðmeltÞ ¼ Fe 2 Fe O4ðmagnetiteÞ ð2Þ
duced (e.g., graphitic), but lesser amounts are expected during interac-
tion with normal crustal rocks. The proportion of Cu lost from the In a relatively oxidized melt, where most of the Fe is present as Fe3+
magma flux in the latter case is similar to the estimated proportion of (dissolved as Fe2O3), the Fe3+/Fe2+ ratio and fO2 will increase in the re-
plutonic to extrusive rocks in arcs, and does not, therefore, indicate maining melt with magnetite crystallization:
that magmas ascending from MASH zones are significantly depleted in
Cu (or that lower crustal cumulate rocks are significantly enriched in 6 Fe
III
¼ 4 Fe
III II
þ O2ðmeltÞ ð3Þ
2 O3ðmeltÞ 2 Fe O4ðmagnetiteÞ
Cu). Few data exist for concentrations of other metals such as Au and
PGE in arc cumulates and magmas, but the higher partition coefficients whereas oxidation state will decrease in relatively reduced melts where
of these elements between sulfide phases and silicate melts suggest that most of the Fe is present as Fe2+ (dissolved as FeO):
evolved arc magmas might indeed be relatively depleted, consistent
with the relatively low abundances of Au and particularly PGE in II III II
6 Fe OðmeltÞ þ O2ðmeltÞ ¼ 2 Fe 2 Fe O4ðmagnetiteÞ ð4Þ
many porphyry Cu deposits, and the occurrence of PGEs in Alaskan-
type layered intrusions (e.g., Batanova et al., 2005; Spandler et al.,
2000; Thakurta et al., 2014). Circumstances that give rise to more Au- The boundary Fe3+/Fe2+ value for pure magnetite fractionation is
rich porphyry deposits either reflect late-stage magmatic and hydro- equivalent to log(Fe2O3/FeO) = 0.26, which is well above the average
thermal processes (e.g., Muntean and Einaudi, 2000; Murakami et al., for arc magmas (–0.21 ± 0.38; Figs. 2 and 3). Thus, typical arc magmas
2010; Sillitoe, 2000) or atypical arc, back-arc, or post-subduction pro- might be expected to become more reduced with magnetite crystalliza-
cesses associated with more alkaline magmas (e.g., Jensen and Barton, tion. The onset of magnetite fractionation in relatively evolved magmas
2000; Richards, 1995, 2009; Solomon, 1990; Wyborn and Sun, 1994). (≥60 wt.% SiO2) has been termed the “magnetite crisis” by Jenner et al.
In particular, Li and Audétat (2013) suggest that the mantle sources of (2010) because of its effect on the Fe3+/Fe2+ ratio, and therefore oxida-
potassic magmas might be saturated in monosulfide solid solution tion state, of the remaining magma. Reduction by this process is argued
(MSS) rather than a sulfide liquid, into which Au has a lower partition to cause reduction of soluble and abundant sulfate to less soluble sul-
coefficient. Such alkaline magmas might therefore have a higher poten- fide, leading to saturation of the magma in sulfide melt and/or minerals,
tial to form magmatic hydrothermal Au deposits upon upper crustal and consequent depletion of the magma in metals. However, while arc
emplacement. magmas do show an overall weakly decreasing trend of Cu concentra-
tions with differentiation (Fig. 4), Fe3+/Fe2+ ratios and ΔFMQ values
4. Fractional crystallization and devolatilization generally increase (as shown in Fig. 3, and observed in individual mag-
matic suites: e.g., Richards et al., 2006, 2013). This, suggests that either
Fractional crystallization, along with other processes such as magma the magnetite effect is not significant, or that other processes are coun-
mixing and wallrock (crustal) contamination, progressively affect the tering it (e.g., Fe2+-rich silicate mineral fractionation), or both.
compositions of arc magmas as they rise towards the surface. The The above arguments assume fractionation of pure magnetite. How-
most profound effects likely occur in deep crustal MASH zones (or ever, magnetite does not appear suddenly on the cotectic during mag-
“hot zones”; Annen et al., 2006), which transform and homogenize matic differentiation (i.e., as a “crisis”), but rather appears as a late-
the flux of basaltic arc magma from the asthenosphere into evolved, crystallizing end-member of the spinel solid-solution series, starting
intermediate-composition magmas (basaltic andesites to andesites, with Cr-spinel in primitive basaltic magmas, followed by Ti-magnetite
and locally more felsic melts derived by crustal melting; Eichelberger in andesites, and magnetite in dacites (e.g., Macdonald et al., 2000;
et al., 2006; Lee and Bachmann, 2014; Mišković and Francis, 2006; Müntener and Ulmer, 2006; Müntener et al., 2001; Nandedkar et al.,
Romick et al., 1992; Smith et al., 2010). This is also where the calc- 2014; Pichavant and Macdonald, 2007; Richards et al., 2006; Spandler
J.P. Richards / Lithos 233 (2015) 27–45 37

et al., 2003; Tollan et al., 2012; Zimmer et al., 2010). Cr- and Ti-rich spi- which on average appears to outweigh the reductive effect of magnetite
nels have lower Fe3+/Fe2+ ratios than magnetite (due to substitution of crystallization.
Ti4+ and Cr3+ for Fe3+; Stamper et al., 2014b; Waychunas, 1991), and The data in Fig. 3 suggest that most arc magmas are moderately ox-
their crystallization would therefore tend to cause an increase of the idized, but can become strongly oxidized (and in some cases reduced;
Fe3+/Fe2+ ratio in moderately oxidized primitive to intermediate com- Kelley and Cottrell, 2012), during differentiation. Both processes will
position magmas (i.e., oxidation) rather than a decrease. For example, in have major effects on the solubility of metals in the magma because of
a volcanic suite from the Neogene Antofalla volcano in Argentina, spi- their effects on magmatic sulfur content and speciation (Jugo et al.,
nels ranged from chromite in basaltic andesites, through Ti-rich magne- 2005b; Métrich and Clocchiatti, 1996). As discussed above, reduction
tite in andesites, to Ti-poor magnetite in dacites, with log(Fe2O3/FeO) will result in voluminous sulfide saturation and loss of potential to
ratios increasing from −0.40 in chromite (significantly lower than the form later magmatic hydrothermal ore deposits; this may explain the
median ratio of − 0.22 in arc magmas) to 0.19 in Ti-poor magnetite. dearth of porphyry deposits in, for example, Japan, where the lower
Magmatic oxidation states (calculated from magnetite-ilmenite pairs) crust contains graphitic lithologies (Sato, 2012; Tomkins et al., 2012).
correspondingly increased from ΔFMQ = ~ 1–1.5 to ~ 2–2.5 over this On the other hand, increasing oxidation state is likely to make sulfide
compositional range (Richards et al., 2006). Thus, early crystallization melt and mineral saturation less likely, whereas sulfate saturation and
of abundant Fe3+-poor spinel corresponds with an increase in oxidation anhydrite crystallization may occur (e.g., Audétat et al., 2004; Baker
state of the magma, a process that continues despite fractionation of and Rutherford, 1996; Carroll and Rutherford, 1985, 1987; Chambefort
(much smaller amounts of) Fe3 +-rich magnetite from more evolved et al., 2008; Imai et al., 1993; Luhr et al., 1984; Parat et al., 2002;
melts. Suffice to say, the interplay between these various redox buffers Streck and Dilles, 1998). Botcharnikov et al. (2011, 2013) have shown
and fractionating mineral compositions is complex and poorly under- that, up to ΔFMQ ≈ +1, gold solubility increases in S-bearing basalts–
stood at this time, but the bulk empirical data suggest that progressive, andesites, but drops off sharply at higher oxygen fugacities as sulfide,
mild oxidation with fractionation of arc magmas is more common than with which Au combines to dissolve in the melt, breaks down to sulfate.
reduction. In contrast, Zajacz et al. (2012) and Sun et al. (2014) have respectively
The process most widely recognized as having an oxidative effect on shown that Cu and Mo solubilities increase steadily with increasing
magmas as they rise towards the surface is devolatilization, especially of fO2, because these metals are dissolved predominantly as oxide species
reduced or weakly oxidized gaseous species such as H2, H2S, and SIVO2 (CuO0.5 and MoO3) rather than as sulfides. Consequently, given that arc
(Burgisser and Scaillet, 2007; Candela, 1986; Gaillard et al., 2001; magmas globally have oxidation states in the range ΔFMQ = + 0.5 to
Holloway, 2004; Mathez, 1984; Zimmer et al., 2010). Of these species, +2 (as reviewed above), they appear to be ideally suited to the trans-
SO2 is likely to be the most abundant in arc magmas, and SO2 is portation of Cu, Mo, Au, as well as other minor metals. In the case of
known to be a major component of volcanic emissions worldwide gold, however, highly oxidized magmas (ΔFMQ N N+1) may have lim-
(Sharma et al., 2004; Wallace, 2001, 2005). ited potential for Au transport and ore formation.
As for Fe, in a relatively oxidized melt in which most of the S is dis-
solved as SVIO24 − (i.e., where the average valence of S in the melt is 5. The magmatic to hydrothermal transition
N+4; ΔFMQ ≥~+1; Beermann et al., 2011; Jugo et al., 2005a), fO2 will
increase with SIVO2 degassing: Water solubility in silicate melts increases with pressure (Baker and
Alletti, 2012; Burnham, 1979). Typical andesitic arc magmas contain
VI IV
MS O4ðmeltÞ ¼ MOðmeltÞ þ S O2ðvaporÞ þ 0:5 O2ðmeltÞ ð5Þ ~4 wt.% upon ascent into the upper crust (Plank et al., 2013), and can
be expected to saturate in H2O as pressure falls to 1–2 kb, or upon rising
(where M is an element such as Ca or Fe; Baker and Rutherford, 1996), to depths of ~3–6 km (Burnham, 1979). This process is sometimes re-
whereas fO2 will decrease in a more reduced melt (i.e., where the aver- ferred to as “first boiling” (due to decompression upon ascent), whereas
age valence of S in the melt is b+4; ΔFMQ ≤ ~ +1): “second boiling” is caused by post-emplacement crystallization and
concentration of water in the residual melt (Candela, 1989). Likely
–II IV these two processes overlap, but it is clear that the bulk of dissolved
MS ðmeltÞ þ 1:5 O2ðmeltÞ ¼ MOðmeltÞ þ S O2ðvaporÞ ð6Þ
magmatic water exsolves during second boiling at relatively shallow
levels in the crust, whereas CO2, which has much lower solubility in sil-
Thus, typical S-rich, moderately oxidized arc magmas are likely to
icate melts, begins to degas much earlier and at depth (i.e., during first
undergo late stage oxidation as a result of SO2 degassing and coupled
boiling; Holloway, 1976; King and Holloway, 2002). It is this shallow-
magnetite crystallization:
level volatile exsolution that is critical to the formation of magmatic-
II VI III II hydrothermal ore deposits, by partitioning metals from the melt into
3 Fe OðmeltÞ þ 3 MS O4ðmeltÞ ¼ Fe 2 Fe O4ðmagnetiteÞ þ 3 MOðmeltÞ
IV
an initially high temperature, saline, S-rich hydrothermal fluid phase,
þ 3 S O2ðvaporÞ þ O2ðmeltÞ ð7Þ in which metals are solvated as chloride, sulfide, or bisulfide complexes
(Candela and Holland, 1984; Candela and Piccoli, 1995; Frank et al.,
An additional late-stage oxidative process that has been proposed by 2011; Hedenquist and Lowenstern, 1994; Heinrich et al., 1999;
Bell and Simon (2011) is the exsolution of saline fluids, in which Fe2+ is Migdisov et al., 2014; Pokrovski et al., 2008; Seo et al., 2009; Simon
more soluble than Fe3+, again having the effect of increasing the Fe3+/ et al., 2005, 2006; Webster, 2004; Williams et al., 1995; Zajacz and
Fe2+ ratio in the residual magma. Halter, 2009). The efficiency of this partitioning process is controlled
Devolatilization of arc magmas begins at deep levels in the crust by several factors, governed primarily by the availability of metals in
with exsolution of a CO2-rich gas phase, into which some SO2 likely par- the melt, the volume of hydrothermal fluid exsolved relative to the vol-
titions (Blundy et al., 2010; Holloway, 1976; Kamenetsky et al., 2001; ume of melt, and the efficiency with which that fluid physically mixes
King and Holloway, 2002; Lowenstern, 2001; Newman et al., 2000; with the melt as bubbles prior to separation as a discrete volatile
Roberge et al., 2009; Wallace, 2005). But SO2 loss increases dramatically plume (Cline and Bodnar, 1991; Cloos, 2001; Henley and McNabb,
upon water saturation at shallower levels in the crust (5–10 km; 1978; Shinohara and Hedenquist, 1997).
Burnham, 1979; Wallace, 2001, 2005), and is almost quantitative upon In their seminal study, Cline and Bodnar (1991) showed that the
eruption at surface (whole-rock S contents of erupted arc magmas are depth of emplacement of the source magma body critically affects the
much lower than estimated magmatic values; Alt et al., 1993; Carroll availability of metals and the salinity of the exsolved fluid. Deeply
and Rutherford, 1985). Consequently, magmas reaching the upper emplaced magmas (~2 kb) exsolve a relatively saline fluid after ~ 60%
crust can be expected to undergo significant oxidation due to SO2 loss, crystallization, which can efficiently scavenge chloride-soluble metals
38 J.P. Richards / Lithos 233 (2015) 27–45

such as Cu from the residual magma (see also Candela, 1992; Chiaradia 1/T x 103
1.5 1.0
et al., 2012; Cline, 1995; Huber et al., 2012). In contrast, shallowly –5
emplaced magmas (0.5–1 kb) initially exsolve low salinity vapors that
are inefficient in dissolving chloride-soluble metals. Salinity of the
exsolved fluid increases at late stages of crystallization (≥90%), but by
this point most of the metal content of the magma has been locked up
at background concentrations in crystals. Gold, on the other hand, –10 I-type
may be solubilized early by dilute fluids because it is thought to be dis- Magmas
solved as bisulfide species rather than as chlorides; this may explain the Porphyry
observation that Cu-poor porphyry Au deposits are characteristically log fO
2 Cu-Au Deposits
emplaced at shallow depths (Muntean and Einaudi, 2000; Murakami
et al., 2010; Sillitoe, 2000; Vila and Sillitoe, 1991). –15
Several authors have suggested that late-stage sulfide saturation
prior to volatile exsolution in the upper crust might be an important MORB
preconcentration step, by first partitioning chalcophile and siderophile S-type
metals into a sulfide melt or crystalline phase. These sulfides would Magmas
then be susceptible to oxidation and/or redissolution by exsolving aque- –20
ous fluids, and transfer of the contained metals to the fluid (Halter et al.,
2002, 2005; Jugo et al., 1999; Keith et al., 1997; Larocque et al., 2000;
Nadeau et al., 2010, 2013; Spooner, 1993; Stavast et al., 2006; Porphyry
Wilkinson, 2013). While the presence of sparse sulfide inclusions in Sn-W Deposits
phenocrysts in plutonic rocks associated with porphyry deposits is not –25
uncommon (op. cit., and Borrok et al., 1999), it is not clear how impor- 400 500 600 700 800 900 1000 1100
tant this process is to enhancing the efficiency of magma-to-fluid trans- T (°C)
fer of metals, or whether this two-step process is more efficient than a
one-step process of partitioning metals directly from the melt into an Fig. 7. Temperature vs. log fO2 diagram, showing the ranges of conditions characteristic of
porphyry Cu-Au deposits, I-type calc-alkaline magmas, S-type magmas, and mid-ocean-
exsolving volatile phase (Audétat and Pettke, 2006; Candela and
ridge basalts (MORB). The main mineral and volatile fO2 buffer curves are also shown.
Holland, 1984; Candela and Piccoli, 1995; Cline and Bodnar, 1991). Modified from Burnham and Ohmoto (1980) and Hedenquist and Richards (1998). Abbre-
Fig. 6 suggests that quite large volumes of sulfide (N~ 1%) would need viations: Fay = fayalite; Hm = hematite; Mt = magnetite; Po = pyrrhotite; Py = pyrite;
to unmix or crystallize from the magma before a significant proportion Qz = quartz.
of the dissolved Cu could be extracted. With the possible exception of
some dikes associated with the Bingham porphyry, Utah (which contain pyrite). Anhydrite is also precipitated in abundance due to the high ac-
up to 0.1 modal % sulfide globules; Keith et al., 1997), such high sulfide tivities of dissolved sulfate (Gustafson and Hunt, 1975; Lowell and
concentrations have not been observed in porphyry-related intrusive Guilbert, 1970). As temperatures continue to cool and Eq. (8) progresses
rocks. It could be argued that the sulfides have been redissolved prior further to the right, the fluids become increasingly acidic from the disso-
to final crystallization, and that the evidence has therefore been lost ciation of H2SO4 (and HCl; Hedenquist and Taran, 2013), leading to the
(e.g., Larocque et al., 2000), but one might still expect to see abundant characteristic formation of sericite (phyllic alteration) and then clays
sulfide inclusions trapped in and protected by phenocrysts if this pro- (argillic alteration) in originally feldspar-stable wallrocks (potassic al-
cess was important, whereas these inclusions are in fact sparse. Instead, teration; Beane and Titley, 1981; Harris and Golding, 2002; Hemley
the relatively high and increasing oxidation state of evolved arc magmas and Jones, 1964; Montoya and Hemley, 1975; Reed et al., 2013). Copper
reaching the upper crust (ΔFMQ = +1 to +2) suggests that the bulk of and Mo precipitate from these fluids in response to: (1) cooling be-
S in these magmas is present as sulfate (Carroll and Rutherford, 1985; tween ~ 425°–350 °C (Crerar and Barnes, 1976; Hemley et al., 1992;
Chambefort et al., 2008), with sparse sulfides being stable only because Hurtig and Williams-Jones, 2014b; Klemm et al., 2007; Landtwing
of high total sulfur fugacity (and therefore also high fH2S). et al., 2005, 2010; Redmond et al., 2004); (2) the increase in fH2S
Hydrothermal fluids exsolved from these magmas will have similar- resulting from SO2 disproportionation (Simon and Ripley, 2011); and
ly high oxidation states, approximately corresponding to the sulfur gas (3) in the case of Mo, decreasing pH (Rempel et al., 2008, 2009; Seo
(SO2–H2S) buffer curve (~ ΔFMQ = + 2; Fig. 7; Einaudi et al., 2003; et al., 2012). Gold may also precipitate with these sulfide minerals
Hedenquist and Richards, 1998; Ohmoto, 1986; Ohmoto and Rye, (Simon et al., 2005), and/or may stay in solution to lower temperatures
1979). With cooling below ~ 450 °C, the SO2–H2S buffer curve crosses as bisulfide or chloride complexes, potentially to be precipitated in the
the magnetite–hematite buffer (Fig. 7), as reflected in the commonly geothermal/epithermal environment (Gammons and Williams-Jones,
observed late-stage oxidation of early hydrothermal magnetite (associ- 1997; Hedenquist et al., 1993, 1998; Heinrich et al., 2004; Hurtig and
ated with chalcopyrite in potassic alteration) to hematite + pyrite Williams-Jones, 2014a; Kouzmanov and Pokrovski, 2012; Pudack
(Einaudi et al., 2003; Sun et al., 2013). A key reaction from the point of et al., 2009; Simmons and Brown, 2007).
view of ore deposition from these magmatic-hydrothermal fluids is
the disproportionation of SO2 (the predominant sulfur species in high- 6. Conclusions
temperature fluids) to H2S and H2SO4, which occurs progressively as
the fluid cools below ~ 400 °C (Holland, 1965; Kusakabe et al., 2000; The purpose of this review has been to explore the roles of one inten-
Whitney, 1988): sive parameter (oxidation state) and one extensive parameter (sulfur
content) in arc magma-genesis and metallogeny. These two parameters
IV –II VI together exercise fundamental controls on the behavior of metals
4 S O2ðaqÞ þ 4 H2 O ¼ H2 S ðaqÞ þ 3 H2 S O4ðaqÞ ð8Þ
during partial melting in the mantle, differentiation in the crust, and dur-
ing volatile exsolution and cooling. The additional essential presence of
This reaction, for the first time in the evolution of typical arc mag- water in these magmatic-to-hydrothermal systems is taken as read
matic to hydrothermal systems, generates high activities of dissolved (e.g., Burnham, 1979; Candela, 1992; Richards, 2011b).
sulfide, which cause the early precipitation of Cu- and Mo-bearing sul- The majority of experimental and analytical observations suggest
fide minerals (chalcopyrite, bornite, and molybdenite, along with that seafloor alteration and sedimentation since the Neoproterozoic
J.P. Richards / Lithos 233 (2015) 27–45 39

Oxygenation Event introduces a flux of oxidized components into the to fall below ~ 400 °C to form H2S and H2SO4. The newly formed H2S
mantle wedge upon subduction and slab devolatilization. Chief among quickly bonds with dissolved metals to precipitate as sulfide minerals
these oxidants is probably sulfate, originally derived from seawater (chalcopyrite, molybdenite, pyrite), while the H2SO4 (and HCl) causes
and fixed in altered seafloor basalts or sediments. Ferric iron is less mo- increasingly acidic forms of wallrock alteration at lower temperatures
bile in hydrothermal fluids, but may cause further oxidation of sulfides (b350 °C; sericite and clay).
to soluble sulfates during prograde metamorphism by in situ conversion In summary, the combination of high water and sulfur contents, and
to ferrous iron. Ferric iron may also be efficiently mobilized by partial moderate oxidation state (ΔFMQ ≈ +1) of typical arc magmas is suffi-
melts of the slab. cient to mobilize metals such as Cu and Mo from the asthenospheric
The process of hydration and oxidation of the mantle wedge above mantle wedge above subduction zones into the upper crust, where
the subducting and devolatilizing slab is likely to be slow and incremen- they can be partitioned into exsolving hydrothermal fluids and
tal, possibly requiring up to 10 m.y. to achieve a steady state condition of reprecipitate upon cooling to form porphyry-type ore deposits. Condi-
1 or 2 log fO2 units above background depleted mantle values (e.g., up to tions that might cause retention of large volumes of sulfide, either in
ΔFMQ ≈ +1). However, this modest increase in oxidation state is suffi- the mantle source or during lower crustal MASH processes, will likely
cient to shift the balance between dissolved sulfide and sulfate compo- reduce the potential for direct porphyry ore formation, although these
nents in the melt towards more soluble sulfates, allowing the magmas sulfides and their contained metals might be remobilized during later
to dissolve much larger amounts of sulfur than reduced melts. The re- (post-subduction) melting events.
quirement for a considerable amount of time before the mantle wedge
becomes “primed” to generate potentially fertile (moderately oxidized,
S-rich, hydrous) melts may explain the rarity of porphyry deposits in Acknowledgements
primitive island arcs, and the late appearance of large porphyry deposits
within magmatic cycles in many continental arcs. This work was supported by a Discovery Grant from the Natural
Although some metals may be carried into the mantle wedge by the Sciences and Engineering Research Council of Canada. I thank Derek
subduction zone fluids, this is not essential, because the mantle already Wyman for inviting this contribution, which is dedicated to the memory
contains substantial background concentrations of base and precious of Robert Kerrich. Rob was my post-doctoral supervisor, and instilled in
metals (e.g., 30 ppm Cu, 1 ppb Au, 6.2 ppb Pt in depleted mantle; me a sense of balanced and rigorous enquiry that I hope this review
Salters and Stracke, 2004). The significance of the flux of water and ox- does justice to. The loss of Rob’s influence and mentorship is a great
idized sulfur from the slab is that it lowers the solidus of peridotite caus- blow to our discipline. I thank Katy Evans and David Cooke for helpful
ing partial melting, and lowers the stability of condensed sulfide phases reviews of the manuscript, and Clayton Deutsch for statistical advice.
(melts or minerals) while maintaining a high fS2. Consequently,
chalcophile and siderophile elements will behave incompatibly
(where residual sulfides are absent) or weakly compatibly (where sul- References
fide abundances are low), and will tend to partition into the silicate
Alt, J.C., Shanks, W.C., Jackson, M.C., 1993. Cycling of sulfur in subduction zones: The
melt, whose sulfur load will be dominantly in the form of sulfate. geochemistry of sulfur in the Mariana Island Arc and back-arc trough. Earth and
Processing of these largely metal-undepleted magmas in lower Planetary Science Letters 119, 477–494.
crustal MASH or “hot zones” under mildly oxidized but marginally Alt, J.C., Garrido, C.J., Shanks, W.C., Turchyn, A., Padrón-Navarta, J.A., Sánchez-Vizcaíno,
V.L., Pugnaire, M.T.G., Marchesi, C., 2012. Recycling of water, carbon, and sulfur during
sulfide-saturated (high fS2) conditions can be expected to leave at subduction of serpentinites: A stable isotope study of Cerro del Almirez, Spain. Earth
least some of the metal load of the magma flux in cumulates and gab- and Planetary Science Letters 327–328, 50–60.
broic plutons, but available data do not indicate that these rocks contain Alt, J.C., Schwarzenbach, E.M., Früh-Green, G.L., Shanks, W.C., Bernasconi, S.M., Garrido,
C.J., Crispini, L., Gaggero, L., Padrón-Navarta, J.A., Marchesi, C., 2013. The role of
much more than background concentrations of metals such as Cu. Data serpentinites in cycling of carbon and sulfur: Seafloor serpentinization and subduc-
for Au and PGE are sparse, and it is possible that these highly siderophile tion metamorphism. Lithos 178, 40–54.
elements may be significantly enriched in residual sulfides, and may Andersen, D.J., Lindsley, D.H., Davidson, P.M., 1993. QUILF: A Pascal program to assess
equilibria among Fe-Mg-Mn-Ti oxides, pyroxenes, olivine, and quartz. Computers
represent a source of metals for later, post-subduction magmatism.
and Geosciences 19, 1333–1350.
Thus, segregation of this material does not seem to significantly deplete Andreani, M., Muñoz, M., Marcaillou, C., Delacour, A., 2013. μXANES study of iron redox
the remaining magma in Cu (or Mo), although Au and PGE may be lost. state in serpentine during oceanic serpentinization. Lithos 178, 70–83.
Annen, C., Blundy, J.D., Sparks, R.S.J., 2006. The genesis of intermediate and silicic magmas
Andesitic magmas rising into the upper crust still typically contain be-
in deep crustal hot zones. Journal of Petrology 47, 505–539.
tween 50–100 ppm Cu, which has been shown to be more than ade- Arai, S., Ishimaru, S., 2008. Insights into petrological characteristics of the lithosphere of
quate to form large porphyry Cu deposits (Cline and Bodnar, 1991). mantle wedge beneath arcs through peridotite xenoliths: a review. Journal of Petrology
Transferring the remaining metal load of the magma into a hydro- 49, 665–695.
Arculus, R.J., 1994. Aspects of magma genesis in arcs. Lithos 33, 189–208.
thermal fluid phase that can generate porphyry Cu-Mo ± Au minerali- Audétat, A., Pettke, T., 2006. Evolution of a porphyry-Cu mineralized magma system at
zation might be either a one- or two-step process, involving either Santa Rita, New Mexico (USA). Journal of Petrology 47, 2021–2046.
direct partitioning of metals from the melt into the exsolving hydrother- Audétat, A., Pettke, T., Dolejš, D., 2004. Magmatic anhydrite and calcite in the ore-forming
quartz-monzodiorite magma at Santa Rita, New Mexico (USA): genetic constraints
mal fluid, or preconcentration of metals in an immiscible or crystallizing on porphyry-Cu mineralization. Lithos 72, 147–161.
sulfide phase, followed by redissolution of these sulfides by hydrother- Baker, D.R., Alletti, M., 2012. Fluid saturation and volatile partitioning between melts and
mal fluid. The high and increasing oxidation states of evolved arc hydrous fluids in crustal magmatic systems: The contribution of experimental mea-
surements and solubility models. Earth-Science Reviews 114, 298–324.
magmas suggests that, while some condensed sulfide may indeed be Baker, L.L., Rutherford, M.J., 1996. Crystallisation of anhydrite-bearing magmas. In: Brown,
present in late-stage magmas, it will not be abundant, and is therefore M., Candela, P.A., Peck, D.L., Stephens, W.E., Walker, R.J., Zen, E-an (Eds.), Origin of
not likely to significantly affect the overall metal budget of the system. granites and related rocks. Geological Society of America Special Paper 315,
pp. 243–250.
Occam’s razor suggest that, in the absence of clear evidence to the con-
Bali, E., Audétat, A., Keppler, H., 2011. The mobility of U and Th in subduction zone fluids:
trary, the simpler one-step magma-to-fluid partitioning process can ef- an indicator of oxygen fugacity and fluid salinity. Contributions to Mineralogy and
ficiently form large porphyry deposits, and the processes involved are Petrology 161, 597–613.
Ballard, J.R., Michael Palin, J., Campbell, I.H., 2002. Relative oxidation states of magmas in-
well established in the literature.
ferred from Ce(IV)/Ce(III) in zircon: application to porphyry copper deposits of
Sulfur content and oxidation state have one last, essential role to northern Chile. Contributions to Mineralogy and Petrology 144, 347–364.
play in the precipitation of sulfide ores from metalliferous magmatic- Ballhaus, C., 1993. Redox states of lithospheric and asthenospheric upper mantle.
hydrothermal fluids. The bulk of the sulfur in moderately oxidized, Contributions to Mineralogy and Petrology 114, 331–348.
Ballhaus, C., Berry, R.F., Green, D.H., 1991. High pressure experimental calibration of the
high temperature hydrothermal fluids is dissolved as SO2, but this mol- olivine-orthopyroxene-spinel oxygen geobarometer: implications for the oxidation
ecule begins to back-react with its solvent (H2O) as temperatures start state of the upper mantle. Contributions to Mineralogy and Petrology 107, 27–40.
40 J.P. Richards / Lithos 233 (2015) 27–45

Barker, S.J., Wilson, C.J.N., Baker, J.A., Millet, M.-A., Rotella, M.D., Wright, I.C., Wysoczanski, Canfield, D.E., Poulton, S.W., Narbonne, G.M., 2006. Late-Neoproterozoic deep-ocean oxy-
R.J., 2013. Geochemistry and petrogenesis of silicic magmas in the intra-oceanic genation and the rise of animal life. Science 315, 92–95.
Kermadec Arc. Journal of Petrology 54, 351–391. Canil, D., 1999. Vanadium partitioning between orthopyroxene, spinel and silicate melt
Batanova, V.G., Pertsev, A.N., Kamenetsky, V.S., Ariskin, A.A., Mochalov, A.G., Sobolev, A.V., and the redox states of mantle source regions for primary magmas. Geochimica et
2005. Crustal evolution of island-arc ultramafic magma: Galmoenan pyroxenite– Cosmochimica Acta 63, 557–572.
dunite plutonic complex, Koryak Highland (Far East Russia). Journal of Petrology Carmichael, I.S.E., 1991. The redox states of basic and silicic magmas: a reflection of their
46, 1345–1366. source regions. Contributions to Mineralogy and Petrology 106, 129–141.
Beane, R.E., Titley, S.R., 1981. Porphyry copper deposits. Part II. Hydrothermal alteration Carmichael, I.S.E., 2002. The andesite aqueduct: Perspectives on the evolution of interme-
and mineralization. Economic Geology 75th Anniversary, 235–263. diate magmatism in west-central (105–99°W) Mexico. Contributions to Mineralogy
Beermann, O., Botcharnikov, R.E., Holtz, F., Diedrich, O., Nowak, M., 2011. Temperature de- and Petrology 143, 641–663.
pendence of sulfide and sulfate solubility in olivine-saturated basaltic magmas. Carmichael, I.S.E., Ghiorso, M.S., 1986. Oxidation–reduction relations in basic magma: a
Geochimica et Cosmochimica Acta 75, 7612–7631. case for homogeneous equilibria. Earth and Planetary Science Letters 78, 200–210.
Behn, M.D., Kelemen, P.B., Hirth, G., Hacker, B.R., Massonne, H.-J., 2011. Diapirs as the Carmichael, I.S.E., Frey, H.M., Lange, R.A., Hall, C.M., 2006. The Pleistocene cinder cones
source of the sediment signature in arc lavas. Nature Geoscience 4, 641–646. surrounding Volcán Colima, Mexico re-visited: eruption ages and volumes, oxidation
Bell, A.S., Simon, A., 2011. Experimental evidence for the alteration of the Fe3+/ΣFe of sil- states, and sulfur content. Bulletin of Volcanology 68, 407–419.
icate melt caused by the degassing of chlorine-bearing aqueous volatiles. Geology 39, Carroll, M.R., Rutherford, M.J., 1985. Sulfide and sulfate saturation in hydrous silicate
499–502. melts. Proceedings of the Fifteenth Lunar and Planetary Science Conference, Part 2.
Bézos, A., Humler, E., 2005. The Fe3+/ΣFe ratios of MORB glasses and their implications for Journal of Geophysical Research v. 90, Supplement, pp. C601–C612.
mantle melting. Geochimica et Cosmochimica Acta 69, 711–725. Carroll, M.R., Rutherford, M.J., 1987. The stability of igneous anhydrite: experimental re-
Bischoff, J.L., Dickson, F.W., 1975. Seawater–basalt interaction at 200°C and 500 bars: im- sults and implications for sulfur behavior in the 1982 El Chichon trachyandesite
plications for origin of sea-floor heavy-metal deposits and regulation of seawater and other evolved magmas. Journal of Petrology 28, 781–801.
chemistry. Earth and Planetary Science Letters 25, 385–397. Castillo, P.R., Janney, P.E., Solidum, R.U., 1999. Petrology and geochemistry of Camiguin
Blatter, D.L., Carmichael, I.S.E., 1998a. Hornblende peridotite xenoliths from central Island, southern Philippines: insights to the source of adakites and other lavas in a
Mexico reveal the highly oxidized nature of subarc upper mantle. Geology 26, complex arc setting. Contributions to Mineralogy and Petrology 134, 33–51.
1035–1038. Cervantes, P., Wallace, P.J., 2003. Role of H2O in subduction-zone magmatism: New in-
Blatter, D.L., Carmichael, I.S.E., 1998b. Plagioclase-free andesites from Zitácuaro sights from melt inclusions in high-Mg basalts from central Mexico. Geology 31,
(Michoacán), Mexico: petrology and experimental constraints. Contributions to Min- 235–238.
eralogy and Petrology 132, 121–138. Chambefort, I., Dilles, J.H., Kent, A.J.R., 2008. Anhydrite-bearing andesite and dacite as a
Blevin, P.L., 2004. Redox and compositional parameters for interpreting the granitoid source for sulfur in magmatic-hydrothermal mineral deposits. Geology 36, 719–722.
metallogeny of Eastern Australia: Implications for gold-rich ore system. Resource Chen, S.-Y., Ambe, S., Takematsu, N., Ambe, F., 1996. The chemical states of iron in marine
Geology 54, 241–252. sediments by means of Mössbauer spectroscopy in combination with chemical
Blundy, J., Cashman, K.V., Rust, A., Witham, F., 2010. A case for CO2-rich arc magmas. Earth leachings. Journal of Oceanography 52, 705–715.
and Planetary Science Letters 290, 289–301. Chiaradia, M., 2014. Copper enrichment in arc magmas controlled by overriding plate
Borrok, D., Kesler, S.E., Vogel, T.A., 1999. Sulfide minerals in intrusive and volcanic rocks of thickness. Nature Geoscience 7, 43–46.
the Bingham-Park City belt, Utah. Economic Geology 94, 1213–1230. Chiaradia, M., Merino, D., Spikings, R., 2009. Rapid transition to long-lived deep crustal
Botcharnikov, R.E., Linnen, R.L., Wilke, M., Holtz, F., Jugo, P.J., Berndt, J., 2011. High gold magmatic maturation and the formation of giant porphyry-related mineralization
concentrations in sulphide-bearing magma under oxidizing conditions. Nature Geo- (Yanacocha, Peru). Earth and Planetary Science Letters 288, 505–515.
science 4, 112–115. Chiaradia, M., Ulianov, A., Kouzmanov, K., Beate, B., 2012. Why large porphyry Cu deposits
Botcharnikov, R.E., Holtz, F., Mungall, J.E., Beermann, O., Linnen, R.L., Garbe-Schonberg, D., like high Sr/Y magmas? Scientific Reports 2, 685. http://dx.doi.org/10.1038/
2013. Behavior of gold in a magma at sulfide-sulfate transition: Revisited. American srep00685.
Mineralogist 98, 1459–1464. Christie, D.M., Carmichael, I.S.E., Langmuir, C.H., 1986. Oxidation states of mid-ocean ridge
Brandon, A.D., Draper, D.S., 1996. Constraints on the origin of the oxidation state of man- basalt glasses. Earth and Planetary Science Letters 79, 397–411.
tle overlying subduction zones: An example from Simcoe, Washington, USA. Cline, J.S., 1995. Genesis of porphyry copper deposits: the behavior of water, chloride, and
Geochimica et Cosmochimica Acta 60, 1739–1749. copper in crystallizing melts. In: Pierce, F.W., Bolm, J.G. (Eds.), Arizona Geological So-
Brounce, B., Kelley, K., Cottrell, E., 2013. Evaluating proxies for oxygen fugacity at the ciety Digest v. 20, pp. 69–82.
Mariana arc. Goldschmidt Conference 2013 Abstracts. Mineralogical Magazine v. 77, Cline, J.S., Bodnar, R.J., 1991. Can economic porphyry copper mineralization be generated
p. 777. by a typical calc-alkaline melt? Journal of Geophysical Research 96, 8113–8126.
Brounce, M.N., Kelley, K.A., Cottrell, E., 2014. Variations in Fe3+/∑Fe of Mariana arc ba- Cloos, M., 2001. Bubbling magma chambers, cupolas, and porphyry copper deposits. In-
salts and mantle wedge fO2. Journal of Petrology 55, 2513–2536. ternational Geology Review 43, 285–311.
Bryant, J.A., Yogodzinski, G.M., Churikova, T.G., 2007. Melt-mantle interactions beneath Core, D.P., Kesler, S.E., Essene, E.J., Dufresne, E.B., Clarke, R., Arms, D.A., Walko, D., Rivers,
the Kamchatka arc: Evidence from ultramafic xenoliths from Shiveluch volcano. M.L., 2005. Copper and zinc in silicate and oxide minerals in igneous rocks from the
Geochemistry, Geophysics, Geosystems 8, Q04007. http://dx.doi.org/10.1029/ Bingham – Park City Belt, Utah: synchrotron X-ray-fluorescence data. Canadian Min-
2006GC001443. eralogist 43, 1781–1796.
Bureau, H., Keppler, H., 1999. Complete miscibility between silicate melts and hydrous Core, D.P., Kesler, S.E., Essene, E.J., 2006. Unusually Cu-rich magmas associated with giant
fluids in the upper mantle: experimental evidence and geochemical implications. porphyry copper deposits: Evidence from Bingham, Utah. Geology 34, 41–44.
Earth and Planetary Science Letters 165, 187–196. Cottrell, E., Kelley, K.A., Lanzirotti, A., Fischer, R.A., 2009. High-precision determination of
Burgisser, A., Scaillet, B., 2007. Redox evolution of a degassing magma rising to the sur- iron oxidation state in silicate glasses using XANES. Chemical Geology 268, 167–179.
face. Nature 445, 194–197. http://dx.doi.org/10.1038/nature05509. Crabtree, S.M., Lange, R.A., 2012. An evaluation of the effect of degassing on the oxidation
Burnham, C.W., 1979. Magmas and hydrothermal fluids. In: Barnes, H.L. (Ed.), Geochem- state of hydrous andesite and dacite magmas: a comparison of pre- and post-eruptive
istry of hydrothermal ore deposits, 2nd edition John Wiley and Sons, New York, Fe2+ concentrations. Contributions to Mineralogy and Petrology 163, 209–224.
pp. 71–136. Crerar, D.A., Barnes, H.L., 1976. Ore solution chemistry V. Solubilities of chalcopyrite and
Burnham, C.W., Ohmoto, H., 1980. Late-stage processes in felsic magmatism: Mining Ge- chalcocite assemblages in hydrothermal solution at 200°C to 350°C. Economic Geol-
ology Special Issue. No. 8, 1–11. ogy 71, 772–794.
Burnham, A.D., Berry, A.J., 2012. An experimental study of trace element partitioning be- Dale, C.W., Macpherson, C.G., Pearson, D.G., Hammond, S.J., Arculus, R.J., 2012. Inter-
tween zircon and melt as a function of oxygen fugacity. Geochimica et Cosmochimica element fractionation of highly siderophile elements in the Tonga Arc due to flux
Acta 95, 196–212. melting of a depleted source. Geochimica et Cosmochimica Acta 89, 202–225.
Burnham, A.D., Berry, A.J., 2014. The effect of oxygen fugacity, melt composition, temper- Davies, J.H., Stevenson, D., 1992. Physical model of source region of subduction zone vol-
ature and pressure on the oxidation state of cerium in silicate melts. Chemical Geol- canics. Journal of Geophysical Research 97, 2037–2070.
ogy 366, 52–60. De Hoog, J.C.M., Mason, P.R.D., van Bergen, M.J., 2001. Sulfur and chalcophile elements in
Campbell, I.H., Naldrett, A.J., 1979. The influence of silicate:sulfide ratios on the geochem- subduction zones: constraints from a laser ablation ICP-MS study of melt inclusions
istry of magmatic sulfides. Economic Geology 74, 1503–1506. from Galunggung Volcano, Indonesia. Geochimica et Cosmochimica Acta 65,
Candela, P.A., 1986. The evolution of aqueous vapor from silicate melts: Effect on oxygen 3147–3164.• Suggests excess S in mantle wedge introduced during slab dehydration;
fugacity. Geochimica et Cosmochimica Acta 50, 1205–1211. S primarily derived from subducted seafloor sediments.
Candela, P.A., 1989. Magmatic ore-forming fluids: thermodynamic and mass transfer cal- De Hoog, J.C.M., Hattori, K.H., Hoblitt, R.P., 2004. Oxidized sulfur-rich mafic magma at
culations of metal concentrations. In: Whitney, J.A., Naldrett, A.J. (Eds.), Ore deposi- Mount Pinatubo, Philippines. Contributions to Mineralogy and Petrology 146,
tion associated with magmas. Soc. Econ. Geol., Reviews in Economic Geology v. 4, 750–761.
pp. 203–221. Debret, B., Andreani, M., Muñoz, M., Bolfan-Casanova, N., Carlut, J., Nicollet, C., Schwartz,
Candela, P.A., 1992. Controls on ore metal ratios in granite-related ore systems: An exper- S., Trcera, N., 2014. Evolution of Fe redox state in serpentine during subduction.
imental and computational approach. Transactions of the Royal Society of Edinburgh: Earth and Planetary Science Letters 400, 206–218.
Earth Sciences 83, 317–326. Deering, C.D., Vogel, T.A., Patino, L.C., Szymanski, D.W., Alvarado, G.E., 2012. Magmatic
Candela, P.A., Holland, H.D., 1984. The partitioning of copper and molybdenum between processes that generate chemically distinct silicic magmas in NW Costa Rica and
silicate melts and aqueous fluids. Geochimica et Cosmochimica Acta 48, 373–380. the evolution of juvenile continental crust in oceanic arcs. Contributions to Mineral-
Candela, P.A., Piccoli, P.M., 1995. Model ore-metal partitioning from melts into vapor and ogy and Petrology 163, 259–275.
vapor/brine mixtures. In: Thompson, J.F.H. (Ed.), Magmas, fluids, and ore deposits. Defant, M.J., Drummond, M.S., 1990. Derivation of some modern arc magmas by melting
Mineralogical Association of Canada, Short Course Series v. 23, pp. 101–127 (ch. 5). of young subducted lithosphere. Nature 347, 662–665.
J.P. Richards / Lithos 233 (2015) 27–45 41

Dilles, J.H., Kent, A.J.R., Wooden, J.L., Tosdal, R.M., Koleszar, A., Lee, R.G., Farmer, L.P., Hedenquist, J.W., Taran, Y.A., 2013. Modeling the formation of advanced argillic lithocaps:
2015. Zircon compositional evidence for sulfur-degassing from ore-forming arc volcanic vapor condensation above porphyry intrusions. Economic Geology 108,
magmas. Economic Geology 110, 241–251. http://dx.doi.org/10.2113/econgeo. 1523–1540.
110.1.241. Hedenquist, J.W., Simmons, S.F., Giggenbach, W.F., Eldridge, C.S., 1993. White Island, New
Droop, G.T.R., 1987. A general equation for estimating Fe3 + concentrations in ferro- Zealand, volcanic-hydrothermal system represents the geochemical environment of
magnesian silicates and oxides from microprobe analyses, using stoichiometric high-sulfidation Cu and Au ore deposition. Geology 21, 731–734.
criteria. Mineralogical Magazine 51, 431–435. Hedenquist, J.W., Arribas Jr., A., Reynolds, J.R., 1998. Evolution of an intrusion-centered hy-
Edmonds, M., Humphreys, M.C.S., Hauri, E.H., Herd, R.A., Wadge, G., Rawson, H., drothermal system: Far Southeast–Lepanto porphyry and epithermal Cu-Au deposits,
Ledden, R., Plail, M., Barclay, J., Aiuppa, A., Christopher, T.E., Giudice, G., Guida, Philippines. Economic Geology 93, 373–404.
R., 2014. Pre-eruptive vapour and its role in controlling eruption style and lon- Heinrich, C.A., Günther, D., Audétat, A., Ulrich, T., Frischknecht, R., 1999. Metal fraction-
gevity at Soufrière Hills Volcano. Geological Society, London, Memoirs 39, ation between magmatic brine and vapor, determined by microanalysis of fluid inclu-
291–315. sions. Geology 27, 755–758.
Eggins, S.M., 1993. Origin and differentiation of picritic arc magmas, Ambae (Aoba), Heinrich, C.A., Dreisner, T., Steffánson, A., Seward, T.M., 2004. Magmatic vapor con-
Vanuatu. Contributions to Mineralogy and Petrology 114, 79–100. traction and the transport of gold from the porphyry environment to epithermal
Eichelberger, J.C., Izbekov, P.E., Browne, B.L., 2006. Bulk chemical trends at arc volcanoes ore deposits. Geology 32, 761–764.
are not liquid lines of descent. Lithos 87, 135–154. Hemley, J.J., Jones, W.R., 1964. Chemical aspects of hydrothermal alteration with empha-
Einaudi, M.T., Hedenquist, J.W., Inan, E.E., 2003. Sulfidation state of fluids in active sis on hydrogen metasomatism. Economic Geology 59, 538–569.
and extinct hydrothermal systems: Transitions from porphyry to epithermal en- Hemley, J.J., Cygan, G.L., Fein, J.B., Robinson, G.R., d’Angelo, W.M., 1992. Hydrothermal ore-
vironments. In: Simmons, S.F., Graham, I. (Eds.), Volcanic, Geothermal, and ore- forming processes in the light of studies in rock-buffered systems: I. Iron-copper-
forming fluids: Rulers and witnesses of processes within the Earth. Society of zinc-lead sulfide solubility relations. Economic Geology 87, 1–22.
Economic Geologists, Special Publication 10, pp. 285–313. Henley, R.W., McNabb, A., 1978. Magmatic vapor plumes and ground-water interaction in
Evans, K.A., 2012. The redox budget of subduction zones. Earth-Science Reviews 113, porphyry copper emplacement. Economic Geology 73, 1–20.
11–32. Hermann, J., Spandler, C.J., 2008. Sediment melts at sub-arc depths: an experimental
Evans, K.A., Tomkins, A.G., 2011. The relationship between subduction zone redox budget study. Journal of Petrology 49, 717–740.
and arc magma fertility. Earth and Planetary Science Letters 308, 401–409. Hildreth, W., Moorbath, S., 1988. Crustal contributions to arc magmatism in the Andes of
Evans, K.A., Elburg, M.A., Kamenetsky, V.S., 2012. Oxidation state of subarc mantle. Geology central Chile. Contributions to Mineralogy and Petrology 98, 455–489.
40, 783–786. Holland, H.D., 1965. Some applications of thermochemical data to problems of ore deposits
Evans, K.A., Tomkins, A.G., Cliff, J., Fiorentini, M.L., 2014. Insights into subduction zone II. Mineral assemblages and the composition of ore forming fluids. Economic Geology
sulfur recycling from isotopic analysis of eclogite-hosted sulfides. Chemical Geol- 60, 1101–1166.
ogy 365, 1–86. Holloway, J.R., 1976. Fluids in the evolution of granitic magmas: Consequences of finite
Foley, S.F., Barth, M.G., Jenner, G.A., 2000. Rutile/melt partition coefficients for trace CO2 solubility. Geological Society of America Bulletin 87, 1513–1518.
elements and an assessment of the influence of rutile on the trace element char- Holloway, J.R., 2004. Redox reactions in seafloor basalts: possible insights into silicic hy-
acteristics of subduction zone magmas. Geochimica et Cosmochimica Acta 64, drothermal systems. Chemical Geology 210, 225–230.
933–938. Hou, Z., Yang, Z., Qu, X., Meng, X., Li, Z., Beaudoin, G., Rui, Z., Gao, Y., Zaw, K., 2009. The
Frank, M.R., Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., 2011. Gold and copper Miocene Gangdese porphyry copper belt generated during post-collisional extension
partitioning in magmatic-hydrothermal systems at 800°C and 100MPa. Geochimica et in the Tibetan Orogen. Ore Geology Reviews 36, 25–51.
Cosmochimica Acta 75, 2470–2482. Hou, Z., Zhang, H., Pan, Z., Yang, Z., 2011. Porphyry Cu (–Mo–Au) deposits related to
Frey, F.A., Chappell, B.W., Roy, S.D., 1978. Fractionation of rare-earth elements in the melting of thickened mafic lower crust: Examples from the eastern Tethyan
Tuolumne Intrusive Series, Sierra Nevada batholith, California. Geology 6, 239–242. metallogenic domain. Ore Geology Reviews 39, 21–45.
Frost, D.J., McCammon, C.A., 2008. The redox state of Earth’s mantle. Annual Review of Huber, C., Bachmann, O., Vigneresse, J.-L., Dufek, J., Parmigiani, A., 2012. A physical
Earth and Planetary Science 36, 389–420. model for metal extraction and transport in shallow magmatic systems. Geo-
Gaillard, F., Scaillet, B., Pichavant, M., Bény, J.-M., 2001. The effect of water and fO2 on the chemistry, Geophysics, Geosystems 13, Q08003. http://dx.doi.org/10.1029/
ferric–ferrous ratio of silicic melts. Chemical Geology 174, 255–273. 2012GC004042.
Gammons, C.H., Williams-Jones, A.E., 1997. Chemical mobility of gold in the porphyry- Hurtig, N.C., Williams-Jones, A.E., 2014a. An experimental study of the transport of gold
epithermal environment. Economic Geology 92, 45–59. through hydration of AuCl in aqueous vapour and vapour-like fluids. Geochimica et
Gill, J.B., 1981. Orogenic andesites and plate tectonics. Springer-Verlag, New York (390 pp.). Cosmochimica Acta 127, 305–325.
Greene, A.R., Debari, S.M., Kelemen, P.B., Blusztajn, J., Clift, P.D., 2006. A detailed geochem- Hurtig, N.C., Williams-Jones, A.E., 2014b. An experimental study of the solubility of MoO3 in
ical study of island arc crust: the Talkeetna Arc section, south-central Alaska. Journal aqueous vapour and low to intermediate density supercritical fluids. Geochimica et
of Petrology 47, 1051–1093. Cosmochimica Acta 136, 169–193.
Grove, T.L., Kinzler, R.J., 1986. Petrogenesis of andesites. Annual Review of Earth and Imai, A., Listanco, E.L., Fuji, T., 1993. Petrologic and sulfur isotopic significance of highly
Planetary Sciences 14, 417–454. oxidized and sulfur-rich magma of Mt. Pinatubo, Philippines. Geology 21, 699–702.
Grove, T.L., Elkins-Tanton, L.T., Parman, S.W., Chatterjee, N., Müntener, O., Gaetani, G.A., Ishihara, S., Sasaki, A., 1989. Sulfur isotopic ratios of the magnetite-series and ilmenite-
2003. Fractional crystallization and mantle-melting controls on calc-alkaline differen- series granitoids of the Sierra Nevada batholith—a reconnaissance study. Geology
tiation trends. Contributions to Mineralogy and Petrology 145, 515–533. 17, 788–791.
Grove, T.L., Chatterjee, N., Parman, S.W., Médard, E., 2006. The influence of H2O on mantle Ishizuka, O., Tani, K., Reagan, M.K., Kanayama, K., Umino, S., Harigane, Y., Sakamoto, I.,
wedge melting. Earth and Planetary Science Letters 249, 74–89. Miyajima, Y., Yuasa, M., Dunkley, D.J., 2011. The timescales of subduction initiation
Grove, T.L., Till, C.B., Krawczynski, M.J., 2012. The role of H2O in subduction zone and subsequent evolution of an oceanic island arc. Earth and Planetary Science Letters
magmatism. Annual Review of Earth and Planetary Sciences 40, 413–439. 306, 229–240.
Gustafson, L.B., Hunt, J.P., 1975. The porphyry copper deposit at El Salvador, Chile. Jagoutz, O., 2013. Were ancient granitoid compositions influenced by contemporaneous
Economic Geology 70, 857–912. atmospheric and hydrosphere oxidation states? Terra Nova 25, 95–101.
Halter, W.E., Pettke, T., Heinrich, C.A., 2002. The origin of Cu/Au ratios in porphyry-type Jégo, S., Dasgupta, R., 2013. Fluid-present melting of sulfide-bearing ocean-crust:
ore deposits. Science 296, 1844–1846. Experimental constraints on the transport of sulfur from subducting slab to mantle
Halter, W.E., Heinrich, C.A., Pettke, T., 2005. Magma evolution and the formation of por- wedge. Geochimica et Cosmochimica Acta 110, 106–134.
phyry Cu-Au ore fluids: evidence from silicate and sulfide melt inclusions. Jégo, S., Dasgupta, R., 2014. The fate of sulfur during fluid-present melting of subducting
Mineralium Deposita 39, 845–863. basaltic crust at variable oxygen fugacity. Journal of Petrology 55, 1019–1050.
Hamada, M., Fujii, T., 2008. Experimental constraints on the effects of pressure and Jégo, S., Pichavant, M., 2012. Gold solubility in arc magmas: Experimental determination
H2O on the fractional crystallization of high-Mg island arc basalt. Contributions of the effect of sulfur at 1000 °C and 0.4 GPa. Geochimica et Cosmochimica Acta 84,
to Mineralogy and Petrology 155, 767–790. 560–592.
Hamlyn, P.R., Keays, R.R., 1986. Sulfur saturation and second-stage melts: Application to Jégo, S., Pichavant, M., Mavrogenes, J.A., 2010. Controls on gold solubility in arc magmas: An
the Bushveld platinum metal deposits. Economic Geology 81, 1431–1445. experimental study at 1000 °C and 4 kbar. Geochimica et Cosmochimica Acta 74,
Hamlyn, P.R., Keays, R.R., Cameron, W.E., Crawford, A.J., Waldron, H.M., 1985. Pre- 2165–2189.
cious metals in magnesian low-Ti lavas: implications for metallogenesis and sul- Jenner, F.E., O'Neill, H.St.C., Arculus, R.J., Mavrogenes, J.A., 2010. The magnetite crisis in the
fur saturation in primary magmas. Geochimica et Cosmochimica Acta 49, evolution of arc-related magmas and the initial concentration of Au, Ag and Cu. Journal
1797–1811. of Petrology 51, 2445–2464.
Hanson, G.N., 1980. Rare earth elements in petrogenetic studies of igneous systems. Jenner, F.E., Arculus, R.J., Mavrogenes, J.A., Dyriw, N.J., Nebel, O., Hauri, E.H., 2012.
Annual Review of Earth and Planetary Sciences 8, 371–406. Chalcophile element systematics in volcanic glasses from the northwestern Lau
Harris, A.C., Golding, S.D., 2002. New evidence of magmatic-fluid–related phyllic alter- Basin. Geochemistry, Geophysics, Geosystems 13, Q06014. http://dx.doi.org/10.
ation: Implications for the genesis of porphyry Cu deposits. Geology 30, 335–338. 1029/2012GC004088.
Hattori, K., 1993. High-sulfur magma, a product of fluid discharge from underlying mafic Jensen, E.P., Barton, M.D., 2000. Gold deposits related to alkaline magmatism. In:
magma: evidence from Mount Pinatubo, Philippines. Geology 21, 1083–1086. Hagemann, S.G., Brown, P.E. (Eds.), Gold in 2000. Reviews in Economic Geology
Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of hydro- v. 13, pp. 279–314.
thermal ore deposits. Nature 370, 519–527. Johnson, M.C., Plank, T., 1999. Dehydration and melting experiments constrain the fate of
Hedenquist, J.W., Richards, J.P., 1998. The influence of geochemical techniques on the subducted sediments. Geochemistry, Geophysics, Geosystems 1, 1007. http://dx.doi.
development of genetic models for porphyry copper deposits. In: Richards, J.P., org/10.1029/1999GC000014.
Larson, P.B. (Eds.), Techniques in hydrothermal ore deposits geology. Reviews Jugo, P.J., 2009. Sulfur content at sulfide saturation in oxidized magmas. Geology 37,
in Economic Geology v. 10, pp. 235–256 (ch. 10). 415–418.
42 J.P. Richards / Lithos 233 (2015) 27–45

Jugo, P.J., Candela, P.A., Piccoli, P.M., 1999. Magmatic sulfides and Au:Cu ratios in porphyry contributions of metals and sulfur to ore-forming fluids. Canadian Mineralogist 38,
deposits: an experimental study of copper and gold partitioning at 850 °C, 100 MPa in 1233–1249.
a haplogranitic melt–pyrrhotite–intermediate solid solution–gold metal assemblage, Laubier, M., Grove, T.L., Langmuir, C.H., 2014. Trace element mineral/melt partitioning for
at gas saturation. Lithos 46, 573–589. basaltic and basaltic andesitic melts: An experimental and laser ICP-MS study with
Jugo, P.J., Luth, R.W., Richards, J.P., 2005a. An experimental study of the sulfur content in application to the oxidation state of mantle source regions. Earth and Planetary Science
basaltic melts saturated with immiscible sulfide or sulfate liquids at 1300 °C and Letters 392, 265–278.
1.0 GPa. Journal of Petrology 46, 783–798. Laurenz, V., Fonseca, R.O.C., Ballhaus, C., Jochum, K.P., Heuser, A., Sylvester, P.J., 2013. The
Jugo, P.J., Luth, R.W., Richards, J.P., 2005b. Experimental data on the speciation of sulfur as solubility of palladium and ruthenium in picritic melts: 2. The effect of sulfur.
a function of oxygen fugacity in basaltic melts. Geochimica et Cosmochimica Acta 69, Geochimica et Cosmochimica Acta 108, 172–183.
497–503. Lee, C.-T.A., Bachmann, O., 2014. How important is the role of crystal fractionation in mak-
Jugo, P.J., Wilke, M., Botcharnikov, R.E., 2010. Sulfur K-edge XANES analysis of natural and ing intermediate magmas? Insights from Zr and P systematics. Earth and Planetary
synthetic basaltic glasses: Implications for S speciation and S content as function of Science Letters 393, 266–274.
oxygen fugacity. Geochimica et Cosmochimica Acta 74, 5926–5938. Lee, C.-T.A., Leeman, W.P., Canil, D., A Li, Z.-X., 2005. Similar V/Sc systematics in MORB and
Kamenetsky, V.S., Binns, R.A., Gemmell, J.B., Crawford, A.J., Mernagh, T.P., Maas, R., Steele, arc basalts: Implications for the oxygen fugacities of their mantle source regions.
D., 2001. Parental basaltic melts and fluids in eastern Manus backarc Basin: implica- Journal of Petrology 46, 2313–2336.
tions for hydrothermal mineralisation. Earth and Planetary Science Letters 184, Lee, C.-T.A., Luffi, P., Le Roux, V., Dasgupta, R., Albaréde, F., Leeman, W.P., 2010. The redox
685–702. state of arc mantle using Zn/Fe systematics. Nature 468, 681–685.
Kawamoto, T., 2006. Hydrous phases and water transport in the subducting slab. Reviews Lee, C.-T.A., Luffi, P., Chin, E.J., Bouchet, R., Dasgupta, R., Morton, D.M., Le Roux, V., Yin, Q.-Z.,
in Mineralogy and Geochemistry 62, 273–289. Jin, D., 2012. Copper systematics in arc magmas and implications for crust-mantle dif-
Kay, R.W., 1978. Aleutian magnesian andesites: melts from subducted Pacific ocean crust. ferentiation. Science 336, 64–68.
Journal of Volcanology and Geothermal Research 4, 117–132. Lepage, L.D., 2003. ILMAT: an Excel worksheet for ilmenite–magnetite geothermometry
Keays, R.R., 1987. Principles of mobilization (dissolution) of metals in mafic and ultramaf- and geobarometry. Computers and Geosciences 29, 673–678.
ic rocks — the role of immiscible magmatic sulphides in the generation of hydrother- Li, Y., Audétat, A., 2012. Partitioning of V, Mn, Co, Ni, Cu, Zn, As, Mo, Ag, Sn, Sb, W, Au, Pb,
mal gold and volcanogenic massive sulphide deposits. Ore Geology Reviews 2, 47–63. and Bi between sulfide phases and hydrous basanite melt at upper mantle conditions.
Keays, R.R., 1995. The role of komatiitic and picritic magmatism and S-saturation in the Earth and Planetary Science Letters 355–356, 327–340.
formation of ore deposits. Lithos 34, 1–18. Li, Y., Audétat, A., 2013. Gold solubility and partitioning between sulfide liquid,
Keith, J.D., Whitney, J.A., Hattori, K., Ballantyne, G.H., Christiansen, E.H., Barr, D.L., Cannan, monosulfide solid solution and hydrous mantle melts: Implications for the formation
T.M., Hook, C.J., 1997. The role of magmatic sulfides and mafic alkaline magmas in the of Au-rich magmas and crust–mantle differentiation. Geochimica et Cosmochimica
Bingham and Tintic mining districts, Utah. Journal of Petrology 38, 1679–1690. Acta 118, 247–262.
Kelley, K.A., Cottrell, E., 2009. Water and the oxidation state of subduction zone magmas. Li, Z.-X.A., Lee, C.-T.A., 2004. The constancy of upper mantle fO2 through time inferred
Science 325, 605–607. from V/Sc ratios in basalts. Earth and Planetary Science Letters 228, 483–493.
Kelley, K.A., Cottrell, E., 2012. The influence of magmatic differentiation on the oxidation Li, C., Ripley, E.M., Thakurta, J., Stifter, E.C., Qi, L., 2013. Variations of olivine Fo–Ni contents
state of Fe in a basaltic arc magma. Earth and Planetary Science Letters 329–330, and highly chalcophile element abundances in arc ultramafic cumulates, southern
109–121. Alaska. Chemical Geology 351, 15–28.
Kepezhinskas, P., Defant, M.J., Widom, E., 2002. Abundance and distribution of PGE and Liu, X., Xiong, X., Audétat, A., Li, Y., Song, M., Li, L., Sun, W., Ding, X., 2014. Partitioning of
Au in the island-arc mantle: implications for sub-arc metasomatism. Lithos 60, copper between olivine, orthopyroxene, clinopyroxene, spinel, garnet and silicate
113–128. melts at upper mantle conditions. Geochimica et Cosmochimica Acta 125, 1–22.
Kesler, S.E., Wilkinson, B.H., 2006. The role of exhumation in the temporal distribution of Lowell, J.D., Guilbert, J.M., 1970. Lateral and vertical alteration-mineralization zoning in
ore deposits. Economic Geology 101, 919–922. porphyry copper ore deposits. Economic Geology 65, 373–408.
Kesler, S.E., Sutter, J.F., Issigonis, M.J., Jones, L.M., Walker, R.L., 1977. Evolution of porphyry Lowenstern, J.B., 2001. Carbon dioxide in magmas and implications for hydrothermal sys-
copper mineralization in an oceanic island arc: Panama. Economic Geology 72, tems. Mineralium Deposita 36, 490–502.
1142–1153. Luhr, J.F., 1992. Slab-derived fluids and partial melting in subduction zones: insights from
Kesler, S.E., Chryssoulis, S.L., Simon, G., 2002. Gold in porphyry copper deposits: its abun- two contrasting Mexican volcanoes (Colima and Ceboruco). Journal of Volcanology
dance and fate. Ore Geology Reviews 21, 103–124. and Geothermal Research 54, 1–18.
Kessel, R., Schmidt, M.W., Ulmer, P., Pettke, T., 2005. Trace element signature of Luhr, J.F., Carmichael, I.S.E., Varekamp, J.C., 1984. The 1982 eruptions of El Chichón
subduction-zone fluids, melts and supercritical liquids at 120–180 km depth. Nature Volcano, Chiapas, Mexico: mineralogy and petrology of the anhydrite-bearing pum-
437, 724–727. ice. Journal of Volcanology and Geothermal Research 23, 69–108.
Kimura, J.-I., Ariskin, A.A., 2014. Calculation of water-bearing primary basalt and estima- Macdonald, R., Hawkesworth, C.J., Heath, E., 2000. The Lesser Antilles volcanic chain: a
tion of source mantle conditions beneath arcs: PRIMACALC2 model for WINDOWS. study in arc magmatism. Earth-Science Reviews 49, 1–76.
Geochemistry, Geophysics, Geosystems 15, 1494–1514. http://dx.doi.org/10.1002/ Macpherson, C.G., Dreher, S.T., Thirlwall, M.F., 2006. Adakites without slab melting: High
2014GC005329. pressure differentiation of island arc magma, Mindanao, the Philippines. Earth and
King, P.L., Holloway, J.R., 2002. CO2 solubility and speciation in intermediate (andesitic) Planetary Science Letters 243, 581–593.
melts: the role of H2O and composition. Geochimica et Cosmochimica Acta 66, Malaspina, N., Poli, S., Fumagalli, P., 2009. The oxidation state of metasomatized mantle
1627–1640. wedge: Insights from C-O-H-bearing garnet peridotite. Journal of Petrology 50,
Klemm, L.M., Pettke, T., Heinrich, C.A., Campos, E., 2007. Hydrothermal evolution of the El 1533–1552.
Teniente deposit, Chile: Porphyry Cu-Mo ore deposit from low-salinity magmatic Malaspina, N., Scambelluri, M., Poli, S., Van Roermund, H.L.M., Langenhorst, F.,
fluids. Economic Geology 102, 1021–1045. 2010. The oxidation state of mantle wedge majoritic garnet websterites
Klemme, S., Prowatke, S., Hametner, K., Gunther, D., 2005. Partitioning of trace elements metasomatised by C-bearing subduction fluids. Earth and Planetary Science
between rutile and silicate melts: Implications for subduction zones. Geochimica et Letters 298, 417–426.
Cosmochimica Acta 69, 2361–2371. Mallmann, G., O'Neill, H.St.C., 2009. The crystal/melt partitioning of V during mantle melt-
Klepeis, K.A., Clarke, G.L., Rushmer, T., 2003. Magma transport and coupling between de- ing as a function of oxygen fugacity compared with some other elements (Al, P, Ca,
formation and magmatism in the continental lithosphere. GSA Today 13, 4–11. Sc, Ti, Cr, Fe, Ga, Y, Zr and Nb). Journal of Petrology 50, 1765–1794.
Kouzmanov, K., Pokrovski, G.S., 2012. Hydrothermal controls on metal distribution in por- Mathez, E.A., 1984. Influence of degassing on oxidation-states of basaltic magmas. Nature
phyry Cu (-Mo-Au) systems. Society of Economic Geologists, Special Publication 16, 310, 371–375.
573–618. McInnes, B.I.A., McBride, J.S., Evans, N.J., Lambert, D.D., Andrew, A.A., 1999. Osmium iso-
Kusakabe, M., Komoda, Y., Takano, B., Abiko, T., 2000. Sulfur isotopic effects in the dispro- tope constraints on ore metal recycling in subduction zones. Science 286, 512–516.
portionation reaction of sulfur dioxide in hydrothermal fluids: implications for the McInnes, B.I.A., Gregoire, M., Binns, R.A., Herzig, P.M., Hannington, M.D., 2001. Hydrous
δ34S variations of dissolved bisulfate and elemental sulfur from active crater lakes. metasomatism of oceanic sub-arc mantle, Lihir, Papua New Guinea: petrology and
Journal of Volcanology and Geothermal Research 97, 287–307. geochemistry of fluid-metasomatised mantle wedge xenoliths. Earth and Planetary
Kushiro, I., Syono, Y., Akimoto, S., 1968. Melting of a peridotite nodule at high pressures Science Letters 188, 169–183.
and high water pressures. Journal of Geophysical Research 73, 6023–6029. Métrich, N., Clocchiatti, R., 1996. Sulfur abundance and its speciation in oxidized alkaline
Labanieh, S., Chauvel, C., Germa, A., Quidelleur, X., 2012. Martinique: a clear case for sed- melts. Geochimica et Cosmochimica Acta 60, 4151–4160.
iment melting and slab dehydration as a function of distance to the trench. Journal of Métrich, N., Bonnin-Mosbah, M., Susini, J., Menez, B., Galoisy, L., 2002. Presence of sulfite
Petrology 53, 2441–2464. (SIV) in arc magmas: Implications for volcanic sulfur emissions. Geophysical Research
Landtwing, M.R., Pettke, T., Halter, W.E., Heinrich, C.A., Redmond, P.B., Einaudi, M.T., Letters 29, 33-1–33-4. http://dx.doi.org/10.1029/2001GL014607.
Kunze, K., 2005. Copper deposition during quartz dissolution by cooling magmatic- Migdisov, A.A., Bychkov, A.Y., Williams-Jones, Q.E., van Hinsberg, V.J., 2014. A predictive
hydrothermal fluids: The Bingham porphyry. Earth and Planetary Science Letters model for the transport of copper by HCl-bearing water vapour in ore-forming
235, 229–243. magmatic-hydrothermal systems: Implications for copper porphyry ore formation.
Landtwing, M.R., Furrer, C., Redmond, P.B., Pettke, T., Guillong, M., Heinrich, C.A., 2010. Geochimica et Cosmochimica Acta 129, 33–53.
The Bingham Canyon porphyry Cu-Mo-Au deposit. III. Zoned copper-gold ore deposi- Mišković, A., Francis, D., 2006. Interaction between mantle-derived and crustal calc-
tion by magmatic vapor expansion. Economic Geology 105, 91–118. alkaline magmas in the petrogenesis of the Paleocene Sifton Range volcanic complex,
Larocque, J., Canil, D., 2010. The role of amphibole in the evolution of arc magmas and Yukon, Canada. Lithos 87, 104–134.
crust: the case from the Jurassic Bonanza arc section, Vancouver Island, Canada. Mitchell, A.H.G., Bell, J.D., 1973. Island-arc evolution and related mineral deposits. Journal
Contributions to Mineralogy and Petrology 159, 475–492. of Geology 81, 381–405.
Larocque, A.C.L., Stimac, J.A., Keith, J.D., Huminicki, M.A.E., 2000. Evidence for open-system Montoya, J.W., Hemley, J.J., 1975. Activity relations and stabilities in alkali feldspar and
behavior in immiscible Fe-S-O liquids in silicate magmas: Implications for mica alteration reactions. Economic Geology 70, 577–594.
J.P. Richards / Lithos 233 (2015) 27–45 43

Moore, G.M., Carmichael, I.S.E., 1998. The hydrous phase equilibria (to 3 kbar) of an andesite basalt from St. Vincent, Lesser Antilles arc. Geochimica et Cosmochimica Acta 66,
and basaltic andesite from western Mexico: constraints on water content and condi- 2193–2209.
tions of phenocryst growth. Contributions to Mineralogy and Petrology 130, 304–319. Piper, D.Z., Basler, J.R., Bischoff, J.L., 1984. Oxidation state of marine manganese nodules.
Mullen, E.K., McCallum, I.S., 2013. Coexisting pseudobrookite, ilmenite, and titanomagnetite Geochimica et Cosmochimica Acta 48, 2347–2355.
in hornblende andesite of the Coleman Pinnacle flow, Mount Baker, Washington: Evi- Plank, T., Kelley, K.A., Zimmer, M.M., Hauri, E.H., Wallace, P.J., 2013. Why do mafic arc
dence for a highly oxidized arc magma. American Mineralogist 98, 417–425. magmas contain ~4 wt% water on average? Earth and Planetary Science Letters 364,
Mungall, J.E., 2002. Roasting the mantle: Slab melting and the genesis of major Au and 168–179.
Au-rich Cu deposits. Geology 30, 915–918. Pokrovski, G.S., Borisova, A.Y., Harrichoury, J.-C., 2008. The effect of sulfur on vapor–liquid
Mungall, J.E., Brenan, J.M., 2014. Partitioning of platinum-group elements and Au between fractionation of metals in hydrothermal systems. Earth and Planetary Science Letters
sulfide liquid and basalt and the origins of mantle-crust fractionation of the 266, 345–362.
chalcophile elements. Geochimica et Cosmochimica Acta 125, 265–289. Portnyagin, M., Hoernle, K., Plechov, P., Mironov, N., Khubunaya, S., 2007. Con-
Mungall, J.E., Su, S., 2005. Interfacial tension between magmatic sulfide and silicate liq- straints on mantle melting and composition and nature of slab components
uids: Constraints on kinetics of sulfide liquation and sulfide migration through sili- in volcanic arcs from volatiles (H2 O, S, Cl, F) and trace elements in melt inclu-
cate rocks. Earth and Planetary Science Letters 234, 135–149. sions from the Kamchatka Arc. Earth and Planetary Science Letters 255, 53–69.
Muntean, J.L., Einaudi, M.T., 2000. Porphyry gold deposits of the Refugio district, Prouteau, G., Scaillet, B., 2013. Experimental constraints on sulphur behaviour in subduc-
Maricunga belt, northern Chile. Economic Geology 95, 1445–1472. tion zones: implications for TTG and adakite production and the global sulphur cycle
Müntener, O., Ulmer, P., 2006. Experimentally derived high-pressure cumulates from since the Archean. Journal of Petrology 54, 183–213.
hydrous arc magmas and consequences for the seismic velocity structure of lower Pudack, C., Halter, W.E., Heinrich, C.A., Pettke, T., 2009. Evolution of magmatic vapor to
arc crust. Geophysical Research Letters 33, L21308. http://dx.doi.org/10.1029/ gold-rich epithermal liquid: The porphyry to epithermal transition at Nevados de
2006GL027629. Famatina, Northwest Argentina. Economic Geology 104, 449–477.
Müntener, O., Kelemen, P.B., Grove, T.L., 2001. The role of H2O during crystallization of Rabbia, O.M., Hernández, L.B., King, R.W., López-Escobar, L., 2002. Discussion on “Giant ver-
primitive arc magmas under uppermost mantle conditions and genesis of igneous sus small porphyry copper deposits of Cenozoic age in northern Chile: adakitic versus
pyroxenites: an experimental study. Contributions to Mineralogy and Petrology normal calc-alkaline magmatism” by Oyarzun et al. (Mineralium Deposita 36:
141, 643–658. 794–798, 2001). Mineralium Deposita 37, 791–794.
Murakami, H., Seo, J.H., Heinrich, C.A., 2010. The relation between Cu/Au ratio and forma- Rapp, R.P., Watson, E.B., 1995. Dehydration melting of metabasalt at 8–32 kbar:
tion depth of porphyry-style Cu–Au ± Mo deposits. Mineralium Deposita 45, 11–21. Implications for continental growth and crust–mantle recycling. Journal of Petrology
Mysen, B.O., Carmichael, I.S.E., Virgo, D., 1985. A comparison of iron redox ratios in silicate 36, 891–931.
glasses determined by wet-chemical and 57Fe Mössbauer resonant absorption Reagan, M.K., Hanan, B.B., Heizler, M.T., Hartman, B.S., Hickey-Vargas, R., 2008. Petrogen-
methods. Contributions to Mineralogy and Petrology 90, 101–106. esis of volcanic rocks from Saipan and Rota, Mariana Islands, and implications for the
Nadeau, O., Williams-Jones, A.E., Stix, J., 2010. Sulphide magma as a source of metals in evolution of nascent island arcs. Journal of Petrology 49, 441–464.
arc-related magmatic hydrothermal ore fluids. Nature Geoscience 3, 501–505. Redmond, P.B., Einaudi, M.T., Inan, E.E., Landtwing, M.R., Heinrich, C.A., 2004. Copper
Nadeau, O., Stix, J., Williams-Jones, A.E., 2013. The behavior of Cu, Zn and Pb during mag- deposition by fluid cooling in intrusion-centered systems: New insights from the
matic–hydrothermal activity at Merapi volcano, Indonesia. Chemical Geology 342, Bingham porphyry ore deposit, Utah. Geology 32, 217–220.
167–179. Reed, M., Rusk, B., Palandri, J., 2013. The Butte magmatic-hydrothermal system: one fluid
Naldrett, A.J., 1989. Magmatic sulfide deposits. Monographs on Geology and Geophysics yields all alteration and veins. Economic Geology 108, 1379–1396.
14. Oxford University Press, Oxford (186 pp.). Rempel, K.U., Williams-Jones, A.E., Migdisov, A.A., 2008. The solubility of molybdenum di-
Nandedkar, R.H., Ulmer, P., Müntener, O., 2014. Fractional crystallization of primitive, hy- oxide and trioxide in HCl-bearing water vapour at 350 °C and pressures up to 160
drous arc magmas: an experimental study at 0.7 Gpa. Contributions to Mineralogy bars. Geochimica et Cosmochimica Acta 72, 3074–3083.
and Petrology 167, 1015. http://dx.doi.org/10.1007/s00410-014-1015-5. Rempel, K.U., Williams-Jones, A.E., Migdisov, A.A., 2009. The partitioning of
Naney, M.T., 1983. Phase equilibria of rock-forming ferromagnesian silicates in granitic molybdenum(VI) between aqueous liquid and vapour at temperatures up to 370
systems. American Journal of Science 283, 993–1033. °C. Geochimica et Cosmochimica Acta 73, 3381–3392.
Newman, S., Stolper, E., Stern, R., 2000. H2O and CO2 in magmas from the Mariana arc and Richards, J.P., 1995. Alkalic-type epithermal gold deposits—a review. In: Thompson, J.F.H.
back arc system. Geochemistry, Geophysics, Geosystems 1, 1013. http://dx.doi.org/ (Ed.), Magmas, fluids, and ore deposits. Mineralogical Association of Canada, Short
10.1029/1999GC000027. Course Series 23, pp. 367–400 (ch. 17).
Nilsson, K., Peach, C.L., 1993. Sulfur speciation, oxidation state, and sulfur concentration in Richards, J.P., 2002. Discussion of “Giant versus small porphyry copper deposits of Cenozoic
backarc magmas. Geochimica et Cosmochimica Acta 57, 3807–3813. age in northern Chile: adakitic versus normal calc-alkaline magmatism” by Oyarzun
Ohmoto, H., 1986. Stable isotope geochemistry of ore deposits. In: Valley, J.W., Taylor, et al. (Mineralium Deposita 36: 794–798, 2001). Mineralium Deposita 37, 788–790.
H.P., O’Neil, J.R. (Eds.), Stable isotopes in high temperature geological processes. Richards, J.P., 2003. Tectono-magmatic precursors for porphyry Cu-(Mo-Au) deposit for-
Mineralogical Society of America, Reviews in Mineralogy 16, pp. 491–599. mation. Economic Geology 98, 1515–1533.
Ohmoto, H., Rye, R.O., 1979. Isotopes of sulfur and carbon. In: Barnes, H.L. (Ed.), Geochem- Richards, J.P., 2005. Cumulative factors in the generation of giant calc-alkaline porphyry Cu
istry of hydrothermal ore deposits, 2nd edition John Wiley and Sons, New York, deposits. In: Porter, T.M. (Ed.), Super porphyry copper and gold deposits: A global per-
pp. 509–567. spective. v. 1. Porter Geoscience Consulting Publishing, Linden Park, South Australia,
Oyarzun, R., Márquez, A., Lillo, J., López, I., Rivera, S., 2001. Giant versus small porphyry pp. 7–25.
copper deposits of Cenozoic age in northern Chile: adakitic versus normal calc- Richards, J.P., 2009. Postsubduction porphyry Cu-Au and epithermal Au deposits:
alkaline magmatism. Mineralium Deposita 36, 794–798. Products of remelting of subduction-modified lithosphere. Geology 37, 247–250.
Parat, F., Dungan, M.A., Streck, M.J., 2002. Anhydrite, pyrrhotite, and sulfur-rich apatite: Richards, J.P., 2011a. Magmatic to hydrothermal metal fluxes in convergent and collided
tracing the sulfur evolution of an Oligocene andesite (Eagle Mountain, CO, USA). margins. Ore Geology Reviews 40, 1–26.
Lithos 64, 63–75. Richards, J.P., 2011b. High Sr/Y arc magmas and porphyry Cu ± Mo ± Au deposits: Just
Park, J.-W., Campbell, I.H., Arculus, R.J., 2013. Platinum-alloy and sulfur saturation in an add water. Economic Geology 106, 1075–1081.
arc-related basalt to rhyolite suite: Evidence from the Pual Ridge lavas, the Eastern Richards, J.P., 2012. Discussion of Sun et al. (2011): The genetic association of adakites
Manus Basin. Geochimica et Cosmochimica Acta 101, 76–95. and Cu-Au ore deposits. International Geology Review 54, 368–369.
Parkinson, I.J., Arculus, R.J., 1999. The redox state of subduction zones: insights from arc- Richards, J.P., 2013. Giant ore deposits form by optimal alignments and combinations of
peridotites. Chemical Geology 160, 409–423. geological processes. Nature Geoscience 6, 911–916.
Parkinson, I.J., Arculus, R.J., Eggins, S.M., 2003. Peridotite xenoliths from Grenada, Lesser Richards, J.P., Kerrich, R., 2007. Adakite-like rocks: Their diverse origins and questionable
Antilles Island Arc. Contributions to Mineralogy and Petrology 146, 241–262. role in metallogenesis. Economic Geology 102, 537–576.
Peach, C.L., Mathez, E.A., Keays, R.R., 1990. Sulfide melt–silicate melt distribution coeffi- Richards, J.P., Mumin, A.H., 2013a. Magmatic-hydrothermal processes within an evolving
cients for noble metals and other chalcophile elements as deduced from MORB: Earth: Iron oxide-copper-gold and porphyry Cu ± Mo ± Au deposits. Geology 41,
Implications for partial melting. Geochimica et Cosmochimica Acta 54, 3379–3389. 767–770.
Peacock, S.M., 1993. Large-scale hydration of the lithosphere above subducting slabs. Richards, J.P., Mumin, A.H., 2013b. Lithospheric fertilization and mineralization by arc
Chemical Geology 108, 49–59. magmas: Genetic links and secular differences between porphyry copper
Peacock, S.M., Rushmer, T., Thompson, A.B., 1994. Partial melting of subducting oceanic ±molybdenum±gold and magmatic-hydrothermal iron oxide-copper-gold deposits.
crust. Earth and Planetary Science Letters 121, 227–244. Society of Economic Geologists, Special Publication 17, 277–299 (Ch. 9).
Peng, R., Zhai, Y., Li, C., Ripley, E.M., 2013. The Erbutu Ni-Cu deposit in the Central Asian Richards, J.P., McCulloch, M.T., Chappell, B.W., Kerrich, R., 1991. Sources of metals in the
orogenic belt: A Permian magmatic sulfide deposit related to boninitic magmatism Porgera gold deposit, Papua New Guinea: evidence from alteration, isotope, and
in an arc setting. Economic Geology 108, 1879–1888. noble metal geochemistry. Geochimica et Cosmochimica Acta 55, 565–580.
Perfit, M.R., Gust, D.A., Bence, A.E., Arculus, R.J., Taylor, S.R., 1980. Chemical characteristics Richards, J.P., Ullrich, T., Kerrich, R., 2006. The Late Miocene–Quaternary Antofalla volcanic
of island-arc basalts: Implications for mantle sources. Chemical Geology 30, 227–256. complex, southern Puna, NW Argentina: Protracted history, diverse petrology, and
Pettke, T., Oberli, F., Heinrich, C.A., 2010. The magma and metal source of giant porphyry- economic potential. Journal of Volcanology and Geothermal Research 152, 197–239.
type ore deposits, based on lead isotope microanalysis of individual fluid inclusions. Richards, J.P., Jourdan, F., Creaser, R.A., Maldonado, G., DuFrane, S.A., 2013. Geology, geo-
Earth and Planetary Science Letters 296, 267–277. chemistry, geochronology, and economic potential of Neogene volcanic rocks in the
Pichavant, M., Macdonald, R., 2007. Crystallization of primitive basaltic magmas at crustal Laguna Pedernal and Salar de Aguas Calientes segments of the Archibarca lineament,
pressures and genesis of the calc-alkaline igneous suite: experimental evidence from northwest Argentina. Journal of Volcanology and Geothermal Research 258, 47–73.
St Vincent, Lesser Antilles arc. Contributions to Mineralogy and Petrology 154, Ridolfi, F., Renzulli, A., Puerini, M., 2010. Stability and chemical equilibrium of amphibole
535–558. in calc-alkaline magmas: an overview, new thermobarometric formulations and ap-
Pichavant, M., Mysen, B.O., Macdonald, R., 2002. Source and H2O content of high-MgO plication to subduction-related volcanoes. Contributions to Mineralogy and Petrology
magmas in island arc settings: an experimental study of a primitive calc-alkaline 160, 45–66.
44 J.P. Richards / Lithos 233 (2015) 27–45

Righter, K., Campbell, A.J., Humayun, M., Hervig, R.L., 2004. Partitioning of Ru, Rh, Pd, Re, Simon, A.C., Frank, M.R., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2005. Gold
Ir, and Au between Cr-bearing spinel, olivine, pyroxene and silicate melts. partitioning in melt-vapor-brine systems. Geochimica et Cosmochimica Acta 69,
Geochimica et Cosmochimica Acta 68, 867–880. 3321–3335.
Righter, K., Chesley, J.T., Caiazza, C.M., Gibson Jr., E.K., Ruiz, J., 2008. Re and Os concentra- Simon, A.C., Pettke, T., Candela, P.A., Piccoli, P.M., Heinrich, C.A., 2006. Copper partitioning in
tions in arc basalts: The roles of volatility and source region fO2 variations. a melt–vapor–brine–magnetite–pyrrhotite assemblage. Geochimica et Cosmochimica
Geochimica et Cosmochimica Acta 72, 926–947. Acta 70, 5583–5600.
Ringwood, A.E., 1977. Petrogenesis in island arc systems. In: Talwani, M., Pitman, W.C. Slansky, E., Johan, Z., Ohnenstetter, M., Barron, L.M., Suppel, D., 1991. Platinum minerali-
(Eds.), Island arcs, deep sea trenches, and back arc basins. American Geophysical zation in the Alaskan-type intrusive complexes near Fifield, N.S.W., Australia. Part 2.
Union [Maurice Ewing Series I], pp. 311–324. Platinum-group minerals in placer deposits at Fifield. Mineralogy and Petrology 43,
Roberge, J., Delgado-Granados, H., Wallace, P.J., 2009. Mafic magma recharge supplies high 161–180.
CO2 and SO2 gas fluxes from Popocatépetl volcano, Mexico. Geology 37, 107–110. Smith, I.E.M., Stewart, R.B., Price, R.C., Worthington, T.J., 2010. Are arc-type rocks the
Rock, N.M.S., Groves, D.I., 1988. Do lamprophyres carry gold as well as diamonds? Nature products of magma crystallisation? Observations from a simple oceanic arc volcano:
332, 253–255. Raoul Island, Kermadec Arc, SW Pacific. Journal of Volcanology and Geothermal Re-
Rohrlach, B.D., Loucks, R.R., 2005. Multi-million-year cyclic ramp-up of volatiles in a lower search 190, 219–234.
crustal magma reservoir trapped below the Tampakan copper-gold deposit by Mio- Sobolev, A., Chaussidon, M., 1996. H2O concentrations in primary melts from supra-
Pliocene crustal compression in the southern Philippines. In: Porter, T.M. (Ed.), subduction zones and mid-ocean ridges: Implications for H2O storage and recycling
Super porphyry copper and gold deposits: A global perspective. v. 2. PGC Publishing, in the mantle. Earth and Planetary Science Letters 137, 45–55.
Adelaide, South Australia, pp. 369–407. Solomon, M., 1990. Subduction, arc reversal, and the origin of porphyry copper-gold de-
Romick, J.D., Kay, S.M., Kay, R.W., 1992. The influence of amphibole fractionation on the posits in island arcs. Geology 18, 630–633.
evolution of calc-alkaline andesite and dacite tephra from the central Aleutians, Alas- Song, S., Su, L., Niu, Y., Lai, Y., Zhang, L., 2009. CH4 inclusions in orogenic harzburgite:
ka. Contributions to Mineralogy and Petrology 112, 101–118. Evidence for reduced slab fluids and implication for redox melting in mantle
Rowe, M.C., Kent, A.J.R., Nielsen, R.L., 2009. Subduction influence on oxygen fugacity and wedge. Geochimica et Cosmochimica Acta 73, 1737–1754.
trace and volatile elements in basalts across the Cascade Volcanic Arc. Journal of Pe- Spandler, C.J., Eggins, S.M., Arculus, R.J., Mavrogenes, J.A., 2000. Using melt inclusions to
trology 50, 61–91. determine parent-magma compositions of layered intrusions: Application to the
Ruscitto, D.M., Wallace, P.J., Cooper, L.B., Plank, T., 2012. Global variations in H2O/Ce: 2. Greenhills Complex (New Zealand), a platinum group minerals–bearing, island-arc
Relationships to arc magma geochemistry and volatile fluxes. Geochemistry, Geo- intrusion. Geology 28, 991–994.
physics, Geosystems 13, Q03025. http://dx.doi.org/10.1029/2011GC003887. Spandler, C.J., Arculus, R.J., Eggins, S.M., Mavrogenes, J.A., Price, R.C., Reay, A.J., 2003.
Rutherford, M.J., Devine, J.D., 1988. The May 18, 1980, eruption of Mount St. Helens. 3. Sta- Petrogenesis of the Greenhills Complex, Southland, New Zealand: magmatic differen-
bility and chemistry of amphibole in the magma chamber. Journal of Geophysical Re- tiation and cumulate formation at the roots of a Permian island-arc volcano. Contri-
search 93, 11,949–11,959. butions to Mineralogy and Petrology 144, 703–721.
Ryerson, F.J., Watson, E.B., 1987. Rutile saturation in magmas: Implications for Ti-Nb-Ta Spooner, E.T.C., 1993. Magmatic sulphide/volatile interaction as a mechanism for produc-
depletion in island-arc basalts. Earth and Planetary Science Letters 86, 225–239. ing chalcophile element enriched, Archean Au-quartz, epithermal Au-Ag and Au
Sadofsky, S.J., Portnyagin, M., Hoernle, K., Bogaard, P., 2008. Subduction cycling of volatiles skarn hydrothermal ore fluids. Ore Geology Reviews 7, 359–379.
and trace elements through the Central American volcanic arc: evidence from melt Stamper, C.C., Melekhova, E., Blundy, J.D., Arculus, R.J., Humphreys, M.C.S., Brooker, R.A.,
inclusions. Contributions to Mineralogy and Petrology 155, 433–456. 2014a. Oxidised phase relations of a primitive basalt from Grenada, Lesser Antilles.
Saito, T., Ishikawa, N., Kamata, H., 2004. Iron-titanium oxide minerals in block-and-ash- Contributions to Mineralogy and Petrology 167, 954. http://dx.doi.org/10.1007/
flow deposits: implications for lava dome oxidation processes. Journal of Volcanology s00410-013-0954-6.
and Geothermal Research 138, 283–294. Stamper, C.C., Blundy, J.D., Arculus, R.J., Melekhova, E., 2014b. Petrology of plutonic xeno-
Salters, V.J.M., Stracke, A., 2004. Composition of the depleted mantle. Geochemistry Geo- liths and volcanic rocks from Grenada, Lesser Antilles. Journal of Petrology 55,
physics Geosystems 5, 1–27. 1353–1387.
Sajona, F.G., Maury, R.C., 1998. Association of adakites with gold and copper mineraliza- Staudigel, H., Plank, T., White, B., Schmincke, H.-U., 1996. Geochemical fluxes during sea-
tion in the Philippines. Comptes Rendus de l’Académie des Sciences, Série II, Sciences floor alteration of the basaltic upper oceanic crust: DSDP sites 417 and 418. In: Bebout,
de la Terre et des Planètes 326, 27–34. G.E., Scholl, D.W., Kirby, S.H., Platt, J.P. (Eds.), Subduction top to bottom: Washington.
Sasaki, A., Ishihara, S., 1979. Sulfur isotopic composition of the magnetite-series and DC, American Geophysical Union, Geophysical Monograph 96, pp. 19–38.
ilmenite-series granitoids in Japan. Contributions to Mineralogy and Petrology 68, Stavast, W.J.A., Keith, J.D., Christiansen, E.H., Dorais, M.J., Tingey, D., 2006. The fate of mag-
107–115. matic sulfides during intrusion or eruption, Bingham and Tintic districts, Utah. Eco-
Sato, K., 2012. Sedimentary crust and metallogeny of granitoid affinity: Implications from nomic Geology 101, 329–345.
the geotectonic histories of the Circum-Japan Sea region, Central Andes and south- Stolper, E., Newman, S., 1994. The role of water in the petrogenesis of Mariana trough
eastern Australia. Resource Geology 62, 329–351. magmas. Earth and Planetary Science Letters 121, 293–325.
Sauerzapf, U., Lattard, D., Burchard, M., Engelmann, R., 2008. The titanomagnetite-ilmenite Streck, M.J., Dilles, J.H., 1998. Sulfur evolution of oxidized arc magmas as recorded in ap-
equilibrium: new experimental data and thermo-oxybarometric application to the crys- atite from a porphyry copper batholith. Geology 26, 523–526.
tallization of basic to intermediate rocks. Journal of Petrology 49, 1161–1185. http://dx. Sun, W.D., Arculus, R.J., Kamenetsky, V.S., Binns, R.A., 2004. Release of gold-bearing fluids
doi.org/10.1093/petrology/egn021. in convergent margin magmas prompted by magnetite crystallization. Nature 431,
Schmidt, M.W., Dardon, A., Chazot, G., Vannucci, R., 2004. The dependence of Nb and Ta 975–978.
rutile-melt partitioning on melt composition and Nb/Ta fractionation during subduc- Sun, W., Zhang, H., Ling, M.-X., Ding, X., Chung, S.-L., Zhou, J., Yang, X.-Y., Fan, W., 2011.
tion processes. Earth and Planetary Science Letters 226, 415–432. The genetic association of adakites and Cu-Au ore deposits. International Geology Re-
Schmidt, A., Weyer, S., John, T., Brey, G.P., 2009. HFSE systematics of rutile-bearing view 53, 691–703.
eclogites: New insights into subduction zone processes and implications for the Sun, W.-d., Liang, H.-y., Ling, M.-x., Zhan, M.-z., Ding, X., Zhang, H., Yang, X.-y., Li, Y.-l.,
earth’s HFSE budget. Geochimica et Cosmochimica Acta 73, 455–468. Ireland, T.R., Wei, Q.-r., Fan, W.-m., 2013. The link between reduced porphyry copper
Seo, J.H., Guillong, M., Heinrich, C.A., 2009. The role of sulfur in the formation of magmatic– deposits and oxidized magmas. Geochimica et Cosmochimica Acta 103, 263–275.
hydrothermal copper–gold deposits. Earth and Planetary Science Letters 282, 323–328. Sun, W., Audétat, A., Dolejš, D., 2014. Solubility of molybdenite in hydrous granitic melts
Seo, J.H., Guillong, M., Heinrich, C.A., 2012. Separation of molybdenum and copper at 800 °C, 100–200 MPa. Geochimica et Cosmochimica Acta 131, 393–401.
in porphyry deposits: the roles of sulfur, redox, and pH in ore mineral deposi- Szabó, C., Falus, G., Zajacz, Z., Kovács, I., Bali, E., 2004. Composition and evolution of litho-
tion at Bingham Canyon. Economic Geology 107, 333–356. sphere beneath the Carpathian–Pannonian Region: a review. Tectonophysics 393,
Shafiei, B., Haschke, M., Shahabpour, J., 2009. Recycling of orogenic arc crust triggers por- 119–137.
phyry Cu mineralization in Kerman Cenozoic arc rocks, southeastern Iran. Takahashi, Y., Kagashima, S.-I., Mikoshiba, M.U., 2005. Geochemistry of adakitic quartz di-
Mineralium Deposita 44, 265–283. orite in the Yamizo Mountains, central Japan: Implications for Early Cretaceous
Shanks, W.C., Bischoff, J.L., Rosenbauer, R.J., 1981. Seawater sulfate reduction and adakitic magmatism in the inner zone of southwest Japan. The Island Arc 14, 150–164.
sulfur isotope fractionation in basaltic systems: interaction of seawater with Tatsumi, Y., 1989. Migration of fluid phases and genesis of basalt magmas in subduction
fayalite and magnetite at 200–350 °C. Geochimica et Cosmochimica Acta 45, zones. Journal of Geophysical Research 94, 4697–4707.
1977–1995. Thakurta, J., Ripley, E.M., Li, C., 2008. Pre-requisites for sulphide-poor PGE and sulphide-
Sharma, K., Blake, S., Self, S., Krueger, A.J., 2004. SO2 emissions from basaltic eruptions, rich Cu-Ni-PGE mineralization in Alaskan-type complexes. Journal of the Geological
and the excess sulfur issue. Geophysical Research Letters 31 (31), L13612. http:// Society of India 72, 611–622.
dx.doi.org/10.1029/2004GL019688. Thakurta, J., Ripley, E.M., Li, C., 2014. Platinum group element geochemistry of sulfide-rich
Shields-Zhou, G., Och, L., 2011. The case for a neoproterozoic oxygenation event: Geo- horizons in the Ural-Alaskan–type ultramafic complex of Duke Island, southeastern
chemical evidence and biological consequences. GSA Today 21, 4–11. Alaska. Economic Geology 109, 643–659.
Shinohara, H., Hedenquist, J.W., 1997. Constraints on magma degassing beneath the Far Thirlwall, M.F., Graham, A.M., Arculus, R.J., Harmon, R.S., Macpherson, C.G., 1996. Resolu-
Southeast porphyry Cu-Au deposit, Philippines. Journal of Petrology 38, 1741–1752. tion of the effects of crustal assimilation, sediment subduction, and fluid transport in
Sillitoe, R.H., 2000. Gold-rich porphyry deposits: Descriptive and genetic models and their island arc magmas: Pb-Sr-Nd-O isotope geochemistry of Grenada, Lesser Antilles.
role in exploration and discovery. Reviews in Economic Geology 13, 315–345. Geochimica et Cosmochimica Acta 60, 4785–4810.
Sillitoe, R.H., 2012. Copper provinces. Society of Economic Geologists, Special Publication Tiepolo, M., Tribuzio, R., 2008. Petrology and U-Pb zircon geochronology of amphibole-
16, 1–18. rich cumulates with sanukitic affinity from Husky Ridge (Northern Victoria Land,
Simmons, S.F., Brown, K.L., 2007. The flux of gold and related metals through a volcanic Antarctica): Insights into the role of amphibole in the petrogenesis of subduction-
arc, Taupo Volcanic Zone, New Zealand. Geology 35, 1099–1102. related magmas. Journal of Petrology 49, 937–970.
Simon, A.C., Ripley, E.M., 2011. The role of magmatic sulfur in the formation of ore de- Tilling, R.I., Gottfried, D., Rowe, J.J., 1973. Gold abundance in igneous rocks: bearing on
posits. Reviews in Mineralogy and Geochemistry 73, 513–578. gold mineralization. Economic Geology 68, 168–186.
J.P. Richards / Lithos 233 (2015) 27–45 45

Timm, C., de Ronde, C.E.J., Leybourne, M.I., Layton-Matthews, D., Graham, I.J., 2012. Widom, E., Kepezhinskas, P., Defant, M., 2003. The nature of metasomatism in the sub-arc
Sources of chalcophile and siderophile elements in Kermadec arc lavas. Economic Ge- mantle wedge: evidence from Re-Os isotopes in Kamchatka peridotite xenoliths.
ology 107, 1527–1538. Chemical Geology 196, 283–306.
Tollan, P.M.E., Bindeman, I., Blundy, J.D., 2012. Cumulate xenoliths from St. Vincent, Lesser Wilkinson, J.J., 2013. Triggers for the formation of porphyry ore deposits in magmatic arcs.
Antilles Island Arc: a window into upper crustal differentiation of mantle-derived ba- Nature Geoscience 6, 917–925.
salts. Contributions to Mineralogy and Petrology 163, 189–208. Williams, T.J., Candela, P.A., Piccoli, P.M., 1995. The partitioning of copper between silicate
Tomkins, A.G., Rebryna, K.C., Weinberg, R.F., Schaefer, B.F., 2012. Magmatic sulfide forma- melts and two-phase aqueous fluids: An experimental investigation at 1 kbar, 800 °C
tion by reduction of oxidized arc basalt. Journal of Petrology 53, 1537–1567. and 0.5 kbar, 850 °C. Contributions to Mineralogy and Petrology 121, 388–399.
Trail, D., Watson, E.B., Tailby, N.D., 2011. The oxidation state of Hadean magmas and im- Wilson, T.R.S., Thomson, J., Colley, S., Hydes, D.J., Higgs, N.C., 1985. Early organic diagene-
plications for early Earth’s atmosphere. Nature 480, 79–82. sis: The significance of progressive subsurface oxidation fronts in pelagic sediments.
Trail, D., Watson, E.B., Tailby, N.D., 2012. Ce and Eu anomalies in zircon as proxies for the Geochimica et Cosmochimica Acta 49, 811–822.
oxidation state of magmas. Geochimica et Cosmochimica Acta 97, 70–87. Wood, B.J., Virgo, D., 1989. Upper mantle oxidation state: ferric iron contents of lherzolite
Tulloch, A.J., Kimbrough, D.L., 2003. Paired plutonic belts in convergent margins and spinels by 57Fe Mössbauer spectroscopy and resultant oxygen fugacities. Geochimica
the development of high Sr/Y magmatism: Peninsular Ranges batholith of Baja- et Cosmochimica Acta 53, 1277–1291.
California and Median batholith of New Zealand. Geological Society of America, Spe- Wyborn, D., Sun, S.-s., 1994. Sulphur-undersaturated magmatism – a key factor for gener-
cial Paper 374, 1–21. ating magma-related copper-gold deposits. Australian Geological Survey Organisa-
Ulmer, P., 2001. Partial melting in the mantle wedge – the role of H2O in the genesis of tion Research Newsletter 21, 7–8.
mantle-derived ‘arc-related’ magmas. Physics of the Earth and Planetary Interiors Wyman, D., Kerrich, R., 1989. Archean shoshonitic lamprophyres associated with Superior
127, 215–232. province gold deposits: distribution, tectonic setting, noble metal abundances, and
Van Hoose, A.E., Streck, M.J., Pallister, J.S., Wälle, M., 2013. Sulfur evolution of the 1991 significance for gold mineralization. Economic Geology, Monograph 6, 651–667.
Pinatubo magmas based on apatite. Journal of Volcanology and Geothermal Research Wysoczanski, R.J., Handler, M.R., Schipper, C.I., Leybourne, M.I., Creech, J., Rotella, M.D.,
257, 72–89. Nichols, A.R.L., Wilson, C.J.N., Stewart, R.B., 2012. The tectonomagmatic source of
Vila, T., Sillitoe, R.H., 1991. Gold-rich porphyry systems in the Maricunga belt, northern ore metals and volatile elements in the southern Kermadec arc. Economic Geology
Chile. Economic Geology 86, 1238–1260. 107, 1539–1556.
Wallace, P.J., 2001. Volcanic SO2 emissions and the abundance and distribution of Yogodzinski, G.M., Lees, J.M., Churikova, T.G., Dorendorf, F., Wöerner, G., Volynets, O.N.,
exsolved gas in magma bodies. Journal of Volcanology and Geothermal Research 2001. Geochemical evidence for the melting of subducting oceanic lithosphere at
108, 85–106. plate edges. Nature 409, 500–504.
Wallace, P.J., 2005. Volatiles in subduction zone magmas: Concentrations and fluxes Zajacz, Z., Halter, W., 2009. Copper transport by high temperature, sulfur-rich magmatic
based on melt inclusion and volcanic gas data. Journal of Volcanology and Geother- vapor: Evidence from silicate melt and vapor inclusions in a basaltic andesite from
mal Research 140, 217–240. the Villarrica volcano (Chile). Earth and Planetary Science Letters 282, 115–121.
Wallace, P.J., Edmonds, M., 2011. The sulfur budget in magmas: Evidence from melt inclu- Zajacz, Z., Candela, P.A., Piccoli, P.M., Wälle, M., Sanchez-Valle, C., 2012. Gold and copper
sions, submarine glasses, and volcanic gas emissions. Reviews in Mineralogy and in volatile saturated mafic to intermediate magmas: Solubilities, partitioning, and im-
Geochemistry 73, 215–246. plications for ore deposit formation. Geochimica et Cosmochimica Acta 91, 140–159.
Wang, J., Hattori, K.H., Kilian, R., Stern, C.R., 2007. Metasomatism of sub-arc mantle peri- Zajacz, Z., Candela, P.A., Piccoli, P.M., Sanchez-Valle, C., Wälle, M., 2013. Solubility and
dotites below southernmost South America: reduction of fO2 by slab-melt. Contribu- partitioning behavior of Au, Cu, Ag and reduced S in magmas. Geochimica et
tions to Mineralogy and Petrology 153, 607–624. Cosmochimica Acta 112, 288–304.
Waychunas, G.A., 1991. Crystal chemistry of oxides and oxyhydroxides. In: Lindsley, D.H. Zhang, Z.-W., Li, W.-Y., Gao, Y.-B., Li, C., Ripley, E.M., Kamo, S., 2014. Sulfide mineralization
(Ed.), Oxide minerals: Petrologic and megnetic significance. Mineralogical Society of associated with arc magmatism in the Qilian Block, western China: zircon U-Pb age
America, Reviews in Mineralogy v. 25, pp. 11–68. and Sr-Nd-Os-S isotope constraints from the Yulonggou and Yaqu gabbroic intru-
Webster, J.D., 2004. The exsolution of magmatic hydrosaline chloride liquids. Chemical sions. Mineralium Deposita 49, 279–292.
Geology 210, 33–48. Zimmer, M.M., Plank, T., Hauri, E.H., Yogodzinski, G.M., Stelling, P., Larsen, J., Singer, B.,
Whitney, J.A., 1988. Composition and activity of sulfurous species in quenched magmatic Jicha, B., Mandeville, C., Nye, C.J., 2010. The role of water in generating the calc-
gases associated with pyrrhotite-bearing silicic systems. Economic Geology 83, alkaline trend: New volatile data for Aleutian magmas and a new Tholeiitic Index.
86–92. Journal of Petrology 51, 2411–2444.

S-ar putea să vă placă și