Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/266677773

Porphyry deposits and oxidized magmas

Article  in  Ore Geology Reviews · March 2015


DOI: 10.1016/j.oregeorev.2014.09.004

CITATIONS READS

388 8,729

11 authors, including:

Weidong Sun Ruifang Huang


Institute of Oceanology Southern University of Science and Technology
514 PUBLICATIONS   38,990 CITATIONS    23 PUBLICATIONS   943 CITATIONS   

SEE PROFILE SEE PROFILE

He Li hu Yongbin
Institute of Oceanology,Chinese Academy of Sciences Chinese Academy of Sciences
55 PUBLICATIONS   1,581 CITATIONS    11 PUBLICATIONS   588 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Formation of Porphyry Copper Deposits View project

Rare-metal Granites View project

All content following this page was uploaded by Lipeng Zhang on 30 October 2017.

The user has requested enhancement of the downloaded file.


Ore Geology Reviews 65 (2015) 97–131

Contents lists available at ScienceDirect

Ore Geology Reviews


journal homepage: www.elsevier.com/locate/oregeorev

Review

Porphyry deposits and oxidized magmas


Weidong Sun a,⁎, Rui-fang Huang b,c, He Li a, Yong-bin Hu a,c, Chan-chan Zhang a,c, Sai-jun Sun a,c,
Li-peng Zhang a,c, Xing Ding b, Cong-ying Li a, Robert E. Zartman a, Ming-xing Ling b
a
CAS Key Laboratory of Mineralogy and Metallogeny, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences, 511 Kehua Street, Wushan, Guangzhou 510640, China
b
State Key Laboratory of Isotope Geochemistry, Guangzhou Institute of Geochemistry, Chinese Academy of Sciences, 511 Kehua Street, Wushan, Guangzhou 510640, China
c
University of the Chinese Academy of Sciences, Beijing 100049, China

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

Article history: Porphyry deposits supply most of the world's Cu and Mo resources. Over 90% of the porphyry deposits are found
Received 17 June 2014 at convergent margins, especially above active subduction zones, with much fewer occurrences at post-
Received in revised form 19 August 2014 collisional or other tectonic settings. Porphyry Cu–(Mo)–(Au) deposits are essentially magmatic–hydrothermal
Accepted 2 September 2014
systems, which are generally initiated by injection of oxidized magmas saturated with metal-rich aqueous fluids,
Available online 16 September 2014
i.e., the parental magmas need to be water rich and oxidized with most of the sulfur appearing as sulfate in the
Keywords:
magma. Sulfur is the most important geosolvent that controls the behavior of Cu and other chalcophile elements,
Porphyry deposit due to high partition coefficients of chalcophile elements between sulfide and silicate melts. Small amount of
Oxidized magmas residual sulfides can hold a large amount of Cu. Therefore, it is essential to eliminate residual sulfides to get
Oxygen fugacity high Cu contents in magmas for the formation of porphyry deposits. Sulfate (SO2− 4 ) is over 10 times more soluble
Adakite than sulfide (S2−), and thus the solubility of sulfur depends strongly on sulfur speciation, which in turn depends
Slab melts on oxygen fugacities. The magic number of oxygen fugacity is log fO2 N FMQ + 2 (i.e., ΔFMQ + 2), where FMQ is
Arc magmas the fayalite–magnetite–quartz oxygen buffer. Most of the sulfur in magmas is present as sulfate at oxygen fugac-
Plate subduction
ities higher than ΔFMQ + 2. Correspondingly, the solubility of sulfur increases from ~1000 ppm up to N 1 wt.%.
Ridge subduction
Oxidation promotes the destruction of sulfides in the magma source, and thereby increases initial chalcophile
element concentrations, forming sulfur-undersaturated magmas that can further assimilate sulfides during as-
cent. Copper, Mo and Au act as incompatible elements in sulfide undersaturated magmas, leading to high
chalcophile element concentrations in evolved magmas. The final porphyry mineralization is controlled by
sulfate reduction, which is usually initiated by magnetite crystallization, accompanied by decreasing pH and
correspondingly increasing oxidation potential of sulfate. Hematite forms once sulfate reduction lowers the pH
sufficiently, driving the oxidation potential of sulfate up to the hematite–magnetite oxygen fugacity (HM) buffer,
which is ~ΔFMQ + 4. Given that ferrous iron is the most important reductant that is responsible for sulfate re-
duction during porphyry mineralization, the highest oxygen fugacity favorable for porphyry mineralization is
the HM buffer. In addition to the oxidation of ferrous iron during the crystallization of magnetite and hematite,
reducing wallrock may also contribute to sulfate reduction and mineralization. Nevertheless, porphyry deposits
are usually mineralized in the whole upper portion of the pluton, whereas interactions with country rocks are
generally restricted at the interface, therefore assimilation of reducing sediments is not likely to be a decisive con-
trolling process. Degassing of oxidized gases has also been proposed as a major process that is responsible for sul-
fate reduction. Degassing, however, is not likely to be a main process in porphyry mineralization that occurs at
2–4 km depths in the upper crust. Sulfide formed during sulfate reduction is efficiently scavenged by aqueous
fluids, which transports metals to shallower depths, i.e., the top of the porphyry and superjacent wallrock. Ac-
cording to traditional views, sulfide saturation and segregation during magma evolution is not favorable for
the formation of porphyry Cu ± Au ± Mo deposits. This is the main difference between porphyry deposits and
Ni–Cu sulfide deposits. Nevertheless, in places with thick sections of reducing sediments, e.g., the western
North America, sulfide saturation and segregation may occur during evolution of the magma, forming Cu-rich cu-
mulates at the base of plutons. These Cu-rich sulfides may evolve into porphyry mineralization or even control
the ore-forming process. Their contribution depends heavily on subsequent oxidation, i.e., a major contribution
can be expected only when the sulfide cumulates are oxidized to sulfate, liberating the chalcophile elements. Sul-
fate reduction and ferrous Fe oxidation form H+, which dramatically lowers the pH values of ore-forming fluids
and causes pervasive alteration zones in porphyry Cu deposits. The amount of H+ released during mineralization
and the alkali content in the porphyry together control the intensity of alterations. In principle, H2 and methane

⁎ Corresponding author.
E-mail address: weidongsun@gig.ac.cn (W. Sun).

http://dx.doi.org/10.1016/j.oregeorev.2014.09.004
0169-1368/© 2014 Published by Elsevier B.V.
98 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

form during the final mineralization process of porphyry deposits, but are mostly oxidized by ferric Fe during
subsequent processes. Some of the reduced gases, however, may survive the highly oxidizing environment to es-
cape from the system, or even to get trapped in fluid inclusions. Therefore, small amount of reduced gases in fluid
inclusions cannot argue against the oxidized feature of the magmas. Reduced magmas are not favorable for por-
phyry mineralization. Reduced porphyry deposits so far reported are just mineralization that has either been re-
duced in host rock away from the causative porphyry or through assimilation of reducing components during
emplacement.
© 2014 Published by Elsevier B.V.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2. Brief introduction of major oxygen buffers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
2.1. Fayalite–magnetite–quartz (FMQ) oxygen buffer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.2. Hematite–magnetite (HM) oxygen buffer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.3. Ni–NiO(NNO) oxygen buffer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
2.4. Pyrite + pyrrhotite + magnetite (PPM) oxygen buffer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102
3. The association of porphyry deposits with oxidized magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.1. Large porphyry deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
3.1.1. Porphyry Cu and Au deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
3.1.2. Porphyry Cu–Mo deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
3.1.3. Porphyry Mo deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.2. Linkage between oxidized magmas and porphyry deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.2.1. Sulfur oxidation, sulfide under saturation and residual sulfide . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
3.2.2. Sulfate reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
3.2.3. Hematite–magnetite intergrowth . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
3.3. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
4. The association of porphyry deposits with reduced magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
4.1. Reduced magmas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113
4.1.1. Evidence for reduced magmas of Catface porphyry deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
4.1.2. Evidence for reduced magma in the Baogutu porphyry deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.2. Source of copper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115
4.3. Formation of reduced porphyry deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
4.3.1. The formation of the Catface porphyry deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
4.3.2. The formation of the Baogutu porphyry deposit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
4.4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5. Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.1. The oxygen fugacities at convergent margins . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.2. The difference between porphyry and epithermal in terms of oxygen fugacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.2.1. Magnetite crisis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
5.2.2. Oxygen fugacity and open systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
5.3. Adakite, slab melting, ridge subduction and porphyry Cu deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.3.1. Adakite and porphyry Cu deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
5.3.2. Ridge subduction and porphyry Cu deposits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
5.4. Alterations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
6. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127

1. Introduction The consensus is that most of the porphyry Cu ± Mo ± Au systems


are initiated by injection of oxidized adakitic magma saturated with
Porphyry deposits are hosts to one of the most important economic aqueous fluids that are S- and metal-rich, i.e., the parental magmas
mineral associations (Cooke et al., 2005; Halter et al., 2005; Heinrich must be water rich and oxidized (e.g., Ballard et al., 2002; Burnham
et al., 2004; Mutschler et al., 2010; Sillitoe, 2010), accounting for ~80% and Ohmoto, 1980; Garrido et al., 2002; Imai, 2002; Liang et al., 2006;
Cu and ~95% Mo of the world's total reserves. It is also an important re- Mungall, 2002; Sillitoe, 2010; Stern et al., 2007; Sun et al., 2013b). It is,
source of Au, Ag, Zn, Sn and W. Most porphyry deposits are found above however, still controversial as regards to: why high oxygen fugacity is
active subduction zones (Fig. 1) (e.g., Chiaradia, 2014; Chiaradia et al., favorable for the mineralization of porphyry deposits, how oxidized
2012; Gonzalez-Partida et al., 2003; Hedenquist et al., 1998; Kesler, the magma could be, whether adakitic magma is essential for porphyry
1997; Lee, 2014; Richards, 1999, 2013; Sillitoe, 2010; Sun et al., 2011; mineralization or whether the porphyry deposits can be associated with
Wilkinson, 2013), with a few occurrences at post-collisional or other normal arc rocks (Fig. 2), and why the pure porphyry Mo deposits are
tectonic settings (Sillitoe, 2010), e.g., porphyry Mo deposits in the east- also closely associated with highly oxidized magmas.
ern Qinling orogenic belt (Chen, 2013; Li et al., 2012a; N. Li et al., 2013) Copper, Au and Mo are chalcophile elements, which are strongly
and, arguably porphyry Cu–Mo deposits in Gangdese belt on the south controlled by the behavior and speciation of sulfur. Therefore, the less
Tibetan Plateau (Hou et al., 2009; Qu et al., 2004; Xiao et al., 2012) the quantity of residual sulfide, the higher the initial Cu contents in pri-
and some porphyry Cu deposits in Iran (Calagari, 2003; Castillo, 2006; mary magmas (Fig. 3) (Lee et al., 2012; Sun et al., 2004a, 2013b). Exper-
Haschke et al., 2010; Shafiei et al., 2009). iments show that sulfate is much more soluble than sulfide in magmas
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 99

Fig. 1. Worldwide distribution of porphyry Cu deposits. Note, most of the porphyry deposits are distributed along convergent margins. Porphyry Mo deposits are not shown.
Modified after Sun et al. (2013b). Data sources: Mutschler et al. (2010).

(Beermann et al., 2011; Jugo, 2009). Therefore, considerably more sulfur et al., 2003a,b), which leads to Cu depletion in the mantle wedge.
is removed in the form of sulfate under higher oxygen fugacity (Lee Also, the moderate mobility of Cu during plate subduction (J.L. Li et al.,
et al., 2012; Liang et al., 2009; Sun et al., 2013b). Meanwhile, sulfide is 2013) can compensate for the depletion caused by previous melting.
kept undersaturated during the evolution of oxidized magmas, such The implication thus is that normal arc rocks, i.e., peridotitic melts, are
that no sulfide segregation is expected (Sun et al., 2012a; Sun et al., not favorable for porphyry Cu mineralization.
2013b). Therefore, Cu, Mo and Au act as moderately incompatible ele- Still other authors have argued that, instead of high oxidation, sul-
ments (McDonough and Sun, 1995; Sun et al., 2003a,b,c), which become fide saturation of the magma and consequent pre-enrichment through
enriched during the early stage of magma evolution, and suddenly go sulfide accumulation are the most important step for porphyry ore de-
into magmatic fluids during magnetite crystallization (Sun et al., posits (Chiaradia, 2014; Lee, 2014; Wilkinson, 2013). This is probably
2004a, 2013a). It has been proposed that oxygen fugacity of log true along the western margin of the North American continent and
fO2 N FMQ + 2 (i.e., ΔFMQ + 2, where FMQ is the fayalite–magnetite– other places where reduced sediments are well developed (see more
quartz oxygen buffer) is the magic number for porphyry mineralization in Section 4). However, this hypothesis raises several other questions.
(e.g., Mungall, 2002; Sun et al., 2013b). Sulfate is the dominant species Why are porphyry deposits usually associated with oxidized magmas,
at log fO2 N FMQ + 2, which is much more soluble than sulfide in which is a situation so different from Ni–Cu sulfide deposits that clearly
magmas (Jugo, 2009). Therefore, residual sulfides are more efficiently experienced sulfide saturation? How does porphyry remain oxidized
destroyed at oxygen fugacities higher than FMQ + 2, releasing their after collecting sulfide from the pre-enriched sulfide accumulates?
chalcophile elements (Sun et al., 2013b). Others have argued that SO2 Why are the grades of porphyry deposits much lower than sulfide satu-
is the main sulfur species dissolved in porphyry magmas (Richards, rated Ni–Cu sulfide deposits? Why is there no Ni in porphyry deposits?
2014; Smith et al., 2012), not sulfate nor sulfide. It was further argued Moreover, some porphyry deposits are apparently associated with
that variations in oxidation state over typical ranges for arc magmas reduced magmas at ΔFMQ − 0.5 to − 3 (Cao et al., 2014; Rowins,
(ΔFMQ = 0 to +2) have no major effects on the potential of magmas 2000; Smith et al., 2012). Although reduced porphyry deposits are
to form porphyry Cu–(Mo) deposits during plate subduction rare, with tonnages much smaller than oxidizing porphyry deposits
(Richards, 2011b). The implication is that ΔFMQ + 2 is not of impor- (Cao et al., 2014), the mechanism needs to be clarified.
tance for porphyry Cu mineralization. Instead, water is the most impor- This contribution focuses on the controlling factors of porphyry min-
tant factor that controls porphyry mineralization (Richards, 2011a). The eralization. Major controversies to be discussed here are: (1) Why do
question then becomes that most arc plutons are water saturated and most of the porphyry deposits associated with oxidized magmas occur
highly oxidized, why do only a very small fraction of special magmas at convergent margins? (2) What is the most favorable oxygen fugacity
(mostly adakitic characterized by high Sr/Y and high Sr) in arc settings range for porphyry mineralization? (3) What is the genetic connection
form porphyry deposits. For example, there are essentially no porphyry between high oxygen fugacity magmas and porphyry deposits?
deposits in Japan (Fig. 1). Note that Japan arc is associated with older (4) Why are some porphyry deposits associated with reducing
and presumably wetter subducting plate, such that arc rocks are pre- magmas? (5) What are the connections between porphyry deposits
sumably wetter than subduction related rocks along the eastern margin and reduced magmas, if there is any?
of the Pacific Ocean (Sun et al., 2012a,b, 2013a,b).
A recent study has suggested that the source region for arc magmas 2. Brief introduction of major oxygen buffers
is probably neither unusually oxidized nor enriched in economic ele-
ments of interest, such as Cu, as has been shown by studies of primitive Oxygen fugacity (fO2) is an important geological parameter affecting
arc magmas (Lee et al., 2010, 2012). This is seemingly consistent with the stability of minerals, the evolution of magmas, and ore-forming pro-
the moderate incompatibility of Cu during mantle magmatism (Sun cesses. It is an equivalent of the partial pressure of oxygen in a particular
100 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Fig. 2. Two different models for porphyry Cu ± Au ± Mo deposits. A. Porphyry deposits are formed in normal arc rocks (after Richards, 2011a). According to this model, even the formation
of giant porphyry deposits is nothing special but optimization of normal ore-forming processes, controlled by distinct tectonic configurations, reactive host rocks, or focused fluid flow that
have helped to enhance the overall process (Richards, 2013). B. Porphyry deposits are associated with slab melts (modified after Wilkinson, 2013), which have high initial Cu contents (Sun
et al., 2011).

environment (atmosphere, magmas, rocks, etc.) corrected for the non- pressure (Fig. 4), i.e., the oxygen fugacity of equilibration at a fixed pres-
ideal character of the gas. Oxygen fugacity is very important to the be- sure is defined by one of the curves in oxygen fugacity versus tempera-
haviors of elements, but it is not a very precise term because in some ture diagrams. The concept of oxygen fugacity and oxygen fugacity
cases, oxidation–reduction reactions do not involve any oxygen and buffer have been well developed and widely used in high pressure
the oxygen fugacity changes with pressure and temperature (Sun experiments.
et al., 2014a,b). “Redox state” is a better term to describe the relative Oxygen fugacity is of critical importance for the behaviors of ele-
proportions of an element among its different oxidation states. Never- ments. Therefore, it is crucial to know the oxygen fugacity of a geochem-
theless, oxygen fugacity is popularly used in Earth sciences. ical process and to control oxygen fugacity during experiments. There
Oxygen fugacity is usually notated as variations relative to a certain are several ways to control oxygen fugacity of experiments in the labo-
oxygen buffer. Oxygen buffer refers to an assemblage of minerals or com- ratory. Gas-mixing techniques have been commonly used for control-
pounds that constrains oxygen fugacity as a function of temperature and ling oxygen fugacity of experiments conducted at high temperature
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 101

Fig. 3. Copper contents during partial melting of mantle peridotite under different oxygen fugacities modeled by Lee et al. (2012). (A) Variation of aggregated melt and residual mantle
composition with degree of partial melting (F) at 2GPa, 1350 °C and fO2 at FMQ + 0, assuming initial S content of 200 ppm, i.e. 0.06 wt.% of sulfide. (B) Copper contents in primary
melt as a function of F and log fO2 (ΔFMQ); same P–T conditions as in (A). (C) Cu content of aggregate liquids versus fO2 for different F. Gray field refers to Cu in primitive MORB and
arc magmas (Lee et al., 2012).

(above 1300 K) and 1 atm pressure (Huebner, 1987). The gas mixtures
mostly used are CO2–CO and CO2–H2. At higher pressure, oxygen fugac-
ity can be controlled by the oxygen buffer technique (Eugster, 1957)
and the Shaw membrane (Shaw, 1963), and measured with the hydro-
gen fugacity sensor technique (Chou, 1978). This oxygen buffer tech-
nique was first developed by Eugster in 1957 to prevent oxidation of
iron during growth of hydrous ferrous silicates and then later to deter-
mine the stability of annite (Eugster, 1957; Eugster and Wones, 1962;
Wones and Eugster, 1965). The oxygen buffers commonly used in
high pressure and high temperature experiments are fayalite–
magnetite–quartz (FMQ), Co–CoO, Ni–NiO (NNO), hematite–magnetite
(HM), and MnO–Mn3O4. These buffers are also useful as notations in the
measurement of the oxygen fugacity of natural rocks and magmas. A
comparison of the log fO2–T relationship for these buffers is shown in
Fig. 4. In experiments using the conventional double-capsule method,
the mixture of buffer materials and water is loaded in the outer capsule
and the sample is sealed in the inner capsule. Hydrogen gas forms
during the reaction between buffer materials and water in the outer cap-
sule, then diffuses through the inner capsule wall and equilibrates with
Fig. 4. Comparison of log fO2–T relationships for the Ni–NiO, Co–CoO, FMQ, MnO–Mn3O4,
and HM oxygen buffers at 1 bar.
the hydrogen in the inner capsule. Capsules with high hydrogen diffusiv-
The log fO2 of Co–CoO, FMQ and HM are taken from Chou (1978). The log fO2 of Ni–NiO ity are usually taken as the inner capsule (e.g., Pt), whereas that with low
and MnO–Mn3O4 are taken from Huebner and Sato (1970). hydrogen diffusivity is used as the outer capsule (e.g., Ag). The
102 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

prerequisite for using the double-capsule technique is that the buffer


materials themselves do not diffuse into the inner capsule. Therefore,
they cannot be used for system with haplogranitic melts and H2SO4
solutions because Co and Ni diffuse into the inner capsule, resulting in
formation of nickel sulfide or cobalt sulfide (Keppler, 2010).
The other way to regulate the oxygen fugacity of sample assem-
blages is to control the hydrogen fugacity using the Shaw membrane,
which introduces variable amount of hydrogen gas into the system
(Shaw, 1963). The advantage of this technique is that the hydrogen fu-
gacity can be varied independently and continuously. Moreover, the
Ag–AgCl–H2O–HCl acid buffer can be used as a fH2 sensor by loading
the acid-buffer assemblage in a small Pt capsule and putting the capsule
in a larger capsule containing the sample assemblages (Chou, 1978). In
this case, the measured concentration of HCl in the sensor indicates the
fH2 of the outer sample system.

2.1. Fayalite–magnetite–quartz (FMQ) oxygen buffer

The FMQ oxygen buffer is widely used in studies of natural samples,


as well as for experiments. Given that the oxygen fugacity of most igne-
ous rocks plot within several log units of the FMQ buffer, the oxygen fu-
gacities of natural samples are often noted as log unit variations from
the FMQ buffer, i.e., in ΔFMQ units. For example, ΔFMQ + 2 means ox-
ygen fugacity 2 log units higher than the values defined by the FMQ
buffer. The FMQ equilibrium has been determined experimentally by
many researchers, e.g., Chou (1978), Myers and Eugster (1983) and
Eugster and Wones (1962). The oxygen fugacity is controlled through
reaction (1):

3Fe2 SiO4 þ O2 ¼ 3SiO2 þ 2Fe3 O4 : ð1Þ

As shown in Fig. 5, the pressure dependence of log fO2 for this buffer
is not significant, but log fH2 changes greatly from 1 bar to 10 kbar.

2.2. Hematite–magnetite (HM) oxygen buffer Fig. 5. Pressure dependence of log fO2 and log fH2 of FMQ oxygen buffer. The log fO2 of Ni–
NiO buffer is taken from Chou (1978). Method for calibrating log fH2 of FMQ buffer is the
same as that of the HM buffer.
The HM (hematite + magnetite) equilibrium has been determined
by several researchers (e.g., Chou, 1978; Eugster and Wones, 1962;
Hemingway, 1990; Myers and Eugster, 1983). Oxygen fugacity is con- Ni þ H2 O ¼ NiO þ H2 : ð5Þ
trolled through reaction (2) or (3):
As shown in Fig. 7, the pressure dependence of log fO2 is not signif-
4Fe3 O4 þ O2 ¼ 6Fe2 O3 ð2Þ icant, but log fH2 increases rapidly at low pressures but less at higher
pressures. The Ni–NiO buffer is commonly used in experiments, bring-
2Fe3 O4 þ H2 O ¼ 3Fe2 O3 þ H2 : ð3Þ ing with it distinct advantages. First, the oxygen fugacity of Ni–NiO is
quite close to the intrinsic oxygen fugacity of the hydrothermal vessel
Eq. (3) better describes the system with water present. As shown in (NiNiO + 0.5). Thus, even quite small amounts of Ni and NiO powder
Fig. 6, the pressure dependence of log fO2 is quite small, while log fH2 in- could last for a long time during experiments, e.g., around 200 mg of
creases rapidly at low pressures but less at higher pressures. For high Ni and NiO is enough for around a 10-day experiment. Moreover,
pressure experiments, the HM buffer is more oxidized than the intrinsic oxygen fugacity controlled by the Ni–NiO buffer is quite important
oxygen fugacity of the hydrothermal vessel. Thus, hematite was con- for the oxidation states of many elements, e.g., sulfur in the fluids is
sumed quite fast, e.g., in some experiments around 3.73 mg Fe2O3 was mostly as H2S at the oxygen fugacity of NNO, but SO2 and SO3 appear
consumed in the buffer assemblage every hour under steady-state con- when the oxygen fugacity increases by 0.5 log unit (Binder and
ditions, and 100 mg of Fe2O3 lasted only 26.8 h (Chou, 1986). Therefore, Keppler, 2011).
the HM buffer is not suitable for experiments conducted in hydrother-
mal vessels with water as the pressure medium. In natural systems, 2.4. Pyrite + pyrrhotite + magnetite (PPM) oxygen buffer
the oxygen fugacity of the HM buffer is usually much higher than that
of magmas. Most of the porphyry deposits, however, reach the HM The PPM (pyrite + pyrrhotite + magnetite) equilibrium controls
buffer value during mineralization. f S2 and fO2 through reactions (6)–(8):

2.3. Ni–NiO(NNO) oxygen buffer 2Fe1−x S þ ð1−2xÞS2 ¼ 2ð1−xÞFeS2 ð6Þ

The Ni–NiO oxygen buffer has been experimentally determined by


Huebner and Sato (1970) and O'Neil and Pownceby (1993). The oxygen 6Fe1−x S þ 4O2 ¼ 2ð1−xÞFe3 O4 þ 3S2 ð7Þ
fugacity is controlled through reactions (4) and (5):

Ni þ 1=2O2 ¼ NiO ð4Þ 3FeS2 þ 2O2 ¼ Fe3 O4 þ 3S2 : ð8Þ


W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 103

Fig. 6. Pressure dependence of log fO2 and log fH2 for the HM buffer. The log fO2 of Fig. 7. Pressure dependence of log fO2 and log fH2 (Ni–NiO oxygen buffer). The logfO2 of
HM buffer is taken from Chou (1978). The log fH2 is calculated using log K = log f H2 + Ni–NiO buffer is taken from Huebner and Sato (1970). Method for calibrating log f H2 of
1/2log fO2, where K is the equilibrium constant of H2O = H2 + 1/2O2 determined by Ni–NiO buffer is the same as that of HM buffer.
Supcrt92 with the database DPRONS2003.

almost always attributable to the injection of an oxidized adakitic


This buffer has been used in hydrothermal experiments, e.g., Spry magma that is saturated with S- and metal-rich aqueous fluids,
and Scott (1986) and Crerar et al. (1978). The fO2 of the PPM buffer i.e., the parental magmas are water rich and oxidized (Ballard et al.,
has been determined by Kishima (1989) and Shi (1992), and their re- 2002; Mungall, 2002; Sillitoe, 2010). Several small “reducing porphyry
sults are in general agreement. Fig. 8. shows a comparison of the fO2–T deposits” constitute the rare exceptions (Cao et al., 2014; Rowins,
relationship of the PPM buffer with those of the NNO and HM oxygen 2000; Sillitoe, 1999) (see detailed discussion in Section 4 below). Inter-
buffers. The log fO2 of the PPM oxygen buffer is located between those growths of magnetite and hematite indicate that oxygen fugacities of
of NNO and HM buffers, e.g., at 700 °C, the log fO2 of PPM oxygen buffer porphyry deposits often reach the values defined by the HM buffer
is around 1 log10 unit lower than that of the HM buffer, but 3.3 log10 unit (Fig. 10). In this section, we review the oxygen fugacities of a number
higher than that of the NNO buffer. As shown in Fig. 8b, log fO2 decreases of famous deposits and then discuss the genetic connections between
slightly with increasing pressure (Kishima, 1989; Shi, 1992). oxidized magmas and porphyry deposits. The main questions discussed
At high temperature, the assemblage PPM is replaced by magnetite in this section include that: whether oxygen fugacity is a controlling
and pyrrhotite (MPo) because pyrite is not stable (Sun et al., 2013a,b; factor that dictates the distribution of porphyry deposits. What is the
Tomkins, 2010). Consequently, SO2 becomes the dominant species, genetic link between high oxygen fugacity magmas and porphyry
and fH2, fH2O, and fH2S decrease abruptly. The breakdown temperature deposits? What is the most favorable oxygen fugacity range for porphyry
of pyrite increases with increasing pressures, e.g., at 2 kbar, the break- mineralization? And how do oxygen fugacities change during porphyry
down temperature is 745 °C, and increases to around 800 °C to mineralization?
10 kbar (Shi, 1992).
3.1. Large porphyry deposits

3. The association of porphyry deposits with oxidized magmas The close association between highly oxidized magmas and porphy-
ry deposits may best be illustrated by a FeO versus fO2 diagram (Fig. 9).
It has long been proposed that most porphyry deposits are closely The highly oxidized nature of porphyry magmas have been reported in
associated with oxidized magmas (Burnham and Ohmoto, 1980; essentially all types of porphyries (Mungall, 2002; Sillitoe, 2010; Sun
Candela, 1992; Hedenquist and Lowenstern, 1994; Liang et al., 2006; et al., 2013b; Vila et al., 1991), including porphyry Cu and Cu–Au
Liang et al., 2009; Mungall, 2002; Sillitoe, 2010; Sun et al., 2012a, deposits (Sillitoe, 2010), porphyry Au deposits (Vila and Sillitoe, 1991;
2013b), also known as the magnetite-series magmas (Ishihara and Vila et al., 1991), to porphyry Cu–Mo deposits (Cuadra and Rojas,
Terashima, 1989) (Fig. 9). These porphyry Cu–(Mo)–(Au) deposits are 2001; Lynch and Ortega, 1997; Stern et al., 2007) and pure porphyry
104 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Ohno, 2005; Imai et al., 1993; Sun et al., 2013b), and from active subduc-
tion zones to post-collisional zones (Hou et al., 2004, 2007b, 2009; Li
et al., 2012a; Xiao et al., 2012).
One common indicator of the high oxygen fugacity of porphyry de-
posits is sulfate, i.e., both magmatic and hydrothermal sulfate,
e.g., hydrothermal anhydrite and magmatic anhydrite and gypsum are
abundant in essentially all large porphyry deposits (Cooke et al., 2011;
Zhang, Ling et al., 2013; Halter et al., 2005; Imai et al., 2007; Kavalieris
et al., 2011; Li et al., 2008; Liang et al., 2009; Stern et al., 2007; Vila
et al., 1991). In addition to sulfate, hypogene hematite and specularite
have also been reported in many porphyry deposits (Baker et al.,
1997; Hedenquist et al., 1998; Imai, 2001; Imai et al., 2007; Li et al.,
2008; Seedorff and Einaudi, 2004b; Sillitoe, 2010; Spry et al., 1996;
Vila and Sillitoe, 1991; Vila et al., 1991). The hematite–magnetite
(HM) oxygen buffer has been taken as the upper limit of oxygen fugac-
ities that are favorable for porphyry mineralization (Sun et al., 2013b).
Previous authors also have proposed that the oxygen fugacities of differ-
ent porphyry deposits are slightly different, in the order of: porphyry Cu
and porphyry Au deposits N porphyry Cu–Mo deposits N porphyry Mo
deposits (Fig. 9) (Thompson et al., 1999).

3.1.1. Porphyry Cu and Au deposits


Porphyry Cu and Au deposits almost do have systematically higher
oxygen fugacities and FeO contents compared to other porphyry de-
posits (i.e., Cu–Mo and Mo). The best examples are the Cenozoic por-
phyry Cu and Au deposits occurring in the southwestern Pacific
islands, and to a less extent, the Paleozoic porphyry Cu–Au deposits in
the Central Asian Orogenic Belt.
From this one could surmise that porphyry Cu–Au deposits are asso-
ciated with island arcs, whereas porphyry Cu–Mo deposits are associat-
ed with continental arcs. For example, Cenozoic porphyry deposits
located in island arcs in the southwestern Pacific are all Cu–Au deposits,
e.g., Grasberg and Batu Hijau in Indonesia; Panguna and Ok Tedi in
Fig. 8. Comparison of the log fO2 of PPM (pyrite–pyrrhotite–magnetite) buffer with those Papua New Guinea; Lepanto–Far South East, Tampakan, Atlas and
of Ni–NiO and hematite–magnetite buffer (A) and pressure dependence of the logfO2 of Sipilay in the Philippines (Cooke et al., 2005). Some researchers have
PPM buffer. even argued that Cu comes from the mantle, whereas Mo comes from
The log fO2 of PPM buffer is calibrated from Shi (1992). the continental crust (Mao et al., 2011). This is not the case, however, al-
though porphyry Cu (±Au and ±Mo) deposits indeed do have higher
Mo deposits (Seedorff and Einaudi, 2004a,b). They are found in different
εNd isotopic values (Hou et al., 2007a), implying less contribution from
tectonic settings, ranging from continental arc (Sillitoe, 2010; Stern
the continental crust. This observation, however, does not necessarily
et al., 2007, 2011) to island arc (Hedenquist et al., 1998; Imai and
support a mantle origin for Cu, either, because both the depleted mantle
and the oceanic crust have high εNd values, whereas enriched mantle
and the continental crust have low εNd values. As argued below
(Section 3.2.1), a strong case can be made that most of the Cu comes
from subducted oceanic crust. The distribution of porphyry deposits
does not support the premise that the Cu comes from the mantle.
Moreover, porphyry Cu–Au deposits are not always formed in island
arc environments, either. For example, there are also many porphyry
Cu–Au deposits (without economic levels of Mo) in the western
American continents, including some of the world's top 25 largest por-
phyry deposits, e.g., Cerro Casale in Chile; Minas Conga in Peru; Bajo
de la Alumbrera in Argentina; Pebble Copper in the USA; and Prosperity
in Canada. Here we give brief introductions to three famous porphyry
Cu–Au deposits.

3.1.1.1. Grasberg, Indonesia. Grasberg in Irian Jaya, Indonesia is one of the


largest high-grade hypogene porphyry Cu–Au deposit (38.32 Mt Cu @
1.12% and 3662 t Au @ 1.07 g/t) in the world, ranking no. 1 among por-
phyry Au deposits and among the top 10 porphyry Cu deposits (Cooke
et al., 2005). It was formed ~3 Ma ago. The presence of anhydrite and
Fig. 9. Schematic plot of Fe content in magmas versus oxidation state (fO2) for calc-alkaline hematite indicates very high oxygen fugacities (Cooke et al., 2005;
to alkaline magmas associated with porphyry Cu, Cu–Mo and Mo deposits and W, Sn de- Mathur et al., 2000), near the HM buffer. This deposit is related to the
posits. The approximate boundary between magnetite- and ilmenite-series magmas
(Ishihara, 1977) is also shown. Nevertheless, the HM buffer is the upper limit of oxygen fu-
closure of a small backarc basin between Indonesia and the South
gacity favorable for porphyry deposits. China Sea. As is characteristic of Cenozoic porphyry deposits in south-
Modified after Thompson et al. (1999). western Pacific, it contains no Mo.
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 105

Fig. 10. Images of magnetite–hematite intergrowths from major porphyry deposits in China, indicating that the oxygen fugacity reached the HM buffer (Zhang, Ling et al., 2013). A. Dexing
(Zhang, Ling et al., 2013), B. Xiongcun, C. Duobuza, D. Yulong (Sun et al., 2013a,b), E. Qulong, F. Zijinshan (Sun et al., 2013a,b).

3.1.1.2 . Far Southeast–Lepanto, Philippines. The Far Southeast–Lepanto magnetite–titanohematite assemblage, which indicates oxygen fu-
porphyry and epithermal Cu–Au deposits, Philippines, are parts of a gacities near the HM buffer at nearly magmatic temperature (Imai,
large epithermal–porphyry Cu–Au body with a total of 5.48 Mt Cu @ 2001).
0.8 wt.% and 973 t Au @ 1.42 g/t (Table 1; Cooke et al., 2005; In summary, essentially, most if not all, porphyry Cu–Au deposits are
Hedenquist et al., 1998). Sulfates, including anhydrite and alunite (likely highly oxidized (Figs. 9, 10). In addition to those mentioned above, he-
related to the epithermal mineralization) are abundant in the minerali- matite flakes in quartz veinlets have been reported in the Waisoi por-
zation system, and hematite formed during chlorite alteration, indicat- phyry Cu deposit (Namosi district), Viti Levu, Fiji (Imai et al., 2007).
ing high oxygen fugacity up to the HM buffer (Hedenquist et al., 1998). The Tongshankou porphyry skarn Cu–Au deposit, in the Lower Yangtze
River belt, central eastern China, also has primary hematite next to sul-
3.1.1.3. Santo Tomas II (Philex), Philippines. The Santo Tomas II (Philex) fides (Li et al., 2008). The euhedral characteristics of the hematite imply
porphyry deposit, Philippines, has total reserves of 1.2 Mt Cu @ a hydrothermal origin.
0.33 wt.% and 233 t Au @ 0.64 g/t (Table 1). It is also associated with Porphyry Au deposits have oxygen fugacities similar to, if not higher
a highly oxidized magma, as indicated by anhydrite veinlets and a than, porphyry Cu deposits. For example, Marte, a large porphyry Au

Table 1
Age, grade and tonnages of major porphyry deposits discussed in the text.

Deposit Province Ref Age Tonnage Au Au Cu Cu Mo Mo


grade grade grade

(Ma) (Mt) (g/t) (t) (wt.%) (Mt) (wt.%) (Mt)

Cu–Au Grasberg Irian Jaya Cooke et al. (2005), Mutschler et al. (2010) 4–3 3409 1.07 3662 1.12 38.3
Lepanto-Far N. Luzon Cooke et al. (2005), Mutschler et al. (2010) 1.5–1.2 685 1.42 973 0.8 5.48
South East
Santo Tomas II Philippines Mutschler et al. (2010) 1 364 0.64 233 0.33 1.2
Cu–Mo El Teniente Central Chile Cooke et al. (2005), Mutschler et al. (2010) 7.1–4.6 11,845 0.0035 437 0.92 109 0.02 2.50
Chuquicamata Nothern Chile Cooke et al. (2005), Mutschler et al. (2010) 33.6 15,052 0.04 301 0.71 106 0.024 1.81
Dexing China Mutschler et al. (2010), Zhang, Ling et al. (2013), 170–148 1500 0.18 19 0.45 8.4 0.01 0.29
Zhang, Sun et al. (2013)
Qulong China Xiao et al. (2012) 14 0.5 10.4 0.03 0.5
Yulong China Mutschler et al. (2010), Liang et al. (2006) Paleogene 850 0.84 7.14 0.03 0.15
Sar-Cheshmeh Iran Cooke et al. (2005), Mutschler et al. (2010) 12 1200 0.27 324 1.20 14.4 0.03 0.36
Batu Hijau Indonesia Mutschler et al. (2010) 5 1644 0.35 572 0.44 7.26
Panguna Bougainville Cooke et al. (2005) 3.5 1415 0.57 799 0.46 6.51
Ok Tedi PNG Mutschler et al. (2010) 1.2–1.1 700 0.64 446 0.64 4.48
Tampakan Philippines Cooke et al. (2005) 3.3–2.2 1400 0.24 336 0.55 7.70
Atlas Philippines Cooke et al. (2005) 61 1380 0.24 331 0.50 6.90
Mo Henderson United States Mutschler et al. (2010) 30–27 727 0.17 1.24
Reduced Catface Canada Mutschler et al. (2010) Middle Eocene 308 0.37 1.14 0.01 0.02
Reduced Baogutu China Cao et al. (2014) 0.1 14 0.28 0.63 0.011 0.018
106 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

deposit located in the Maricunga belt of the Andean Cordillera, northern involvement of such reducing agents as organic matter seemingly ex-
Chile, contains several percents of hematite, magnetite, anhydrite and plains the system's lower oxygen fugacity. The most famous porphyry
gypsum (Vila et al., 1991), clearly reaching the HM buffer (Figs. 10, 11). Cu–Mo deposits are the Cenozoic ones located along the eastern Pacific
It has also been argued that iron oxides–copper–gold (known as margin.
IOCG) deposits are the low S version of porphyry Cu–Au deposits, con-
trolled by secular changes in oceanic sulfate content and the geothermal 3.1.2.1. El Teniente. The supergiant El Teniente deposit in Chile is the
gradients at the end of the Precambrian (Richards and Mumin, 2013). largest porphyry Cu–Mo–Au deposit in the world, with total reserves
This is still elusive. Nevertheless, IOCG deposits are all highly oxidized, of ~ 109 Mt Cu @0.92%, 2.5 Mt Mo @ 0.02% and 437 t Au @0.035 g/t
up to the HM buffer. (Table 1; Cooke et al., 2005; Mutschler et al., 2010). The deposit spatially
corresponds to the Juan Fernandez Ridge (Cooke et al., 2005), and thus
has been attributed to slab melting during ridge subduction (Sun et al.,
3.1.2. Porphyry Cu–Mo deposits 2010). Others argued that ridge subduction is not responsible for mak-
In general, porphyry Cu–Mo deposits have FeO contents systemati- ing the supergiant porphyry deposit based on the inferred migration
cally lower than porphyry Cu, but higher than pure porphyry Mo de- history of the arc (Kay et al., 2005). El Teniente is a nested porphyry sys-
posits, with oxygen fugacities also falling between them (Fig. 8). Many tem that, according to different authors, was active for ~ 1 Ma (Baker
porphyry Cu–Mo deposits have high sulfate contents, with much less et al., 2011; Cannell et al., 2003) up to more than 7 Ma (14.2–6.5 Ma)
magnetite. No hypogene hematite has yet been reported in some of (Barra, 2011; Stern et al., 2011; Vry et al., 2010). High Fe2O3/FeO values
the supergiant porphyry Cu–Mo deposits, partially due to severe super- of 1 to 3 (Garrido et al., 2002), quartz anhydrite veins and anhydrite-
gene oxidation, which makes it difficult to identify primary hematite. cemented breccias as well as gypsum clearly indicate high oxygen fu-
Porphyry Cu–Mo deposits have Cu and Au grades comparable to gacities (Klemm et al., 2007; Vry et al., 2010). Although it is not spelled
those of porphyry Cu ± Au deposits, with low Mo grades, ranging out, the high Fe2O3/FeO ratios of 1 to 3 (Garrido et al., 2002) indicate
from 0.01 to 0.03 wt.% (Table 1) (Cooke et al., 2005). Some of the super- abundant hematite, reaching the HM buffer. Hydrothermal rutile also
giant deposits may have large Mo reserves of several million tons indicates high oxygen fugacity (NNO + 1.3) (Rabbia et al., 2009),
(Cooke et al., 2005). The Cu–Au and Mo mineralizations that occur in which is however much lower than the HM buffer, though. Neverthe-
the same ore bodies usually do not happen at the same time, indicating less, the oxygen fugacity undoubtedly fluctuated during porphyry min-
different sources and/or different mineralization processes. Consistent- eralization (Liang et al., 2009; Sun et al., 2013b), and the hydrothermal
ly, the Mo/Cu ratios of these deposits are generally more than 20 times rutile only records the oxygen fugacities under which it crystallized.
higher than the primitive mantle ratio value (McDonough and Sun, Most of the Cu in El Teniente was emplaced during the late magmat-
1995), suggesting that Mo has been added to the porphyry deposits ic stage (Cannell et al., 2005). A small sulfur- (N 3 wt.%) and copper-rich
from different sources. (N0.5 wt.%) fine-grained igneous rock known as “Porphyry A” stock
One of the most important geologic processes that enriches Mo is the (b1 km3, 6.09 ± 0.18 Ma) (Stern et al., 2011) has abundant igneous an-
oxidation–reduction cycle operating during chemical weathering at the hydrite, with varied textures ranging from interstitial to poikilitic and
Earth's surface, i.e., Mo is mobilized during oxidation and then enriched corresponding modal abundances from 10 to 20%, respectively. These
in organic-rich sediments due to reduction (Li et al., 2012a,b). The igneous anhydrite grains with planar crystal boundaries, occur along
with fresh and unaltered biotite, feldspars, quartz, and Fe-oxides
(Stern et al., 2007). Because the anhydrite-rich stock is isotopically sim-
ilar to all the other igneous rocks of the late Miocene deposit (Baker
et al., 2011; Stern et al., 2007) formed at a time of regional compressive
deformation, it has been proposed that an oxidized parent magma in the
large productive magma chamber was undergoing igneous fraction-
ation at that time. During a period of recharge by mantle-derived
mafic magmas into the base of the chamber and associated volatile
transfer and concentration near its roof, the opportunity arose to pro-
duce the Cu- and S-rich magmas that formed the anhydrite-bearing in-
trusive rocks (Stern et al., 2007).

3.1.2.2. Chuquicamata. Chuquicamata is another supergiant porphyry


deposit in Chile, with a total reserve of ~ 106 Mt Cu @ 0.71%, 1.81 Mt
Mo @ 0.024% and 301 t Au @ 0.04 g/t (Table 1. Cooke et al., 2005;
Mutschler et al., 2010). Similar to El Teniente, the formation of the
Chuquicamata porphyry Cu deposits also lasted for several million
years, between 36 and 31 Ma (Ballard et al., 2001; Ossandon et al.,
2001). It is closely associated with highly oxidized adakite (Oyarzun
et al., 2001, 2002). Because quantitative oxygen barometers based on
Fe–Ti oxides are prone to resetting, and primary whole rock Fe(III)/
Fe(II) ratios and anhydrite, if originally present, are unlikely to have
survived the hydrothermal alteration and surficial weathering of this
deposit, the zircon Ce4+/Ce3+ ratio has been used to indicate its high
Fig. 11. Stability domains of the trisulfur ion S−
3 , sulfate, and sulfide in an aqueous solution,
oxygen fugacity (Fig. 12) (Ballard et al., 2002).
as a function of oxygen fugacity (log10 fO2) and acidity (pH = −log10 mH+, in mol per kg)
at 350 °C and 0.5 GPa illustrating sulfate reduction (modified after Pokrovski and
Dubrovinsky, 2011; Sun et al., 2013a,b). Also shown is the oxygen fugacity of the major 3.1.2.3. Dexing. The Dexing porphyry Cu deposit is located in southeast-
mineral buffers (HM, thick horizontal orange line; NNO and FMQ, horizontal dashed ern China, with total reserves of 8.4 Mt Cu @ 0.45%, 0.29 Mt Mo @ 0.01%
lines) and the neutrality point of pure water (the vertical dashed line). The orange field be- and 19 t Au @ 0.18 g/t (Table 1. Zhang, Ling et al., 2013; Zhu et al., 1983).
tween HM and ΔFMQ +2 is the optimum condition for porphyry Cu, Au, Mo mineraliza-
tion (Sun et al., 2013b). Lines E1, E2, E3 and E4, show trajectories for sulfur reduction (Sun
The deposit consists of a cluster of three porphyries, with Tongchang as
et al., 2013b). Stability domains of trisulfur ion S− 3 at total dissolved sulfur concentrations
the largest in the center, Fujiawu the second in the southeast and
of 1 wt.% and 0.1 wt.% are shown. Zhushahong, a small deposit in the northwest (Li et al., 2007; X.F. Li
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 107

Eurasian continents, Qulong is taken as a typical post-collisional por-


phyry deposit. Nevertheless, the Indian plate is still subducting north-
ward. It is not clear whether the portion of the subducting slab
underneath Qulong was continental or oceanic. Intergrowths of hema-
tite and magnetite (Fig. 10) indicate the oxygen fugacity of Qulong
reached the HM buffer. Adakites found in the Qulong deposit are mix-
tures of continental and slab melts (Sun et al., 2012a).

3.1.2.5. Yulong. The Yulong porphyry Cu–Au deposit belt is distributed


along the northwestern extension of the Red River–Ailao Shan fault sys-
tem, at the eastern margin of the Tibetan Plateau (Hou et al., 2007b;
Jiang et al., 2006; Liang et al., 2006), and covers an area of ~ 300 km
long and ~ 20 km wide. Five major porphyries, which contain most of
the Cu reserves so far discovered in the belt, are located in a narrow,
elongated domain of approximately 50 km long and 10 km wide and
Fig. 12. Ce4+/Ce3+ and (Eu/Eu*)N ratios of zircon grains for Dexing and Shapinggou are closely associated with Cenozoic high potassium intrusive rocks
porphyry deposits as well as ore-bearing and ore-barren porphyries from Chile. High (Liang et al., 2006, 2007). The Yulong porphyry is the largest one in
Ce4+/Ce3+ signifies high oxygen fugacity. Data of ore-bearing and ore-barren samples in the Yulong copper deposit belt, with reserves of N7.14 Mt of Cu @
Chile from Ballard et al. (2002), Dexing from Zhang, Ling et al. (2013) and Shapinggou 0.84% and 0.15 Mt of Mo @ 0.028% (Table 1; Liang et al., 2006;
from Zhang, Li et al. (2013).
Mutschler et al., 2010). All five known porphyry deposits all together
contain a total of more than 8 million tons of Cu and Mo reserves. Zircon
U–Pb dating shows that the formation ages of deposits within the
et al., 2013; Zhou et al., 2013). The ore minerals comprise pyrite, chalco- Yulong ore belt range from 41.2 to 36.9 Ma, extending over a period of
pyrite, molybdenite, minor tetrahedrite, bornite and chalcocite. Gangue similar to 4.3 Ma, with formation ages decreasing systematically from
minerals include quartz, muscovite, chlorite, calcite, minor epidote and northwest to southeast (Liang et al., 2006). Zircon grains from the
anhydrite. The porphyry is an adakite formed at ~170 Ma (Zhang, Ling Yulong ore-bearing porphyries have higher Ce4 +/Ce3 + values than
et al., 2013; Wang et al., 2006a,b). Another poly-metal deposit, Yinshan, those from barren porphyries in the region (Liang et al., 2006). Abun-
formed roughly at the same time and may be paragenetically related dant magnetite and hematite suggest that the ore-bearing porphyries
(Wang et al., 2013). The Dexing porphyry deposit is closely associated are highly oxidized, reaching the HM buffer (Figs. 10, 11) (Sun et al.,
with highly oxidized magmas (Zhang, Ling et al., 2013; Li and Sasaki, 2013b).
2007), e.g., it is famous for abundant specular hematite, which mostly
formed during late stage hydrothermal alteration (Fig. 13). Inter-
growths of magnetite and hematite indicate its oxygen fugacity reached
the magnetite–hematite buffer (Fig. 11). Hematite is also found in fluid
inclusions in Dexing (Liu et al., 2011).

3.1.2.4. Qulong. The Qulong porphyry Cu–Mo deposit is now the largest
porphyry-type deposit in China, with reserves of 10.4 Mt Cu @ 0.5%
and 0.5 Mt Mo @ 0.03% (Table 1. Xiao et al., 2012). It is located in the
Gangdese orogenic belt in southern Tibet. In addition to abundant hy-
drothermal anhydrite of up to 10% or more, magmatic anhydrite is
also reported in unaltered granodiorite porphyry. These anhydrite oc-
currences indicate that the Qulong magmatic–hydrothermal system
was highly oxidized and sulfur-rich, with abundant sulfates (Xiao
et al., 2012; Yang et al., 2009) (Fig. 14). Given that it formed at
~ 14 Ma, which post-date the initial collision between Indian and

Fig. 13. An image of specularite from the Tongchang deposit in the Dexing porphyry Cu Fig. 14. Images of magmatic anhydrite from the Qulong giant porphyry deposit in the
deposits. Specularite cutting across carbonate veins, indicating that it was formed very Gangdese porphyry belt, south Tibetan Plateau. Bi = biotite; Anh = anhydrite; Kf =
late, likely through hydrothermal alteration. K-feldspar; Py = pyrite; Cpy = chalcopyrite; Ap = apatite.
108 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

3.1.2.6. Sar-Cheshmeh. The Sar-Cheshmeh is one of the top 20 largest very high oxygen fugacity, comparable to those of Chuquicamata
porphyry Cu–Au–Mo deposit in the world with reserves of 14.4 Mt Cu (Fig. 12) (Zhang, Li et al., 2013).
@ 1.2%, 0.36 Mt Mo @ 0.03%, and 324 t Au @ 0.27 g/t (Cooke et al.,
2005). It is located in southwestern Iran and is associated with several 3.2. Linkage between oxidized magmas and porphyry deposits
intrusive pulses of Miocene stocks (~12.2 Ma) (Mutschler et al., 2010),
ranging in composition from diorite through granodiorite to quartz– Copper, Au and Mo are all chalcophile elements, the behaviors of
monzonite (Hezarkhani, 2006). Molybdenum enrichment and deposi- which are mainly controlled by reduced sulfur, e.g., sulfide, hydrosulfide

tion took place before Cu. Anhydrite is a popular mineral identified in complexes (Sun et al., 2004a), or polysulfide (e.g., S2− 2 , S3 ) complexes
most veins (Hezarkhani, 2006). Hematite has been reported, but it oc- (Sun et al., 2013b). Although most geologists agree that porphyry de-
curs as a secondary supergene ore forming mineral (Shahabpour, posits are closely associated with oxidized magmas (Sillitoe, 2010;
1991). The Sar-Cheshmeh and several other giant Miocene porphyry Sun et al., 2013b), it is still hotly debated as regarding to why oxidized
deposits of Iran and Pakistan are located in a region undergoing conti- magmas favor porphyry mineralization and how oxidized the ore-
nent–continent collision. The geodynamic and architectural controls forming magmas should be. Major disagreements include: (1) What is
on porphyry formation in this complex tectonic zone are unclear the most favorable oxygen fugacity for porphyry mineralization?
(Cooke et al., 2005). The ore-forming magmas in this section of the Some workers proposed that ΔFMQ + 2 is a magic number for porphyry
Tethys margin have high Sr contents (N 500 ppm) and Sr/Y (N 50) mineralization (Mungall, 2002), whereas ΔFMQ + 2 to +4 is the most
(Haschke et al., 2010), which are clearly adakitic (Defant and favorable range of oxygen fugacity for porphyry deposits (Sun et al.,
Drummond, 1990). Previous studies suggested that these magmas are 2013b). Yet others claimed that variations in oxidation state over typical
a product of continental arc-style magmatism (Cooke et al., 2005). ranges for arc magmas formed during plate subduction (ΔFMQ = 0
They were attributed to either the attempted subduction of the Arabian to +2) have no major effects on the potential to form porphyry Cu ±
plate beneath the Eurasian plate, or the change from subduction of Au ± Mo deposits during plate subduction (Richards, 2011b). It is
oceanic to continental crust (Cooke et al., 2005), or partial melting of a even argued that some porphyry deposits formed in reduced magmas
fertile copper- and sulfur-enriched arc crustal keel (Haschke et al., (Cao et al., 2014; Rowins, 2000; Smith et al., 2012). (2) What is the
2010). Further studies are needed on the formation of Sar-Cheshmeh main sulfur species in porphyry? Some workers concluded that sulfate
and other Cenozoic porphyry deposits in the belt. is the predominant sulfur species in ore-forming porphyries (Cooke
et al., 2011; Field et al., 2005; Liang et al., 2009; Sotnikov et al., 2004;
3.1.3. Porphyry Mo deposits Sun et al., 2013a,b), whereas others argue that the main sulfur species
Porphyry Mo deposits generally have higher SiO2 and lower FeO is SO2 dissolved in porphyry magmas (Richards, 2014; Smith et al.,
contents. It also has been proposed that porphyry Mo deposits have sys- 2012), i.e., neither sulfate nor sulfide. (3) Is sulfide saturation during
tematically lower oxygen fugacities than porphyry Cu (Au) and Cu–Mo magma evolution important for porphyry mineralization? Some
deposits (Fig. 9). There are three major porphyry Mo deposit belts in the workers proposed that sulfide undersaturation is important for porphy-
world, the Henderson–Climax (Klemm et al., 2008; Pettke et al., 2010; ry mineralization, i.e., no residual sulfide remains in the source (Li et al.,
Seedorff and Einaudi, 2004a,b; Singer, 2008), Qinling–Dabie (Chen, 2012b; Liang et al., 2009; Sun et al., 2013b), whereas others argued that
2013; Chen et al., 2000; Li et al., 2012a; Mao et al., 2008; Stein et al., sulfide saturation and accumulation are of critical importance for por-
1997; Yang et al., 2013; Zeng et al., 2013; Zhang, Li et al., 2013) and phyry mineralization (Wilkinson, 2013).
Xing'an–Mongolia belts (Wu et al., 2011; Y. Zhang et al., 2013; Zeng In general, the porphyry mineralization process consists of 3 major
et al., 2013). The oxygen fugacity can also reach the HM buffer during phases: (1) source, i.e., extraction of chalcophile elements from the
sulfate reduction and mineralization (Fig. 10), likely because of the source; (2) transportation and concentration of ore-forming elements;
lower FeO contents. and (3) deposition and fixation of the ore forming elements into
orebodies. The relationship between oxidized magmas and porphyry
deposits during these three steps are discussed below.
3.1.3.1. Henderson. Henderson porphyry Mo deposit is located in the Cli-
max–Henderson Mo belt, Colorado. It consists of 12 Oligocene rhyolitic 3.2.1. Sulfur oxidation, sulfide under saturation and residual sulfide
stocks in three centers, Henderson (oldest), Seriate, and Vasquez Copper, Au and Mo are all chalcophile elements, with high partition
(deepest and youngest), at depths of ~ 1 km below the surface at the coefficients between sulfide and melts, e.g., DCu = 1334 ± 210 (Patten
Red Mountain (Seedorff and Einaudi, 2004a), with a reserve of et al., 2013), DAu = 4500 –11,200 (Mungall and Brenan, 2014), and
1.24 Mt Mo @ 0.17 wt.% (Table 1). Previous studies proposed that, in DMo = 0.15–5.15 (Li and Audetat, 2013), respectively. Some experi-
the Henderson porphyry Mo deposit, Fe was leached at high to moder- ments suggest that DMo increases with decreasing oxygen fugacities
ately high temperatures and then fixed in the rock at lower tempera- (Li and Audetat, 2013). Therefore, the behaviors of Cu and Au are
tures, first mainly as magnetite and then as pyrite and minor strongly controlled by sulfide, whereas Mo is much less sensitive to sul-
pyrrhotite and specular hematite (Seedorff and Einaudi, 2004a). Other fide, especially at high oxygen fugacities. During mantle melting, Cu as
porphyry deposits in the Climax–Henderson belt also contain hematite, well as Au and Mo are all moderately incompatible elements, with par-
indicating these deposits reached the HM buffer. tition coefficients ranging between those of Yb and Ce (McDonough and
Sun, 1995; Sun et al., 2003a,b, 2004a). The estimated Cu abundance in
3.1.3.2. Shapinggou. The Shapinggou porphyry Mo deposit is the largest the primitive mantle is 30 ppm (McDonough and Sun, 1995), whereas
Climax-type Mo deposit in the world, with total proven Mo reserves the Cu concentrations in MORB so far published range from 70 to
of over 2.2 million metric tons @ 0.17 wt.% (Zhang, Li et al., 2013). It is 150 ppm (Hofmann, 1988; Sun et al., 2003b). When there is no residual
located in the western Dabie Mountains, along the east extension of sulfide in the mantle source, Cu and also other chalcophile elements be-
the East Qinling Mo mineralization belt (Chen, 2013; Chen et al., 2013; come highly incompatible (Lee et al., 2012; Liu et al., 2014). Partial melt-
Gao et al., 2010; H.Y. Li et al., 2012; Han et al., 2013; N. Li et al., 2013; ing of mantle peridotite may form melts with high Cu contents up to
Stein et al., 1997; Yang et al., 2013; Zeng et al., 2013; Zhu et al., 2010), 350 ppm at high oxygen fugacities, at very low degree of partial melting
to the north of the Triassic suture between the north and south China (Fig. 3) (Lee et al., 2012). Nevertheless, most arc magmas formed at
blocks (Sun et al., 2002). Both Re–Os isochron and U–Pb zircon dating much N 10% partial melting, resulting in much lower Cu contents
give the same age of ~111 Ma (Zhang, Li et al., 2013), which assigns it (e.g., 100 ppm), which is not favorable than reducing magmas for por-
to the third mineralization pulse of the East Qinling Mo belt (Li et al., phyry mineralization. In contrast to mantle peridotite, MORB has sulfur
2012a; Mao et al., 2008, 2011). High zircon Ce4+/Ce3+ ratios indicate abundances of about 1000 ppm, the effects of oxygen fugacity on
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 109

residual sulfides, and, consequently, the Cu content in the melts, be-


comes much more obvious when considering partial melting of
subducted oceanic slabs.
The S abundance in the primitive mantle is estimated to be about
200–250 ppm (McDonough and Sun, 1995), and about 150 ppm in de-
pleted mantle (Lorand, 1990; Mavrogenes and O'Neill, 1999). Mainly,
the sulfur speciation in magmas is controlled by oxygen fugacity. Sulfide
is the predominant sulfur species at oxygen fugacities lower than the
FMQ buffer. Sulfate proportions start to climb up above the FMQ buffer.
Most of the sulfur in the magmas is present as sulfate at ΔFMQ + 2
(Fig. 11) (Jugo, 2009; Jugo et al., 2005, 2010). Experiments show that
sulfate is much more soluble than sulfide in magmas. The sulfur content
at sulfide saturation (SCSS) increases solely as a function of increasing
fO2 (Jugo, 2009):

h i
2–
SCSS ¼ S ð1 þ expð2:23ΔFMQ –2:89ÞÞ ð9Þ

and

 h i h i
6þ 2–
ΔFMQ C ¼ 1:29 þ 0:45 ln S – ln S ð10Þ

where ΔFMQC refers to the critical fO2 for simultaneous saturation of


sulfide and sulfate (Jugo, 2009).
The predicted SCSS content at sulfide saturation in basalts ranges
from 1300 ppm at ΔFMQ − 1 and 1500 ppm at ΔFMQ + 0.5, to
7500 ppm at ΔFMQ + 2 and 1.4 wt.% at ΔFMQ + 2.3 (Fig. 15) (Jugo,
2009). At oxygen fugacities of ΔFMQ 0 to + 2.5, the more oxidizing
the system, the higher the S contents in the magmas, and thus high ox-
ygen fugacity is the most efficient way to eliminate residual sulfides
(Lee et al., 2012; Sun et al., 2013b). The solubility of sulfide in magmas
is independent of oxygen fugacity below the FMQ buffer (Mavrogenes
and O'Neill, 1999), implying that most of the sulfur is removed in the
form of sulfate under high oxygen fugacity (Jugo, 2009), and with its re-
moval releasing much more of the chalcophile elements (Lee et al.,
2012; Liang et al., 2009; Sun et al., 2004b).
The average S concentration in MORB is ~ 1000 ppm (O'Neill and
Mavrogenes, 2002), while laboratory experimental measurements
have determined the S content at sulfide saturation in basalts to be Fig. 15. Sulfur speciation curve vs. fO2 for basaltic glasses based on the S6+/ΣS estimates of
1300 ppm at oxygen fugacities of ΔFMQ 0 (Jugo, 2009). The small differ- XANES (in bold), EPMA (dashed line) (Jugo, 2009; Jugo et al., 2005, 2010). Also shown are
fields of different tectonic settings, (A) Japan arc and Mexican arc; (B) MORB, OIB and
ence between the average MORB and experimental values is likely due
mantle wedge, and arc. MORB = mid-ocean ridge basalt; IAB = Island arc basalt;
to slight differences in oxygen fugacities and pressure effects, i.e., sulfide BABB = Back arc basin basalt; OIB = Oceanic island basalt.
is less soluble under higher pressures (Mavrogenes and O'Neill, 1999).
Given that most MORB forms through ~ 10% partial melting of the
depleted MORB mantle, it extracts about 100 ppm of sulfur from
the mantle source, leaving behind ~ 150 ppm of sulfur as residual sul-
fide. In case all the sulfur is present as sulfide, ~ 25% partial melting is Nevertheless, normal arc rocks are not favorable materials for por-
needed to eliminate all the residual sulfides from the mantle source phyry Cu mineralization. Arc magmas are probably neither unusually
at ΔFMQ 0 (Fig. 3) (Lee et al., 2012). Such large degrees of partial oxidized nor enriched in economic elements of interest, such as Cu
melting would dilute the released Cu and produce picrite or even (Lee et al., 2010; Lee et al., 2012). Instead, most porphyry Cu deposits
komatiite, not basalt. are closely associated with adakite (Oyarzun et al., 2001; Sajona and
For the mantle wedge above subducting slabs, the estimated S con- Maury, 1998; Sun et al., 2010, 2012a; Thieblemont et al., 1997), or so
tent at sulfide saturation ranges from 1500 ppm (at ΔFMQ + 0.4) to called high Sr/Y porphyries (Chiaradia et al., 2012; Richards, 2011a).
4500 ppm (at ΔFMQ + 1.7) (Jugo, 2009). As shown in Fig. 16, the Most of the ore forming adakites were formed through slab melting
highest oxygen fugacity in arc magmas may be NΔFMQ + 3, which (Sun et al., 2011, 2012a), deriving from the subducting oceanic crust
allows up to 1.4 wt.% of S at sulfide saturation (Jugo, 2009). At oxygen itself rather than the overlying mantle wedge. MORB has higher Cu
fugacities below ΔFMQ + 1, 10% or less partial melting of the mantle contents (70–150 ppm) (Hofmann, 1988; Sun et al., 2003b) than
wedge allows preservation of residual sulfides, assuming that the S other portions of the subducted oceanic slab, and thus is likely the
abundances in the mantle wedge is also 250 ppm. At higher oxygen fu- main contributor to mineralization. Therefore, partial melting of
gacity, no residual sulfides can be expected unless there is additional subducted oceanic slabs forms magmas with high Cu contents (Sun
sulfur added from the subducting slab, which in actuality is probably et al., 2011, 2012b), which plausibly explains the close associations
the case. During metamorphism of the subducting slab, S is mobile between adakites, especially formed during ridge subductions, and por-
from metamorphic rocks ranging in grade from blueschist to amphibo- phyry Cu deposits. These papers, however, did not consider the effects
lite (Sun et al., 2013a; Tomkins, 2010), such that they can release S and of oxygen fugacity and dramatically underestimated the contributions
chalcophile elements to the mantle wedge (J.L. Li et al., 2013). of slab melting.
110 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

porphyry deposits have very small surface exposures, ranging in area


from less than 1 km2 (mostly) to ~ 5 km2 (in rare cases). Thus, the
magma chamber would need to be unrealistically thick, e.g., thicker
than the continental crust. Alternatively, the magma must scavenge
laterally for distances over 10 km and concentrates the Cu at one rel-
atively small spot. None of these scenarios are practical. For adakitic
melts with initial Cu contents of ~ 1000 ppm, the amount of magma
needed measures only 60 km3 for the supergiant El Teniente. A
magma chamber extending to a depth of several kilometers is
sufficient.
Slab melts need to pass through the mantle lithosphere before they
eventually are emplaced in the upper crust, and thus may lose or collect
chalcophile elements by interactions with the mantle. Sulfide saturated
melts would lose some chalcophile elements. For slab melts formed at
melting degrees higher than that require for SCSS, they would be under-
saturated in sulfide and thus could assimilate more sulfides and
chalcophile elements from the surrounding mantle while ascending, de-
pending on the partitioning relationships between melt and residual
sulfide in the mantle. For example, 10% partial melting of MORB with
a S content of 1000 ppm at ΔFMQ N 2.3, would form a melt with a S con-
tent of ~1 wt.%, which is only half of its SCSS. This melt then would have
the capacity to assimilate large amounts of sulfide. This plausibly ex-
plains the common association between high-Mg adakites and porphy-
ry Cu deposits.
Meanwhile, sulfide is kept undersaturated during the evolution
of the oxidized magma, such that no sulfide segregation occurs,
and Cu, Mo, and Au act as moderately incompatible elements.
They would become enriched at an early stage of magma evolution,
and dissolve into fluids following reduction accompanying magne-
tite crystallization (Sun et al., 2004a, 2013b), or other reduction
processes.
It has long been proposed that oxygen fugacities higher than ΔFMQ +
2 have reached the magic number for porphyry mineralization
(Mungall, 2002; Sun et al., 2013b). As discussed above, this is exactly
Fig. 16. Oxygen fugacity of different tectonic settings. Note the range of oxygen fugacity for the point at which residual sulfide is eliminated by ~10% partial melting
continental peridotite is much less than that of Bryant et al. (2007), after removing perido- of a subducted slab, forming melts with initial Cu contents up to
tite samples from arc settings. The oxygen fugacities of convergent margin magmas are
1000 ppm. In contrast, at oxygen fugacity even only slightly lower,
systematically higher than those of intraplate settings. Magmas from intraplate settings
without influence from plate subduction are too reduced for mineralization (Sun et al., e.g., ΔFMQ + 1.7, the highest Cu contents that can be reached through
2013b). slab melting is ~450 ppm.
Modified after Bryant et al. (2007). The systematically low oxygen fugacities in Japan arc volcanic rocks
(bΔFMQ + 2) compared to those from the western American conti-
nents (up to ΔFMQ + 3) (Fig. 16), is an important factor controlling
Slab melts formed at oxygen fugacities higher than ΔFMQ + 2 are fa- the distribution of Cu porphyry deposits, i.e., abundant porphyry Cu de-
vorable for porphyry Cu mineralization. MORB has an average S content posits are located in the western American continents, whereas none
of ~1000 ppm (O'Neill and Mavrogenes, 2002), from which it is easy to occur in Japan (Fig. 1) (Sun et al., 2012b, 2013b).
get the 22% and 10% partial melting needed to eliminate residual sul-
fides at ΔFMQ + 1.7 and ΔFMQ + 2, respectively. Correspondingly, 3.2.2. Sulfate reduction
the Cu contents of such melts would be 450 and 1000 ppm, respectively. At oxygen fugacities higher than ΔFMQ + 2, most of the sulfur in
Slab melts formed at ΔFMQ + 2 would have Cu contents more than magmas is present as sulfate, and such magmas usually have higher con-
twice as high as those formed at ΔFMQ + 1.7, and thus are considerably tents of total S and chalcophile elements. Copper however, is a
more favorable for Cu porphyry mineralization than mantle wedge chalcophile element, and its final deposition during mineralization
melts. Only four times greater enrichments of Cu are needed to reach should be controlled mainly by the behavior of reduced sulfur (Liang
economic porphyry deposit proportions of ~ 4000 ppm. This ought to et al., 2009; Sun et al., 2004a, 2013b). Therefore, the final stage of miner-
be easily achieved through magma evolution plus hydrothermal pro- alization inevitably requires the reduction of sulfate (S6+: HSO− 2−
4 /SO4 )
2− − 2−
cesses. At oxygen fugacities of ΔFMQ + 2.3 or higher, only 5% partial in the oxidized source magmas to sulfides (S : H2S/HS /S ) or

melting is needed to remove all the residual sulfides, forming melts polysulfides (e.g., S2−
2 , S3 ) (Sun et al., 2013b).
with initial Cu contents of ~ 2000 ppm. In contrast, unrealistic N50% In principle, there are several ways to lower the oxygen fugacities
partial melting is needed to eliminate all residual sulfides at oxygen during magma evolution. Among them are degassing of oxidized vola-
fugacities lower than ΔFMQ + 1, corresponding to a Cu content of tile species (i.e. CO2, SO3), assimilation of reducing country rocks
less than 200 ppm in the melt. Residual sulfide would exist until (Ishihara and Matsuhisa, 1999; Smith et al., 2012), and reduction of sul-
the entire slab melted, in which case sulfide would remain the pre- fate by other elements, e.g., fractional crystallization of magnetite and
dominant sulfur species in both melt and slab (under reducing con- even hematite (Liang et al., 2009; Sun et al., 2004a, 2013b).
ditions of b ΔFMQ 0). It has also been proposed that degassing of SO2 may lower the oxy-
In order to form supergiant porphyry deposits, e.g., El Teniente, up to gen fugacity (Kelley and Cottrell, 2012). As illustrated in Eq. (11),
600 km3 of magma with a Cu content of 100 ppm would be needed to degassing of SO2 indeed consumes oxygen, such that it may lower the
get the necessary amount of metal (Stern et al., 2011). Note that most oxygen fugacity of the system. However, degassing of SO2 cannot reduce
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 111

sulfate to sulfide in the magmas. Instead, it drives the reaction (Eq. (11))
to the right, which leads to the oxidation of sulfide.

2H2 S þ 3O2 ¼ 2H2 O þ 2SO2 ð11Þ

Degassing of CO2 may reduce sulfate to sulfide, through a process


similar to that described by Eq. (12).

2C þ SO3 þ H2 O ¼ 2CO2 þ H2 S ð12Þ

As shown in Fig. 17, however, the C–CO2 buffer is lower than the SSO
buffer (Mungall, 2002), therefore there is little C in the system at SSO
buffer. Moreover, degassing of any type is not likely to be a key process
during porphyry mineralization, which is usually occurring at depths of
~2–4 km (Sillitoe, 2010).
Assimilation of reducing sediments may lead to sulfate reduction as
proposed by previous authors (Shen and Pan, 2013; Smith et al., 2012)
(also see Section 4). Nevertheless, assimilation occurring during the
emplacement of porphyries would result in heterogeneous reduction,
because reduction is mainly focused along the porphyry–country rock
interface. This kind of reduction process has been reported for Baogutu
(Shen and Pan, 2013), which was classified as a reduced porphyry Cu
deposit associated with ilmenite-series magmas (Cao et al., 2014) (see
detailed discussion in Section 4.3). Assimilation during magma chamber
development may be better mixed and more homogenized. However, it
would result in early sulfide saturation and segregation from the melt.
Because sulfide is denser than silicate melt, it would tend to sink,
which is more likely to occur during the mineralization of Cu–Ni
deposits, rather than porphyry deposits. Fig. 18. Ferric/ferrous Fe ratios and estimated oxygen fugacities of Manus glasses (Sun
Elements that can exist in variable oxidation states and that are pres- et al., 2004a). Oxygen fugacity of Manus glasses was calculated using the following equa-
 
ent in sufficient abundances to affect the redox state of the silicate Earth tion: ln XFe 2 O3
XFeO ¼ a lnƒO2 þ bT þ c þ ΣdiXi, where, a = 0.196, b = 1.1492 × 104, c =
are C, H, S, and Fe (Mungall, 2002). In highly oxidized porphyry −6.675, dAl2O3 = −2.243, dFeO = − 1.828, dCaO = 3.201, dNa2O = 5.854, dK2O =
6.215 (Kress and Carmichael, 1991).
magmas, sulfate, CO2 and H2O are the dominant species, which are al-
ready oxidized and thus cannot reduce sulfate to sulfide, leaving ferrous
iron as the only probable reducing agent (Sun et al., 2013b). Iron is a ferrous iron is coupled with the reduction of sulfate to sulfide
major element, with about 20 to 30% of the total Fe in a typical porphyry (Eqs. (13), (14)). Magnetite crystallization at Manus and many other
occurring as ferrous iron, corresponding to an oxygen fugacity of less arc volcanic rock series is coupled with dramatic decreases in Cu and
than ΔFMQ + 2 (Fig. 18). Magnetite contains 66.7% ferric iron, whereas Au (Sun et al., 2004a), which is an important factor in understanding
hematite contains 100% ferric iron. Therefore, the crystallization of mag- Au and Cu mineralization (Sun et al., 2004a). The reduction in the
netite and hematite tends to lower the proportions of ferric iron in the amount of Cu and Au during magma evolution is well-known in arc
magma, which apparently would lower the oxygen fugacity. In the magmas (Moss et al., 2001; Sun et al., 2011; Togashi and Terashima,
case of the Manus deposit in volcanic rocks, the Fe3 +/(total Fe) did 1997), and has been referred to as the magnetite crisis (Jenner et al.,
not change much during magnetite crystallization (Fig. 18). This was at- 2010). Indeed, magnetite crystallization/alteration is often taken as
tributed to sulfate buffering (Sun et al., 2004a), i.e., the oxidation of the controlling process for porphyry Cu and Au mineralization by caus-
ing sulfate reduction and presumably also oxygen fugacity fluctuations
(Liang et al., 2009).
There are several ways to describe the oxidation of ferrous Fe. In
magma system, this is better described as oxidation of FeO by sulfate
or another oxidant (Eq. (13)) (Sun et al., 2004a):
2− 2−
SO4 þ 8FeO ¼ 4Fe2 O3 þ S : ð13Þ

Eq. (13) is simplified, Fe2O3 represents the ferric Fe in Fe3O4. This re-
action can be more precisely described by Eq. (14):
2− 2−
SO4 þ 12FeO ¼ 4Fe3 O4 þ S : ð14Þ

These reactions have no effects on the pH value of the system. This is


one reason why the oxygen fugacity of the Manus magma did not
change much (Sun et al., 2004a), or even decrease slightly during the
crystallization of magnetite and sulfate reduction (Jenner et al., 2010)
(see also Section 5).

Fig. 17. Sulfate reduction and oxygen buffers. CCO = carbon dioxide–carbon oxide buffer;
3.2.3. Hematite–magnetite intergrowth
SSO = sulfide–sulfur oxide buffer. Interestingly, in addition to abundant magnetite in porphyry de-
After Mungall (2002) and Sun et al. (2014a,b). posits (Astudillo et al., 2010; Audetat et al., 2004; Baker et al., 1997;
112 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Dilles et al., 2011; Imai, 2005; Liu et al., 2011; Yang et al., 2002), hema- The porphyry mineralization process requires a continuous reduc-
tite and specularite (a hydrothermal variety of hematite) are common tion of sulfate to sulfide, coupled with consumption of oxygen and
in porphyry copper deposits (Figs. 10, 13) (Zhang, Ling et al., 2013; even reduction of CO2 to CH4. Given that porphyry magmas are highly
Sillitoe, 2010; Sun et al., 2013b). As mentioned previously, primitive he- oxidized, reaction (21) is not likely to occur. No significant amounts of
matite–magnetite intergrowths have been reported in many porphyry methane have been detected in most cases, except in reduced porphy-
Cu deposits all over the world (Fig. 10), including nearly all major por- ries (see details in Section 4).
phyry deposits in China (Sun et al., 2013b), as well as some deposits Based on high pressure experiments, it was proposed that the tri-
in South America (Ballard et al., 2002; Patricio and Gonzalo, 2001; Vila sulfur ion S−
3 is an important sulfur species (Fig. 11) at pressures and
and Sillitoe, 1991) and the Southwest Pacific islands (Hedenquist temperatures (Pokrovski and Dubrovinsky, 2011) that are common
et al., 1998; Imai et al., 2007), and Mongolia (Khashgerel et al., 2008). for porphyry systems (Hedenquist and Lowenstern, 1994; Seo et al.,
Most of the hematite and/or specularite formed during the late stages 2009; Sillitoe, 2010; Sun et al., 2014a,b). It has been further proposed
of porphyry mineralization (Sillitoe, 2010; Sun et al., 2013b). that in general geological processes where the tri-sulfur ion may have
The occurrence of primitive hematite in closely association with been involved need to be reconsidered (Pokrovski and Dubrovinsky,
magnetite strongly suggests a very high oxygen fugacity (Fig. 11), 2011), and this would require the investigation of different reaction
such that reaching the HM oxygen fugacity buffer (equivalent to pathways (Sun et al., 2013b). In fluids, the reactions can be described
~ΔFMQ + 4) during porphyry Cu mineralization. This suggests that, in by Eqs. (22) and (23)
contrast to hydrothermal systems in the Manus backarc basin (Jenner
2− 2þ − þ
et al., 2010; Sun et al., 2004a), the oxygen fugacity of porphyries 6SO4 þ 52H2 O þ 57Fe ¼ 2S3 þ 19Fe3 O4 þ 104H ð22Þ
increases during magnetite crystallization (Sun et al., 2013b).
The further oxidation of magnetite is controlled by the pH values of − 2þ 2− þ
the system (Sun et al., 2013b). In contrast to the magma system, ferrous 2S3 þ 20H2 O þ 15Fe ¼ 6S þ 5Fe3 O4 þ 40H : ð23Þ
Fe is mainly present as Fe2+ in aqueous fluids. In this case, the oxidation
of ferrous Fe is better described by Eq. (15): Both of these reactions (Eqs. (22), (23)) also would lower the pH of
the fluids, which in turn would elevate the oxidation potential of sulfate.
2− 2þ 2− þ Such conditions will promote the formation of hematite and would be
SO4 þ 12Fe þ 12H2 O ¼ 4Fe3 O4 þ S þ 24H : ð15Þ
favorable for porphyry mineralization. The oxidation potential of the

SO2−
4 –S3 reaction also rises with decreasing pH (Fig. 11), so that mag-
As shown in Eqs. (2) and (1), HM and FMQ oxygen fugacity buffers
netite may be further oxidized to hematite by SO24 −, releasing OH−
do not change with pH. The oxidation potential of sulfate, however, de-
(Eq. (24)). This in turn increases the pH and promotes magnetite forma-
pends strongly on pH (Figs. 11). The oxidation of ferrous iron by sulfate
tion (Sun et al., 2013b), which may explain magnetite and hematite in-
leads to a decreases in pH in ore-forming fluids within the porphyry
tergrowths (Fig. 10). In addition to direct oxidation of magnetite,
system (Eq. (15)), which drives up the oxidation potential of sulfate.
hematite and specularite may form directly from fluids (Eq. (25)), re-
Consequently, the fO2 increases during the reduction of sulfate
leasing H+, which lowers pH and promotes further oxidation of ferrous
(Fig. 11) (Pokrovski and Dubrovinsky, 2011; Sun et al., 2013b). It is es-
Fe. Pure hydrothermal hematite and specularite can often be deter-
timated that the amount of H+, released during the reduction of sulfate
mined to have formed during the late stages of porphyry mineralization
and oxidation of ferrous iron, may lower the pH down to +2, depending
(Sillitoe, 2010), e.g., the specularite occurring in late stage veins at
on temperatures, and if there are no pH buffers in the system to signif-
Dexing (Fig. 13).
icantly elevate the oxygen fugacity (Sun et al., 2014a,b) and oxidize
magnetite or ferrous Fe in hydrothermal fluids to hematite and/or 2−
38Fe3 O4 þ 6SO4 þ 5H2 O ¼ 57Fe2 O3 þ 2S3 þ 10OH
− −
ð24Þ
specularite (Eq. (16)).

2þ 2− − þ
SO4
2−
þ 8Fe

þ 8H2 O ¼ 4Fe2 O3 þ S
2−
þ 16H
þ
ð16Þ 38Fe þ 6SO4 þ 33H2 O ¼ 19Fe2 O3 þ 2S3 þ 66H ð25Þ

In magmas, in contrast to fluids, FeO is more abundant than Fe2+. If


Or further oxidation of magnetite (Eq. (17)): S−
3 is indeed stable as proposed (Pokrovski and Dubrovinsky, 2011),
then the reduction of sulfate in magmas first leads to an increase in
2− 2−
SO4 þ 8Fe3 O4 ¼ 12Fe2 O3 þ S : ð17Þ pH (Eq. (26)), which is compensated by further reduction of S− 3
(Eq. (27)). Therefore, S−
3 may indeed cause more fluctuations in oxygen
The formation of hydrothermal hematite will further lower the pH, fugacities through a different reaction pathway, but it has no major in-
whereas further oxidation of magnetite will not change the pH. All the fluences on the final results (Sun et al., 2014a,b).
sulfate reduction reactions provide S2 −, which promotes mineraliza-
2− − −
tion, forming chalcopyrite (Eq. (18)) and more pyrite (Eq. (19)) 6SO4 þ 5H2 O þ 57FeO ¼ 2S3 þ 19Fe3 O4 þ 10OH ð26Þ
(Heinrich, 1990) and releasing H2. Hydrogen may further react with
CO2 and/or O2 (Eqs. (20), (21)) or escape from the porphyry system
− 2− þ
during degassing or diffusive loss (Sun et al., 2014a,b). 2S3 þ 5H2 O þ 15FeO ¼ 6S þ 5Fe3 O4 þ 10H ð27Þ

þ 2þ þ For large porphyry deposits, the reduction of sulfate inevitably re-


Cu þ Fe þ 2H2 S ¼ CuFeS2 þ 3H þ §H2 ð18Þ
sults in the oxidation of magnetite to hematite as the pH is lowered
(Figs. 10, 11). Therefore, the hematite –magnetite intergrowths may
2þ þ be taken as a possible ore indicator in future prospecting for copper
Fe þ 2H2 S ¼ FeS2 þ 2H þ H2 ð19Þ
deposits.

2H2 þ O2 ¼ 2H2 O ð20Þ


3.3. Summary

Sulfur is one of the most important geosolvents that controls the be-
4H2 þ CO2 ¼ CH4 þ 2H2 O ð21Þ havior of copper and other chalcophile elements, therefore it is essential
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 113

to the understanding of mineralization processes of copper and a variety


of other metal resources. Sulfate is the dominant sulfur mineral species
in porphyries associated with large copper deposits. This is because the
oxidation of sulfide to sulfate during partial melting is essential for the
efficient extraction of chalcophile elements out of the source region,
especially from subducted oceanic slabs, which have more than
5 times more sulfur than the mantle. In contrast, metals of porphyry
Cu deposits are hosted in sulfides, which require low oxygen fugacity
for their stabilization during the final stage of mineralization. Therefore,
the key process of porphyry mineralization is oxidation and reduction of
sulfur, controlled by ferrous/ferric Fe and pH values. Sulfate reduction in
hydrothermal fluids lowers the pH and consequently elevates the oxy-
gen fugacity of the system up to the HM buffer. The acid released during
sulfate reduction causes alterations. In contrast, sulfate reduction in
magmas does not change the pH, such that the oxygen fugacity is slight-
ly reduced. All types of porphyry deposits, ranging from Cu + Au,
Cu + Mo, and Mo deposits, may be oxidized sufficiently to reach the
redox states of the HM buffer.

4. The association of porphyry deposits with reduced magmas

Although most of the porphyry deposits are closely associated with


oxidized magmas, a small group of porphyry deposits are reported as
apparently related to reduced magmas with oxygen fugacities ranging
from ΔFMQ − 0.5 to − 3 (Fig. 19) (Cao et al., 2014; Rowins, 2000;
Smith et al., 2012). Reduced porphyry deposits so far reported are: 17
Mile Hill, Western Australia; the Minãs de San Anton, Mexico (Rowins,
2000); the Baogutu, northwestern China (Cao et al., 2014; Shen and
Pan, 2013; Shen et al., 2010a, 2012); North Fork Deposit, west Central
Cascades, western North America (Smithson and Rowins, 2005);
Catface; Vancouver Island, British Columbia (Smith et al., 2012); and a
few other small porphyry deposits (Rowins, 2000). Three criteria have
been proposed for the classification of reduced porphyry Cu deposits:
(1) lack of primary hematite and sulfate minerals, (2) rich in hypogene
pyrrhotite and CH4, and (3) oxides belongs to the ilmenite-series,
reduced, I-type granitoids (Rowins, 2000).
Instead of primary hematite, magnetite, and sulfate (e.g., anhydrite,
gypsum) that are frequently seen in oxidized porphyries, these
“reduced” porphyry Cu–Au deposits contain abundant hypogene pyr-
rhotite, and commonly have carbonic-rich ore fluids with substantial
proportions of CH4, clearly indicating a reducing condition (Cao et al.,
2014; Rowins, 2000; Smith et al., 2012). More importantly, they are as-
sociated with ilmenite-bearing, reduced I-type granitoids (Cao et al.,
2014; Rowins, 2000), in direct contrast to oxidized porphyries that are
related to a magnetite-series mineralogy (Ishihara and Sasaki, 1991; Fig. 19. Diagram of log fO2 versus temperature and 1000/T illustrating the oxygen
Thompson et al., 1999). For example, ilmenite is an end-member of fugacities of (A) oxidized porphyry deposits and (B) reduced deposits, Catface and
Baogutu. The systematic difference in oxygen fugacities is likely not primary. Baogutu is
the hematite–ilmenite solid solutions with a very low hematite compo- hosted in ilmenite-series diorites, containing native antimony (An and Zhu, 2010),
nent, varying from XHem (0.01–0.1), for the Catface phase and XHem hypogenepyrrhotite, and methane-rich fluid inclusions. Previous studies suggested that
(0.04–0.09) for the Hecate Bay phase (Smith et al., 2012). Magnetite is assimilation occurring during the emplacement of porphyries resulted in reduction,
nearly pure Fe3O4 with less than 2% Fe2TiO4 component for Catface which is mainly focused along the porphyry–country rock interface (Shen and Pan, 2013).
Modified after Oyarzun et al. (2001) and Cao et al. (2014) and Smith et al. (2012),
(Smith et al., 2012) and Baogutu (An and Zhu, 2010; Cao et al., 2014;
respectively.
Shen and Pan, 2013; Shen et al., 2010a,b). Ilmenite is more abundant
than magnetite in the reduced porphyries (Cao et al., 2014; Smith
process needs to be clarified. Some major questions to be answered
et al., 2012), e.g., all intrusive phases at Catface have accessory FeTi
are: How reduced magmas formed in an arc environment, where oxi-
oxides with ilmenite/magnetite ratios of ~9:1 (Smith et al., 2012), indi-
dized magmas are common, e.g., in western North America? In contrast,
cating a relatively reduced oxidation state, i.e., belonging to the
why are there no porphyry deposits in Japan, where the ilmenite-series
ilmenite-series (Ishihara, 2004). Experiments with apatite have shown
is well-developed (Ishihara and Murakami, 2006)? How do Cu and Au
that SO3 contents reflect the oxygen fugacity of the magmas (Peng
get enriched in reduced magmas? What's the relation between oxidized
et al., 1997). The uniform and low SO3 contents in apatite from reduced
and reduced porphyries?
magma deposits were used to argue that the low oxygen fugacity of
these porphyries is primary (Fig. 20) (Cao et al., 2014; Smith et al.,
2012). 4.1. Reduced magmas
The reduced porphyry deposits range in age from the Late Archean
to the Oligocene (Rowins, 2000). Although the tonnages of these re- Arc magmas usually have high oxygen fugacities (Fig. 16) (Arculus,
duced porphyry deposits are generally much smaller than oxidized por- 1994; Ballhaus, 1993; Carmichael, 1991; Kelley and Cottrell, 2009; Sun
phyry deposits, the mechanism that controls these mineralization et al., 2012a, 2013a,b), a condition that has been attributed to a variety
114 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

(2) evolvement of subducted reducing sediments, and (3) assimilation


of reduced sediments during magma emplacement (Smith et al., 2012).
Previous studies suggested that a slab window introduces hot, up-
welling asthenospheric mantle in the subduction zone environment,
forming non-arc-like alkalic and adakitic magmatism in the volcanic
arc (Abratis and Worner, 2001; Groome and Thorkelson, 2009; H. Li
et al., 2011, 2012; Kinoshita, 1997). The Kula–Farallon ridge in the
northeastern Pacific began descending under the North American conti-
nent in Alaska in the Late Cretaceous (Madsen et al., 2006; Scharman
et al., 2012), and then between 62 and 11 Ma migrated southward
(Cole and Stewart, 2009; Cole et al., 2006; Liu et al., 2008), forming a
semi-continuous forearc magmatic belt from Alaska to Oregon
(Madsen et al., 2006). Ridge subduction forms oxidized adakites by
slab melting (Defant and Drummond, 1990; Ling et al., 2009, 2013),
followed by reduced A-type granites (Abratis and Worner, 2001; H. Li
et al., 2012; Thorkelson and Breitsprecher, 2005; Yogodzinski et al.,
1994). Interestingly, the Catface porphyry deposit is closely associated
with Mt Washington, which is an adakite (Smith et al., 2012) and seem-
ingly would support the slab window model. Nevertheless, adakites
Fig. 20. Diagram of Cl versus SO3 in apatite indicating the oxygen fugacities of Catface and formed by slab melting have much higher initial Cu contents than
Baogutu. The SO3 contents of Catface apatite are all lower than 0.15 wt.%, corresponding to mantle-derived melts, and thus are favorable for porphyry mineraliza-
oxygen fugacities below the NNO buffer. In contrast, the SO3 contents of Baogutu apatite
range from less than 0.1 wt.% to much higher than 0.2 wt.%, corresponding to varied
tion (Sun et al., 2010, 2011), suggesting that the reduced Catface por-
oxygen fugacities, ranging from below the NNO buffer to close to the NNO +1. HDP = phyry could be a country rock hosting the ore deposit, rather than the
hornblende diorite porphyry; GDP = Granodiorite porphyry; D = diorite. causative porphyry.
Modified after Cao et al. (2014) and Smith et al. (2012). Subduction of reducing sediments also reduces the oxygen fugacity of
arc magmas, as seen in the Japan arc (Takagi, 2004). Carbon rich sedi-
of mechanisms. The addition of subduction released oxidizing compo- ments in the descending plate react with H2O to produce CH4 and CO2
nents to the mantle wedge or directly to arc magmas is proposed as (Ballhaus, 1993; Takagi, 2004), or directly releases CH4 by devolatilization
the most straightforward way, e.g., water (Kelley and Cottrell, 2009), (Song et al., 2009). This process, however, was excluded from further
oxidizing fluids and/or melts (Brandon and Draper, 1996), fluids with consideration based on the low 87Sr/86Sr of Catface (Smith et al.,
oxidizing components, e.g., hematite, sulfate (X.M. Sun et al., 2007), 2012). Methane, however, may decouple from Sr, such that Sr isotopes
slab melts with high Fe3+/Fe2+ratios (Mungall, 2002), as well as other cannot give a conclusive answer. Nevertheless, the oceanic plate
oxidized components, e.g., carbonates, or other oxidized sediments. subducting underneath Japan is Cretaceous in age, which has experi-
Arc magmas may also get oxidized during their evolution and ascent enced eight Ocean Anoxic Events (Jenkyns, 2010), and thus contains
due to the following mechanisms (Ballhaus, 1993; Lee et al., 2005, abundant organic rich sediments. In contrast, the oceanic plates
2010): (1) Changing oxygen buffers from graphite CO2 (CCO) equilibria subducting underneath the western North American continent are
in the mantle source to Fe3 +–Fe2+ equilibria after graphite has been younger and have experienced only one Ocean Anoxic Event. Moreover,
eliminated by partial melting (Ballhaus, 1993); (2) Fractional crystalli- ocean ridges are young and generally sediment-poor. Therefore, sub-
zation of olivine and other minerals that favors ferrous iron over ferric duction of reducing sediments is not a plausible explanation for the
iron (Carmichael, 1991); (3) Assimilation of oxidizing country rocks low fO2 shown by the Catface porphyry.
(Lee et al., 2005); (4) Degassing of reduced volatile species (e.g., H2, The third mechanism proposed for the reduced condition of the
H2S and CH4) (Ballhaus, 1993; Lee et al., 2005); (5) Changing pH Catface porphyry was by assimilation of reduced sediments. The Catface
(e.g., lowering pH in a sulfate–sulfide system) (Sun et al., 2013a,b); porphyry may have ascended through and interacted with graphite-rich
(6) Magma recharging which preferentially enriches ferric Fe (Lee reducing country rock known to exist in the region (Smith et al., 2012).
et al., in press). The reduction of ferric Fe through reaction with reducing materials,
There are also several mechanisms that, in principle, may lower the e.g., graphite, in sediments was proposed as a potential factor responsi-
oxygen fugacity of magmas: (1) Addition of reducing sediments from ble for the low fO2 of the parental Catface magma. This mechanism was
plate subduction (Takagi, 2004); (2) Assimilation of reducing country excluded based on Sr isotopes (see details below) (Smith et al., 2012).
rocks (Ishihara and Matsuhisa, 1999; Smith et al., 2012); (3) Degassing The authors then proposed that the reduced Catface porphyry was
of oxidized volatile species (i.e. CO2, SO3); (4) Addition of subduction re- due to degassing of SO2 from the magma. This argument is based on
leased reducing fluids (Song et al., 2009); and (5) Fractional crystalliza- the low S content in the rock's apatite and a recent work by Kelley
tion of magnetite and even hematite (Liang et al., 2009; Sun et al., and Cottrell (2012), which proposed that fractional crystallization
2004a, 2013a,b). As discussed previously (Section 3.2.2), the magnetite coupled with degassing of S in a sub volcanic arc magma chamber can
crystallization reduces the oxygen fugacity only in magmatic systems result in a reduction of N 2 log fO2 units in the magma. This calculated re-
(reactions (13) and (14)), and may only lower the oxygen fugacity to duction was based on the rock's ferric iron content and S in melt inclu-
the FMQ oxygen buffer. Here we discuss two reduced porphyry deposits sions, and would hold true for magmas varying in composition from
in detail. basaltic andesite to dacite (Kelley and Cottrell, 2012). Degassing of
SO2 indeed has a significant influence on the oxygen fugacity in volcanic
4.1.1. Evidence for reduced magmas of Catface porphyry deposit systems. It is, however, not likely to be a major process in plutons, which
Catface is the largest reduced porphyry Cu deposit so far reported, should experience much less degassing.
emplaced at 40.4–41 Ma, with an indicated reserve of 56.9 Mt @ 0.4% Moreover, apatite from the Catface has a homogenous S content
Cu and an additional inferred resource of 262.4 million tons @ 0.38% throughout the crystal (Smith et al., 2012), which argues against
Cu. Three potential mechanisms for the formation of Catface reducing major degassing induced S loss during magma evolution. Experiments
magmas had been discussed before degassing was assigned as the caus- on felsic liquids showed that the SO3 content in apatite increase with in-
ative process: (1) upwelling of the reducing asthenospheric mantle in- creasing fO2 from 0.04 wt.% at the FMQ buffer, up to N1.0 wt.% at the HM
duced by the opening of a slab window during ridge subduction, buffer (Peng et al., 1997). As pointed out by Smith et al. (2012), apatite
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 115

crystallizes below 950 °C in silicic calc-alkaline magmas (Green and degassing cannot explain the decoupling between δD and O isotope
Watson, 1982). If the magma at Catface was reduced during degassing compositions. Moreover, the degassing of porphyries should be much
of SO3, there should be compositional zonation in the apatite, which is less pronounced than the degassing of volcanic rocks.
not the case (Fig. 20) (Smith et al., 2012). The systematically lighter H coupled with a slightly lighter O can
best be explained by the addition of meteoritic water. Baogutu is located
in the center of the Eurasian continent. Meteoritic water there should be
4.1.2. Evidence for reduced magma in the Baogutu porphyry deposit very light in H and O isotope composition. Addition of such water
Baogutu is located in West Junggar, Central Asian Orogenic Belt, and should dramatically reduce the δD of the magma with a much lesser
is the second largest reduced porphyry Cu deposit so far reported, effect on O because magmas have low hydrogen contents and abundant
with controlled reserves of 0.63 Mt Cu @ 0.28 wt.%, 14 t Au @ oxygen contents.
0.1 ppm, 1.8 × 104 t Mo @ 0.011 wt.% and 390 t Ag @ 1.8 ppm (Cao
et al., 2014). It contains native antimony (An and Zhu, 2010), abundant 4.2. Source of copper
hypogene pyrrhotite, and methane-rich fluid inclusions. The low apatite
SO3 content, whole rock Fe2O3/FeO, and fluid compositions indicate a For oxidized magmas, as discussed in Section 3, excess sulfur is pre-
low fO2 of ~ NNO for the magma and NNO–NNO − 2 for associated cipitated in the form of sulfate, leaving behind less residual sulfide, such
hydrothermal fluid (Cao et al., 2014; Shen et al., 2010b). that Cu, Au and other chalcophile elements become enriched in magmas
Based on H–O and sulfide S–Pb isotope data, it has been argued that (Lee et al., 2012; Sun et al., 2012a, 2013a,b). In contrast, reduced
the methane-rich ore-forming fluids were derived from a deep mantle magmas usually retain residual sulfides, such that they have low prima-
source with little contamination from penetrated sediments (Cao ry Cu contents (Fig. 3) (Lee et al., 2012). Then, how do reduced magmas
et al., 2014). The assimilation model, however, does not contradict get enriched in Cu?
with a deep mantle source for the S, Pb, nor Cu. In fact, most Cu should As pointed out by previous authors, porphyry deposits associated
have come from slab melts (Sun et al., 2011, 2012b), as also should have with reduced magmas are generally small (Cao et al., 2014). This was at-
come the S and Pb, which are present in abundances similar to mantle tributed to the originally low Cu and Au contents of the magmas, less
derived magmas. The H isotopes of fluid inclusions from Baogutu have magmatic fluids released due to deep emplacement, or tectonic settings
very low δD values (Shen et al., 2012; Zhang et al., 2010), which are that are not favorable for porphyry mineralization (Cao et al., 2014).
much lower than for either magmatic fluids or metamorphic fluids They also argued that significantly lower sulfur solubility in reduced
(Fig. 21). This was interpreted as the main evidence for degassing melt (Jugo, 2009; Jugo et al., 2005, 2010) keeps S2− as the dominant
(Cao et al., 2014). Degassing, however, should result in heavier hydro- sulfur species, which potentially isolates sulfides from the magma dur-
gen and oxygen isotopes in the residual magmas. The authors used pub- ing its migration to the site of final emplacement, thus producing only
lished heavy H values from circum-Pacific volcanic gases (Giggenbach, relatively small chalcophile endowments (Cao et al., 2014). None of
1992b) to argue that degassing should leave behind fluids with even these arguments are convincing to us. First, lower Cu and Au contents
lighter H than contained in the magmatic and metamorphic fluids would first result in lower grade, not necessarily smaller tonnages.
(Cao et al., 2014). The heavy H of circum-Pacific volcanic gases, howev- The grades of porphyry Cu and Au deposits associated with reduced
er, was originally explained as having arisen from the addition of seawa- magmas are comparable to, if not higher than, those associated with
ter to the gases (Giggenbach, 1992a). Basic principles of isotope oxidized magmas (Cooke et al., 2005; Rowins, 2000). Second, there is
geochemistry dictate instead that in degassing, i.e., evaporating, the re- no systematic difference in terms of emplacement depths and tectonic
maining liquid becomes isotopically heavier in H and O. In addition, settings between reduced and oxidized porphyry deposits. More impor-
tantly, compared to oxidized magmas, reduced magmas indeed have
lower sulfur contents with S2− occurring as the dominant sulfur species
(Jugo et al., 2010), but this does not necessarily mean higher S2− in re-
duced magmas as claimed by Cao et al. (2014). Given that the solubility
of sulfide is independent of oxygen fugacity, but increases with decreas-
ing pressure under reducing conditions (Mavrogenes and O'Neill,
1999), the high proportions of S2− cannot “isolate sulfides from the
magma during migration” as proposed.
Based on a synthesis of theoretical, experimental, and field data, it
has been proposed that Cu and Au can be transported via the vapor
phase to distal sites as far as several kilometers away from the causative
porphyry due to fluid boiling or immiscible phase separation. Conse-
quently, the source porphyry becomes a low-grade sub-economic Cu–
Au core or failed porphyry Cu system (Rowins, 2000), which is more
or less similar to an epithermal ore system. Experiments find that the
main transporting agents of Cu at the porphyry level are brines and
that models based on transporting copper in the vapor phase are incor-
rect (Lerchbaumer and Audetat, 2012). It is further demonstrated, using
experimental studies, that brine–vapor separation in porphyry deposits
does not cause selective Cu transfer to the vapor, but is more likely to
destabilize Cu complexes and promote copper ore deposition during de-
Fig. 21. δD versus δ18O diagram and calculated isotopic composition of waters in hydro-
compression and unmixing of the two fluid phases. In contrast, Au may
thermal fluids derived from measured isotopic composition of quartz and its fluid inclu-
sion for the Baogutu deposit from Cao et al. (2014). Reference lines and boxes are as be selectively transferred into the vapor phase, allowing its transport
follow: meteoric water line (Craig, 1961), felsic magmatic water (Taylor, 1992), residual from the deeper porphyry copper deposits to form shallower
water in intrusion after degassing and crystallization (Taylor, 1974), low-salinity vapor epithermal gold deposits (Seo and Heinrich, 2013). This explains the
discharges from high-temperature volcanic fumaroles (Giggenbach, 1992b), primary Au-rich capping feature of many reduced porphyry deposits. Mean-
magmatic water area, metamorphic water area and multiple formation water area
(Hoefs, 2004). H–O isotopic data are from (Zhang et al., 2010) and (Shen et al., 2012).
while, Cu may be transported to distal sites through normal fluids.
Low-salinity vapor shows strong seawater signals, which cannot represent “degassing” Such a reduced porphyry Cu–Au mineralization model does not contra-
fluids. dict the current understanding of porphyry Cu–Au formation. In fact,
116 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

the recognition of reduced porphyry Cu–Au systems encourages a than MORB (~ 100 ppm) (Sun et al., 2003b), such that melt derived
search for distal sites that are favorable for focusing and precipitating from them should have much lower Cu contents and be less favorable
Au and Cu-rich vapors (Rowins, 2000). This implies that such porphyry for porphyry Cu mineralization. Therefore, the reduced Catface porphy-
deposits themselves are not genetically related to reduced magmas, ry is likely to be a country rock that only hosts the deposit.
i.e., they are not the causative magma, but the host rock. Based on the above discussion, we propose the following model for
the formation of the Catface porphyry. Intrusions in the Vancouver
4.3. Formation of reduced porphyry deposits Island range in age from 51 to 35.5 Ma (Madsen et al., 2006), spanning
the time during which the Kula–Farallon ridge collided with the conti-
Porphyry deposits related to reduced magmas have two other dis- nent. Accompanying the oblique subduction of the ridge (Madsen
tinct characteristics. They are associated with carbonic-rich ore forming et al., 2006), its two limbs separated, with a slab window opened in be-
fluids (Cao et al., 2014; Rowins, 2000; Smith et al., 2012), and multiple tween. The first limb formed early through partial melting of the hot
intrusive events (Cao et al., 2014; Rowins, 2000; Smith et al., 2012). In subducting plate. The resulting adakites may then have been reduced
addition to the transportation of Cu and Au from the causal porphyry through reaction with the thick (maximum thickness of 4 km)
to distal sites of deposition (Rowins, 2000), we propose two other carbon-rich Cretaceous Nanaimo Group sediments (Madsen et al.,
ways that may form reduced porphyry deposits: (1) by reduction of 2006) as they migrated upward. Consequently, sulfate is reduced to
oxidized magmas during their ascent in the crust, and (2) by using re- sulfide, leaving behind Cu-rich sulfide accumulations in the lower
duced magmatic rocks that are present only as country rocks to host crust. This is followed by emplacement of the Catface porphyry at 41–
ore deposits originating from underlying oxidized magmas. 40.4 Ma (Smith et al., 2012), which took place at a time that the slab
window was open. The mantle derived parent magmas of the Catface
4.3.1. The formation of the Catface porphyry deposit porphyry are expected to have been more reduced (near the FMQ
As discussed above (Section 3.2.2), assimilation of reduced sedi- oxygen fugacity buffer) and drier than the adakite, and to have a
ments may have reduced the oxygen fugacity of the Catface porphyry, lower initial Cu contents. Nevertheless, it became wetter and even
a possibility that was excluded based on Sr isotopes. The 87Sr/86Sr more reduced (ΔFMQ − 0.3 to − 3) with a higher Cu content after
ratio of the Catface intrusions is 0.704, whereas those of the graphite- assimilating the Cu-rich sulfide accumulations in the lower crust. Mean-
rich metasediments of the Pacific Rim terrane range from 0.706 to while, ridge subduction induced compression and consequent uplift
0.708 (Smith et al., 2012). We find such small differences in Sr isotopes and erosion occurred, which favored the exposure of the porphyry de-
to be negligible, and thus do not argue strongly against the assimilation posit. Oxidized Mt Washington adakite was emplaced at 41–35.3 Ma
model. Graphite and especially methane need not be considered during the subduction of the west limb of the ridge, bringing more ore
coupled with silicate minerals, such that the reducing action of the forming fluids into the Catface porphyry. We infer that large propor-
graphite and methane is not proportionally related to the assimilation tions of these adakites are still buried due to the dramatically lessened
of silicates nor sulfides, thereby decoupling them from the Sr, Pb, and compression, uplifting and erosion following ridge subduction.
S isotopes. Relationship between Sr and C aside, the amount of C needed
is very small. The total iron (expressed as Fe2O3) content ranges from
2.25 wt.% to 5.53 wt.% for the Catface porphyry. One carbon atom re- 4.3.2. The formation of the Baogutu porphyry deposit
duces 4 ferric Fe atoms (Eq. (28)), assuming all the Fe in the parental Based on previously obtained fluid inclusion H–O isotope data and
Catface magmas was ferric Fe, and all the ferric Fe was reduced by sulfide S–Pb isotope data, it was proposed that the methane-rich ore
graphite, then only 0.04 to 0.1 wt.% of graphite is required, which may forming fluids in the Baogutu porphyry deposit were derived from a
only have had a limited influence on Sr isotopes. More interestingly, deep mantle source with little contamination from sedimentary compo-
the Catface intrusions have fO2 values nearly identical to the C–CO– nents (Cao et al., 2014). As discussed above (Section 4.1.2, Fig. 21), the
CO2 buffer at similar pressure and temperature conditions (Smith authors' understanding about H–O isotopes is in error, whereas only
et al., 2012), which would support the assimilation of graphite-rich sed- small amount of C is enough to lower the oxygen fugacity of the Baogutu
iments. low Fe magmas. Therefore, S–Pb isotopes cannot place any decisive con-
straints on the role of assimilation in forming the deposit. As pointed out
C þ 2Fe2 O3 ¼ 4FeO þ CO2 ð28Þ by the authors, detailed studies are needed to clarify the origin of the
CH4.
Reduced sediments usually also contain methane (Cao et al., 2014),
More importantly, there are three intrusive phases at Baogutu,
which is a more efficient reductant, i.e., 1 methane molecule reduces 8
which from old to young are: (1) the main diorite phase, (2) dikes of di-
ferric Fe atoms (Eq. (29)), such that only ~0.02 to 0.05 wt.% of methane
orite porphyry and granodiorite porphyry intruding the early diorite
is needed. Therefore, the parental Catface magmas may have acquired a
stock, and (3) dikes of hornblende diorite porphyry intruding all the
low fO2 during their ascent through the crust. The homogenous SO3 in
three phases. All the samples with the exception of two were collected
apatite may simply be due to an early reduction of the magmas, before
from the diorite, which are not porphyry at all. The reduced diorite
apatite crystallization.
magma may have no relation with the porphyry mineralization, except
CH4 þ 4Fe2 O3 ¼ 8FeO þ CO2 þ 2H2 O ð29Þ to act as a host rock.
We propose that the reduced features of the Baogutu porphyry de-
Considering that the Catface porphyry is not adakitic, it may well be posit are secondary and occurred during emplacement, thus having no
simply the country rock that hosts the deposit. The Catface porphyry major influence on the mineralization process. A similar model has
deposit is closely associated with a nearby pluton on Mt Washington been proposed based on geochemical and mineralogical studies by
(35–41 Ma), which is an adakite (Smith et al., 2012). The mineralization Shen and Pan (2013), in which mineral composition data suggest that
(40.9 Ma) and the emplacement age of the Catface porphyry (40.4– the primary magma of the Baogutu porphyry deposit is oxidized. The
41 Ma) (Smith et al., 2012) are both within the age range of the Mt heterogeneous and reduced characteristics of the deposit are attributed
Washington adakite. Adakites along the eastern Pacific rim are mostly to significant country-rock contamination after emplacement (Shen
formed by slab melting (Liu et al., 2010; Sun et al., 2012a,b), which are and Pan, 2013) in an arc setting (Shen et al., 2013a,b).
likely to have high initial Cu contents, and are thus favorable for porphy- Similar to Catface, there are also adakites in Baogutu. The Baogutu
ry mineralization (Sun et al., 2010, 2011). In contrast, the Cu contents of adakites have been attributed to the mixing of ~ 95% slab melt with
the asthenosphere (~ 30 ppm) (McDonough and Sun, 1995) and the ~ 5% sediment-derived melt in the Late Carboniferous close to a
continental crust (~27 ppm) (Rudnick and Gao, 2003) are much lower subducting spreading ridge (Tang et al., 2010). Therefore, more work
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 117

on porphyry dykes are needed to clarify the origin of the Baogutu por- from graphite–CO2 (CCO) equilibria in the mantle source to Fe3 +–
phyry deposit. Fe2+ equilibria as the magma ascends (Ballhaus, 1993). Consistently,
it becomes more oxidizing with decreasing depths and pressures
4.4. Summary (Stagno and Frost, 2010; Stagno et al., 2013). Moreover, most mantle
minerals, such as olivine, favor ferrous Fe over ferric Fe, and fractional
Reduced magmas are not favorable for porphyry mineralization. The crystallization of these minerals elevates the Fe3+/Fe2+ (Carmichael,
reduced porphyry deposits so far reported are either simply distal host 1991). Given that ferric Fe is highly incompatible, magma recharging
rocks located as far as several kilometers away from or above the caus- will further enrich the ferric Fe (Lee et al., in press), resulting in higher
ative porphyry buried underneath, or originally oxidized magmas that oxygen fugacities. In addition, assimilation of oxidizing country rocks
were reduced through assimilation of reducing components. Degassing (Lee et al., 2005), degassing of reduced volatile species (e.g., H2, H2S
of SO2 is not likely to be a main process for forming porphyry deposits and CH4) (Ballhaus, 1993; Lee et al., 2005) and lowering of pH values
because it cannot explain the features of reduced porphyries. Careful in a sulfate–sulfide dominated system (Sun et al., 2013a,b) may also el-
studies are needed to identify the causative porphyries. Oxidized evate the oxygen fugacity of the magmas. Based on similar Zn/FeT ratios
adakitic (slab melts) rocks are overwhelmingly the most favorable can- of mantle peridotite and primitive arc magmas, it has been argued that
didates for porphyry mineralizations. primary arc magmas are not necessarily oxidized (Lee et al., 2010). In-
creasing Zn/FeT ratios with decreasing MgO content argues that the
high oxygen fugacity of arc magmas is acquired through magma evolu-
5. Discussion
tion (Lee et al., 2010), indicating that continuing chemical evolution
during magma transport and emplacement may have major effects on
5.1. The oxygen fugacities at convergent margins
the oxidized characteristics of arc magmas. This may equally explain
the diversity in oxygen fugacities at convergent margins. Nevertheless,
The oxygen fugacity of the mantle and volcanic arc has been under
high oxygen fugacities developed during magma evolution may provide
study for a long time (Ballhaus, 1993; Carmichael, 1991; Cottrell and
little or no contribution to eliminating residual sulfides, and thus little
Kelley, 2013; Kelley and Cottrell, 2009; Lee et al., 2005, 2010;
contribution to porphyry mineralization. This decoupling may partially
Parkinson and Arculus, 1999). The consensus is that the oxygen fugacity
explain why most arc magmas are highly oxidized, whereas only a
of arc magmas is systematically higher than that of MORB (Fig. 16)
small portion of them form porphyry Cu deposits.
(Carmichael, 1991; Kelley and Cottrell, 2009; Parkinson and Arculus,
1999; Sun et al., 2012a, 2013a,b). Nevertheless, it is still hotly debated
5.2. The difference between porphyry and epithermal in terms of oxygen
as regards to how arc magmas get oxidized. As mentioned in
fugacity
Section 3, a variety of mechanisms have been proposed.
The most straightforward way to elevate the oxygen fugacity of arc
Most Cu porphyry deposits form at depths of 2–4 km, and are usually
magmas is by the addition of oxidizing components to the mantle
associated with epithermal deposits at shallow depths if not removed
wedge or directly to the magmas. Water is the most abundant volatile
by erosion (Figs. 22d, 23a, b) (Cooke et al., 2011; Hedenquist et al.,
component released during plate subduction. It has been proposed
1998; Heinrich, 2005; Heinrich et al., 2004; Hollings et al., 2005;
that H2O reacts with FeO, forming Fe2O3 with the release of H2
Sillitoe, 2010). This seemingly implies that porphyry and epithermal de-
(Eq. (30)) (Brandon and Draper, 1996; Kelley and Cottrell, 2009).
posits are closely related. Many epithermal deposits, however, are not
Most of the ferric Fe will be transferred into the melt because ferric Fe
linked to porphyry deposits. For example, epithermal deposits formed
is more incompatible than ferrous Fe (Lee et al., in press). This reaction,
in the mid-ocean ridge and backarc basins usually do not have any asso-
however, is not controlled by water. Instead, it is controlled by ferrous
ciated porphyry deposit. This is probably mainly because the crust in
Fe, which is more stable under high pressure. Recent experiments
backarc basins is too thin. Regardless, the oxygen fugacities of
show that water and H2 can coexist as two immiscible phases. This
epithermal deposits have a much larger ranges than porphyry deposits.
immiscibility implies that water is stable in the mantle under high
pressure (Bali et al., 2013).
5.2.1. Magnetite crisis
More accurately, the ferrous Fe in fluids reacts with water, releasing
The magnetite crisis (Jenner et al., 2010) refers to the dramatic de-
H2 and H+ (Eq. (31)). This is supported by the abundance of oxidized
creases in Cu and Au during magnetite crystallization in arc volcanic
components, e.g. magnetite–hematite and sulfates, found in the subduc-
rocks (Sun et al., 2004a), which is a common phenomenon (Moss
tion released fluids of ultrahigh pressure quartz veins, which may ele-
et al., 2001; Sun et al., 2011; Togashi and Terashima, 1997). This magne-
vate the oxygen fugacity of the mantle wedge (X.M. Sun et al., 2007).
tite crisis is taken as a main process that leads to the formation of ore-
It has also been proposed that slab melts are the most efficient way to
forming fluids, responsible for the Au and Cu mineralization of both
transfer high Fe3+/Fe2+ ratios responsible for high oxygen fugacity in
epithermal and porphyry deposits (Liang et al., 2009; Sun et al., 2004a,
adakites (Mungall, 2002) and arc magmas (Brandon and Draper, 1996).
2013a,b) (see also Section 3): Sulfate is reduced to sulfide during mag-
netite crystallization, forming hydrosulfide complexes, which scavenge
H2 O þ 2FeO ¼ Fe2 O3 þ H2 ð30Þ
chalcophile elements into fluids and subsequently transport the metals
to favorable places for forming deposits (Sun et al., 2004a).
2þ þ
3H2 O þ 2Fe ¼ Fe2 O3 þ H2 þ 4H ð31Þ This mineralization model has been challenged by some later
studies. Although using the same set of samples (Jenner et al., 2012),
The addition of other oxidized materials, e.g., Fe3+, C4+, and S6 +, and essentially repeating the results of the previous study (Sun et al.,
from subducted sediments and oceanic crust to the mantle wedge 2004a), it was proposed that magnetite fractionation triggers sulfide
may elevate the redox states of the mantle as well (Fig. 18) (Evans saturation (Jenner et al., 2010). This conclusion was drawn based main-
et al., 2012). Arc magmas may also get oxidized during their evolution ly on the behavior of Se, which was assumed as a proxy that follows S
and ascent by processes accompanying melting, crystallization, assimi- closely during magmatic evolution except that it is not lost during
lation, degassing, etc (Ballhaus, 1993; Lee et al., 2005, 2010). For exam- low-pressure (near sea-floor) degassing (Jenner et al., 2010). Fluid ex-
ple, graphite and diamond are stable in the deep mantle. During mantle traction, however, is different from degassing, and thus Se provides no
melting, CO2 is incompatible whereas C is compatible, and graphite may constraints on the process. In fact, as shown in the supplementary infor-
be consumed through oxidation melting (Stagno et al., 2013). Most of mation, sulfide is undersaturated in Manus glasses (Fig. 24) (Sun et al.,
the C in magma occurs as CO2, therefore the oxygen buffer changes 2004a).
118 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Fig. 22. Different models for the mineralization of porphyry deposits. After (A) Lee (2014): Mantle-derived magmas (red) intrude into the cold upper plate (black) of subduction zones,
generating crystallizing magma chambers. Thick continental arcs are more favorable for porphyry mineralization because of accumulation of copper-rich sulfide cumulates; (B) after
Wilkinson (2013), which also proposes that sulfide saturation and accumulation are very important for the formation of porphyry Cu deposits; (C) after Richards (2011b), illustrating
porphyry and epithermal mineralizations in arc and postcollisional settings; (D) after Richards (2013), highlighting features or processes that may result in supercharging of these systems
to form giant deposits.
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 119

Fig. 22 (continued).

There are indeed sulfide inclusions in mineral phenocryst from Ni and Pt are controlled by sulfide complexes in fluid. This may ex-
Manus glasses (Fig. 25) (Sun et al., 2004a). The sulfide phase contains plain why there is no Ni and Pt in most porphyry and epithermal
high Au and Ag contents in addition to Cu in sulfides, but no other deposits.
chalcophile elements, such as Ni, Re, and Pt (Jenner et al., 2010). This The lack of Re in these sulfides can be plausibly interpreted by the
was used to argue that the sulfide phase is crystalline rather than an im- high oxygen fugacity, under which most of the Re is Re6+, and thus be-
miscible sulfide melt (Jenner et al., 2010). haves as lithophile rather than chalcophile elements. As shown in
Crystallization of sulfide cannot explain this phenomenon, because Fig. 26, in contrast to Cu and Au (Sun et al., 2004a) and also V, Co, Pt,
Ni and Pt are even more chalcophile than Cu, with partition coefficients and Se (Jenner et al., 2010), Re concentrations in Manus glasses keep in-
between sulfide and silicate melts of several hundreds and more than creasing when magnetite starts to crystallize, and then decrease gradu-
20 thousand, respectively (Table 2). We propose that these sulfides ally (Sun et al., 2003a). This can best be explained by the reduction of
formed directly from magmatic fluids. As shown in Fig. 25, sulfides in Re6 + to Re4 + during magnetite crystallization. Correspondingly, Re
phenocrysts are associated with fluid inclusions (Sun et al., 2004a). changes from incompatible (Sun et al., 2003a,b,c), to compatible
It is very likely that the preferential enrichments of Cu and Au over (Mallmann and O'Neill, 2007).
120 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Fig. 23. Cartoons illustrate the relationship between some epithermal deposits with porphyry deposits underneath. Modified after: (A) Hedenquist and Lowenstern (1994) and
(B) Heinrich (2005). Note, not all epithermal deposits are associated with porphyry deposits.

5.2.2. Oxygen fugacity and open systems of ~ 500 (Keppler, 2010). Sulfides formed through sulfate reduction
In contrast to the dramatically elevated oxygen fugacities during are released into magmatic fluids, where they scavenge chalcophile
magnetite crisis, the oxygen fugacity of the Manus magmas did not elements (Jenner et al., 2010; Sun et al., 2004a). Meanwhile, the
change much, or even lessened slightly (Jenner et al., 2010) during the reactions of Eqs. (16) and (17) are driven to the right, promoting
crystallization of magnetite and sulfate reduction (Sun et al., 2004a). sulfate reduction.
This can be interpreted as a magmatic process, i.e., the magnetite crys- Hydrothermal magnetite and even hematite may also form in fluids
tallization and sulfate reduction recorded in glasses occurred during released from arc volcanic magmas, e.g., Manus glasses, releasing H+.
magma evolution. As mentioned in Section 3.2.2, magnetite crystalliza- Nevertheless, volcanic magma systems are much more open than por-
tion in magmas reduces sulfate without changing the pH values. The phyry systems, so that the formation of hydrothermal iron oxides does
solubility of sulfides in fluids is very high, with partition coefficients not necessarily affect the oxygen fugacity of the arc magmas.
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 121

sulfides. This explains why epithermal deposits are distributed in a vari-


ety of tectonic settings, ranging from mid-ocean ridges and backarc ba-
sins to arcs.

5.3. Adakite, slab melting, ridge subduction and porphyry Cu deposits

Adakite was initially named for rocks formed through partial melt-
ing of subducted young oceanic crust (b25 Ma, represented by mid-
ocean ridge basalt, MORB) (Defant and Drummond, 1990; Kay, 1978).
In contrast to most other rock types, adakite is defined by geochemical
compositions (e.g., SiO2 ≥ 56 wt.%, Al2O3 ≥ 15 wt.%, Y ≤ 18 ppm,
Yb ≤ 1.9 ppm and Sr ≥ 400 ppm) without detailed petrographic con-
straints. Therefore, (1) both eruptive and intrusive rocks can be classi-
fied as adakites; and (2) adakites may be produced simply by partial
melting of mafic rocks in the presence of garnet and absence of plagio-
clase. Different mechanisms have been proposed to produce adakites,
e.g., partial melting of the lower continental crust (Chung et al., 2003;
Gao et al., 2004; Guo et al., 2006; Xu et al., 2002, 2006; Zhang et al.,
2001b) or underplated new crust (Hou et al., 2009; Martin, 1999), or
Fig. 24. Sulfide contents in Manus submarine volcanic glasses, showing that sulfide is far
below saturation (after Sun et al., 2004a). Therefore, sulfide segregation induced by sulfide by fractional crystallization of normal arc magmas (Castillo, 2006;
saturation as proposed by previous authors (Jenner et al., 2010) is not a feasible way to ex- Macpherson et al., 2006; Richards and Kerrich, 2007). Given that the
plain the magnetite crisis. Filled and open circles represent volcanic glasses analyzed by oceanic crust is very different from continental crust, slab melts can be
different authors, filled triangles represent melt inclusions. distinguished from lower continental crust melts using geochemical
criteria (Ling et al., 2011; Liu et al., 2010; Sun et al., 2012a).
Given that volcanic systems are far more open, the initial Cu, Au con-
tents are not of critical importance to porphyry mineralization. Metals 5.3.1. Adakite and porphyry Cu deposits
can be leached out as far as there is enough water circulation. Therefore, Most porphyry Cu deposits are associated with adakites (Oyarzun
it does not require very high oxygen fugacities to eliminate residual et al., 2001; Sajona and Maury, 1998; Sun et al., 2011, 2012a,b, 2013a,
b; Thieblemont et al., 1997), but the association is not true vice versa.
Many adakites, e.g., those from the Dabie Mountains, are not mineral-
ized at all (Huang et al., 2008; Ling et al., 2013; Liu et al., 2012; Wang
et al., 2007a). Partial melting of thickened eclogitic lower continental
crust (Wang et al., 2006a,b, 2007a,b; Zhang et al., 2001a) and fractional
crystallization of garnet (Macpherson et al., 2006) or amphibole
(Richards and Kerrich, 2007) may also form high Sr/Y magmas. The
lower continental crust melts have much lower Cu abundance and
lower oxygen fugacity than subducting slabs, such that they are not fa-
vorable for forming porphyry Cu deposits. Fractional crystallization of
garnet and/or amphibole plays no positive role in Cu mineralization, ei-
ther (Sun et al., 2011, 2012a).
Adakite formation through slab melting, on the other hand, does
favor porphyry Cu mineralization (Mungall, 2002; Sajona and Maury,
1998; Sun et al., 2011, 2012a; Thieblemont et al., 1997). Several differ-
ent explanations for the fertility of adakitic slab melts have been
proposed:

Oxidized. It has been argued that slab melts might be unusually oxi-
dized and rich in sulfur (Oyarzun et al., 2001), due to high Fe3+ con-
tent from oxidative sea-floor alteration. As a consequence, this
elevated fO2 state causes oxidation of chalcophile metal-bearing sul-
fide phases in the mantle wedge, releasing metals to the silicate melt
phase (Mungall, 2002). However, although oxidation is crucial to
porphyry mineralization, adakites are not systematically more oxi-
dized than normal arc magmas (Ballhaus, 1993; X.M. Sun et al.,
2007; Sun et al., 2013a,b). Normal arc melts are not systematically
more enriched in Cu than MORB, either (Lee et al., 2012).
Water: It has been proposed that slab melts might be unusually
water rich (Sajona and Maury, 1998). High water contents in
magmas may suppress the crystallization of plagioclase, and pro-
mote the formation of amphibole, resulting in high Sr/Y signatures
Fig. 25. Images of sulfide inclusion in an olivine phenocryst of Manus glass under (Richards, 2011a, 2012). The problem with this model is that no
(A) reflected light and (B) and transparent light. Only one big sulfide globule is clearly studies have ever demonstrated that adakites are more hydrous
identified under reflected light, which seemingly indicates sulfide saturation. This grain than normal arc rocks. Moreover, the crystallization of amphibole
is actually associated with fluid inclusions as shown under transparent light. In addition,
plays no role in Cu mineralization. Copper is incompatible in most
there are several sulfide grains, which are all associated with fluid inclusions (Sun et al.,
2004a). The unique composition of sulfides in Manus glasses (low Ni, Pt, etc.) may be plau- major silicate minerals, but could be compatible in amphibole, de-
sibly interpreted by sulfides crystallized from magmatic fluids. pending on the composition of the magmas and amphibole (Sun
122 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Table 2
Summary of sulfide/silicate partition coefficients.

Cu Au Ag Ni Pt Re Mo References

212–278 – – 240–308 – – – Rajamani and Naldrett (1978)


1383 1.5E+4–1.9E+4 – 500–900 – – – Peach et al. (1990)
250–970 – – 540–4400 – – – Gaetani and Grove (1997)
330–1070 7E+3–1.3E+4 300–1680 210–1270 – – 0.20–14.48 Li and Audetat (2012, 2013)
1040–1624 – 853–1528 646–965 – – – Patten et al. (2013)
1050–1850 4.1E+03–1.12E+04 – – 2.41E+05–4.19E+05 361–960 – Mungall and Brenan (2014)

et al., 2012b, 2014a). Therefore, none of the arguments concerning magma in a non-erupting, closed-system pluton where sulfur
water abundance can be plausibly interpreted to explain an associa- might precipitate as hydrothermal sulfides and sulfates instead of
tion between adakites and porphyry deposits. Because of these un- being degassed as SO2 (Oyarzun et al., 2001, 2002). In addition, com-
certainties, it has been asserted that there is no single, obvious pression also results in uplifting and erosions, which are later favor-
reason why slab melts have an unusually high potential to form por- able for the exposure of porphyry deposits. The question is again
phyry deposits (Richards, 2013). why normal arc rocks in compressed environments do not form por-
Felsic: Adakite is formed by partial melting of basaltic rocks, such phyry deposits.
that they should logically be more felsic than peridotite melts High Cu contents. Oceanic crust has a much higher Cu abundance
(Sajona and Maury, 1998). It has been further proposed that adakites (~ 100 ppm) (Sun et al., 2003a,b) than the mantle (30 ppm)
might more readily crystallize as intrusive plutons because of their (McDonough and Sun, 1995) or the continental crust (~ 27 ppm)
viscous felsic nature, leading to the generation of a more efficient (Rudnick and Gao, 2003). It has been proposed that partial melting
crustal magmatic–hydrothermal systems (Sajona and Maury, of the subducted oceanic crust forms adakites with systematically
1998). In fact, adakites associated with porphyry Cu deposits are higher Cu initial contents—favorable for porphyry Cu mineralization
mostly intermediate in composition, not felsic. (Sun et al., 2011, 2012a,b). Previous studies, however, did not quan-
Compression. Adakites are often generated by flat subduction of titatively model the influence of oxygen fugacity, although its posi-
young oceanic crust with associated compressional stress in the tive effect has been emphasized (Sun et al., 2011, 2013a,b).
upper plate. Such an environment should be favorable for trapping
The effects of oxygen fugacity on normal arc magmas have been
nicely modeled (Fig. 3) (Lee et al., 2012). Normal arc rocks form by par-
tial melting of peridotite from the mantle wedge. The primitive mantle
contains ~250 ppm of S (McDonough and Sun, 1995), whereas depleted
mantle peridotite contains ~ 150 ppm of S (O'Neill and Mavrogenes,
2002). Partial melting of mantle peridotite can easily eliminate residual
sulfide even under reducing conditions, e.g., by ~20% partial melting at
ΔFMQ 0 (Fig. 3) (Lee et al., 2012). In contrast, oceanic crust has sulfur
abundances of over 1000 ppm (O'Neill and Mavrogenes, 2002), which
is about 4 times greater than the primitive mantle value (McDonough
and Sun, 1995). As discussed in Section 3, oxygen fugacities higher
than ΔFMQ + 2 are of critical importance to eliminate residual sulfides
during slab melting. However, at oxygen fugacities higher than ΔFMQ
+ 2, ~5–10% of partial melting is enough to eliminate residual sulfide
from the subducted oceanic crust, forming adakitic melts with high Cu
and S contents up to 2000 ppm and percent levels, respectively. Such el-
evated contents in the magma, of course, are consistent with the high
Cu, S contents present in ore bearing porphyries.

5.3.2. Ridge subduction and porphyry Cu deposits


More than half of the world porphyry Cu deposits are located along
the western coasts of the North and South American continents com-
prising the eastern Pacific margin. The total resources there are estimat-
ed to be N1.8 billion tons, accounting for about 60% of the world's total
Cu resource estimation (Mutschler et al., 2010). Twenty out of the
world's top 25 giant porphyry Cu deposits are located there. In contrast,
there are essentially no porphyry Cu deposits located along the north-
western Pacific margin, e.g., Japan.
Many large porphyry Cu–Au deposits are connected to subduction of
spreading and aseismic ridges (Fig. 27) (Cooke et al., 2005; Sun et al.,
2010). As discussed above, slab melting is usually involved in creating
magmas with high Cu contents and adequately high oxygen fugacity,
which are the two main controlling factors for porphyry Cu mineraliza-
Fig. 26. Diagrams of SiO2 versus Cu and Re. Note, in contrast to Cu, Re keeps increasing tion (Section 3). Subduction of young ridges, both spreading and
when Cu drops suddenly at the point when magnetite starts to crystallize. Rhenium is aseismic, in particular, produces adakites with high oxygen fugacity,
much less chalcophile than Cu under high oxygen fugacities. It is present mainly in the
form of Re6+ and acts as an incompatible element before magnetite crystallization and
making it the best geological process for porphyry Cu deposits (Sun
was gradually reduced to Re4+ during magnetite crystallization. et al., 2010, 2013a,b). There are several subducting ridges along the
Modified after Sun et al. (2003a, 2004a). east Pacific margin, e.g., in Chile and Peru in the South America. These
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 123

Fig. 27. Many large porphyry Cu deposits are closely associated with ridge subduction, because subduction of young ridges is the most favorable geologic process for slab melting in the
Phenozoic, forming highly oxidized melt with high initial Cu contents.
Modified after Sun et al. (2010).

ridges were mostly younger than 25 Ma when they began to subduct, melting in the Phanerozoic. Geochemical signatures of ridge subduction
and are closely associated with large porphyry Cu–Au deposits (Fig. 27). are important exploration targets for large porphyry Cu–Au deposits.
There are also several ridges (most of which are aseismic, i.e., island
chains) on the western Pacific plate (Fig. 28). Most of these ridges, how- 5.4. Alterations
ever, are older than 100 Ma (W.D. Sun et al., 2007), and are not likely to
form adakites. In addition, the oxygen fugacities in Japan, Izu–Bonin– Porphyry deposits have very well developed alteration zones that
Mariana, and other arcs along the northwestern Pacific margins, are sys- typically affect several cubic kilometers of rock (Lowell and Guilbert,
tematically lower than ΔFMQ + 2 (Fig. 16a). For these reasons, we find 1970; Sillitoe, 2010; Titley, 1981) (Fig. 30), which is of critical impor-
it not surprising that no economically viable porphyry deposits associat- tance to understanding porphyry mineralization processes and improv-
ed with volcanic arcs are known in the northwestern Pacific region. ing their exploration.
There are porphyry ore deposits located at the southwestern Pacific Alteration is mainly controlled by pH values of the ore-forming fluids
margin, which, however, are much less developed in terms of tonnage (Sillitoe, 2010). The amount of H+ released during mineralization
and the number of deposits. It is noteworthy that these deposits are as- (e.g., Eqs. (15), (16), (18), (19), (22), (23), (25)) together with alkali con-
sociated with the closure of backarc basins younger than 25 Ma tents in the porphyry together control advanced argillic lithocap
(Fig. 29). formation and alteration (Sillitoe, 2010). For example, sericitic and
Considering the geotherm's concave downward shape of subducting advanced argillic alteration are much less well developed in porphyry
slabs, ridge subduction is the most favorable tectonic setting for slab Cu deposits associated with alkaline than with calc-alkaline intrusions
124 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Fig. 28. Distributions of aseismic ridges (island chains) in the Western Pacific. These aseismic ridges are all much older than 25 Ma, such that do not get melted during subduction.
Modified after W.D. Sun et al. (2007).

Fig. 29. Ages of backarc basins in the southwestern Pacific. Subduction of young backarc basin crust forms adakite, which is favorable for porphyry mineralization.
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 125

þ þ
(Lang et al., 1995; Sillitoe and Burrows, 2002), reflecting control of the 3KAlSi3 O8 þ2H ¼ KAl2 Si3 AlO10 ðOHÞ2 þ2K þ 6SiO2 ð33Þ
K+/H+ ratio by magma chemistry (Sillitoe, 2010). As summarized recent- Potassium feldspar sericite

ly (Sun et al., 2014b), the main alteration reactions are Eqs. (32)–(37):
þ þ þ
3NaAlSi3 O8 þK þ 2H ¼ KAl2 ½AlSi3 O10 ðOHÞ2 þ6SiO2 þ 3Na ð34Þ
þ Sodiumfeldspar sericite
2KðMg; FeÞ3 AlSi3 O10 ðOHÞ2 þ4H
Biotite
2þ þ þ þ
¼ AlðMg; FeÞ5 AlSi3 O10 ðOHÞ8 þðMg; FeÞ þ 2K þ 3SiO2 ð32Þ 2KAl3 Si3 O10 ðOHÞ2 þ2H þ 3H2 O ¼ 3Al2 Si2 O5 ðOHÞ4 þ2K ð35Þ
chlorite Sericite kaolinite

Fig. 30. Alteration patterns of porphyry deposits. After: (A) Lowell and Guilbert (1970); (B) Sillitoe (2010) and (C), Richards (2011b).
126 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Fig. 30 (continued).

sericite (fine-grained muscovite) and/or illite, and magnetite to hema-


KFe3 AlSi3 O10 ðOHÞ2 þ1=2O2 ¼ KAlSi3 O8 þFe3 O4 þ H2 O ð36Þ
Biotite potassium feldspar tite (martite and/or specularite), along with deposition of pyrite and
chalcopyrite (Sillitoe, 2010).
Sericitic alteration in porphyry Cu deposits normally overprints the
potassic and chlorite–sericite assemblages. It may be subdivided into
CaAl2 Si2 O8 þ2KCl þ 4SiO2 ¼ 2KAlSi3 O8 þCaCl2 : ð37Þ
Anorthite potassium feldspar two different types—a less common, early greenish to greenish-gray in
color alteration, and a far more common white alteration. The sericitic
alteration is commonly pyrite dominated, implying effective removal
The alteration zone in porphyry Cu deposits starts from barren, early
of the Cu (±Au) present in the former chlorite–sericite and/or potassic
sodic–calcic upward through potentially ore-grade potassic, chlorite–
assemblages. It may also constitute ore with Cu either in the form of
sericite, and sericitic, to advanced argillic, and finally the lithocap
chalcopyrite or as high sulfidation-state assemblages (Sillitoe, 2010).
(Sillitoe, 2010). In general, the alteration–mineralization zones become
The lower portion of argillic lithocaps may overprint the upper parts
progressively younger upward (Fig. 30), consequently the shallower al-
of the porphyry Cu deposits, whereas the sericitic alteration transforms
teration–mineralization zones overprint deeper ones.
upwardly to quartz–pyrophyllite. The advanced argillic alteration pref-
Sodic–calcic alteration is typically sulfide and metal poor (except for
erentially affects lithologic units with low acid-buffering capacities
Fe as magnetite) but can host mineralization in Au-rich porphyry Cu de-
(Sillitoe, 2010).
posits. It is commonly located in the immediate wallrocks of the por-
phyry intrusion, or are found as a centrally located zone of some
porphyry Cu stocks (Sillitoe, 2010). 6. Conclusion
Potassic alteration is located in the center and deeper portions of the
porphyry. Dominant mineral changes from biotite in relatively mafic The key processes of porphyry mineralization are oxidation and re-
porphyry intrusions and host rocks, to K-feldspar in more felsic, grano- duction of sulfur. Most of the porphyry deposits are closely associated
dioritic to quartz monzonitic settings (Sillitoe, 2010). Quartz-K ± Na- with oxidized magmas, with sulfate as the dominant sulfur mineral
feldspar overprints may destroy the more typical potassic assemblages. species. Sulfur is one of the most important geosolvents that controls
The chalcopyrite ± bornite ore in many porphyry Cu deposits is largely the behaviors of copper and other chalcophile elements, therefore
confined to potassic zones. Potassic-altered wallrocks may attain N 1 km knowledge of its geochemical behavior is essential to the understanding
thickness. The potassic alteration generally becomes less intense from of mineralization processes for copper and a variety of other metal re-
the older to younger porphyry phases (Sillitoe, 2010). This is likely sources. Given that for most chalcophile elements, the partition coeffi-
due to a lowering of the pH as mineralization continues. cient between sulfide and melt is very high, elimination of residual
Chlorite–sericite alteration produces pale-green rocks and is wide- sulfide is essential for the extraction of chalcophile elements from the
spread in the shallower parts of some porphyry Cu deposits, source and thus the formation of porphyry deposits. The solubility of
overprinting preexisting potassic assemblages. The alteration is charac- sulfur depends strongly on sulfur speciation, which in turn depends
terized by transformation of mafic minerals to chlorite, plagioclase to on oxygen fugacities. Sulfate is over 10 times more soluble than sulfide.
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 127

At oxygen fugacities higher than ΔFMQ + 2, most of the sulfur in melts Astudillo, N., Roperch, P., Townley, B., Arriagada, C., Chauvin, A., 2010. Magnetic polarity
zonation within the El Teniente copper–molybdenum porphyry deposit, central
is present as sulfate, such that the solubility of sulfur increases from Chile. Mineral. Deposita 45 (1), 23–41.
~1300 ppm to 2 wt.%. Consequently, slab melts becomes sulfur under- Audetat, A., Pettke, T., Dolejs, D., 2004. Magmatic anhydrite and calcite in the ore-forming
saturated at N5% partial melting, with Cu contents of over 1000 ppm. quartz–monzodiorite magma at Santa Rita, New Mexico (USA): genetic constraints
on porphyry–Cu mineralization. Lithos 72 (3–4), 147–161.
Therefore, ΔFMQ + 2 is often considered the magic number for porphy- Baker, T., Ash, C.H., Thompson, J.F.H., 1997. Geological setting and characteristics of the
ry mineralization. Red Chris porphyry copper–gold deposit, northwestern British Columbia. Explor.
Metals of porphyry Cu deposits are hosted in sulfides, which require Min. Geol. 6 (4), 297–316.
Baker, M.J., Cooke, D.R., Hollings, P.N., 2011. LA-ICP-MS U–Pb zircon ages and Nd isotope data
reduction of sulfate to sulfide during the final stage of mineralization. In for the Teniente Plutonic Complex: implications for the evolution of the El Teniente
principle, sulfate can also be reduced by assimilation of reducing sedi- Porphyry Cu–Mo Deposit. Let's Talk Ore Deposits vols. I and Ii, (853–855 pp.).
ments or degassing of oxidizing gases. Porphyry deposits are usually Bali, E., Audetat, A., Keppler, H., 2013. Water and hydrogen are immiscible in Earth's man-
tle. Nature 495 (7440), 220–222.
mineralized throughout the whole pluton, whereas interactions with
Ballard, J.R., Palin, J.M., Williams, I.S., Campbell, I.H., Faunes, A., 2001. Two ages of porphy-
country rocks occur mainly at the interface, such that assimilation of re- ry intrusion resolved for the super-giant Chuquicamata copper deposit of northern
ducing sediments is not likely to be the controlling process. Degassing is Chile by ELA-ICP-MS and SHRIMP. Geology 29 (5), 383–386.
not the main process in deeply emplaced porphyry bodies, such that it is Ballard, J.R., Palin, J.M., Campbell, I.H., 2002. Relative oxidation states of magmas inferred
from Ce(IV)/Ce(III) in zircon: application to porphyry copper deposits of northern
not likely to be a major process for sulfate reduction, either. Ferrous iron Chile. Contrib. Mineral. Petrol. 144 (3), 347–364.
is the most important reductant that is responsible for sulfate reduction Ballhaus, C., 1993. Oxidation states of the lithospheric and asthenospheric upper mantle.
during porphyry mineralization. The highest oxygen fugacity favorable Contrib. Mineral. Petrol. 114, 331–348.
Barra, F., 2011. Assessing the longevity of porphyry Cu–Mo deposits: examples from the
for porphyry mineralization is the HM buffer. Otherwise, there is no fer- Chilean Andes. Let's Talk Ore Deposits vols. I and Ii, (145–147 pp.).
rous Fe in the system. The reduction of sulfate and oxidation of ferrous Beermann, O., Botcharnikov, R.E., Holtz, F., Diedrich, O., Nowak, M., 2011. Temperature de-
Fe lower the pH value. This, in turn, elevates the oxidation potential of pendence of sulfide and sulfate solubility in olivine-saturated basaltic magmas.
Geochim. Cosmochim. Acta 75 (23), 7612–7631.
sulfate, driving the oxygen fugacities up to the HM buffer. These Binder, B., Keppler, H., 2011. The oxidation state of sulfur in magmatic fluids. Earth Planet.
behaviors explain the popular occurrence of hypogene magnetite and Sci. Lett. 301, 190–198.
hematite (and specularite) in porphyry deposits. Brandon, A.D., Draper, D.S., 1996. Constraints on the origin of the oxidation state of
mantle overlying subduction zones: an example from Simcoe, Washington, USA.
Low pH fluids cause formation of pervasive alteration zones in por- Geochim. Cosmochim. Acta 60 (10), 1739–1749.
phyry Cu deposits, starting from barren, early sodic–calcic alteration up- Bryant, J.A., Yogodzinski, G.M., Churikova, T.G., 2007. Melt–mantle interactions beneath
ward through potentially ore-grade potassic, chlorite–sericite, and the Kamchatka arc: evidence from ultramafic xenoliths from Shiveluch volcano.
Geochem. Geophys. Geosyst. 8 (4), Q04007. http://dx.doi.org/10.1029/2006GC
sericitic alterations, to advanced argillic alterations, and finally forming
001443.
the lithocaps. The amount of H+ released during mineralization and the Burnham, C.W., Ohmoto, H., 1980. Late-stage processes of felsic magmatism. Min. Geol.
alkali content in the porphyry together control advanced argillic Spec. Issue 8, 1–12.
lithocap formation and alterations. Calagari, A.A., 2003. Stable isotope (S, O, H and C) studies of the phyllic and potassic–
phyllic alteration zones of the porphyry copper deposit at Sungun, East Azarbaidjan,
Hydrogen and methane form during the final mineralization process Iran. J. Asian Earth Sci. 21 (7), 767–780.
of porphyry deposits. Most of the hydrogen and methane should have Candela, P.A., 1992. Controls on ore metal ratios in granite-related ore systems — an ex-
been oxidized by ferric Fe. In special cases, some of the reduced gases perimental and computational approach. Trans. R. Soc. Edinb. Earth Sci. 83, 317–326.
Cannell, J., Cooke, D.R., Stein, H.J., Markey, R., 2003. New paragenetically constrained Re–
may escape from the system, and even get trapped in fluid inclusions. Os molybdenite ages for El Teniente Cu–Mo porphyry deposit, Central Chile. Mineral
Therefore, small amount of reduced gases in fluid inclusions cannot Exploration and Sustainable Development vols. 1 and 2, (255–258 pp.).
argue against the oxidized feature of the magmas. Cannell, J., Cooke, D.R., Walshe, J.L., Stein, H., 2005. Geology, mineralization, alteration, and
structural evolution of the El Teniente porphyry Cu–Mo deposit. Econ. Geol. 100 (5),
Reduced magmas are not favorable for porphyry mineralization. 979–1003.
There are indeed several small porphyry deposits that appear to be re- Cao, M., Qin, K., Li, G., Jin, L., Evans, N.J., Yang, X., 2014. Baogutu: an example of reduced
lated to reduced porphyries. In our opinion, the reduced porphyry de- porphyry Cu deposit in western Junggar. Ore Geol. Rev. 56, 159–180.
Carmichael, I.S.E., 1991. The redox states of basic and silicic magmas — a reflection of their
posits so far reported are either host rocks at distal sites as far as
source regions. Contrib. Mineral. Petrol. 106 (2), 129–141.
several kilometers away from or above the causative porphyry lying Castillo, P.R., 2006. An overview of adakite petrogenesis. Chin. Sci. Bull. 51 (3), 258–268.
deep underneath, or a consequence of fluid transportation. Some of Chen, Y.J., 2013. The development of continental collision metallogeny and its application.
Acta Petrol. Sin. 29 (1), 1–17.
the reduced porphyries were originally associated with oxidized
Chen, Y.J., Li, C., Zhang, J., Li, Z., Wang, H.H., 2000. Sr and O isotopic characteristics of por-
magmas but were reduced through assimilation of reducing compo- phyries in the Qinling molybdenum deposit belt and their implication to genetic
nents during emplacement. Degassing of SO2 is not likely to be a main mechanism and type. Sci. China. Ser. D Earth Sci. 43, 82–94.
process for porphyry deposit formation, and it doesn't reduce sulfate Chen, W., Xu, Z.W., Lu, X.C., Yang, X.N., Li, H.C., Qu, W.J., Chen, J.Q., Wang, H., Wang, S.H.,
2013. Petrogenesis of the Bao'anzhai granite and associated Mo mineralization, west-
to sulfide, either. Therefore, degassing of SO2 cannot explain the features ern Dabie orogen, east-central China: constraints from zircon U–Pb and molybdenite
of reduced porphyries. Given that a large portion of metals in a porphyry Re–Os dating, whole-rock geochemistry, and Sr–Nd–Pb–Hf isotopes. Int. Geol. Rev. 55
deposit are hosted in wall rocks, attention is needed to identify the caus- (10), 1220–1238.
Chiaradia, M., 2014. Copper enrichment in arc magmas controlled by overriding plate
ative porphyries. thickness. Nature Geosci. 7, 43–46.
Chiaradia, M., Ulianov, A., Kouzmanov, K., Beate, B., 2012. Why large porphyry Cu deposits
like high Sr/Y magmas? Sci. Rep. 2. http://dx.doi.org/10.1038/srep00685.
Chou, I.M., 1978. Calibration of oxygen buffers at elevated P and T using the hydrogen fu-
Acknowledgments gacity sensor. Am. Mineral. 63, 690–703.
Chou, I.M., 1986. Permeability of precious metals to hydrogen at 2 kbar total pressure and
We would like to thank Professor Cin-ty A. Lee for constructive elevated temperatures. Am. J. Sci. 286, 638–658.
Chung, S.L., Liu, D.Y., Ji, J.Q., Chu, M.F., Lee, H.Y., Wen, D.J., Lo, C.H., Lee, T.Y., Qian, Q., Zhang,
discussions. This study is supported by the National Natural Science
Q., 2003. Adakites from continental collision zones: melting of thickened lower crust
Foundation of China (nos. 41090374, 41121002, 41172080). This is con- beneath southern Tibet. Geology 31 (11), 1021–1024.
tribution no. IS-1939 from GIGCAS. Cole, R.B., Stewart, B.W., 2009. Continental margin volcanism at sites of spreading ridge
subduction: examples from southern Alaska and western California. Tectonophysics
464 (1–4), 118–136.
Cole, R.B., Nelson, S.W., Layer, P.W., Oswald, P.J., 2006. Eocene volcanism above a de-
References pleted mantle slab window in southern Alaska. Geol. Soc. Am. Bull. 118 (1–2),
140–158.
Abratis, M., Worner, G., 2001. Ridge collision, slab-window formation, and the flux of Cooke, D.R., Hollings, P., Walsh, J.L., 2005. Giant porphyry deposits: characteristics, distri-
Pacific asthenosphere into the Caribbean realm. Geology 29 (2), 127–130. bution, and tectonic controls. Econ. Geol. 100 (5), 801–818.
An, F., Zhu, Y.F., 2010. Native antimony in the Baogutu gold deposit (west Junggar, NW Cooke, D.R., Deyell, C.L., Waters, P.J., Gonzales, R.I., Zaw, K., 2011. Evidence for magmatic–
China): its occurrence and origin. Ore Geol. Rev. 37 (3–4), 214–223. hydrothermal fluids and ore-forming processes in epithermal and porphyry deposits
Arculus, R.J., 1994. Aspects of magma genesis in arcs. Lithos 33 (1–3), 189–208. of the Baguio District, Philippines. Econ. Geol. 106 (8), 1399–1424.
128 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Cottrell, E., Kelley, K.A., 2013. Redox heterogeneity in mid-ocean ridge basalts as a func- Hollings, P., Cooke, D., Clark, A., 2005. Regional geochemistry of tertiary igneous rocks in
tion of mantle source. Science 340 (6138), 1314–1317. central Chile: implications for the geodynamic environment of giant porphyry copper
Craig, H., 1961. Isotopic variations in meteoric waters. Science 133, 1702–1703. and epithermal gold mineralization. Econ. Geol. 100 (5), 887–904.
Crerar, D.A., Susak, N.J., Borcsik, M., Schwartz, S., 1978. Solubility of the buffer assemblage Hou, Z.Q., Gao, Y.F., Qu, X.M., Rui, Z.Y., Mo, X.X., 2004. Origin of adakitic intrusives gener-
pyrite–pyrrhotite–magnetite in NaCl solutions from 200 to 350 °C. Geochim. ated during mid-Miocene east–west extension in southern Tibet. Earth Planet. Sci.
Cosmochim. Acta 42, 1427–1437. Lett. 220 (1–2), 139–155.
Cuadra, P., Rojas, G., 2001. Oxide mineralization at the Radomiro Tomic porphyry copper Hou, Z.Q., Pan, X.F., Yang, Z.M., Qu, X.M., 2007a. Porphyry Cu–(Mo–Au) deposits no relat-
deposit, northern Chile. Econ. Geol. Bull. Soc. Econ. Geol. 96 (2), 387–400. ed to oceanic-slab subduction: examples from Chinese porphyry deposits in conti-
Defant, M.J., Drummond, M.S., 1990. Derivation of some modern arc magmas by melting nental settings. Geoscience 21 (2), 332–351 (in Chinese with English abstract).
of young subducted lithosphere. Nature 347, 662–665. Hou, Z.Q., Xie, Y.L., Xu, W.Y., Li, Y.Q., Zhu, X.K., Khin, Z., Beaudoin, G., Rui, Z.Y., Wei, H.A.,
Dilles, J.H., Tomlinson, A.J., Garcia, M., Alcota, H., 2011. The geology of the Fortuna Grano- Ciren, L., 2007b. Yulong deposit, eastern Tibet: a high-sulfidation Cu–Au porphyry
diorite Complex, Chuquicamata district, Northern Chile: relation to porphyry copper copper deposit in the eastern Indo-Asian collision zone. Int. Geol. Rev. 49 (3),
deposits. Let's Talk Ore Deposits. vols I and Ii, (398–400 pp.). 235–258.
Eugster, H.P., 1957. Heterogeneous reactions involving oxidation and reduction at high Hou, Z.Q., Yang, Z.M., Qu, X.M., Meng, X.J., Li, Z.Q., Beaudoin, G., Rui, Z.Y., Gao, Y.F., Zaw, K.,
pressures and temperatures. J. Chem. Phys. 26, 1160. 2009. The Miocene Gangdese porphyry copper belt generated during post-collisional
Eugster, H.P., Wones, D.R., 1962. Stability relations of the ferruginous biotite, annite. J. extension in the Tibetan Orogen. Ore Geol. Rev. 36 (1–3), 25–51.
Petrol. 3, 82–125. Huang, F., Li, S.G., Dong, F., He, Y.S., Chen, F.K., 2008. High-Mg adakitic rocks in the Dabie
Evans, K.A., Elburg, M.A., Kamenetsky, V.S., 2012. Oxidation state of subarc mantle. Geol- orogen, central China: implications for foundering mechanism of lower continental
ogy 40 (9), 783–786. crust. Chem. Geol. 255 (1–2), 1–13.
Field, C.W., Zhang, L., Dilles, J.H., Rye, R.O., Reed, M.H., 2005. Sulfur and oxygen isotopic Huebner, J.S., 1987. Use of gas mixtures at low pressure to specify oxygen and other fu-
record in sulfate and sulfide minerals of early, deep, pre-Main Stage porphyry Cu– gacities of furnace atmospheres. Hydrothermal Experimental Techniques, pp. 20–60.
Mo and late Main Stage base-metal mineral deposits, Butte district, Montana. Huebner, J.S., Sato, M., 1970. The oxygen fugacity–temperature relationships of manga-
Chem. Geol. 215 (1–4), 61–93. nese oxide and nickel oxide buffers. Am. Mineral. 55, 934–952.
Gaetani, G.A., Grove, T.L., 1997. Partitioning of moderately siderophile elements among Imai, A., 2001. Generation and evolution of ore fluids for porphyry Cu–Au mineralization
olivine, silicate melt, and sulfide melt: constraints on core formation in the Earth of the Santo Tomas II (Philex) deposit, Philippines. Resour. Geol. 51 (2), 71–96.
and Mars. Geochim. Cosmochim. Acta 61 (9), 1829–1846. Imai, A., 2002. Metallogenesis of porphyry Cu deposits of the western Luzon arc,
Gao, S., Rudnick, R.L., Yuan, H.L., Liu, X.M., Liu, Y.S., Xu, W.L., Ling, W.L., Ayers, J., Wang, X.C., Philippines: K–Ar ages, SO3 contents of microphenocrystic apatite and significance
Wang, Q.H., 2004. Recycling lower continental crust in the North China craton. Nature of intrusive rocks. Resour. Geol. 52 (2), 147–161.
432 (7019), 892–897. Imai, A., 2005. Evolution of hydrothermal system at the Dizon porphyry Cu–Au deposit,
Gao, Y.L., Zhang, J.M., Ye, H.S., Meng, F., Zhou, K., Gao, Y., 2010. Geological characteristics Zambales, Philippines. Resour. Geol. 55 (2), 73–90.
and molybdenite Re–Os isotopic dating of Shiyaogou porphyry molybdenum deposit Imai, A., Ohno, S., 2005. Primary ore mineral assemblage and fluid inclusion study of the
in the East Qinling. Acta Petrol. Sin. 26 (3), 729–739. Batu Hijau porphyry Cu–Au deposit, Sumbawa, Indonesia. Resour. Geol. 55 (3),
Garrido, I., Cembrano, J., Sina, A., Stedman, P., Yanez, G., 2002. High magma oxidation state 239–248.
and bulk crustal shortening: key factors in the genesis of Andean porphyry copper Imai, A., Listanco, E.L., Fujii, T., 1993. Petrologic and sulfur isotopic significance of highly
deposits, central Chile (31–34° S). Rev. Geol. Chile 29 (1), 43–54. oxidized and sulfur-rich magma of Mt-Pinatubo, Philippines. Geology 21 (8),
Giggenbach, W.F., 1992a. Isotopic shifts in waters from geothermal and volcanic systems 699–702.
along convergent plate boundaries and their origin. Earth Planet. Sci. Lett. 113 (4), Imai, A., Ohbuchi, Y., Tanaka, T., Morita, S., Yasunaga, K., 2007. Characteristics of porphyry
495–510. Cu mineralization at Waisoi (Namosi district), Viti Levu, Fiji. Resour. Geol. 57 (4),
Giggenbach, W.F., 1992b. SEG distinguished lecture — magma degassing and mineral de- 374–385.
position in hydrothermal systems along convergent plate boundaries. Econ. Geol. Ishihara, S., 1977. The magnetite-series and ilmenite-series granitic rock. Min. Geol. 27,
Bull. Soc. Econ. Geol. 87 (7), 1927–1944. 930–941.
Gonzalez-Partida, E., Levresse, G., Carrillo-Chavez, A., Cheilletz, A., Gasquet, D., Jones, D., Ishihara, S., 2004. The redox state of granitoids relative to tectonic setting and earth his-
2003. Paleocene adakite Au–Fe bearing rocks, Mezcala, Mexico: evidence from geo- tory: the magnetite–ilmenite series 30 years later. Trans. R. Soc. Edinb. Earth Sci. 95,
chemical characteristics. J. Geochem. Explor. 80 (1), 25–40. 23–33.
Green, T.H., Watson, E.B., 1982. Crystallization of apatite in natural magmas under high- Ishihara, S., Matsuhisa, Y., 1999. Oxygen isotopic constraints on the geneses of the Mio-
pressure, hydrous conditions, with particular reference to orogenic rock series. cene Outer Zone granitoids in Japan. Lithos 46 (3), 523–534.
Contrib. Mineral. Petrol. 79 (1), 96–105. Ishihara, S., Murakami, H., 2006. Fractionated ilmenite-series granites in Southwest Japan:
Groome, W.G., Thorkelson, D.J., 2009. The three-dimensional thermo-mechanical signature source magma for REE–Sn–W mineralizations. Resour. Geol. 56 (3), 245–256.
of ridge subduction and slab window migration. Tectonophysics 464 (1–4), 70–83. Ishihara, S., Sasaki, A., 1991. Ore-deposits related to granitic magmatism in Japan — a
Guo, F., Fan, W., Li, C., Zhang, S.-H., Zhao, Y., Song, B., Yang, Z., Hu, J.-M., Wu, H., 2006. Geo- magmatic viewpoint. Episodes 14 (3), 286–292.
chemistry of late Mesozoic adakites from the Sulu belt, eastern China: magma genesis Ishihara, S., Terashima, S., 1989. Carbon contents of the magnetite-series and ilmenite-
and implications for crustal recycling beneath continental collisional orogens: series granitoids in Japan. Geochem. J. 23 (1), 25–36.
Carboniferous granitic plutons from the northern margin of the North China Block: Jenkyns, H.C., 2010. Geochemistry of oceanic anoxic events. Geochem. Geophys. Geosyst.
implications for a late Palaeozoic active continental margin. Geol. Mag. 143 (1), 1–13. 11, Q03004. http://dx.doi.org/10.1029/2009GC002788.
Halter, W., Heinrich, C., Pettke, T., 2005. Magma evolution and the formation of porphyry Jenner, F.E., O'Neill, H.S.C., Arculus, R.J., Mavrogenes, J.A., 2010. The magnetite crisis in the
Cu–Au ore fluids: evidence from silicate and sulfide melt inclusions. Mineral. evolution of arc-related magmas and the initial concentration of Au, Ag and Cu.
Deposita 39 (8), 845–863. J. Petrol. 51 (12), 2445–2464.
Han, Y.G., Zhang, S.H., Pirajno, F., Zhou, X.W., Zhao, G.C., Qu, W.J., Liu, S.H., Zhang, J.M., Jenner, F.E., O'neill, H.S.C., Arculus, R.J., Mavrogenes, J.A., 2012. The magnetite crisis in the
Liang, H.B., Yang, K., 2013. U–Pb and Re–Os isotopic systematics and zircon Ce4+/ evolution of arc-related magmas and the initial concentration of Au, Ag and Cu (vol
Ce3+ ratios in the Shiyaogou Mo deposit in eastern Qinling, central China: insights 51, pg 2445, 2010). J. Petrol. 53 (5), 1089-1089.
into the oxidation state of granitoids and Mo (Au) mineralization. Ore Geol. Rev. Jiang, Y.H., Jiang, S.Y., Ling, H.F., Dai, B.Z., 2006. Low-degree melting of a metasomatized
55, 29–47. lithospheric mantle for the origin of Cenozoic Yulong monzogranite-porphyry, east
Haschke, M., Ahmadian, J., Murata, M., McDonald, I., 2010. Copper mineralization Tibet: geochemical and Sr–Nd–Pb–Hf isotopic constraints. Earth Planet. Sci. Lett.
prevented by arc-root delamination during Alpine–Himalayan collision in central 241 (3–4), 617–633.
Iran. Econ. Geol. 105 (4), 855–865. Jugo, P.J., 2009. Sulfur content at sulfide saturation in oxidized magmas. Geology 37 (5),
Hedenquist, J.W., Lowenstern, J.B., 1994. The role of magmas in the formation of hydro- 415–418.
thermal ore-deposits. Nature 370 (6490), 519–527. Jugo, P.J., Luth, R.W., Richards, J.P., 2005. Experimental data on the speciation of sulfur as a
Hedenquist, J.W., Arribas, A., Reynolds, T.J., 1998. Evolution of an intrusion-centered function of oxygen fugacity in basaltic melts. Geochim. Cosmochim. Acta 69 (2),
hydrothermal system: Far Southeast–Lepanto porphyry and epithermal Cu–Au 497–503.
deposits, Philippines. Econ. Geol. Bull. Soc. Econ. Geol. 93 (4), 373–404. Jugo, P.J., Wilke, M., Botcharnikov, R.E., 2010. Sulfur K-edge XANES analysis of natural and
Heinrich, C.A., 1990. The chemistry of hydrothermal tin (–tungsten) ore deposition. Econ. synthetic basaltic glasses: implications for S speciation and S content as function of
Geol. 85, 457–481. oxygen fugacity. Geochim. Cosmochim. Acta 74 (20), 5926–5938.
Heinrich, C.A., 2005. The physical and chemical evolution of low-salinity magmatic fluids Kavalieris, I., Rinchin, O., Gombojav, J., Gombosuren, N., Crane, D., Orssich, C., Bat-Erdene,
at the porphyry to epithermal transition: a thermodynamic study. Mineral. Deposita K., 2011. Characteristics of the Oyu Tolgoi porphyry deposits, South Gobi, Mongolia.
39, 864–889. Let's Talk Ore Deposits vols. I and Ii, (342–344 pp.).
Heinrich, C.A., Driesner, T., Stefansson, A., Seward, T.M., 2004. Magmatic vapor contraction Kay, R.W., 1978. Aleutian magnesian andesite: melts from subducted Pacific Ocean crust.
and the transport of gold from the porphyry environment to epithermal ore deposits. J. Volcanol. Geotherm. Res. 4, 117–132.
Geology 32 (9), 761–764. Kay, S.M., Godoy, E., Kurtz, A., 2005. Episodic arc migration, crustal thickening, subduction
Hemingway, B.S., 1990. Thermodynamic properties for bunsenite, NiO, magnetite, Fe3O4, erosion, and magmatism in the south-central Andes. Geol. Soc. Am. Bull. 117 (1–2),
and hematite, Fe2O3 with comments on selected oxygen buffer reactions. Am. Miner- 67–88.
al. 75, 781–790. Kelley, K.A., Cottrell, E., 2009. Water and the oxidation state of subduction zone magmas.
Hezarkhani, A., 2006. Hydrothermal evolution of the Sar–Cheshmeh porphyry Cu–Mo de- Science 325 (5940), 605–607.
posit, Iran: evidence from fluid inclusions. J. Asian Earth Sci. 28 (4–6), 409–422. Kelley, K.A., Cottrell, E., 2012. The influence of magmatic differentiation on the oxidation
Hoefs, J. (Ed.), 2004. Stable Isotope Geochemistry. Springer, Germany. state of Fe in a basaltic arc magma. Earth Planet. Sci. Lett. 329, 109–121.
Hofmann, A.W., 1988. Chemical differentiation of the Earth: the relationship between Keppler, H., 2010. The distribution of sulfur between haplogranitic melts and aqueous
mantle, continental crust, and oceanic crust. Earth Planet. Sci. Lett. 90, 297–314. fluids. Geochim. Cosmochim. Acta 74 (2), 645–660.
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 129

Kesler, S.E., 1997. Metallogenic evolution of convergent margins: selected ore deposit Liang, H.Y., Campbell, I.H., Allen, C.M., Sun, W.D., Yu, H.X., Xie, Y.W., Zhang, Y.Q.,
models. Ore Geol. Rev. 12 (3), 153–171. 2007. The age of the potassic alkaline igneous rocks along the Ailao Shan–Red
Khashgerel, B.E., Kavalieris, I., Hayashi, K., 2008. Mineralogy, textures, and whole-rock River shear zone: implications for the onset age of left-lateral shearing. J. Geol.
geochemistry of advanced argillic alteration: Hugo Dummett porphyry Cu-Au depos- 115 (2), 231–242.
it, Oyu Tolgoi mineral district. Mongolia. Mineral. Deposita 43, 913–932. Liang, H.Y., Sun, W.D., Su, W.C., Zartman, R.E., 2009. Porphyry copper–gold mineralization
Kinoshita, O., 1997. Slab window-related magmatism caused by the Kula–Pacific ridge at Yulong, China, promoted by decreasing redox potential during magnetite alter-
subduction beneath the Eurasia continent in the Cretaceous. Episodes 20 (3), ation. Econ. Geol. 104 (4), 587–596.
185–187. Ling, M.X., Wang, F.Y., Ding, X., Hu, Y.H., Zhou, J.B., Zartman, R.E., Yang, X.Y., Sun, W.D.,
Kishima, N., 1989. A thermodynamic study on the pyrite–pyrrhotite–magnetite–water 2009. Cretaceous ridge subduction along the Lower Yangtse River Belt, Eastern
system at 300–500 °C with relevance to the fugacity/concentration quotient of aque- China. Econ. Geol. 104 (2), 303–321.
ous H2S. Geochim. Cosmochim. Acta 53, 2143–2155. Ling, M.X., Wang, F.Y., Ding, X., Zhou, J.B., Sun, W.D., 2011. Different origins of adakites
Klemm, L.M., Pettke, T., Heinbicth, C.A., Campos, E., 2007. Hydrothermal evolution of the from the Dabie Mountains and the Lower Yangtze River Belt, eastern China:
El Teniente deposit, Chile: Porphyry Cu–Mo ore deposition from low-salinity mag- geochemical constraints. Int. Geol. Rev. 53 (5–6), 727–740.
matic fluids. Econ. Geol. 102 (6), 1021–1045. Ling, M.X., Li, Y., Ding, X., Teng, F.Z., Yang, X.Y., Fan, W.M., Xu, Y.G., Sun, W.D., 2013. Destruc-
Klemm, L.M., Pettke, T., Heinrich, C.A., 2008. Fluid and source magma evolution of tion of the North China Craton induced by ridge subductions. J. Geol. 121 (2), 197–213.
the Questa porphyry Mo deposit, New Mexico, USA. Mineral. Deposita 43 (5), Liu, L.J., Spasojevic, S., Gurnis, M., 2008. Reconstructing Farallon plate subduction beneath
533–552. North America back to the Late Cretaceous. Science 322 (5903), 934–938.
Kress, V.C., Carmichael, I.S.E., 1991. The compressibility of silicate liquids containing Fe2O3 Liu, S.A., Li, S.G., He, Y.S., Huang, F., 2010. Geochemical contrasts between early Cretaceous
and the effect of composition, temperature, oxygen fugacity and pressure on their ore-bearing and ore-barren high-Mg adakites in central-eastern China: implications
redox states. Contrib. Mineral. Petrol. 108, 82–92. for petrogenesis and Cu–Au mineralization. Geochim. Cosmochim. Acta 74,
Lang, J.R., Lueck, B., Mortensen, J.K., Russell, J.K., Stanley, C.R., Thompson, J.F.H., 1995. 7160–7178.
Triassic–Jurassic silica-undersaturated and silica-saturated alkalic intrusions in the Liu, X., Fan, H.R., Hu, F.F., Hu, B.G., Zhu, X.Y., 2011. SEM-EDS investigation of daughter min-
cordillera of British-Columbia — implications for arc magmatism. Geology 23 (5), erals of fluid inclusions at the Dexing porphyry Cu–Mo deposit, Jiangxi Province,
451–454. China. Acta Petrol. Sin. 27 (5), 1397–1409.
Lee, C.-T.A., 2014. Copper conundrums. Nat. Geosci. 7, 10–11. Liu, S.A., Li, S.G., Guo, S.S., Hou, Z.H., He, Y.S., 2012. The Cretaceous adakitic–basaltic–
Lee, C.T.A., Leeman, W.P., Canil, D., Li, Z.X.A., 2005. Similar V/Sc systematics in MORB and granitic magma sequence on south-eastern margin of the North China Craton: impli-
arc basalts: implications for the oxygen fugacities of their mantle source regions. J. cations for lithospheric thinning mechanism. Lithos 134, 163–178.
Petrol. 46 (11), 2313–2336. Liu, X., Xiong, X., Audétat, A., Li, Y., Song, M., Li, L., Sun, W., Ding, X., 2014. Partitioning of
Lee, C.T.A., Luffi, P., Le Roux, V., Dasgupta, R., Albarede, F., Leeman, W.P., 2010. The redox copper between olivine, orthopyroxene, clinopyroxene, spinel, garnet and silicate
state of arc mantle using Zn/Fe systematics. Nature 468 (7324), 681–685. melts at upper mantle conditions. Geochim. Cosmochim. Acta 125, 1–22.
Lee, C.T.A., Luffi, P., Chin, E.J., Bouchet, R., Dasgupta, R., Morton, D.M., Le Roux, V., Yin, Q.Z., Lorand, J.P., 1990. Are spinel lherzolite xenoliths representative of the abundance of sulfur
Jin, D., 2012. Copper systematics in arc magmas and implications for crust–mantle in the upper mantle. Geochim. Cosmochim. Acta 54 (5), 1487–1492.
differentiation. Science 336 (6077), 64–68. Lowell, J.D., Guilbert, J.M., 1970. Lateral and vertical alteration–mineralization zoning in
Lee, C.-T.A., Lee, T.C., Wu, C.-T., 2014. Modeling the compositional evolution of recharging, porphyry ore deposits. Econ. Geol. 65 (4) (373-).
evacuating, and fractionating (REFC) magma chambers: implications for differentia- Lynch, G., Ortega, S., 1997. Hydrothermal alteration and tourmaline–albite equilibria at
tion of arc magmas. Geochim. Cosmochim. Acta http://dx.doi.org/10.1016/j.gca. the Coxheath porphyry Cu–Mo–Au deposit, Nova Scotia. Can. Mineral. 35, 79–94.
2013.08.009 (in press). Macpherson, C.G., Dreher, S.T., Thirlwall, M.F., 2006. Adakites without slab melting: high
Lerchbaumer, L., Audetat, A., 2012. High Cu concentrations in vapor-type fluid inclusions: pressure differentiation of island arc magma, Mindanao, the Philippines. Earth Planet.
an artifact? Geochim. Cosmochim. Acta 88, 255–274. Sci. Lett. 243 (3–4), 581–593.
Li, Y., Audetat, A., 2012. Partitioning of V, Mn, Co, Ni, Cu, Zn, As, Mo, Ag, Sn, Sb, W, Au, Pb, Madsen, J.K., Thorkelson, D.J., Friedman, R.M., Marshall, D.D., 2006. Cenozoic to Recent
and Bi between sulfide phases and hydrous basanite melt at upper mantle conditions. plate configurations in the Pacific Basin: ridge subduction and slab window
Earth Planet. Sci. Lett. 355, 327–340. magmatism in western North America. Geosphere 2 (1), 11–34.
Li, Y., Audetat, A., 2013. Gold solubility and partitioning between sulfide liquid, Mallmann, G., O'Neill, H.S.C., 2007. The effect of oxygen fugacity on the partitioning of Re
monosulfide solid solution and hydrous mantle melts: implications for the formation between crystals and silicate melt during mantle melting. Geochim. Cosmochim. Acta
of Au-rich magmas and crust–mantle differentiation. Geochim. Cosmochim. Acta 118, 71 (11), 2837–2857.
247–262. Mao, J.W., Xie, G.Q., Bierlein, F., Qu, W.J., Du, A.D., Ye, H.S., Pirajno, F., Li, H.M., Guo, B.J., Li,
Li, X.F., Sasaki, M., 2007. Hydrothermal alteration and mineralization of middle Jurassic Y.F., Yang, Z.Q., 2008. Tectonic implications from Re–Os dating of Mesozoic molybde-
Dexing porphyry Cu–Mo deposit, southeast China. Resour. Geol. 57 (4), 409–426. num deposits in the East Qinling–Dabie orogenic belt. Geochim. Cosmochim. Acta 72
Li, X.F., Watanabe, Y., Mao, J.W., Liu, S.X., Yi, X.K., 2007. Sensitive high-resolution ion mi- (18), 4607–4626.
croprobe U–Pb Zircon and 40Ar–39Ar muscovite ages of the Yinshan deposit in the Mao, J.W., Pirajno, F., Xiang, J.F., Gao, J.J., Ye, H.S., Li, Y.F., Guo, B.J., 2011. Mesozoic molyb-
northeast Jiangxi province, South China. Resour. Geol. 57 (3), 325–337. denum deposits in the east Qinling–Dabie orogenic belt: characteristics and tectonic
Li, J.W., Zhao, X.F., Zhou, M.F., Vasconcelos, P., Ma, C.Q., Deng, X.D., de Souza, Z.S., Zhao, Y. settings. Ore Geol. Rev. 43 (1), 264–293.
X., Wu, G., 2008. Origin of the Tongshankou porphyry-skarn Cu–Mo deposit, eastern Martin, H., 1999. Adakitic magmas: modern analogues of Archaean granitoids. Lithos 46
Yangtze craton, Eastern China: geochronological, geochemical, and Sr–Nd–Hf isotopic (3), 411–429.
constraints. Mineral. Deposita 43 (3), 315–336. Mathur, R., Ruiz, J., Titley, S., Gibbins, S., Margotomo, W., 2000. Different crustal sources
Li, H., Zhang, H., Ling, M.X., Wang, F.Y., Ding, X., Zhou, J.B., Yang, X.Y., Tu, X.L., Sun, W.D., for Au-rich and Au-poor ores of the Grasberg Cu–Au porphyry deposit. Earth Planet.
2011. Geochemical and zircon U–Pb study of the Huangmeijian A-type granite: impli- Sci. Lett. 183 (1–2), 7–14.
cations for geological evolution of the Lower Yangtze River belt. Int. Geol. Rev. 53 Mavrogenes, J.A., O'Neill, H.S.C., 1999. The relative effects of pressure, temperature and
(5–6), 499–525. oxygen fugacity on the solubility of sulfide in mafic magmas. Geochim. Cosmochim.
Li, C.Y., Wang, F.Y., Hao, X.L., Zhang, H., Ling, M.X., Zhou, J.B., Li, Y.L., Fan, W.M., Sun, W.D., Acta 63 (7–8), 1173–1180.
2012a. Formation of the world's largest molybdenum metallogenic belt: a plate- McDonough, W.F., Sun, S.S., 1995. The composition of the Earth. Chem. Geol. 120 (3–4),
tectonic perspective on the Qinling molybdenum deposits. Int. Geol. Rev. 54 (9), 223–253.
1093–1112. Moss, R., Scott, S.D., Binns, R.A., 2001. Gold content of eastern Manus basin volcanic rocks:
Li, C.Y., Zhang, H., Wang, F.Y., Liu, J.Q., Sun, Y.L., Hao, X.L., Li, Y.L., Sun, W.D., 2012b. The for- implications for enrichment in associated hydrothermal precipitates. Econ. Geol. Bull.
mation of the Dabaoshan porphyry molybdenum deposit induced by slab rollback. Soc. Econ. Geol. 96 (1), 91–107.
Lithos 150, 101–110. Mungall, J.E., 2002. Roasting the mantle: slab melting and the genesis of major Au and Au-
Li, H., Ling, M.X., Li, C.Y., Zhang, H., Ding, X., Yang, X.Y., Fan, W.M., Li, Y.L., Sun, W.D., 2012c. rich Cu deposits. Geology 30 (10), 915–918.
A-type granite belts of two chemical subgroups in central eastern China: indication of Mungall, J.E., Brenan, J.M., 2014. Partitioning of platinum-group elements and Au between
ridge subduction. Lithos 150, 26–36. sulfide liquid and basalt and the origins of mantle–crust fractionation of the
Li, H.Y., Wang, X.X., Ye, H.S., Yang, L., 2012d. Emplacement ages and petrogenesis of the chalcophile elements. Geochim. Cosmochim. Acta 125, 265–289.
molybdenum-bearing granites in the Jinduicheng area of East Qinling, China: con- Mutschler, F.E., Ludington, S., Bookstrom, A.A., 2010. Giant porphyryrelated metal camps
straints from zircon U–Pb ages and Hf isotopes. Acta Geol. Sin. Eng. Ed. 86 (3), of the world—a database. USGS Open-File Report: 99-556. , (6 pp. (http://geopubs.wr.
661–679. usgs.gov/open-file/of99-556/)).
Li, J.L., Gao, J., John, T., Klemd, R., Su, W., 2013a. Fluid-mediated metal transport in subduc- Myers, J., Eugster, H., 1983. The system Fe–Si–O: oxygen buffer calibrations to 1500 K.
tion zones and its link to arc-related giant ore deposits: constraints from a sulfide- Contrib. Mineral. Petrol. 82, 75–90.
bearing HP vein in lawsonite eclogite (Tianshan, China). Geochim. Cosmochim. Acta O'Neil, H.St.C., Pownceby, M.I., 1993. Thermodynamic data from redox reactions at high
120, 326–362. temperatures. Ι. An experimental and theoretical assessment. Contrib. Mineral. Petrol.
Li, N., Chen, Y.J., Pirajno, F., Ni, Z.Y., 2013b. Timing of the Yuchiling giant porphyry Mo sys- 114, 296–314.
tem, and implications for ore genesis. Mineral. Deposita 48 (4), 505–524. O'Neill, H.S.C., Mavrogenes, J.A., 2002. The sulfide capacity and the sulfur content at sul-
Li, X.F., Hu, R.Z., Rusk, B., Xiao, R., Wang, C.Y., Yang, F., 2013c. U–Pb and Ar–Ar geochronol- fide saturation of silicate melts at 1400 °C and 1 bar. J. Petrol. 43 (6), 1049–1087.
ogy of the Fujiawu porphyry Cu–Mo deposit, Dexing district, Southeast China: impli- Ossandon, G., Freraut, R., Gustafson, L.B., Lindsay, D.D., Zentilli, M., 2001. Geology of the
cations for magmatism, hydrothermal alteration, and mineralization. J. Asian Earth Chuquicamata mine: a progress report. Econ. Geol. Bull. Soc. Econ. Geol. 96 (2),
Sci. 74, 330–342. 249–270.
Liang, H.Y., Campbell, I.H., Allen, C., Sun, W.D., Liu, C.Q., Yu, H.X., Xie, Y.W., Zhang, Y.Q., Oyarzun, R., Marquez, A., Lillo, J., Lopez, I., Rivera, S., 2001. Giant versus small porphyry
2006. Zircon Ce4+/Ce3+ ratios and ages for Yulong ore-bearing porphyries in eastern copper deposits of Cenozoic age in northern Chile: adakitic versus normal calc-
Tibet. Mineral. Deposita 41 (2), 152–159. alkaline magmatism. Mineral. Deposita 36 (8), 794–798.
130 W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131

Oyarzun, R., Marquez, A., Lillo, J., Lopez, I., Rivera, S., 2002. Reply to discussion on “Giant Shen, P., Shen, Y., Pan, H., Li, X.H., Dong, L., Wang, J., Zhu, H., Dai, H., Guan, W., 2012.
versus small porphyry copper deposits of Cenozoic age in northern Chile: adakitic Geochronology and isotope geochemistry of the Baogutu porphyry copper deposit
versus normal calc-alkaline magmatism” by Oyarzun R, Marquez A, Lillo J, Lopez I, in the West Junggar region, Xinjiang, China. J. Asian Earth Sci. 49, 99–115.
Rivera S (Mineralium Deposita 36: 794–798, 2001). Mineral. Deposita 37 (8), Shen, P., Pan, H.D., Xiao, W.J., Chen, X.H., Eleonorad, S., Shen, Y.C., 2013a. Two
795–799. geodynamic–metallogenic events in the Balkhash (Kazakhstan) and the West
Parkinson, I.J., Arculus, R.J., 1999. The redox state of subduction zones: insights from arc- Junggar (China): Carboniferous porphyry Cu and Permian greisen W–Mo mineraliza-
peridotites. Chem. Geol. 160 (4), 409–423. tion. Int. Geol. Rev. 55 (13), 1660–1687.
Patricio, C.C., Gonzalo, R.S., 2001. Oxide mineralization at the Radomiro Tomic porphyry Shen, P., Pan, H.D., Xiao, W.J., Li, X.H., Dai, H.W., Zhu, H.P., 2013b. Early Carboniferous
copper deposit, northern Chile. Econ. Geol. 96, 387–400. intra-oceanic arc and back-arc basin system in the West Junggar, NW China. Int.
Patten, C., Barnes, S.J., Mathez, E.A., Jenner, F.E., 2013. Partition coefficients of chalcophile Geol. Rev. 55 (16), 1991–2007.
elements between sulfide and silicate melts and the early crystallization history of Shi, P.F., 1992. Fluid fugacities and phase equilibria in the Fe–Si–O–H–S system. Am. Min-
sulfide liquid: LA-ICP-MS analysis of MORB sulfide droplets. Chem. Geol. 358, eral. 77, 1050–1066.
170–188. Sillitoe, R.H., 1999. Styles of high-sulphidation gold, silver and copper mineralisation in
Peach, C.L., Mathez, E.A., Keays, R.R., 1990. Sulfide melt silicate melt distribution coeffi- porphyry and epithermal environments. Pacrim'99: International Congress on
cients for noble-metals and other chalcophile elements as deduced from MORB — Earth Science, Exploration and Mining Around the Pacific Rim, Proceedings. 99
implications for partial melting. Geochim. Cosmochim. Acta 54 (12), 3379–3389. (29–44 pp.).
Peng, G.Y., Luhr, J.F., McGee, J.J., 1997. Factors controlling sulfur concentrations in volcanic Sillitoe, R.H., 2010. Porphyry copper systems. Econ. Geol. 105 (1), 3–41.
apatite. Am. Mineral. 82 (11–12), 1210–1224. Sillitoe, R.H., Burrows, D.R., 2002. New field evidence bearing on the origin of the El Laco
Pettke, T., Oberli, F., Heinrich, C.A., 2010. The magma and metal source of giant porphyry- magnetite deposit, northern Chile. Econ. Geol. Bull. Soc. Econ. Geol. 97 (5),
type ore deposits, based on lead isotope microanalysis of individual fluid inclusions. 1101–1109.
Earth Planet. Sci. Lett. 296 (3–4), 267–277. Singer, D.A., 2008. Mineral deposit densities for estimating mineral resources. Math.
Pokrovski, G.S., Dubrovinsky, L.S., 2011. The S− 3 ion is stable in geological fluids at elevated Geosci. 40 (1), 33–46.
temperatures and pressures. Science 331 (6020), 1052–1054. Smith, C.M., Canil, D., Rowins, S.M., Friedman, R., 2012. Reduced granitic magmas in an arc
Qu, X., Hou, Z., Li, Y., 2004. Melt components derived from a subducted slab in late oro- setting: the Catface porphyry Cu–Mo deposit of the Paleogene Cascade Arc. Lithos
genic ore-bearing porphyries in the Gangdese copper belt, southern Tibetan plateau. 154, 361–373.
Lithos 74 (3–4), 131–148. Smithson, D.M., Rowins, S.M., 2005. Reduced I-type magmatism and porphyry Cu–Au
Rabbia, O.M., Hernandez, L.B., French, D.H., King, R.W., Ayers, J.C., 2009. The El Teniente mineralization in the west Central Cascades, WA: the ca. 37 Ma North Fork Deposit.
porphyry Cu–Mo deposit from a hydrothermal rutile perspective. Mineral. Deposita Geochim. Cosmochim. Acta 69 (10), A566-A566.
44 (8), 849–866. Song, S.G., Su, L., Niu, Y.L., Lai, Y., Zhang, L.F., 2009. CH4 inclusions in orogenic harzburgite:
Rajamani, V., Naldrett, A.J., 1978. Partitioning of Fe, Co, Ni, and Cu between sulfide liquid evidence for reduced slab fluids and implication for redox melting in mantle wedge.
and basaltic melts and composition of Ni–Cu sulfide deposits. Econ. Geol. 73 (1), Geochim. Cosmochim. Acta 73 (6), 1737–1754.
82–93. Sotnikov, V.I., Ponomarchuk, V.A., Pertseva, A.P., Berzina, A.P., Berzina, A.N., Girnon, V.O.,
Richards, J.P., 1999. Tectonomagmatic controls on localization of porphyry copper de- 2004. Evolution of sulfur isotopes in porphyry Cu–Mo ore-magmatic systems of
posits, Chile. Mineral Deposits: Processes to Processing vols. 1 and 2, (425–428 pp.). Siberia and Mongolia. Geol. Geofiz. 45 (8), 963–974.
Richards, J.P., 2011a. High Sr/Y arc magmas and porphyry Cu ± Mo ± Au deposits: just Spry, P.G., Scott, S.D., 1986. The stability of zincian spinels in sulfide systems and their po-
add water. Econ. Geol. 106 (7), 1075–1081. tential as exploration guides for metamorphosed massive sulfide deposits. Econ. Geol.
Richards, J.P., 2011b. Magmatic to hydrothermal metal fluxes in convergent and collided 81, 1446–1461.
margins. Ore Geol. Rev. 40 (1), 1–26. Spry, P.G., Paredes, M.M., Foster, F., Truckle, J.S., Chadwick, T.H., 1996. Evidence for a ge-
Richards, J.P., 2012. Discussion of Sun et al. (2011): ‘The genetic association of adakites netic link between gold–silver telluride and porphyry molybdenum mineralization
and Cu–Au ore deposits’. Int. Geol. Rev. 368–369. at the Golden Sunlight deposit, Whitehall, Montana: fluid inclusion and stable isotope
Richards, J.P., 2013. Giant ore deposits formed by optimal alignments and combinations of studies. Econ. Geol. Bull. Soc. Econ. Geol. 91 (3), 507–526.
geological processes. Nat. Geosci. 6 (11), 911–916. Stagno, V., Frost, D.J., 2010. Carbon speciation in the asthenosphere: experimental mea-
Richards, J.P., 2014. Discussion of : the link between reduced porphyry copper deposits surements of the redox conditions at which carbonate-bearing melts coexist with
and oxidized magmas. Geochim. Cosmochim. Acta 126, 643–645. graphite or diamond in peridotite assemblages. Earth Planet. Sci. Lett. 300 (1–2),
Richards, J.P., Kerrich, R., 2007. Adakite-like rocks: their diverse origins and questionable 72–84.
role in metallogenesis. Econ. Geol. 102 (4), 537–576. Stagno, V., Ojwang, D.O., McCammon, C.A., Frost, D.J., 2013. The oxidation state of the
Richards, J.P., Mumin, A.H., 2013. Magmatic–hydrothermal processes within an evolving mantle and the extraction of carbon from Earth's interior. Nature 493 (7430)
Earth: iron oxide–copper–gold and porphyry Cu ± Mo ± Au deposits. Geology 41 (84–+).
(7), 767–770. Stein, H.J., Markey, R.J., Morgan, J.W., Du, A., Sun, Y., 1997. Highly precise and accurate Re–
Rowins, S.M., 2000. Reduced porphyry copper–gold deposits: a new variation on an old Os ages for molybdenite from the East Qinling molybdenum belt, Shaanxi Province,
theme. Geology 28 (6), 491–494. China. Econ. Geol. Bull. Soc. Econ. Geol. 92 (7–8), 827–835.
Rudnick, R.L., Gao, S., 2003. Composition of the continental crust. In: Heinrich, D.H., Stern, C.R., Funk, J.A., Skewes, M.A., Arevalo, A., 2007. Magmatic anhydrite in plutonic
Turekian, K.K. (Eds.), Treatise on Geochemistry. Pergamon, Oxford, pp. 1–64. rocks at the El Teniente Cu–Mo deposit chile, and the role of sulfur- and copper-
Sajona, F.G., Maury, R.C., 1998. Association of adakites with gold and copper mineraliza- rich magmas in its formation. Econ. Geol. 102 (7), 1335–1344.
tion in the Philippines. C. R. Acad. Sci. IIA Earth Planet. Sci. 326 (1), 27–34. Stern, C.R., Skewes, M.A., Arevalo, A., 2011. Magmatic evolution of the Giant El Teniente
Scharman, M.R., Pavlis, T.L., Ruppert, N., 2012. Crustal stabilization through the processes Cu–Mo Deposit, Central Chile. J. Petrol. 52 (7–8), 1591–1617.
of ridge subduction: examples from the Chugach metamorphic complex, southern Sun, W.D., Williams, I.S., Li, S.G., 2002. Carboniferous and triassic eclogites in the western
Alaska. Earth Planet. Sci. Lett. 329, 122–132. Dabie Mountains, east-central China: evidence for protracted convergence of the
Seedorff, E., Einaudi, M.T., 2004a. Henderson porphyry molybdenum system, Colorado: I. North and South China Blocks. J. Metamorph. Geol. 20 (9), 873–886.
Sequence and abundance of hydrothermal mineral assemblages, flow paths of Sun, W.D., Arculus, R.J., Bennett, V.C., Eggins, S.M., Binns, R.A., 2003a. Evidence for rheni-
evolving fluids, and evolutionary style. Econ. Geol. Bull. Soc. Econ. Geol. 99 (1), 3–37. um enrichment in the mantle wedge from submarine arc-like volcanic glasses (Papua
Seedorff, E., Einaudi, M.T., 2004b. Henderson porphyry molybdenum system. Colorado: II. New Guinea). Geology 31 (10), 845–848.
Decoupling of introduction and deposition of metals during geochemical evolution of Sun, W.D., Bennett, V.C., Eggins, S.M., Arculus, R.J., Perfit, M.R., 2003b. Rhenium systemat-
hydrothermal fluids. Econ. Geol. Bull. Soc. Econ. Geol. 99 (1), 39–72. ics in submarine MORB and back-arc basin glasses: laser ablation ICP-MS results.
Seo, J.H., Heinrich, C.A., 2013. Selective copper diffusion into quartz-hosted vapor Chem. Geol. 196 (1–4), 259–281.
inclusions: evidence from other host minerals, driving forces, and consequences for Sun, W.D., Bennett, V.C., Eggins, S.M., Kamenetsky, V.S., Arculus, R.J., 2003c. Enhanced
Cu–Au ore formation. Geochim. Cosmochim. Acta 113, 60–69. mantle-to-crust rhenium transfer in undegassed arc magmas. Nature 422 (6929),
Seo, J.H., Guillong, M., Heinrich, C.A., 2009. The role of sulfur in the formation of mag- 294–297.
matic–hydrothermal copper–gold deposits. Earth Planet. Sci. Lett. 282 (1–4), Sun, W.D., Arculus, R.J., Kamenetsky, V.S., Binns, R.A., 2004a. Release of gold-bearing fluids
323–328. in convergent margin magmas prompted by magnetite crystallization. Nature 431
Shafiei, B., Haschke, M., Shahabpour, J., 2009. Recycling of orogenic arc crust triggers por- (7011), 975–978.
phyry Cu mineralization in Kerman Cenozoic arc rocks, southeastern Iran. Mineral. Sun, W.D., Bennett, V.C., Kamenetsky, V.S., 2004b. The mechanism of Re enrichment in arc
Deposita 44 (3), 265–283. magmas: evidence from Lau Basin basaltic glasses and primitive melt inclusions.
Shahabpour, J., 1991. Some secondary ore formation features of the Sar–Cheshmeh por- Earth Planet. Sci. Lett. 222 (1), 101–114.
phyry copper–molybdenum deposit, Kerman, Iran. Mineral. Deposita 26 (4), Sun, W.D., Ding, X., Hu, Y.H., Li, X.H., 2007. The golden transformation of the Cretaceous
275–280. plate subduction in the west Pacific. Earth Planet. Sci. Lett. 262 (3–4), 533–542.
Shaw, H.T., 1963. Hydrogen–water vapor mixtures: control of hydrothermal atmospheres Sun, X.M., Tang, Q., Sun, W.D., Xu, L., Zhai, W., Liang, J.L., Liang, Y.H., Shen, K., Zhang, Z.M.,
by hydrogen osmosis. Science 139, 1220–1222. Zhou, B., Wang, F.Y., 2007. Monazite, iron oxide and barite exsolutions in apatite ag-
Shen, P., Pan, H.D., 2013. Country-rock contamination of magmas associated with the gregates from CCSD drillhole eclogites and their geological implications. Geochim.
Baogutu porphyry Cu deposit, Xinjiang, China. Lithos 177, 451–469. Cosmochim. Acta 71 (11), 2896–2905.
Shen, P., Shen, Y.C., Pan, H.D., Wang, J.B., Zhang, R., Zhang, Y.X., 2010a. Baogutu porphyry Sun, W.D., Ling, M.X., Yang, X.Y., Fan, W.M., Ding, X., Liang, H.Y., 2010. Ridge subduction
Cu–Mo–Au deposit, West Junggar, northwest China: petrology, alteration, and miner- and porphyry copper–gold mineralization: an overview. Sci. China Ser. D Earth Sci.
alization. Econ. Geol. 105 (5), 947–970. 53 (4), 475–484.
Shen, P., Shen, Y.C., Wang, J.B., Zhu, H.P., Wang, L.J., Meng, L., 2010b. Methane-rich fluid Sun, W.D., Zhang, H., Ling, M.X., Ding, X., Chung, S.L., Zhou, J.B., Yang, X.Y., Fan, W.M., 2011.
evolution of the Baogutu porphyry Cu–Mo–Au deposit, Xinjiang, NW China. Chem. The genetic association of adakites and Cu–Au ore deposits. Int. Geol. Rev. 53 (5–6),
Geol. 275 (1–2), 78–98. 691–703.
W. Sun et al. / Ore Geology Reviews 65 (2015) 97–131 131

Sun, W.D., Ling, M.X., Chung, S.L., Ding, X., Yang, X.Y., Liang, H.Y., Fan, W.M., Goldfarb, R., South China: implications for ore genesis and exploration. J. Asian Earth Sci. 74,
Yin, Q.Z., 2012a. Geochemical constraints on adakites of different origins and copper 343–360.
mineralization. J. Geol. 120 (1), 105–120. Wilkinson, J.J., 2013. Triggers for the formation of porphyry ore deposits in magmatic arcs.
Sun, W.D., Ling, M.X., Ding, X., Chung, S.L., Yang, X.Y., Fan, W.M., 2012b. The genetic asso- Nat. Geosci. 6 (11), 917–925.
ciation of adakites and Cu–Au ore deposits: a reply. Int. Geol. Rev. 54 (3), 370–372. Wones, D.R., Eugster, H.P., 1965. Stability of biotite: experiement, theory, and application.
Sun, W.D., Li, S., Yang, X.Y., Ling, M.X., Ding, X., Duan, L.A., Zhan, M.Z., Zhang, H., Fan, W.M., Am. Mineral. 50, 1228–1272.
2013a. Large-scale gold mineralization in eastern China induced by an Early Creta- Wu, H.Y., Zhang, L.C., Wan, B., Chen, Z.G., Zhang, X.J., Xiang, P., 2011. Geochronological and
ceous clockwise change in Pacific plate motions. Int. Geol. Rev. 55 (3), 311–321. geochemical constraints on Aolunhua porphyry Mo–Cu deposit, northeast China, and
Sun, W.D., Liang, H.Y., Ling, M.X., Zhan, M.Z., Ding, X., Zhang, H., Yang, X.Y., Li, Y.L., Ireland, its tectonic significance. Ore Geol. Rev. 43 (1), 78–91.
T.R., Wei, Q.R., Fan, W.M., 2013b. The link between reduced porphyry copper deposits Xiao, B., Qin, K.Z., Li, G.M., Li, J.X., Xia, D.X., Chen, L., Zhao, J.X., 2012. Highly oxidized
and oxidized magmas. Geochim. Cosmochim. Acta 103, 263–275. magma and fluid evolution of Miocene Qulong giant porphyry Cu–Mo deposit, South-
Sun, W.D., Huang, R.F., Liang, H.Y., Ling, M.X., Li, C.Y., Ding, X., Zhang, H., Yang, X.Y., ern Tibet, China. Resour. Geol. 62 (1), 4–18.
Ireland, T., Fan, W.M., 2014a. Magnetite–hematite, oxygen fugacity, adakite and por- Xu, J.F., Shinjo, R., Defant, M.J., Wang, Q.A., Rapp, R.P., 2002. Origin of Mesozoic adakitic in-
phyry copper deposits: reply to Richards. Geochim. Cosmochim. Acta 126, 646–649. trusive rocks in the Ningzhen area of east China: Partial melting of delaminated lower
Sun, W.D., et al., 2014b. The genetic association between magnetite–hematite and por- continental crust? Geology 30 (12), 1111–1114.
phyry copper deposits: reply to Pokrovski. Geochim. Cosmochim. Acta 126, 639–642. Xu, W.L., Wang, Q.H., Wang, D.Y., Guo, J.H., Pei, F.P., 2006. Mesozoic adakitic rocks from
Takagi, T., 2004. Origin of magnetite- and ilmenite-series granitic rocks in the Japan Arc. the Xuzhou–Suzhou area, eastern China: evidence for partial melting of delaminated
Am. J. Sci. 304 (2), 169–202. lower continental crust. J. Asian Earth Sci. 27 (4), 454–464.
Tang, G.J., Wang, Q., Wyman, D.A., Li, Z.X., Zhao, Z.H., Jia, X.H., Jiang, Z.Q., 2010. Ridge sub- Yang, X.Y., Wang, K.R., Yang, X.M., Sun, L.G., 2002. Characteristics of mineralization and
duction and crustal growth in the Central Asian Orogenic Belt: evidence from Late gold occurrence in Shaxi porphyry copper–gold deposit, central Anhui, China.
Carboniferous adakites and high-Mg diorites in the western Junggar region, northern Neues Jahrb. Mineral. Abh. 177 (3), 293–320.
Xinjiang (west China). Chem. Geol. 277 (3–4), 281–300. Yang, Z.M., Hou, Z.Q., White, N.C., Chang, Z.S., Li, Z.Q., Song, Y.C., 2009. Geology of the post-
Taylor, H.P., 1974. The application of oxygen and hydrogen isotope studies to problems of collisional porphyry copper–molybdenum deposit at Qulong, Tibet. Ore Geol. Rev. 36
hydrothermal alteration and ore deposition. Econ. Geol. 69, 843–883. (1–3), 133–159.
Taylor, B.E., 1992. Degassing of H2O from rhyolite magma during eruption and shallow Yang, L., Chen, F.K., Liu, B.X., Hu, Z.P., Qi, Y., Wu, J.D., He, J.F., Siebel, W., 2013. Geochemistry
inrusion, and the isotopic composition of magmatic water in hydrothermal systems. and Sr–Nd–Pb–Hf isotopic composition of the Donggou Mo-bearing granite porphy-
Geol. Surv. Jpn. Rep. 279, 190–194. ry, Qinling orogenic belt, central China. Int. Geol. Rev. 55 (10), 1261–1279.
Thieblemont, D., Stein, G., Lescuyer, J.L., 1997. Epithermal and porphyry deposits: the Yogodzinski, G.M., Volynets, O.N., Koloskov, A.V., Seliverstov, N.I., Matvenkov, V.V., 1994.
adakite connection. C.R. Acad. Sci., Ser. IIa: Sci. Terre Planets 325 (2), 103–109. Magnesian andesites and the subduction component in a strongly calc-alkaline series
Thompson, J.F.H., Sillitoe, R.H., Baker, T., Lang, J.R., Mortensen, J.K., 1999. Intrusion-relate at Piip Volcano, Far Western Aleutians. J. Petrol. 35 (1), 163–204.
gold deposits associated with tungsten–tin provinces. Mineral. Deposita 34, 323–334. Zeng, Q.D., Liu, J.M., Qin, K.Z., Fan, H.R., Chu, S.X., Wang, Y.B., Zhou, L.L., 2013. Types, char-
Thorkelson, D.J., Breitsprecher, K., 2005. Partial melting of slab window margins: genesis acteristics, and time–space distribution of molybdenum deposits in China. Int. Geol.
of adakitic and non-adakitic magmas. Lithos 79 (1–2), 25–41. Rev. 55 (11), 1311–1358.
Titley, S.R., 1981. Porphyry copper. Am. Sci. 69 (6), 632–638. Zhang, Q., Wang, Y., Wang, Y.L., 2001a. Preliminary study on the components of the lower
Togashi, S., Terashima, S., 1997. The behavior of gold in unaltered island arc tholeiitic crust in east China Plateau during Yanshanian Period: constraints on Sr and Nd isoto-
rocks from Izu-Oshima, Fuji, and Osoreyama volcanic areas, Japan. Geochim. pic compositions of adakite-like rocks. Acta Petrol. Sin. 17 (4), 505–513.
Cosmochim. Acta 61 (3), 543–554. Zhang, Q., Wang, Y., Qian, Q., Yang, J.H., Wang, Y.L., Zhao, T.P., Guo, G.J., 2001b. The char-
Tomkins, A.G., 2010. Windows of metamorphic sulfur liberation in the crust: implications acteristics and tectonic–metallogenic significances of the adakites in Yanshan period
for gold deposit genesis. Geochim. Cosmochim. Acta 74 (11), 3246–3259. from eastern China. Acta Petrol. Sin. 17 (2), 236–244 (in Chinese with English
Vila, T., Sillitoe, R.H., 1991. Gold-rich porphyry systems in the Maricunga Belt, Northern abstract).
Chile. Econ. Geol. Bull. Soc. Econ. Geol. 86 (6), 1238–1260. Zhang, Z.X., Yang, F.Q., Yan, S.H., Zhang, R., Chai, F.M., Liu, F.L., Geng, X.X., 2010. Sources of
Vila, T., Sillitoe, R.H., Betzhold, J., Viteri, E., 1991. The porphyry gold deposit at Marte, ore-forming fluids and materials of the Baogutu porphyry copper deposit in Xinjiang:
Northern Chile. Econ. Geol. Bull. Soc. Econ. Geol. 86 (6), 1271–1286. constrains from sulfur–hydrogen–oxygen isotopes geochemistry. Acta Geol. Sin. 26,
Vry, V.H., Wilkinson, J.J., Seguel, J., Millan, J., 2010. Multistage intrusion, brecciation, and 707–716 (in Chinese with English abstract).
veining at El Teniente, evolution of a nested porphyry system. Econ. Geol. 105 (1), Zhang, H., Ling, M.X., Liu, Y.L., Tu, X.L., Wang, F.Y., Li, C.Y., Liang, H.Y., Yang, X.Y., Arndt, N.T.,
119–153. Sun, W.D., 2013. High oxygen fugacity and slab melting linked to Cu mineralization:
Wang, Q., Wyman, D.A., Xu, J.F., Zhao, Z.H., Jian, P., Xiong, X.L., Bao, Z.W., Li, C.F., Bai, Z.H., evidence from Dexing porphyry copper deposits, southeastern China. J. Geol. 121 (3),
2006a. Petrogenesis of Cretaceous adakitic and shoshonitic igneous rocks in the 289–305.
Luzong area, Anhui Province (eastern China): implications for geodynamics and Zhang, Y., Sun, J.G., Chen, Y.J., Zhao, K.Q., Gu, A.L., 2013. Re–Os and U–Pb geochronology of
Cu–Au mineralization. Lithos 89 (3–4), 424–446. porphyry Mo deposits in central Jilin Province: Mo ore-forming stages in northeast
Wang, Q., Xu, J.F., Jian, P., Bao, Z.W., Zhao, Z.H., Li, C.F., Xiong, X.L., Ma, J.L., 2006b. Petro- China. Int. Geol. Rev. 55 (14), 1763–1785.
genesis of adakitic porphyries in an extensional tectonic setting, Dexing, South Zhang, H., Li, C.Y., Yang, X.Y., Sun, Y.L., Deng, J.H., Liang, H.Y., Wang, R.L., Wang, B.H., Wang,
China: implications for the genesis of porphyry copper mineralization. J. Petrol. 47 Y.X., Sun, W.D., 2013. Shapinggou: the largest Climax-type porphyry Mo deposit in
(1), 119–144. China. Int. Geol. Rev. 56, 313–331.
Wang, Q., Wyman, D.A., Xu, J., Jian, P., Zhao, Z., Li, C., Xu, W., Ma, J., He, B., 2007a. Early Cre- Zhou, Q., Jiang, Y.H., Zhang, H.H., Liao, S.Y., Jin, G.D., Zhao, P., Jia, R.Y., Liu, Z., 2013. Mantle
taceous adakitic granites in the Northern Dabie Complex, central China: implications origin of the Dexing porphyry copper deposit, SE China. Int. Geol. Rev. 55 (3),
for partial melting and delamination of thickened lower crust. Geochim. Cosmochim. 337–349.
Acta 71 (10), 2609–2636. Zhu, X., Huang, C., Rui, Z., Zhou, Y., Zhu, X., Hu, C., Mei, Z., 1983. The Geology of Dexing
Wang, Q., Wyman, D.A., Xu, J.F., Zhao, Z.H., Jian, P., Zi, F., 2007b. Partial melting of thick- Porphyry Copper Ore Field. Geological Publishing House, Beijing (in Chinese).
ened or delaminated lower crust in the middle of eastern China: implications for Zhu, L.M., Zhang, G.W., Guo, B., Lee, B., Gong, H.J., Wang, F., 2010. Geochemistry of the
Cu–Au mineralization. J. Geol. 115 (2), 149–161. Jinduicheng Mo-bearing porphyry and deposit, and its implications for the
Wang, G.G., Ni, P., Wang, R.C., Zhao, K.D., Chen, H., Ding, J.Y., Zhao, C., Cai, Y.T., 2013. Geo- geodynamic setting in East Qinling, PR China. Chem. Erde 70 (2), 159–174.
logical, fluid inclusion and isotopic studies of the Yinshan Cu–Au–Pb–Zn–Ag deposit,

View publication stats

You might also like