Vous êtes sur la page 1sur 4

Roasting the mantle: Slab melting and the genesis of major Au and

Au-rich Cu deposits
James E. Mungall*

Department of Geology, University of Toronto, 22 Russell Street, Toronto, Ontario M5S 3B1, Canada

ABSTRACT
The generation of large deposits of Au and Cu in suprasubduction-zone settings depends
upon a combination of factors, including availability of the chalcophile elements to arc
magmas in their mantle source regions, and the operation of suitable hydrothermal systems in the upper crust where the deposits eventually form. The removal of chalcophile
elements from the mantle wedge into arc magmas can only occur if sulfide is absent from
the melted source rock, requiring oxidation of the mantle wedge to values of log fO2 .
FMQ 1 2, where fO2 is oxygen fugacity and FMQ is the fayalite-magnetite-quartz oxygen
buffer. The only agent capable of effecting this change is ferric iron, carried in solution
by slab-derived partial melts or supercritical fluids. Arc magmas with high potential to
generate Au and Cu deposits will have certain geochemical characteristics; they will have
fO2 more than two log units above FMQ and they will have either adakitic, sodic-alkaline,
or potassic-ultrapotassic affinities. Favorable tectonic settings include subduction of very
young lithosphere or very slow or oblique convergence, flat subduction, and the cessation
of subduction.
Keywords: mineral deposits, genesis, subduction zones, Au, Cu, mantle.
INTRODUCTION
Many of the worlds major deposits of Au
and Cu owe their origins to the emission of
hydrothermal fluids from cooling magmas emplaced above subduction zones; however,
most arc magmas do not produce economic
deposits. Magma associations recognized to
be particularly fertile include potassic calcalkaline magmas (Muller and Groves, 1993),
adakitic magmas (i.e., arc magmas with trace
element signatures of garnet and hornblende
in their source and possibly generated by slab
melting; Kay et al., 1995; Drummond et al.,
1996; Thieblemont et al., 1997; Sajona and
Maury, 1998), and alkaline arc magmas (e.g.,
Richards, 1990; McInnes and Cameron,
1994). Whereas normal subduction processes
do not appear to favor the generation of large
Au and Cu deposits, the tectonic settings recognized as favorable include the onset and
termination of periods of flat subduction
(Kay and Mpodozis, 2001), the reversal of
polarity of a subduction zone (Solomon,
1990), and the cessation of subduction due to
collision (McInnes and Cameron, 1994; Sillitoe, 1997).
In this article I attempt to show that simple
geochemical arguments can be used to account for such disparate observations within a
single framework, elucidating and building
upon the suggestion of Sillitoe (1997) that sulfide oxidation is the key process in the liberation of chalcophile elements from the arc
mantle.
ROLE OF SULFIDE IN MANTLE
SOURCE REGIONS OF ARC MAGMAS
Sulfide melt sequesters the chalcophile elements during partial melting of the mantle,
*E-mail: mungall@geology.utoronto.ca.

and complete dissolution of sulfide into silicate melt with S solubility ;1000 ppm (Mavrogenes and ONeill, 1999) at normal mantle
S contents of ;250 ppm would require .25%
partial melting (e.g., Hamlyn and Keays,
1986). In order for a silicate magma to contain
significant amounts of Au and Cu, sulfide melt
must be absent from its mantle source. Furthermore, unless the magma remains sulfide
undersaturated it will lose Cu and Au during
its evolution (Wyborn and Sun, 1994). If S is
added to the mantle by the same fluids that
induce partial melting in the subduction zone
(Alt et al., 1993; Metrich et al., 1999), then
the degree of partial melting required to destroy residual sulfide might easily exceed
40%. Such abnormally high aggregate degrees
of melting at moderate pressures generate
boninitic magmas, which contain high abundances of the chalcophile elements, including
the platinum group elements (Hamlyn and
Keays, 1986). However, many deposits are associated with alkaline magmas, the compositions of which reflect very low degrees of partial melting of the mantle (e.g., Richards,
1990; McInnes and Cameron, 1994).
Sulfide can also be removed by the oxidation of sulfur from S22 to S16:
FeS(liq) 1 2O2(fluid)
5 FeO(melt) 1 SO3(melt) .

(1)

In mantle peridotite the main reservoir for Ca


is pyroxene, so upon anhydrite saturation
equation 1 can be rewritten as:
FeS(sul) 1 CaFeSi2O6(cpx) 1 2O2(fluid)
5 CaSO4(An) 1 2FeSiO3(opx) .

(2)

Figure 1. Oxygen buffers in log f O2 vs. temperature space (see text). Only addition of
Fe2O3 can bring mantle assemblages higher
than SSO buffer. Undersaturation of magma
in sulfide requires either very high degrees
of partial melting or generation in unshaded
portion of diagram. SSOsulfide-sulfur oxide buffer; CCOcarbon dioxidecarbon oxide buffer; FMQfayalite-magnetite-quartz
oxygen buffer; MHmagnetite-hematite oxygen buffer.

Because the activities of ferrosilite, hedenbergite, and FeS in their respective host phases
vary only through limited ranges, the assemblage anhydrite 1 sulfide melt will constitute
an oxygen buffer (here named SSO; i.e., the
sulfide-sulfur oxide buffer).
OXIDATION OF THE MANTLE
ABOVE SUBDUCTION ZONES
The oxidation state of upper mantle rocks
from many settings has been assessed through
consideration of reactions like
6Fe2SiO4(ol) 1 O2(fluid)
5 2Fe3O4(sp) 1 3Fe2Si2O6(opx) ,

(3)

where fayalite, magnetite, and ferrosilite appear as components in solid solution within
olivine, spinel, and orthopyroxene (ONeill
and Wall, 1987). Oxygen fugacity thus measured in mantle xenoliths not related to subduction zones is generally found to fall within
one log unit of the fayalite-magnetite-quartz
(FMQ) oxygen buffer, at which it is thought
to be held by reactions like the carbon dioxide
carbon oxide (CCO) buffer (Fig. 1; Blundy et
al., 1991):
CO2(fluid) 5 C(gr) 1 O2(fluid).

(4)

Samples of peridotite gathered from volcanoes tapping the mantle above subduction
zones record equilibration at higher log f O2,

q 2002 Geological Society of America. For permission to copy, contact Copyright Permissions, GSA, or editing@geosociety.org.
Geology; October 2002; v. 30; no. 10; p. 915918; 3 figures.

915

typically ranging from FMQ to FMQ 1 2


(i.e., still within the stability field of sulfide
melt; Carroll and Rutherford, 1987) due to the
action of oxidized slab-derived metsomatizing
fluids or melts (e.g., Brandon and Draper,
1996; Luhr and Aranda-Gomez, 1997; Parkinson and Arculus, 1999).
The subducted oceanic crust has variable
but locally very high intrinsic f O2 due to
equilibration with seawater during hydrothermal alteration and deposition of terrigenous
sediment. Melts or fluids derived from the slab
will carry this oxidizing potential up into the
overlying mantle (e.g., McInnes and Cameron,
1994). The only elements existing in variable
oxidation states and present in sufficient abundances to affect the redox state of the upper
mantle are C, H, S, and Fe.
The effect of addition of CO2 to a mantle
assemblage containing graphite can be seen by
combining equations 3 and 4:
6Fe2SiO4(ol) 1 CO2(fluid)
5 2Fe3O4(sp) 1 C(gr) 1 3Fe2Si2O6(opx) . (5)
As long as the fO2 remains below the CCO
buffer, continuous addition of CO2 will lead
to continuous increase in the amount of both
graphite and ferric iron in the assemblage.
Once sufficient iron has been oxidized by the
reduction of CO2 to graphite to bring the assemblage to the CCO buffer, further addition
of CO2 will have no effect on fO2 because it
coexists with stable graphite but cannot oxidize graphite. The metasomatic addition of
CO2 cannot raise fO2 to the SSO buffer (Fig.
1).
The dissociation of water is represented by
H2O(fluid) 5 H2(fluid) 1 0.5O2(fluid).

(6)

The equilibrium constant for this reaction at


upper mantle conditions is so small that the
amount of free oxygen available to participate
in reactions like equation 4 is of the order of
10214 molal during reduction of water from
log fO2 of FMQ 1 2 to FMQ (Frost and Ballhaus, 1998). Water alone cannot act as an oxidizing agent unless the ambient fO2 of the
mantle is in the stability field of H2. Spontaneous physical separation of H2 and O2 into
distinct, spatially separated, reduced, and oxidized reservoirs (cf. Brandon and Draper,
1996) in a formerly homogeneous system
would violate the second law of
thermodynamics.
At ambient f O2 below SSO, 1 mol of added
sulfate would be reduced to sulfide, while oxidizing 8 mol of iron:

916

13Fe2SiO4(ol) 1 SO3(fluid)
5 4Fe3O4(sp) 1 FeS(sul)
1 6.5Fe2Si2O6(opx) .

(7)

If fO2 is at or above the SSO buffer, SO3


will be stable, so that continued addition of
sulfate cannot cause any further oxidation of
iron to raise the fO2 above the stability field
of sulfide. The process of raising fO2 from
CCO to SSO by reduction of the introduced
sulfate causes a net addition of sulfide to the
mineral assemblage. The addition of slabderived sulfate will firmly buffer the fO2 of
the mantle wedge exactly at SSO while increasing the total amount of sulfide melt present. Sulfate metasomatism therefore effectively precludes the subsequent generation of
fertile magmas rich in the chalcophile
elements.
The addition of Fe2O3 will adjust the mantle fO2 along a continuously sliding scale according to reactions like the following, as long
as the mantle remains below the magnetitehematite buffer:
Fe2O3(fluid) 1 Fe2SiO4(ol)
5 Fe3O4(sp) 1 FeSiO3(opx) ,

(8)

where fluid is either aqueous fluid or silicate


melt (Fig. 1). When Fe2O3 addition brings the
fO2 of an assemblage containing sulfide melt
up to the SSO buffer, the fO2 cannot increase
further until all sulfide has been oxidized to
sulfate according to a reaction like
FeS(sul) 1 4Fe2O3(fluid) 1 9FeSiO3(opx)
5 SO3(fluid) 1 9Fe2SiO4(ol) .

Figure 2. A: Concentrations of aqueous species as functions of pressure. Estimated


concentrations of ferric iron complexes
(FeO1aq, HFeO2aq) remain well below those
of ferrous iron (FeOaq) at all pressures below
30 kbar. Sulfate concentration remains several orders of magnitude higher than ferric
iron solubility at all pressures. B: Concentration of Fe2O3 in silicate melts as function
of f O2 (expressed as difference from FMQ in
log units). Note scale difference between A
and B.

(9)

Once sulfide has been fully oxidized there


is no further bar to the continued increase in
fO2 resulting from Fe2O3 addition to the mantle wedge via equation 8.
METASOMATIC AGENTS
I have estimated the concentrations of sulfate and ferric iron in subduction fluids by using thermochemical data and empirical models for aqueous solutes and minerals (Johnson
et al., 1992; E. Shock, 2000, personal commun.) combined with thermochemical and dielectric models for water (Harvey et al., 1996;
Fernandez et al., 1997) to calculate Gibbs free
energies of reaction for dissolution and complexation reactions at 600 8C and pressures to
30 kbar. These calculations are extrapolations
of low-pressure data and results must be regarded with caution.
Calculated concentrations of several aqueous complexes in fluids equilibrated with spinel (Fe species) or anhydrite (sulfate) are
shown in Figure 2A as functions of pressure.

All activity coefficients are assumed to be unity and the activity of magnetite in spinel is
assumed to be 0.1. The estimated concentration of FeO(aq) at 30 kbar is ;250 ppm, similar to total iron concentration measured by
Schneider and Eggler (1986) at 1520 kbar
and temperatures from 750 to 850 8C.
The first-order result of this modeling exercise is that the solubility of anhydrite remains several orders of magnitude greater
than that of ferric iron over the entire pressure
range considered. Whereas aqueous fluids are
able to transport significant amounts of sulfate, they are exceedingly poor carriers of ferric iron.
During reduction of its log fO2 from FMQ
1 4 (cf. McInnes and Cameron, 1994) to
FMQ, an aqueous fluid equilibrated with spinel will donate ;2.5 3 1025 mol Fe13 per
mole of aqueous solution to redox reactions in
metasomatized mantle. To oxidize 1 mol of
sulfide to sulfate, 8 mol of Fe13 are required;
by converting to weight fractions and assumGEOLOGY, October 2002

ing that normal asthenospheric mantle initially


contains 250 ppm S, I find that 1 g of aqueous
fluid can oxidize the sulfide contained in 0.4
g of mantle.
Figure 2B shows the dependence of Fe2O3
concentrations on fO2 estimated by using the
MELTS thermochemical model of Ghiorso
and Sack (1995) in two different examples of
melt thought to have been present in metasomatized mantle wedge peridotites above partially melted oceanic plates (McInnes and
Cameron, 1994; Schiano et al., 1995). The reduction of silicate melts from FMQ 1 4 to
FMQ consumes ;0.01 mol of Fe13 per mole
of silicate melt. Following the same calculation as for aqueous fluid, as previously discussed, I find that 1 g of silicate melt derived
by slab melting can oxidize the sulfide contained in 160 g of mantle peridotite.
Silicate melts thus have a carrying capacity
for Fe2O3 some 400 times greater than have
aqueous fluids. The effectiveness of slab melting is further augmented by the fact that the
mass of aqueous fluid that can be released during slab dehydration is less than or equal to
the amount of water originally present in the
slab, whereas slab melt containing 4% dissolved water will have a mass 25 times the
mass of water originally present. In terms of
the total flux of Fe2O3 per unit mass of water
brought into the subduction zone by the downgoing slab, slab melts will thus carry as much
as 10 000 times more Fe2O3 than can dehydration fluids.
Metasomatism of the upper mantle by the
aqueous fluid released during slab dehydration
will introduce abundant sulfate but only trace
quantities of ferric iron. Slab dehydration will
oxidize the mantle wedge while augmenting
the amount of sulfide present. Melting of peridotite metasomatized by aqueous fluid will
be expected to produce typical sulfide-saturated,
water-rich arc magmas with intrinsic fO2 between FMQ and SSO and minimal potential
for the generation of chalcophile element deposits. In contrast, the flux of Fe2O3 via slab
melts can lead to sulfide-undersaturated melting of fertile asthenosphere and the generation
of Au- and Cu-rich arc magmas.
It is critical to emphasize that the only plausible source of highly oxidized melts or fluids
in the upper mantle is the upper oceanic crust
in a subducted slab. Although melts similar to
slab melts may form by remelting of basaltic
or gabbroic crustal underplates originally
sourced in normal arc mantle, these melts will
retain the low fO2 of their sources (Carmichael, 1991). Basaltic magmas emplaced at
Earths surface with fO2 . SSO must contain
a significant component of melt from subducted oxidized crustal rocks.
TECTONIC CONTROLS
Figure 3 is a schematic section of a convergent margin showing several possible scenarios
GEOLOGY, October 2002

Figure 3. Composite sketch of convergent plate margins (SL is sea level). Scales are in
kilometers, without vertical exaggeration. Inset diagram shows fields in which breakdown
of hydrous phases produces melt, aqueous fluid, or supercritical fluid (SC) from subducted
oceanic crust. Smaller inset diagrams show pressure-temperature (P-T) paths associated
with various tectonic settings. Only those inset diagrams labeled with Au (i.e., B, C, D, and
E) represent tectonic settings capable of generating oxidized Au- and Cu-rich magmas.

for the generation of slab melts or aqueous fluids


for both normal (solid lines) and flat (dashed
lines) subduction. The inset diagram is a simplified sketch of pressure- and temperaturedependant phase relations in hydrous metabasaltic rocks representative of the hydrothermally
altered oceanic crust on the subducted slab
(e.g., Peacock et al., 1994; Schmidt, 1996). In
the field labeled fluid, minerals such as talc,
serpentine, and actinolite undergo progressive
dehydration reactions to release aqueous fluids. At high pressure there are no important
dehydration reactions above the fluid field
until the mica breakdown curve is reached. At
moderate pressures the same minerals break
down to produce slab melts (melt field). At
temperatures above the critical curve there is
no distinction between a dehydration reaction
and a melting reaction because silicate melt
and aqueous fluid are a single thermodynamic
phase (Paillat et al., 1992). The critical curve
shown corresponds approximately to the measured critical end points in the haplograniteor dacite-water system (Bureau and Keppler,
1999). The upper left margin of the supercritical field marks the breakdown of phengite mica to an anhydrous residue and a supercritical potassic fluid (Schmidt, 1996).
Dense supercritical fluids are capable of dissolving large (wt%) loads of otherwise immobile elements such as Al2O3 and Fe2O3.
The six small reproductions of the inset
phase diagram each show the pressuretemperature (P-T) path followed by a subducted slab in various tectonic settings. In normal contemporary subduction zones the
subducted crust never melts (Fig. 3, box A;
see also Peacock et al., 1994), and magmas
produced by hydrous fluxing of the astheno-

spheric mantle wedge will be sulfide saturated


and poor in Au and Cu.
Box B shows the fate of a slab which may
be a: young and hot; b: the leading edge of a
newly formed subduction zone; or c: subducted very slowly or obliquely. It follows a P-T
path across the amphibole solidus before the
mineral assemblage has been dehydrated, permitting small degrees of partial melting of the
subducted slab. Mantle peridotite fluxed by
slab melts will generate adakitic magmas
(Drummond et al., 1996), which may be highly oxidized and potentially fertile for Au and
Cu.
Slab P-T paths intermediate between cases
A and B will allow early, shallow dehydration
and generation of ordinary arc magmas, followed by the release of potassic supercritical
fluids at great depths of ;300 km (labeled SC;
Schmidt, 1996) to generate rear-arc potassic or
ultrapotassic magmas with high fO2 and comparatively high potential to host Au and CuAu deposits (box D; Muller and Groves,
1993).
If a slab stalls in the subduction zone then
its temperature will increase continuously
while it remains at a constant pressure. In this
case the upper portions of the slab will pass
directly into the melt field, whereas the deeper
portions will pass into the supercritical field
(box C). Panguna in Bougainville, Ok Tedi in
New Guinea, Baguio in Luzon (Thieblemont
et al., 1997; Sajona and Maury, 1998), Porgera in New Guinea, and Ladolam on Lihir
(Solomon, 1990; McInnes and Cameron,
1994) are all deposits generated after reversal
of subduction direction led to the stalling of a
slab. The process may lead directly to eruption
of adakitic or potassic magmas, or it may
917

cause emplacement of oxidized veins in the


mantle that do not actually melt until a later
extensional tectonic event produces small volumes of alkaline magmas.
During flat subduction a slab passes under
the overriding plate for great distances at constant pressure. If the process has just begun or
if the slab is beginning to fall back down to a
normal angle, then it will be passing under hot
asthenospheric mantle while remaining at low
pressure for extended periods, possibly leading to slab melting under thin lithosphere (cf.
Gutscher et al., 2000). Under more realistic
lithospheric thicknesses than those modeled
by Gutscher et al. (2000), the result would be
release of oxidized supercritical fluids rather
than adakitic magmas (box E). Major deposits
that have been suggested to be related to slab
melting and adakitic or potassic magmatism
in this setting include El Indio, El Salvador,
Quebrada, and Chuquicamata of the Chilean
Copper Belt, and Bingham, United States
(Thieblemont et al., 1997). If there is no asthenosphere present above the slab, then it
will follow a very cold path, allowing dehydration to occur without fluxing of the mantle,
and causing cessation of magmatism until normal steep subduction resumes (box F; Gutscher et al., 2000; Kay and Mpodozis, 2001).
The importance of slab melting in subduction zones, particularly in relation to flat subduction, remains a contentious issue (Peacock
et al., 1994; Martin, 1999; Gutscher et al.,
2000; Kay and Mpodozis, 2002). Current geochemical probes of arc magma petrogenesis
focus on differences between fluid and melt
transfer from slab to mantle, and are poorly
suited to detection of small slab melt volumes
or the possible presence of supercritical meltfluid mixtures.
I suggest here that a powerful new discriminant is magma oxidation state. If an arc magma has log fO2 . SSO, then it must contain
a component of melted oceanic crust, because
no vector other than slab-derived melt or supercritical fluid can carry such high redox potential into the upper mantle. If the concepts
presented here are correct, then large Au and
Cu deposits may by their very existence provide evidence for slab melting or the release
of supercritical fluids in the source regions of
their parental magmas. Conversely, the recognition of slab melting signatures in arc magmas should be considered a promising first
step in the definition of Au-Cu exploration targets at the regional scale.
ACKNOWLEDGMENTS
The author was funded by a research grant from the
Natural Sciences and Engineering Research Council of

918

Canada during this study. Thanks to Everett Shock for advice on high-pressure extrapolation of Supcrt92.

REFERENCES CITED
Alt, J.C., Shanks, W.C., and Jackson, M.C., 1993, Cycling
of sulfur in subduction zones: The geochemistry of
sulfur in The Mariana Island Arc and back-arc
trough: Earth and Planetary Science Letters, v. 119,
p. 477494.
Blundy, J.D., Brodholt, J.P., and Wood, B.J., 1991, Carbonfluid equilibria and the oxidation state of the upper
mantle: Nature, v. 349, p. 321324.
Brandon, A.D., and Draper, D.S., 1996, Constraints on the
origin of the oxidation state of mantle overlying subduction zones: An example from Simcoe, Washington, USA: Geochimica et Cosmochimica Acta, v. 60,
p. 17391749.
Bureau, H., and Keppler, H., 1999, Complete miscibility
between silicate melts and hydrous fluids in the upper
mantle: Experimental evidence and geochemical implications: Earth and Planetary Science Letters,
v. 165, p. 187196.
Carmichael, I.S.E., 1991, The redox states of basic and silicic magmasA reflection of their source regions:
Contributions to Mineralogy and Petrology, v. 106,
p. 129141.
Carroll, M.R., and Rutherford, M.J., 1987, The stability of
igneous anhydrite: Experimental results and implications for sulfur behaviour during the 1982 El Chichon trachyandesite and other evolved magmas:
Journal of Petrology, v. 28, p. 781801.
Drummond, M.S., Defant, M.J., and Kepezhinskas, P.K.,
1996, Petrogenesis of slab-derived trondhjemite-tonalite-dacite/adakite magmas: Royal Society of Edinburgh Transactions, Earth Sciences, v. 87,
p. 205215.
Fernandez, D.P., Goodwin, A.R.H., Lemmon, E.W., Levelt
Sengers, J.M.H., and Williams, R.C., 1997, A formulation for the static permittivity of water and
steam at temperatures from 238 K to 873 K at pressures up to 1200 MPa, including derivatives and Debye-Huckel coefficients: Journal of Physical and
Chemical Reference Data, v. 26, p. 11251166.
Frost, B.R., and Ballhaus, C., 1998, Constraints on the origin of the oxidation state of mantle overlying subduction zones: An example from Simcoe, Washington, USA: Comment: Geochimica et Cosmochimica
Acta, v. 62, p. 329331.
Ghiorso, M.S., and Sack, R.O., 1995, Chemical mass transfer in magmatic processes IV: A revised and internally consistent thermodynamic model for the interpolation and extrapolation of liquid-solid equilibria
in magmatic systems at elevated temperatures and
pressures: Contributions to Mineralogy and Petrology, v. 119, p. 197212.
Gutscher, M.A., Maury, R., Eissen, J.P., and Bourdon, E.,
2000, Can slab melting be caused by flat subduction?: Geology, v. 28, p. 535538.
Hamlyn, P.R., and Keays, R.R., 1986, Sulfur saturation and
second-stage melts: Application to the Bushveld platinum metal deposits: Economic Geology, v. 81,
p. 14311445.
Harvey, A.H., Peskin, A.P., and Klein, S.A., 1996, NIST/
ASME steam formulation for general and scientific
use: National Institute of Standards and Technology
Standard Reference Database 10, Version 2.2.
Johnson, J.W., Oelkers, E.H., and Helgeson, H.C., 1992,
Supcrt92: A software package for calculating the
standard molal thermodynamic properties of minerals, gases, aqueous species, and reactions from 1 to
5000 bar and 0 to 1000 8C: Computers and Geosciences, v. 18, p. 899947.
Kay, S.M., and Mpodozis, C., 2001, Central Andean ore
deposits linked to evolving shallow subduction systems and thickening crust: GSA Today, v. 11, no. 3,
p. 49.
Kay, S.M., and Mpodozis, C., 2002, Magmatism as a probe
to the Neogene shallowing of the Nazca plate beneath the modern Chilean flat-slab: Journal of South
American Earth Sciences, v. 15, p. 3957.
Kay, S.M., Kurtz, A., and Godoy, E., 1995, Tertiary magmatic and tectonic framework of the El Teniente cop-

per deposit, southwestern Chile (348S to 358S): Geological Society of America Abstracts with Programs,
v. 27, no. 6, p. A409.
Luhr, J.F., and Aranda-Gomez, J.J., 1997, Mexican peridotite xenoliths and tectonic terranes: Correlations
among vent location, texture, temperature, pressure
and oxygen fugacity: Journal of Petrology, v. 38,
p. 10751112.
Martin, H., 1999, Adakitic magmas: Modern analogues of
Archaean granitoids: Lithos, v. 46, p. 411429.
Mavrogenes, J.A., and ONeill, H.S., 1999, The relative
effects of pressure, temperature and oxygen fugacity
on the solubility of sulfide in mafic magmas: Geochimica et Cosmochimica Acta, v. 63, p. 11731180.
McInnes, B.I.A., and Cameron, E.M., 1994, Carbonated,
alkaline hybridizing melts from a sub-arc environment: Mantle wedge samples from the Tabar-LihirTanga-Feni arc, Papua New Guinea: Earth and Planetary Science Letters, v. 122, p. 125141.
Metrich, N., Schiano, P., Clocchiatti, R., and Maury, R.C.,
1999, Transfer of sulfur in subduction settings: An
example from Batan Island (Luzon volcanic arc,
Philippines): Earth and Planetary Science Letters,
v. 167, p. 114.
Muller, D., and Groves, D.I., 1993, Direct and indirect associations between potassic igneous rocks, shoshonites and gold-copper deposits: Ore Geology Reviews,
v. 8, p. 383406.
ONeill, H.S., and Wall, V.J., 1987, The olivine-spinel oxygen geobarometer, the nickel precipitation curve
and the oxygen fugacity of the upper mantle: Journal
of Petrology, v. 28, p. 11691192.
Paillat, O., Elphick, S.C., and Brown, W.L., 1992, The solubility of water in NaAlSi3O8 meltsA reexamination of Ab-H2O phase relationships and critical behavior at high pressures: Contributions to Mineralogy
and Petrology, v. 112, p. 490500.
Parkinson, I.J., and Arculus, R.J., 1999, The redox state of
subduction zones: Insights from arc-peridotites:
Chemical Geology, v. 160, p. 409423.
Peacock, S.M., Rushmer, T., and Thompson, A.B., 1994,
Partial melting of subducting oceanic crust: Earth and
Planetary Science Letters, v. 121, p. 227244.
Richards, J.P., 1990, Petrology and geochemistry of alkaline
intrusives at the Porgera gold deposit, Papua New
Guinea: Journal of Geochemical Exploration, v. 35,
p. 141199.
Sajona, F.G., and Maury, R.C., 1998, Association of adakites with gold and copper mineralization in the Philippines: Paris, Academie de Sciences Comptes Rendus, Sciences de la Terre et les Plane`tes, v. 326,
p. 2734.
Schiano, P., Clocchiatti, R., Shimizu, N., Maury, R.C., Jochum, K.P., and Hofmann, A.W., 1995, Hydrous, silica-rich melts in the sub-arc mantle and their relationship with erupted arc magmas: Nature, v. 377,
p. 595600.
Schmidt, M., 1996, Experimental constraints on recycling
of potassium from subducted oceanic crust: Science,
v. 272, p. 19271930.
Schneider, M.E., and Eggler, D.H., 1986, Fluids in equilibrium with peridotite minerals: Implications for mantle metasomatism: Geochimica et Cosmochimica
Acta, v. 50, p. 711724.
Sillitoe, R.H., 1997, Characteristics and controls of the largest porphyry copper-gold and epithermal gold deposits in the circum-Pacific region: Australian Journal of
Earth Sciences, v. 44, p. 373388.
Solomon, M., 1990, Subduction, arc reversal, and the origin
of porphyry copper-gold deposits in island arcs: Geology, v. 18, p. 630633.
Thieblemont, D., Stein, G., and Lescuyer, J.-L., 1997, Gisements epithermaux et porphyriques: La connexion
adakite: Academie de Sciences Paris, Comptes Rendus, v. 325, p. 103109.
Wyborn, D., and Sun, S.-S., 1994, Sulphur-undersaturated
magmatismA key factor for generating magma-related copper-gold deposits: AGSO Research Newsletter, v. 21, p. 78.
Manuscript received April 2, 2002
Revised manuscript received June 20, 2002
Manuscript accepted June 25, 2002
Printed in USA

GEOLOGY, October 2002

Vous aimerez peut-être aussi