You are on page 1of 4

Modeling of gold scavenging by bismuth melts

coexisting with hydrothermal fluids


Blake Tooth1, Jol Brugger1,2, Cristiana Ciobanu1,2, Weihua Liu3
School of Earth and Environmental Sciences, University of Adelaide, Adelaide, South Australia 5005, Australia
2
South Australian Museum, North Terrace, Adelaide, South Australia 5000, Australia
3
CSIRO Exploration & Mining, School of Geosciences, Monash University, Clayton, Victoria 3168, Australia

ABSTRACT
The effect of gold scavenging by bismuth melts is investigated using equilibrium thermodynamic modeling of an aqueous solutionmineralmelt system. The calculations for the
Au-Bi-Na-Cl-S-H-O system, performed at temperatures between 300 and 450 C, demonstrate
that Au concentrations in the melt are several orders of magnitude higher than in the coexisting
fluid, indicating the possible formation of economic gold deposits from undersaturated aqueous
fluids, in which mineralization would not be expected in the absence of a bismuth melt. The
model applies to any deposit where a Bi melt is stable and coexists with a hydrothermal fluid;
examples of such deposits are known from skarn, intrusion-related, orogenic, and volcanogenic
massive sulfide (VMS) gold systems. In sulfur-poor systems the partitioning curves presented
here can be used directly to correlate the gold concentration in the fluid and the Au grade in
the ore (e.g., Escanaba Trough VMS deposit). These results also illustrate important principles
generally applicable to understanding magmatic-hydrothermal and metamorphic deposits that
may have contained significant volumes of more complex polymetallic melts.
Keywords: hydrothermal fluids, melts, thermodynamics, numerical modeling, gold, bismuth, ore
deposits.
INTRODUCTION
There is now significant evidence of a genetic role for polymetallic
melts in metamorphosed ore deposits, where they concentrate metals as
melt components by diffusion and dissolution during melt mobilization
(e.g., Tomkins et al., 2007). In hydrothermal systems affected by lower
grade metamorphism, this can occur by partitioning of aqueous fluid components into coexisting metallic melts (e.g., Douglas et al., 2000). Bismuth
is a major component of low melting point assemblages, particularly in
variations of the Au-Bi-Te-S system, common in magmatic-hydrothermal
gold deposits. For example, an Au-Bi correlation is distinctive in intrusionrelated gold (IRG) systems with correlation coefficients of 0.70.9 commonly reported (e.g., Baker et al., 2005). Au skarns also display strong
Au-Bi correlations (e.g., Meinert, 2000), as do some orogenic systems
(e.g., Hattu Schist Belt, Finland; Nurmi and Sorjonen-Ward, 1993).
Native bismuth (melting point 271 C) and Bi-rich polymetallic assemblages (e.g., Au-Bi with a eutectic of 241 C; Fig. 1) are molten at temperatures that overlap with the formation conditions of a large range of gold
deposits. The implications of this were illustrated by Douglas et al. (2000),
who presented preliminary experiments in which Au was scavenged from
hydrothermal solutions at temperature, T ~300 C by Bi melt, a process
they termed the liquid bismuth collector model. This model has been considered for interpreting Au-Bi deposits across the magmatic-hydrothermal
spectrum, including: (1) IRG veins at Pogo, Fort Knox (Tintina Belt, Alaska;
McCoy, 2000); (2) epithermal-porphyry transition at Larga (South Apuseni
Mountains, Romania; Cook and Ciobanu, 2004); (3) Au skarns as in the
Ortosa and El Valle in the Rio Narcea Gold Belt (Asturias, Spain; Cepedal
et al., 2006); and (4) recent volcanic massive sulfide (VMS) system in the
Escanaba Trough (Southern Gorda Ridge; Toermanen and Koski, 2005).
The phase relationships of the Au-Bi binary system have been well
described in metallurgical literature (see the GSA Data Repository1).

The Au-Bi binary phase diagram illustrates the potential of Bi melt as


a gold scavenger (Fig. 1). An Au-Bi melt has a eutectic with ~19 wt%
Au at 241 C, and thus is able to incorporate much higher Au concentrations than aqueous fluids at this and higher temperatures. This solubility difference means that Au is expected to partition into Bi melts even
from undersaturated aqueous solutions. However, this model has not been
thermodynamically assessed due to a lack of capability of most geochemical thermodynamic modeling programs to handle complex melts.
In this paper we assess Au scavenging by the Bi melt using equilibrium thermodynamic modeling involving an aqueous electrolyte solution, Au-Bi minerals, and a non-ideal Au-Bi melt. This model is used to
address two important questions related to the genetic role of melts in
hydrothermal systems: under which conditions a fluid and melt coexist,

1
GSA Data Repository item 2008209, thermodynamic model and properties
for the Au-Bi melt under hydrothermal conditions, is available online at www.
geosociety.org/pubs/ft2008.htm, or on request from editing@geosociety.org or
Documents Secretary, GSA, P.O. Box 9140, Boulder, CO 80301, USA.

Figure 1. Phase diagram for Au-Bi system at 1 bar, calculated using


HCh thermodynamic model developed in this study (see Data Repository [see footnote 1]). Experimental data points on liquidus composition by Nathans and Leider (1962) are shown as solid circles.

450

Gold + Melt

350

Melt
Maldonite
+ Gold

Maldonite + Melt

300
250

Bismuth
+ Melt

200
150

Maldonite + Bismuth

Temperature (C)

400

100

Mole % Bi

2008 The Geological Society of America. For permission to copy, contact Copyright Permissions, GSA, or editing@geosociety.org.
GEOLOGY,
October
2008
Geology,
October
2008;
v. 36; no. 10; p. 815818; doi: 10.1130/G25093A.1; 4 figures; Data Repository item 2008209.

90

50
80

70

60

50

40

30

20

Au

10

Gold + Bismuth
Bi

815

THERMODYNAMIC MODELING OF AU SCAVENGING


Equilibrium among the minerals + melt + aqueous fluid system was
computed using the Gibbs free energy minimization algorithm employed
in the HCh package (Shvarov and Bastrakov, 1999). The Bi-Au melt was
introduced as a distinct phase in the model, and its Gibbs free energy
described using the non-random two liquids (NRTL) equation (Renon and
Prausnitz, 1968). The three empirical NRTL parameters were fitted to the
Au-Bi binary diagram. The modeled system was Au-Bi-Na-Cl-S-H-O,
and the calculations included the minerals maldonite, native bismuth, bismuthinite, and native gold. To avoid the additional chemical and numerical
complexity related to the addition of extra components, mineral buffering
reactions likely to control pH [e.g., CO2 (g) , silicates] and f O2(g) and f S2(g)
(e.g., magnetite-pyrite-pyrrhotite) are excluded. The pH was therefore
adjusted using individual components such as NaOH and HCl, and the
effect of redox was calculated at different f O2(g) values. For full details of
the thermodynamic model and properties used, see the Data Repository.
The central problem, i.e., how much Au exists in a fluid at equilibrium with an Au-Bi melt, was explored by equilibrium calculations for
a closed system consisting initially of a hydrothermal fluid and liquid

816

20

300 C, 500 bar


aAu = 103
aS = 101
~5% NaCl

AuCl2

25

10

300 C, 500 bar


aBi = 104.3
aS = 101
~5% NaCl

15
20
3

BiCl6

25

AuOH(aq)

30

Bi(OH)3(aq)

30

Gold

40

300 C

Au
Cl
4

10

15

AuCl 2

10

12

450 C, 500 bar


aAu = 103
aS = 101
~5% NaCl

20

14

Bi-melt

40

Po

AuHS(aq)

20

10

12

Mt

Po

Bi-melt

30

pH

14

14

Hm

Bsm Py

450 C

450 C
0

12

Bi(OH)3(aq)

BiCl6

CH4/CO 2

Au(HS)2 Gold

10

450 C,1000 bar


aBi = 104.3
aS = 101
~5% NaCl

pH 6.0

25

30

300 C
6

15

a
c

10

AuOH(aq)
25

Mt
Py

KMQ buffer

Au(OH)2

Gold
0

CH4/CO 2

Hm

Bism

35

Au(HS)2

Au
H

35

15

Au(OH)2

KMQ buffer

Au
Cl
4

10

S(
aq
)

log a O2(aq)

COMPARATIVE GEOCHEMISTRY OF Au AND Bi


There is broad agreement about Au speciation under conditions most
relevant to deposit formation: Au occurs as Au+ aqueous complexes, of
which AuHS(aq) , Au(HS)2, and AuCl2 are of particular importance depending on the concentration of chloride and sulfur in the fluids (Figs. 2A and
2C) (e.g., Stefansson and Seward, 2003, 2004). For example, Au bisulfide
complexes predominate in the low-salinity fluids with moderate sulfur
contents, typical of orogenic Au systems (e.g., Mikucki, 1998), and Au
chloride complexes are important in the sulfur-poor, saline to hypersaline
fluids typical of gold skarn deposits (Meinert, 2000).
In contrast, limited information is available about the geochemistry
of Bi at temperatures relevant to ore formation. The most important oxidation state of Bi in aqueous solutions is Bi3+, but only two experimental
studies are dedicated to the hydrothermal transport of Bi up to 300 C,
close to the minimum temperature relevant for scavenging by Bi melts
(Kolonin and Laptev, 1982; Wood et al., 1987). Bismuth complexes with
sulfide ligands may be important, but the only available properties have
been extrapolated based on comparison with the other metalloids, Sb and
As, and remain to be experimentally confirmed (Skirrow and Walshe,
2002; Wood et al., 1987). The speciation model outlined by Skirrow and
Walshe (2002) is the most complete to date and forms the basis of the
speciation for Bi used in the present study (see the Data Repository).
However, the lack of experimental studies on Bi hydrothermal geochemistry means that the predicted Bi mineral solubilities must be considered
as semiquantitative only.
The diagrams in Figure 2 show the solubility and aqueous speciation
for both Au and Bi at temperatures relevant for the formation of the main
types of Au deposits (e.g., Mikucki, 1998; Cooke and Simmons, 2000;
Meinert, 2000; Baker et al., 2005) but also above the melting point of
native bismuth (271 C). At 300 C, the fields of Au aqueous bisulfide
complexes and solid bismuthinite (melting point 775 C; Lin et al., 1996)
overlap, thus no Bi melt is predicted to coincide with high Au solubility
in the fluid (Figs. 2A and 2B). However, if the hydrothermal system is
reduced, i.e., pyrrhotite stability or CH4(g) rich, Bi melts can coexist with
Au-carrying fluids. At 450 C the bismuthinite stability field shrinks, while
at the same temperature Au chloride complexes become important for
acidic to near neutral fluids (Fig. 2B). Therefore, at higher temperatures
there is more scope for coexistence of a Bi melt and Au-bearing fluids.

log a O2(aq)

and what are the resultant equilibrium concentrations of Au in the melt


and fluid? The modeling will help produce accurate predictions of the
physico-chemical conditions under which the Bi collector may occur and
of the ore grades arising from such a mechanism.

10

12

14

pH

Figure 2. fO2(aq) versus pH diagrams showing stability fields of Au-Bi


minerals and aqueous complexes for 300 and 450 C. For reference,
predominance fields for common Fe minerals and the CH4(g) /CO2(g)
[fCO2(g) = fCH4(g) ] boundary are shown with pH values for the KMQ
(K-feldsparmuscovitequartz) buffer (aK+ = 0.1 and 0.01). Circled
letters on diagrams correspond to conditions for which corresponding partitioning curves are drawn in Figure 3.

bismuth, to which additional small amounts of Au are added. The initial


melt composition was 100% Bi, and final melt composition was that of the
Au-Bi solvus at the chosen temperature, and is the point where Bi melt is
saturated with Au, i.e., it coexists with an Au mineral (either maldonite or
native gold). The resulting Au partitioning curves between fluid and melt
are shown in Figure 3 for several fluid compositions at 450 C and for a
specific example at 300 C. These conditions are chosen to illustrate the
effect of various parameters on the partitioning of Au between the melt
and aqueous phases.
The process of Au incorporation into Bi melt can be illustrated by
reactions with fluids that carry the predominant types of Au aqueous complexes in the hydrothermal systems discussed in the previous section. The
most important variables can be summarized by the following equilibria for the case of chloride- and sulfur-rich fluids, representing the predominant melt partitioning processes in skarn and orogenic Au deposits,
respectively:
4AuCl2 + 2H2O = 4Au0 (in Bi melt) + 8Cl + 4H+ + O2(g) ,

(1)

Cl-bearing fluid (e.g., skarn);


4Au(HS)2 + 4H+ + 2H2O = 4Au0 (in Bi melt) + 8H2S(aq) + O2(g) , (2)
S-bearing fluid (e.g., orogenic Au).
The equilibrium gold concentration is influenced by the redox potential of the fluid [ f O2(g)], pH, and the concentrations of chloride and bisulfide ions in solution. For equation 1, low pH, high salinity, and f O2(g) favor
Au solubility in the fluid. The main difference between equilibria 1 and 2
is in the effect of pH in controlling HS concentration in a reduced environment, and consequently Au solubility, which is at a maximum where
pH is equal to the first dissociation constant (pK1) of H2S(aq) .
Along curves a and b in Figure 3, the fluid has the same alkaline pH
(7.6) but 3 units difference in log f O2(g) . Taking an arbitrary Au weight

GEOLOGY, October 2008

Au in melt (wt%)
0

10

20

30

40

100

Au in fluid (ppb)

10

0.1
2

10

10
10

10

13 19 20

pH 6

0.1

10

10

pH 5
102
103

Figure 3. Calculation of Au partitioning between Bi-rich melt and


aqueous fluid. Plot of log concentration of Au (ppb) in hydrothermal
fluid versus Au in coexisting melt (wt%) for system Au-Bi-Na-ClS-H-O at 450 C (main diagram) and 300 C (inset). Concentration
of chloride is 1 molal and total sulfur concentration is 0.001 molal,
unless otherwise stated. Other conditions for each curve are as follows. apH = 7.6, fO2(g) = 25; bpH = 7.6, fO2(g) = 28; cpH = 5.6,
fO2(g) = 28; dpH = 7.6, fO2(g) = 28, no S; epH = 10.5, fO2(g) = 28.
Circles represent saturation with native gold, stars represent saturation with maldonite. Inset: partitioning curves drawn for conditions
of Escanaba Trough fluids; T = 300 C, fO2(g) = 40, concentration of
H2S = 0.01 molal, concentration of Cl = 0.5 molal, and pH values = 5
and 6. Range of observed melt compositions for given range of fluid
characteristics is between these curves.

percent value in the Au-Bi melt (e.g., 20 wt%), the Au concentration


between the two fluids is nearly an order of magnitude higher for the
more oxidized case (curve a). This means that a Bi melt can incorporate
the same amount of gold from a less concentrated fluid at more reducing
conditions, e.g., pyrrhotite stability or at the CO2 /CH4 buffer (curve b)
versus pyrite stability (curve a) (compare Figs. 2C and 2D). The same
melt composition will result from a fluid with two orders of magnitude
less Au if a similar reduced fluid as in curve b is considered without
sulfur (curve e). The effect of the change in pH in the partitioning of Au
between an S-bearing fluid and the melt is illustrated by curves c, b, and
d, of Figure 3, which define a profile across Figures 2C and 2D from
acidic to highly alkaline conditions, at the same reduced f O2(g) value
(note that this comparison is qualitative only because of different sulfur
activities between the curves and the activity diagrams). As the condition where pH = pK1 of H2S(aq) is approached (curve b in Fig. 3), Au
solubility in the fluid is highest, close to an order of magnitude higher
than curve c and two orders of magnitude higher than curve d. We have
deliberately excluded consideration of the very acidic case, because as
can be observed in Figure 2, bismuth solubility as chloride complexes is
large at low pH, destabilizing Bi melt under these conditions.
Another important result of these calculations is that a fluid that is
undersaturated with respect to native gold will coexist with a bismuth melt
that contains large amounts of Au. For example, at saturation a hydrothermal fluid corresponding to curve b (Fig. 3) contains ~20 ppb Au and
coexists with native gold and a melt with 42 wt% Au. Under such conditions, a fluid undersaturated by an order of magnitude (2 ppb Au) would
coexist with a Bi melt that still contains ~18 wt% Au. Even at 0.2 ppb,
this fluid coexists with a melt containing ~5 wt% Au. This means that a
deposit containing only 100 ppm Bi deposited as Bi melt in the presence
of an excess of such a fluid would display Au grades of 5 ppm. These

GEOLOGY, October 2008

values are well in the range found in many deposits displaying the
Au-Bi association: IRG deposits commonly display Bi grades >100 ppm
(in many cases >1000 ppm; Thompson et al., 1999). The most important
consequence of the melt model is that economic deposits can form even
when Au is undersaturated by several orders of magnitude in the fluid.
DISCUSSION
A Bi-rich melt may coexist with an aqueous fluid at temperatures
higher than the melting point of native bismuth. The precipitation of a
Bi-rich melt instead of bismuthinite is favored mainly by high temperatures (400 C) or by reducing conditions ( f CH /f CO > 1; pyrrhotite
4
2
stable). These conditions correspond to the reported occurrences of Au
scavenging by Bi melts. The Au-Bi signature recognized in deposits of
magmatic affinity such as skarns (Meinert, 2000) and some of the IRG
deposits (Baker et al., 2005) may suggest that they are best suited for
application of the model presented here. However, given the geochemical
affinity between Au and Bi (e.g., Spooner, 1993), such an Au-Bi signature
may simply reflect the partitioning of Au and Bi at comparable high values from the magmatic source. Thus, understanding of phase relationships
among minerals is essential for assuming a melt scenario.
The Escanaba Trough VMS system is one such example best suited
for application of the Bi-collector model (Toermanen and Koski, 2005).
On the basis of well-preserved mineral assemblages, Bi melts were interpreted to form during the circulation of hydrothermal fluids through
pyrrhotite-rich massive sulfide lenses within organic-matter rich sediments near an active ocean ridge system. The conditions of mineralization [T ~300 C; CH4(g) -rich fluids equilibrating with Po] correspond to
conditions in Figures 2A and 2B, in which Au-bearing fluids coexist with
Bi melt. If we assume the Au in the Escanaba Trough was enriched solely
though the Bi-melt collector, model predictions can be tested by comparing Au and Bi grades. The average Bi concentration is 65 ppm and Au
concentrations range from 1.4 to 10.1 ppm (Toermanen and Koski, 2005).
Assuming that the bulk of both Au and Bi precipitated as Au-Bi melt, the
calculated proportion of Au in these melts is in the range 213 wt% Au.
Under the conditions of formation of this deposit (pyrrhotite stable,
pH = 56, T = 275325 C), Au solubility as AuHS2 is up to 1 ppb, and the
liquidus composition of the Au-Bi melt is ~2024 wt% Au. Hence,
the hydrothermal fluid at Escanaba may have been undersaturated by ~12
orders of magnitude with respect to maldonite (Fig. 3 inset), and in the
absence of the Bi collector, no Au mineralization would have occurred.
The Bi-mineral associations observed in Au deposits are often
more complex than those considered here, usually including S- and/or
Te-bearing minerals. Primary phase relationships and textures may also
be obliterated due to interaction with aqueous fluids postdating melt
precipitation. One common case of overlapping mineralizing events is
illustrated by metamorphic terranes where magmatic intrusions predate
or postdate an orogenic event, e.g., at Maldon (Victoria, Australia). The
typical ore shows the coexistence of maldonite with native Bi (Fig. 4)
representing the eutectic association formed at 241 C in the Au-Bi system. Maldonite partially decomposes into symplectites of native Au and
Bi during cooling below ~116 C. However, the same ores include Bitellurides and/or sulfotellurides (e.g., joseite B, Fig. 4), jonassonite, and
bismuthinite (Ciobanu et al., 2007). Of these, only bismuthinite shows
clear overprint of former assemblages. The others may be interpreted
either as (1) the result of interaction between later, S- and Tebearing
fluids and minerals from the Au-Bi association; or (2) crystallization
from Au-Bi-Te-S melts formed from either orogenic or magmatically
derived fluids. The latter implies that the applicability of the thermodynamic model to natural systems and the accuracy of the predictions
can be improved by considering additional components. However, the
more complex a melt composition, e.g., 45 components such as Au-Bi(Pb)-Te-S, the more difficult it is to recognize equilibrium associations
(eutectics), given the difficulty in obtaining them experimentally.

817

Figure 4. Microphotograph (oil immersion) of typical gold ore at


Maldon (Victoria, Australia) illustrating association of maldonite
with native bismuth that represents eutectic assemblage at 241 C
in Au-Bi system. Note areas with symplectites showing decomposition of maldonite into native gold and bismuth and grain of joseite B
at left side margin of this patch, suggesting melt composition in the
Au-Bi-Te-S system.

A more diverse melt composition has important effects, including


melting point depression and changing stability of the melt with respect to
important physico-chemical parameters. For example, the addition of Te
would likely increase the f O2(g) range for Bi-containing melts coexisting
with Au-bearing fluids. Native tellurium stability is coincident with hematite
(e.g., McPhail, 1995), and observations of natural assemblages support this
hypothesis (e.g., Ciobanu et al., 2005). This will extend the application of a
Bi-melt collector model to epithermal systems that are otherwise excluded
given their rather oxidized nature (e.g., Cooke and Simmons, 2000).
In general, the model is useful for understanding the formation of
any deposit where a Bi melt coexists with a hydrothermal fluid, if equilibrium can be assumed. Because aqueous fluids are ubiquitous in many
environments that may also contain low-temperature polymetallic melts
(200400 C), the interaction between such melts and fluids is probably a
more significant mechanism of ore enrichment and geochemical exchange
than currently recognized.
ACKNOWLEDGMENTS
We are grateful to Bill Birch for providing the Maldon specimen (see Fig. 4),
to John Mavrogenes and Dick England for valuable insight into previous work,
and to Andy Tomkins and Adam Simon for thorough reviews. This work was
enabled by Australian Research Council fellowships DP0208323 to Brugger and
DP0560001 to Ciobanu.
REFERENCES CITED
Baker, T., Pollard, P.J., Mustard, R., Mark, G., and Graham, J.L., 2005, A comparison of granite-related tin, tungsten, and gold-bismuth deposits: Implications for exploration: Society of Economic Geologists Newsletter, v. 61,
p. 517.
Cepedal, A., Fuertes-Fuente, M., Martin-Izard, A., Gonzalez-Nistal, S., and
Rodriguez-Pevida, L., 2006, Tellurides, selenides and Bi-mineral assemblages from the Ro Narcea Gold Belt, Asturias, Spain: Genetic implications in Cu-Au and Au skarns: Mineralogy and Petrology, v. 87, p. 277304,
doi: 10.1007/s0071000601277.
Ciobanu, C.L., Cook, N.J., and Pring, A., 2005, Bismuth tellurides as gold
scavengers, in Mao, J.W., and Bierlein, F.P., eds., Mineral deposit research:
Meeting the global challenge: Berlin, Heidelberg, New York, Springer,
p. 13871390.
Ciobanu, C.L., Birch, W., Pring, A., and Cook, N.J., 2007, Au-Bi-Te-S assemblages from Maldon gold deposit, Victoria, Australia: Geological Society of
America Abstracts with Programs, v. 39, no. 6, p. 626.

818

Cook, N.J., and Ciobanu, C.L., 2004, Bismuth tellurides and sulfosalts from
the Larga hydrothermal system, Metaliferi Mts, Romania: Paragenesis
and genetic significance: Mineralogical Magazine, v. 68, p. 301321, doi:
10.1180/0026461046820188.
Cooke, D.R., and Simmons, S.F., 2000, Characteristics and genesis of epithermal
gold deposits: Reviews in Economic Geology, v. 13, p. 221244.
Douglas, N., Mavrogenes, J., Hack, A., and England, R., 2000, The liquid bismuth
collector model: An alternative gold deposition mechanism, in Skilbeck,
C.G., and Hubble, T.C.T., eds., Understanding planet Earth; searching for
a sustainable future; on the starting blocks of the third millennium: Australian Geological Convention, 15th, Sydney, Abstracts: Sydney, Geological
Society of Australia, 135 p.
Kolonin, G.R., and Laptev, Y.V., 1982, Study of process of dissolution of a-Bi2O3
(bismite) and complex formation of bismuth in hydrothermal solutions:
Geokhimiya, v. 11, p. 16211631.
Lin, J.C., Sharma, R.C., and Chang, Y.A., 1996, The Bi-S (bismuth-sulfur) system: Journal of Phase Equilibria and Diffusion, v. 17, p. 132139, doi:
10.1007/BF02665790.
McCoy, D.T., 2000, Mid-Cretaceous plutonic-related gold deposits of interior
Alaska: Metallogenesis, characteristics, gold associative mineralogy and
geochronology: [Ph.D. thesis]: Fairbanks, University of Alaska, 245 p.
McPhail, D.C., 1995, Thermodynamic properties of aqueous tellurium species
between 25 and 350C: Geochimica et Cosmochimica Acta, v. 59, p. 851
866.
Meinert, L.D., 2000, Gold in skarns related to epizonal intrusions: Reviews in
Economic Geology, v. 13, p. 347375.
Mikucki, E.J., 1998, Hydrothermal transport and depositional processes in
Archean lode-gold systems: A review: Ore Geology Reviews, v. 13, p. 307
321, doi: 10.1016/S01691368(97)000255.
Nathans, M.W., and Leider, M.J., 1962, Studies on bismuth alloys. I. Liquidus
curves of the bismuth-copper, bismuth-silver, and bismuth-gold systems:
Physical Chemistry, v. 66, p. 20122015, doi: 10.1021/j100816a043.
Nurmi, P.A., and Sorjonen-Ward, P., eds., 1993, Geological development, gold
mineralization and exploration methods in the late Archean Hattu Schist
Belt, Ilomantsi, eastern Finland: Geological Survey of Finland Special
Paper 17, 392 p.
Renon, H., and Prausntiz, J.M., 1968, Local compositions in thermodynamic
excess functions for liquid mixtures: American Institute of Chemical Engineers Journal, v. 14, p. 135144.
Stefansson, A., and Seward, T.M., 2003, Stability of gold(I) chloride complexes
in aqueous solutions from 300 to 600C and from 500 to 1800 bar: Geochimica et Cosmochimica Acta, v. 67, p. 45594576, doi: 10.1016/S0016
7037(03)003910.
Stefansson, A., and Seward, T.M., 2004, Gold(I) complexing in aqueous sulphide
solutions to 500C at 500 bar: Geochimica et Cosmochimica Acta, v. 68,
p. 41214143, doi: 10.1016/j.gca.2004.04.006.
Shvarov, Y., and Bastrakov, E., 1999, HCh: A software package for geochemical
modelling. Users guide: AGSO Record, v. 1999/25, p. 60.
Skirrow, R.G., and Walshe, J.L., 2002, Reduced and oxidised Au-Cu-Bi iron
oxide deposits of the Tenant Creek Inlier, Australia: An integrated geologic
and chemical model: Economic Geology and the Bulletin of the Society of
Economic Geologists, v. 97, p. 11671202.
Spooner, E.T.C., 1993, Magmatic sulphide/volatile interaction as a mechanism
for producing chalcophile element enriched, Archean Au-quartz, epithermal
AuAg and Au skarn hydrothermal ore fluids: Ore Geology Reviews, v. 7,
p. 359379.
Thompson, J.F.H., Sillitoe, R.H., Baker, T., Lang, J.R., and Mortensen, J.K.,
1999, Intrusion related gold deposits associated with tungsten-tin provinces: Mineralium Deposita, v. 34, p. 323334.
Toermanen, T.O., and Koski, R.A., 2005, Gold enrichment and the Bi-Au association in pyrrhotite-rich massive sulfide deposits, Escanaba Trough, Southern
Gorda Ridge: Economic Geology and the Bulletin of the Society of Economic Geologists, v. 100, p. 11351150.
Tomkins, A., Pattison, D., and Frost, R., 2007, On the initiation of metamorphic sulfide anatexis: Journal of Petrology, v. 48, p. 511535, doi: 10.1093/
petrology/egl070.
Wood, S.A., Crerar, D.A., and Borcsik, M.P., 1987, Solubility of the assemblage
pyrite-pyrrhotite-magnetite-sphalerite-galena-gold-stibnite-bismuthiniteargentite-molybdenite in H2O-NaCl-CO2 solutions from 200350C: Economic Geology and the Bulletin of the Society of Economic Geologists,
v. 82, p. 18641887.
Manuscript received 25 April 2008
Revised manuscript received 1 July 2008
Manuscript accepted 2 July 2008
Printed in USA

GEOLOGY, October 2008

You might also like