Sie sind auf Seite 1von 29

13.

2 The Chemistry of Metal Transport and Deposition by Ore-Forming


Hydrothermal Fluids
TM Seward, Victoria University of Wellington, Wellington, New Zealand
AE Williams-Jones and AA Migdisov, McGill University, Montréal, QC, Canada
ã 2014 Elsevier Ltd. All rights reserved.

13.2.1 Introduction 29
13.2.1.1 Compositions of Ore-Forming Fluids 30
13.2.2 Hydrothermal Ore Solution Chemistry – The Main Dissolved Components 33
13.2.2.1 Water Solvent at Hydrothermal Conditions 33
13.2.2.2 NaCl – The Main Dissolved Electrolyte Component 35
13.2.2.3 Ion Hydration, Association, and Water Activity 35
13.2.2.4 Weak Acid/Base Equilibria in Hydrothermal Systems 36
13.2.3 Mineral Solubility in Water and Salt Solutions at High Temperature and Pressure 37
13.2.4 Ore Metal Transport and Deposition 40
13.2.4.1 Ore Fluids with Liquid-Like Densities 40
13.2.4.1.1 Ligands in hydrothermal ore solutions 40
13.2.4.1.2 Metal chloride complexing 41
13.2.4.1.3 Complexing with other halide ligands 42
13.2.4.1.4 Metal complexes with hydroxide and other oxygen electron donor ligands 43
13.2.4.1.5 Complexing with hydrosulfide/sulfide ligands 45
13.2.4.1.6 Thioanions 46
13.2.4.1.7 Complexing with other sulfur-containing ligands 47
13.2.4.1.8 Other complexing ligands 48
13.2.4.1.9 Ore fluids with gas-like density 48
13.2.5 Epilogue 50
Acknowledgments 50
References 50

13.2.1 Introduction example, the fluids venting through chimneys at spreading


centers and subduction zones are currently forming deposits
Economically exploitable deposits of metallic minerals (ore with the characteristics of volcanogenic massive sulfide (VMS)
deposits) form in the Earth’s crust through a variety of geologic base metal deposits (Scott, 1997). There is strong evidence
processes. These involve extraction of metals at low that, in many cases, geothermal fluids of the type employed
concentration and transport of these metals to sites of deposi- in energy generation are responsible for the formation of low-
tion where they accumulate in very much higher concentra- sulfidation epithermal precious metal deposits (Barnes and
tion. Hydrothermal fluids are by far the most important agents Seward, 1997; Clark and Williams-Jones, 1990; Krupp and
of metal transport. They may be formational waters of meteoric Seward, 1987, 1990; Weissberg et al., 1979). The same is true
or seawater origin, metamorphic fluids produced during devo- for oil-field brines, which are analogous to the basinal brines
latilization of hydrous minerals, or magmatic fluids released interpreted to form Mississippi Valley-type (MVT) lead–zinc
during decompression and/or crystallization of magmas. Their deposits. Although they are not ore fluids, per se, volcanic
physical state may be that of liquid, vapor, or a supercritical gases are representative of the fluids that at greater pressure
fluid, and chemically, they have highly variable concentrations may form high-sulfidation epithermal precious metal and por-
of dissolved components, including charged and uncharged phyry copper deposits. Indeed, the ore fluids forming these
species containing elements such as Na, K, Ca, Fe, Si, Cl, C, deposits actually may be gases, albeit of greater density than
H, O, and S and, of course, the ore metals. The challenge, if volcanic gases (Williams-Jones and Heinrich, 2005). The fluids
we are to understand the processes involving the hydrothermal from these active hydrothermal systems have been sampled
extraction, transport, and deposition of metals, is to determine extensively during the past 50 years and have provided
the nature of these fluids, the properties that allow them to people with a wealth of data on the chemistry of ore-forming
dissolve metals, and the conditions under which dissolution fluids and, in some cases (e.g., VMS and low-sulfidation
and deposition are optimized. epithermal deposits), the pressure–temperature conditions of
Active hydrothermal systems are an important source of potential ore formation. The other important sources of infor-
information on the composition of ore fluids either because mation on the composition of ore-forming fluids and the
they are forming deposits analogous to those that are mined or conditions of ore formation are fluid inclusions. With the
there is strong evidence linking them to such deposits. For recent development of tools capable of analyzing their element

Treatise on Geochemistry 2nd Edition http://dx.doi.org/10.1016/B978-0-08-095975-7.01102-5 29


30 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

concentrations down to ppm levels, fluid inclusions are now (0.25 m) and F contents (0.4 mm), and higher concentrations
starting to provide chemical data comparable in quality to of K (up to 0.1 m). These differences are attributed to the
those available from active hydrothermal systems. They are addition of a large magmatic component to the hydrothermal
also the principal source of data on the pressure–temperature fluid, possibly as much as 25% (Reeves et al., 2011).
conditions of ore formation. The only ligands likely to be present in concentrations
sufficient to form the complexes necessary for metal transport
are Cl, CO2 species, reduced sulfur species, and F (in arc
13.2.1.1 Compositions of Ore-Forming Fluids
environments) (Table 1). Ammonium ions are generally not
Hydrothermal fluids venting on the seafloor were first discov- detected, except in sediment-covered settings (German and
ered in 1977 (Weiss, et al., 1977). Since then, large numbers of Von Damm, 2003), and in these settings, it seems likely that
studies, mainly along mid-ocean ridges, have documented the they are incorporated at relatively shallow depths of several
chemistry of these fluids (Table 1) and shown that those at hundred meters (Lizarralde et al., 2010) and thus are not likely
T  350  C are forming massive sulfide base metal (and gold) to play a significant role in metal extraction, which is thought
deposits similar to VMS deposits mined in terrestrial settings to mainly take place at greater depth (Lowell et al., 1995;
(Scott, 1997; Tivey, 2005). Studies of mid-ocean ridge systems Tivey, 2005).
have established that the composition of the fluids (for the Geothermal well fluids have been extensively sampled and
most part, hydrothermal liquids) is largely the result of inter- analyzed for over 50 years. Until recently, this sampling had
action of seawater with basalt; there are also minor additions of been restricted to the wellhead, and because the fluid generally
magmatic fluids (Shanks, 2001; Yang and Scott, 1996, 2006). boils (a requirement for conventional geothermal energy pro-
During this interaction, there is a sharp decrease in ex situ pH duction) and precipitates minerals before reaching the surface,
(from 7.8 to between 3 and 4) due to removal of magnesium the composition of the fluid differs considerably from that of
to form secondary silicate minerals (Seyfried, 1987). Chlorine the reservoir fluid. Geothermal fluids display a broad range of
is conserved, and consequently, the chlorinity of the hydro- compositions, reflecting the diverse geology of terrestrial set-
thermal fluids is generally very similar to that of seawater, that tings (Table 1). For example, the chlorinity of the high-
is,  0.5 m (German and Von Damm, 2003). Sodium, which is temperature fluids (>250  C) in the Taupo volcanic zone, a
the overwhelmingly dominant cation in the hydrothermal young rifted arc dominated by andesites and rhyolites, ranges
fluids, is also largely conserved, in most cases, retaining its between 0.02 and 0.05 m, an order of magnitude lower than that
seawater concentration of  0.45 m; some Na may be lost to of seawater (Ellis, 1979a,b; Ellis and Mahon, 1977). At the other
albitization. Among the major elements, only Mg and S are lost extreme, the high-temperature fluids (>250  C) in the Salton Sea
from the seawater, Mg almost completely so. The sulfur con- geothermal system, which is located in a closed evaporitic,
tent drops from nearly 30 mm in seawater to <10 mm in the sedimentary basin (Imperial Valley), commonly have a chlorin-
vent fluid. However, the major change is in its oxidation state. ity between 3 and 4 m; immediately to the south, in the same
Whereas in seawater, sulfur occurs as sulfate, in the vent fluid it basin, geothermal fluids at Cerro Prieto have a chlorinity of 0.2
is present dominantly as H2S (Shanks, 2001). The carbon to 0.5 m (Mercado and Hurtado, 1992; Williams and McKibben,
content, mainly in the form of CO2, is typically two to five 1989). In both the Taupo volcanic zone and the Imperial Valley,
times its seawater content of  2.3 mm. The other major ele- the fluid is predominantly meteoric water with a small contri-
ments, that is, K and Ca, are also added (Ca through the bution of magmatic water. Geothermal systems may also be
breakdown of the anorthite component in plagioclase), typi- dominated by seawater, as is the case for the Reykjanes field in
cally reaching concentrations two to three times their seawater Iceland where the chlorinity is very similar to that of seawater
concentration of  0.01 m. Silicon, which is a minor constitu- (Arnórsson, 1978). The other major components of geothermal
ent of seawater, becomes a major element with a concentration fluids, for example, Na, K, Ca, and CO2, also vary considerably in
of  0.02 m. Iron, which is present in seawater at very low concentration; the seawater-dominated, basalt-hosted Reykjanes
concentrations (<1 nm), may also become a major element fluids are compositionally very similar to seafloor vent fluids
in the hydrothermal fluid; in some cases, its concentration may (Table 1). Until relatively recently, it has not been possible to
exceed 0.1 m. The concentrations of the main ore elements directly measure the metal concentrations of preboiled geo-
(Cu, Zn, and Pb) are highly variable, generally ranging from thermal fluids; however, such data are now starting to become
a few tens to a few thousand ppb, but are thousands to tens of available through the development of ‘downhole’ devices that
thousands of times higher than in seawater; gold concentra- can sample the reservoir fluids directly (e.g., Simmons and
tions have not been measured (Table 1). Brown, 2006). For example, in the case of the geothermal fields
Although most of the information on the composition of of the Taupo volcanic zone, these data show that the reservoir
seafloor hydrothermal fluids comes from systems at mid-ocean fluid contains up to 23 ppb Au, 2400 ppb Ag, and 4850 ppb As;
ridges, data on the compositions of hydrothermal fluids in arcs the H2S content ranges up to 0.007 m (Ellis, 1979a,b; Seward,
and back-arc basins are being increasingly reported (Table 1). 1989). Lower concentrations of these metals were measured
The back-arc fluids are very similar in composition to those of in the Reykjanes reservoir fluid using the same method, that
mid-ocean ridges and are also mainly products of seawater– is, up to 6 ppb Au, 34 ppb Ag, and <1 ppb As; concentrations
basalt interaction (Reeves et al., 2011). However, in the arc and for Cu, Zn, and Pb were 16, 26, and <1 ppb, respectively
near-arc environment, where the magmas are much more si- (Hardardottir et al., 2009). By comparison, concentrations of
licic, their compositions differ considerably from those of hy- Cu, Zn, and Pb reported for the very high-salinity Salton Sea
drothermal fluids at mid-ocean ridges. The main differences geothermal fluid, although sampled at the wellhead, are up to 7,
are significantly lower pH (2–3 vs. 3–4), very much higher CO2 518, and 107 ppm, respectively; the H2S concentration is
Table 1 Selected analyses of the composition of fluids in active hydrothermal systems

Seafloor vent fluids Geothermal fluids Oil-field brines

Seawater Spreading Back-arc basin Arc Rotokawa Salton Sea Cerro Prieto Reykjanes Ladolam gold Mississippi
center deposit

Temp ( C) 2 350 282 268 320 330 337 296 275 107
pH (25  C) 7.8 3.400 4.1 2.6 6.03 5.1 8 – 8.01 5.650
Na (mm) 469.003 436.334 540.637 449.600 12.139 2382.609 419.980 393.000 1154.697 76.257
K (mm) 9.804 23.021 24.023 86.290 2.046 453.846 77.064 38.000 125.641 17.686
Ca (mm) 10.204 16.010 82.270 15.009 0.012 712.500 10.059 47.000 0.227 774.343
Mg (mm) 52.768 – 0.000 0.000 0.000 2.042 0.002 0.390 0.004
Mn (mm) – 0.960 0.348 3.117 – 27.273 0.010 52 000.000 – 1.163
Fe (mm) – 1.664 0.109 2.404 – 30.536 0.004 0.430 – 6.198
Si (mm) 0.200 18.010 15.007 16.008 23.409 9.800 48.169 10.000 19.238 0.712
Ba (mm) 0.140 8.000 17.000 91.001 – 2576.642 – 71.000 – 1056.004
Cu (ppm) – 2.224 0.127 2.288 – 6.800 0.005 16.586 4.450 0.020
Zn (ppm) – 6.931 0.654 7.520 – 507.000 0.006 24.987 0.185 222.000
Pb (ppm) – 63.818 0.829 1450.400 – 102.000 0.005 0.001 0.023 53.200
Mo (ppm) – – – – – – – 0.015 0.043
Sn (ppm) – – – – – – – 0.001 0.840
Cd (ppm) – 0.180 – – – 2.300 – 0.135 – 0.830
Ag (ppb) – 0.037 – – 1100.000 1400.000 4.000 34.626 6.000
Au (ppb) – – – – 7.800 – 4.000 6.106 16.000
As (ppm) – – – – 5.400 – 2.000 0.112 17.000
Sb (ppm) – – – – 1.200 – 0.400 0.025 0.004
Cl (mm) 551.626 497.669 730.500 583.889 14.620 4500.000 526.019 524.000 613.923 5614.690
Br (mm) 0.808 0.855 1.030 1.000 – 1.388 – 0.780 0.451 13.029
F (mm) 0.064 – 0.023 0.116 0.300 – – – – 0.042
B (mm) 0.426 0.548 0.240 1.620 9.000 24.636 2.220 0.709 12.213 8.049
SO4 (mm) 22.366 0.400 1.880 0.420 0.041 0.552 – – 410.375 0.187
H2S (mm) 0.000 7.303 1.800 6.803 7.310 0.294 22.617 0.900 –
HCO3 (mm) 2.931 7.266 7.649 40.854 1.072 35.909 1.072 – 48.261
References Seawater EPR 21 N Vienna Woods PACMANUS Rotokawa Salton Sea Cerro Prieto Reykjanes Ladolam gold Oil-field brines
(Hannington (Hannington (Hannington basin (Weissberg (Williams (Weissberg (Hardardottir deposit (Kharaka et al.,
et al., 2005) et al., 2005) et al., 2005) (Hannington et al., 1979; and et al., 1979; et al., 2009) (Simmons 1987)
et al., 2005) Krupp and McKibben, Clark and and Brown,
Seward, 1989) Williams- 2007)
1990; Jones, 1990)
Simmons
and Brown,
2007)
Table 2 Compositions of selected fluid inclusions from a variety of ore deposit types

Deposit type Sn granite Sn granite Porphyry Cu Porphyry Cu Porphyry Cu Porphyry Cu Zn skarn REE granite MVT MVT MVT

Fluid type Brine Vapor Brine Brine Brine Vapor Brine Brine Brine Brine Brine
Salinity 35.20 4.00 31.20 33.10 49.90 9.50 72.50 19.30 23.70 19.70
Temp ( C) 453 650 720 471 395 372 595 112 114 110
Na (ppm) 74 000 14 000 75 000 67 900 80 000 30 000 95 461 222 600 60 000 74 000 67 000
K (ppm) 57 000 3300 61 300 58 300 99 000 16 000 41 530 92 600 <1040 1400 7200
Ca (ppm) – – – – – – – – 23 000 – 14 000
Mg (ppm) – – – – – – – – 2110 – 1400
Mn (ppm) 22 000 1600 n.a. 14 200 23 000 1500 – 17 900 <32.7 – –
Fe (ppm) 73 000 4100 54 700 74 700 130 000 29 000 – 43 700 – –
Cu (ppm) 2300 4600 2200 2300 5500 30 000 – 230 50 6 –
Zn (ppm) 3600 680 n.a. 3400 9900 6500 5930 2500 n.d. 13 –
Pb (ppm) 3400 190 780 800 2400 620 4350 470 240 – 180
Mo (ppm) – – n.a. <120 90 n.a. – 81 – – –
Sn (ppm) 390 – n.a. <350 n.a. n.a. – 63 – – –
W (ppm) 56 – 57 <67 <80 <50 – 30 – – –
Ag (ppm) 290 5 7 <67 n.a. <70 50 3 – – –
As (ppm) 120 34 8 <400 50 n.a. 248 25 – – –
Sb (ppm) 110 31 – – – – 365 – – –
Ce (ppm) 2 – 6 15 210 <2 – 300 – – –
Bi (ppm) 10 3 n.a. n.a. – – – 5 – – –
Cl (ppm) – – – – – – – – 117 000 140 000 120 000
References Mole granite Mole granite Santa Rita Santa Rita Bajo de la Bajo de la El Mochito Capitan REE Ozark MVT Pb– North Tri-State MVT
(Audétat (Audétat porphyry Cu porphyry Cu Alumbrera Alumbrera Zn–Ag granite Zn deposits Arkansas Pb–Zn
et al., et al., 2000) deposit deposit porphyry Cu–Au porphyry skarn (Banks (Wilkinson MVT Pb–Zn deposits
2000) (Audétat (Audétat deposit (Ulrich Cu–Au deposit (Samson et al., et al., 2009) deposits (Stoffell
et al., 2008) et al., 2008) et al., 2002) (Ulrich et al., et al., 1994) (Stoffell et al., 2004)
2002) 2008) et al., 2004)
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 33

0.0005 m (Williams and McKibben, 1989). The giant appreciable metallic flux on the order of thousands of tonnes
Ladolam gold deposit (a low-sulfidation epithermal deposit) is per annum to the atmosphere (Calabrese et al., 2011).
an actively forming ore deposit on Lihir Island, Papua New Fluid inclusions provide the most direct source of data on
Guinea, which contains 1300 t of gold. The fluids forming the compositions of ore fluids in extinct hydrothermal systems,
this deposit contain up to 16 ppb Au and 4450 ppb Cu; the and analyses of individual fluid inclusions (e.g., using LA-
chloride content is 0.6 m, the CO2 content 0.2 m, and the H2S ICPMS) are the most reliable sources for these data (Table 2).
content 0.0006 m (Simmons and Brown, 2006). With a few exceptions, studies reporting compositions of indi-
The ligands that could be important for metal transport in vidual inclusions have focused on magmatic hydrothermal ore
geothermal fluids vary considerably with the environment in deposits, and data are available for both the single-phase fluids
which the geothermal system is located. In systems like the that exsolved from the magma and the later brine and vapor
Salton Sea, Cerro Prieto, Reykjanes, and Lihir Island, chloride is that were the products of phase separation (Klemm et al., 2007;
clearly an important complexing ligand for some metals such as Samson et al., 2008). These data show that supercritical mag-
Pb. However, in systems like those of the Taupo volcanic zone, matic fluid inclusions typically contain 0.1–0.8 m Cl, 0.1–
the chloride concentrations are typically low (e.g., 0.02–0.06 m), 0.6 m Na, up to 0.3 m K, up to 0.1 m Fe, on average
although some of these systems have quite elevated reduced 3500 ppm Cu, 200 ppm Zn, 40 ppm Pb, and 10 ppm Mo.
sulfur concentrations, such as at Ohaaki (0.002 m) and Roto- Gold concentrations are typically at high ppb levels. The com-
kawa (0.007 m) (Krupp and Seward, 1987, 1990; Seward, 1989). positions of magmatic vapor inclusions are similar, except that
Other species reported to be present in active geothermal systems Cu concentrations can be very much higher, over 1 wt%, and
that might be important for metal transport are ammonia and gold concentrations up to 10 ppm have been reported (Ulrich
sulfate, the latter reaching 0.02 m at Salton Sea and >0.3 m at et al., 1999). Magmatic brine inclusions have much higher
Lihir Island (Simmons and Brown, 2006; Williams and Mckib- concentrations of salts, commonly >6 m Cl, >3 m Na,
ben, 1989). Ammonia (i.e., NH3 þ NH4þ) occurs ubiquitously in >1.5 m K, >1.5 m Fe, and on average 1500 ppm Cu,
geothermal fluids at concentrations up to and/or >0.002 m (see 3500 ppm Zn, 2000 ppm Pb, and 80 ppm Mo, and up to
Ellis, 1979a,b) and forms as an expected consequence of the 3000 ppb Au. Some compositional data have been reported by
redox equilibria involving N2 and H2 (Seward, 1974a,b). Stoffell et al. (2004) for fluid inclusions from MVT deposits.
Oil-field brines, as noted earlier, have long been considered These show that the fluids contain an average of 3.4 m Cl, 2.6 m
to be modern analogues of the fluids interpreted to form MVT Na, 0.2 m K, 0.6 m Ca, 212 ppm Zn, and 94 ppm Pb.
lead–zinc deposits (Hall and Friedman, 1963). Although the
origin of these basinal brines is still a matter of conjecture, for
example, connate or evaporated seawater, variably modified by 13.2.2 Hydrothermal Ore Solution Chemistry –
mixing with meteoric water and diagenetic reaction (Wilson The Main Dissolved Components
and Long, 1993), they are able to transport significant concen-
13.2.2.1 Water Solvent at Hydrothermal Conditions
trations of metals at relatively low temperature (85–150  C).
Their chlorinity ranges from about 2 to nearly 6 m, and Naþ is The review of the chemistry of ore-forming fluids has shown
the dominant cation (up to 3.5 m), followed by Ca2þ (up to clearly that they are multicomponent aqueous electrolyte solu-
2 m; Table 1). The sulfate content ranges up to 0.004 m and that tions containing both simple ionic and more complex molec-
of NH4þ to 0.01 m, but H2S and inorganic carbon concentra- ular species. In this section, the properties of water solvent that
tions are very low, <1  107 and <5  104 m, respectively enable it to be an effective agent for the transport of ore metals
(Carpenter et al., 1974; Kharaka et al., 1987). In addition, oil- and associated elements are considered.
field brines contain significant proportions of dissolved organic At ambient temperature and pressure, water is the archetype
species, such as acetate (Table 1). The principal ore metals are protic solvent that exhibits various unusual properties that may
Zn and Pb, which can reach concentrations of up to 350 ppm be attributed to the large dipole moment of water and the
and 100 ppm, respectively (Carpenter et al., 1974; Kharaka associated hydrogen bonding among neighboring molecules.
et al., 1987). Copper concentrations are typically <200 ppb. Dissolved ionic and molecular species are ‘hydrated’ and inter-
On the other hand, gold concentrations as high as 18 ppb act with water solvent (Figure 1). As temperature and pressure
have been reported (Saunders and Swann, 1990). increase at conditions defined by the two-phase curve (i.e.,
Volcanic gases are the surface representatives of magmatic liquid–vapor equilibrium), liquid water expands as the posi-
hydrothermal fluids, and although they are considerably less tional and orientational constraints alter and the extent of
dense than the vapors that are believed to be ore fluids in some hydrogen bonding diminishes. The increasing molar volume
settings, they do provide information about the compositions (decreasing density) of water liquid as one proceeds to the
of these fluids. In addition to water vapor, volcanic gases can critical point is thus accompanied by concomitant decreases
contain >30 mol% CO2, >20 mol% SO2, >5 mol% H2S, in its dielectric permittivity and other properties, such as vis-
>2mol% HCl, and >0.2 mol% HF; the average contents of cosity. The decrease in the dielectric constant from 80.10 to
these gases for 19 representative volcanoes are 10.6, 5.2, 1.1, 7.22 for liquid water as temperature increases from 20 to
0.8, and 0.1 mol%, respectively (Halmer, 2002). Metal con- 373  C (at the equilibrium vapor pressure) (Fernandez et al.,
tents of volcanic gases range up to 6 ppm Cu, 12 ppm Pb, 1997; Uematsu and Franck, 1980) gives some premonition of
11 ppm Zn, 7 ppm Sn, 250 ppb Ag, and 24 ppb Au (e.g., the changing nature of water dipole–ion interaction, which
Symonds et al., 1987; Wahrenberger et al., 2002; Williams- results in enhanced ion pairing/association and metal complex
Jones et al., 2002), and the total metal content of discharging formation. Increasing pressure acts in an opposing way, caus-
volcanic gases from a single volcano may constitute an ing an increase in the static permittivity (Fernandez et al.,
34 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

ordering diminishes, suggesting that water might be con-


sidered as a hydrogen-bonded continuum. But recent X-ray
Raman scattering data (Wernet et al., 2005) and NDIS mea-
49.9 ⬚ surements and simulations (Bernabei et al., 2008) of supercrit-
ical water indicate that such a view is too simplistic. The X-ray
Raman spectra are consistent with a structural model for su-
Å percritical water comprising small domains of water molecules
2.952 in various tetrahedral configurations in a gas-phase-like mo-
48.5 ⬚
lecular continuum. Extensive hydrogen bonding persists with
increasing temperature and pressure, as demonstrated by com-
putational/theoretical and spectroscopic (Raman and NMR)
studies (e.g., Chialvo and Cummings, 1994; Hoffmann and
-
Conradi, 1997; Kalinichev et al., 1999; Walrafen et al., 1996).
For example, the NMR data of Hoffmann and Conradi (1997)
indicate that at 400  C and 400 bar, 29% of the hydrogen
bonding still exists, compared to ambient conditions. Never-
theless, the detailed interpretation of such results on a molec-
ular level is awaited. The current perspective on the structure
of liquid water from X-ray scattering, spectroscopic and dif-
fraction data, as well as molecular dynamics (MD) and ab
initio approaches, is elegantly summarized by Nilsson and
+ Petterssen (2011).
Self-ionization is an important property of water and aque-
ous solutions, which involves the loss of a proton from the
water molecule, leaving behind a hydroxide ion (OH). Spon-
taneous reaction of the hydrogen nucleus with another water
molecule produces the hydronium ion, H3Oþ. The value of
pKw (i.e.,  log Kw of the equilibrium ion product constant)
depends on temperature, pressure, and, in the case of an aque-
Figure 1 Structure of water in the vicinity of a metal cation. Water
ous solution, ionic strength. It decreases with increasing tem-
molecules immediately adjacent to the ion are arranged so that their
perature to a minimum at between 250 and 350  C
negatively charged oxygen atoms are directed inward to form an inner
hydration shell; these molecules are surrounded by an outer shell of (depending on pressure), above which it increases. For exam-
water molecules that are similarly directed but with less consistency and ple, the pKw at 250  C and saturated water vapor pressure
locally share hydrogen bonds with the bulk water solvent. The (swvp) is 11, whereas it is 14 at 25  C and swvp. The equilib-
development of hydration shells facilitates dissolution by separating rium pH therefore drops from 7 at 25  C to 5.5 at 250  C at
cations from anions. swvp. Increasing pressure decreases the pKw (due to the higher
stability of ionized species) and shifts the temperature mini-
1997), thus encouraging the stability of some ionic species mum to higher temperature (Marshall and Franck, 1981).
under supercritical conditions. The self-ionization of liquid Addition of salts to the water may either increase or decrease
water as manifested by the ion product constant, Kw, changes the ion product, depending on the nature and concentration of
as a function of temperature and pressure as the energetics of the salt present in the solution (Kron et al., 1995). For example,
deprotonation of water molecules changes in response to re- measurements of the ion product constant of water in solu-
lated changes in the hydrogen bonding array. Steam is domi- tions of NaCl, KCl, KNO3, and NaNO3 at 25  C and at swvp
nated by the presence of water clusters that increasingly demonstrate that addition of these electrolytes progressively
interact with each other via hydrogen bonding as the density decreases Kw of water to a minimum value at an ionic strength
of the vapor phase increases to liquid-like densities as the characteristic for the particular electrolyte; further addition of
critical point is approached. As expected, therefore, the viscos- these electrolytes increases the ion product, potentially to
ity of steam increases with increasing temperature at liquid– values well above 14 (Figure 2). Although the pKw minima
vapor equilibrium. for the salts in question are at concentrations <1 m, those for
The molecular structure and nature of the hydrogen bond- other salts may occur at significantly higher concentrations
ing of liquid water and supercritical water at liquid-like densi- (Kron et al., 1995). As most interactions of water with rock
ties continues to be an important and ongoing research area of (i.e., with silicate minerals) occur through pH-dependent re-
fundamental importance to hydrothermal geochemistry and actions and concentrations of many aqueous species also de-
ore solution chemistry. It has been generally considered that pend on pH, pKw is one of the major parameters controlling
liquid water exhibits a short-range tetrahedral ordering with speciation in hydrothermal solutions. For example, the species
respect to the nearest neighbor water molecules, apparently AlOH2þ forms via the reaction Al3þ þ OH ⇆ AlOH2þ, and its
inheriting some remnant configurational aspects from ice. stability therefore depends on the concentration of the hy-
Above about 150  C, the ‘evidence’ from X-ray diffraction droxyl ion and, in turn, the solvent self-ionization.
and neutron diffraction isotopic dilution (NDIS) (Gorbaty As discussed earlier, the ability of water to stabilize ionized
and Kalinichev, 1995; Soper, 2000) for the nearest neighbor (charged) species is controlled by the development of
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 35

14.6 pressures (e.g., P  500 bar), sodium chloride solutions be-


come progressively more associated (to form simple NaCl0
ion pairs), as was elegantly demonstrated by the earlier con-
ductance study of Quist and Marshall (1968). For example, a
14.4
1.0 m NaCl solution at 300  C and at the equilibrium vapor
pressure is about 50% associated. They also showed that in-
creasing pressure up to 4000 bar has the opposite effect and
14.2 causes the disproportionation of associated species, thus favor-
ing the formation of the simple hydrated ions, Naþ and Cl,
pKw

whose partial molar volumes are electrostrictively enhanced


14.0 relative to the associated moieties.
A number of recent MD and ab initio/MD studies have also
shed light on the molecular nature of sodium chloride solu-
13.8 tions at elevated temperatures and pressures (e.g., Dong et al.,
2008; Driesner et al., 1998; Guardia et al., 2006a,b; Sherman
and Collings, 2002; Yui et al., 2010). Driesner et al. (1998)
showed that for a 1 m NaCl solution at 380  C and solution
13.6 density of 0.55 g cm3, the simple hydrated ions (i.e., Naþ and
0 1 2 3 4 5 6 7
Cl) comprise only about 13% of the species in solution, with
NaCl (mol kg-1)
the remainder of the solute consisting of associated species such
Figure 2 The effect of salinity on the ionization constant of water (pKw) as NaCl0, Na2Clþ, NaCl2, and Na2Cl2 but with the dominant
at 25  C; the experimental points are shown by the open circles and associated species being the simple NaCl0 ion pair. Some
calculated values by the solid line. Modified from Kron I, Marshall SL, short-lived, larger associated species, such as Na3Cl4, were
May PM, Hefter G, and Konigsberger E (1995) The ionic product of water also observed in the simulations but have lifetimes of 2 ps.
in highly concentrated aqueous electrolyte solutions. Monatshefte für In summary then, the main electrolyte salt component of
Chemie 126: 819–837. With permission from Springer.
hydrothermal fluids in the Earth’s crust is, with few exceptions,
NaCl. The bulk electrolyte salt association chemistry at high
hydrogen-bonded networks of water molecules. The expansion temperatures and pressures plays a fundamental role in the
associated with increasing temperature distorts these networks complexing and transport/deposition chemistry of many tran-
and leads to a decrease in the relative permittivity and associ- sition metals in hydrothermal ore-forming systems. Thus, in a
ated destabilization of ionized species. In the extreme, water cooling hydrothermal solution ascending buoyantly through
can expand to a fluid with gas-like density or separate a vapor, the Earth’s crust at lower pressures (e.g., P  500 bar), more
and in this state, there is complete destruction of the network. chloride ligands become available for metal complexing as
Water molecules cease to be part of a continuum but instead associated sodium chloride solute species disproportionate
form isolated clusters with small numbers of molecules with decreasing temperature. Conversely, a hydrothermal ore
(Kalinichev and Churakov, 1999). Consequently, the ability solution decompressing isothermally will lead to instability of
of gas-like aqueous fluids to stabilize ionized species is lower some metal complexes as chloride ligands are consumed by
(see the succeeding text), and metal solubility in these fluids is association of the bulk electrolyte.
predominantly, but not exclusively, molecular (i.e., dissolved
species have no charge). Nevertheless, these species commonly
disobey the rules for simple gaseous mixtures. In particular, 13.2.2.3 Ion Hydration, Association, and Water Activity
hydration still occurs and neutrally charged species may be Of particular interest, as well, is the changing nature of ion
incorporated into water clusters. Moreover, there is evidence hydration in hydrothermal fluids as a function of temperature
that the clusters can even support charged species (e.g., and pressure. Insight into ion–solvent interaction at elevated
Likholyot et al., 2005, 2007). temperatures and pressures comprising ion hydration coordi-
nation numbers, ion–water bond lengths, and solvated water
residence times has come relatively recently from spectro-
13.2.2.2 NaCl – The Main Dissolved Electrolyte Component
scopic, neutron, and X-ray diffraction and computational stud-
Hydrothermal ore fluids are multicomponent electrolyte solu- ies, many of which are summarized and discussed by Seward
tions in which NaCl is generally the dominant salt component. and Driesner (2004). More recently, computational studies
Of course, natural hydrothermal systems are more complex (ab initio, MD, and ab initio/MD) have added significantly
with numerous interrelated heterogeneous equilibria deter- to our knowledge of ion hydration under hydrothermal con-
mining the hydrolytic (pH) and redox state of the fluids that ditions, especially with respect to NaCl solutions (e.g.,
undergo phase separation as they migrate and ascend buoy- Bondarenko et al., 2006; Chialvo et al., 2009; Dong et al.,
antly though the Earth’s crust. Sodium chloride is the classic 2008; Fedotova, 2008; Guardia et al., 2006a,b; Nahtigal and
strong electrolyte whose aqueous solutions are considered to Svishchev, 2009). Driesner et al. (1998) studied (using MD)
be extensively ionized into hydrated Naþ and Cl ions at Naþ ion hydration for both cases of infinite dilution and 1.0 m
ambient temperature and pressure, although concentrated so- from 30 to 590 K and at several densities. They showed that
lutions (saturation ¼ 6.013 m at 20  C) may exhibit some as- there is a contraction (i.e., 2.26 to 2.23 Å) of first-shell waters
sociation. However, with increasing temperature and at lower around Naþ, which is also accompanied by a decrease in the
36 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

number of coordinated water molecules with increasing tem- are involved in various, weak acid hydrolytic equilibria and
perature. Moreover, the residence times of first-shell waters play a role in metal transport by fluids in the Earth’s crust. CO2
around Naþ and Cl (in a 1.0 m solution) at 298 K are found may occur in high concentrations greater than 20 mol%
to be 22.5 and 12.0 ps, respectively, but decrease by about an in H2O þ CO2 þ NaCl fluids associated with metamorphic
order of magnitude to 2.2 and 2.4 ps with increasing temper- reactions; however, in the majority of hydrothermal ore fluids
ature to 650 K (Guardia et al., 2006b). in the upper crust, the CO2 (as H2CO3 þ HCO3 þCO32)
In addition, a number of systematic X-ray absorption spec- would be in the range from 0.01 to 1.0 m. CO2 being a volatile
troscopic studies (i.e., EXAFS (extended X-ray absorption fine component means that carbonate equilibria and pH are sensi-
structure)) of Agþ and Sr2þ hydration up to 350  C and at tive to phase separation and boiling in hydrothermal systems.
equilibrium saturated vapor pressures (Seward et al., 1996, In active geothermal/hydrothermal systems, it is well known
1999) have also demonstrated a contraction of the cation– that CO2 partitioning into the vapor (steam) phase during
oxygen (water) distances with increasing temperature as well boiling leads to an increase in the residual liquid-phase pH
as a decrease in the number of the first-shell water molecules. In by as much as two pH units (e.g., from, say, 6 to 8). This affects
the case of Agþ, for example, the Agþ–oxygen (water) distance the stability of metal carbonate complexes as well as the stabil-
undergoes an appreciable contraction of 0.1 Å as temperature ity of any other pH-dependent metal complex equilibria, such
increases from 25 to 350  C, with an accompanying decrease in as hydroxide and/or hydrosulfide complexes.
the number of first-shell waters from 4 to 3. However, data for Of particular importance is H2S and its first deprotonation
Cd2þ hydration up to 300  C and at equilibrium vapor pres- equilibrium in aqueous media,
sures indicate a decrease in the Cd2þ–oxygen (water) distance
with increasing temperature but with no change in the number H2 S ¼ Hþ þ HS
of first-shell waters (Seward et al., 2013). In the case of the the equilibrium constant, K1, for which has been accurately
trivalent cation, In3þ, the hydration environment remains determined up to 360  C at the equilibrium vapor pressure
unchanged (six first-shell waters bound to the In3þ at a distance (Suleimenov and Seward, 1997) as well as at 500 bar and up
of 2.14 Å) over the same temperature range (i.e., 25 to 300  C) to 500  C (Stefansson and Seward, 2004). Note that the equi-
at saturated vapor pressures (Seward et al., 2000). librium constant, K2, for the second ionization of H2S,
The nature of ion hydration (i.e., ion–solvent interaction) is
complex and varies with the changing dielectric properties of HS ¼ Hþ þ S2
water solvent, which depend fundamentally on the nature is small (i.e., pK2 > 17.1 at 24  C and >12.5 at 200 to 270  C;
and extent of hydrogen bonding, which is a function of temper- Giggenbach, 1971a), and hence, the sulfide ion, S2, is present
ature and pressure. However, NaCl is the most concentrated at negligible concentrations in natural hydrothermal solutions.
salt component in hydrothermal fluids in the Earth’s crust, and However, the hydrosulfide ligand, HS, plays a fundamental
its hydration behavior is therefore of importance in determining role in the transport and deposition chemistry of some
the water activity in hydrothermal ore solutions. Changes in elements, such as gold in hydrothermal ore solutions, and
the hydration of Naþ and Cl ions as well as the formation the availability of reduced sulfur is crucial for the precipitation
of associated species, such as NaCl0 (weakly hydrated), can of sulfide ore minerals. During boiling and/or phase separa-
fundamentally affect water activity and hence the solubility of tion, H2S and volatile components, such as CO2, NH3, and H2,
minerals. Thus, a hydrothermal ore solution migrating through partition into the vapor (steam) or less dense, volatile-rich
the Earth’s crust will experience a range of physical and chemical phase, which affects the stability of various metal complexes
phenomena as it cools and decompresses (e.g., Heinrich with respect to ligand availability (e.g., H2S loss) and increas-
et al., 2004), which include element partitioning, mineral ing redox potential of the residual liquid (H2 loss) or more
precipitation, and changing metal complex stabilities, which dense supercritical fluid.
may, in some cases, be temporarily enhanced as a result of the The volatile species NH3 and N2 and H2 are involved in
myriad of changes occurring in a complex multicomponent heterogeneous redox equilibria and ammonium ion stability
system. As shown, for example, by the solubility of a simple through the reactions
gangue mineral such as quartz in sodium chloride and other
electrolyte salt solutions (Newton and Manning, 2000, 2010; N2ðgÞ þ H2ðgÞ ¼ NH3ðgÞ
Shmulovich et al., 2006), the dissolved silica concentration var-
and
ies not only with temperature and pressure but also with the salt
concentration and hence the extent and molecular nature of
NH3ðaqÞ þ Hþ ¼ NH4 þ
association (ion pairing) and ion–solvent interaction with con-
sequent effects on water solvent activity. The changing value of The stability of simple amino complexes (involving the NH3
pK1 of silicic acid as a function of temperature and pressure as ligand) of transition metals under hydrothermal conditions is
well as bulk salt hydrolysis may also be of importance. still poorly known. The ammonium ion also participates in
wall rock alteration reactions in hydrothermal systems, giving
rise to the formation of ammonium-containing silicate min-
13.2.2.4 Weak Acid/Base Equilibria in Hydrothermal
erals such as buddingtonite and tobelite.
Systems
Thus, it is against this background of changing water prop-
In addition to the main electrolyte salt, NaCl, there are other erties (e.g., hydrogen bonding) with associated changes in ion–
important solute components in hydrothermal ore solutions solvent interaction and ion pairing/association and changes in
in the Earth’s crust such as CO2, H2S, and NH3, which water activity in the presence of ligand-supplying weak acid/
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 37

base systems that metal complex equilibria, metal transport


mechanisms, and ore mineral precipitation chemistry at ele- 20 MPa
vated temperatures and pressures relevant to hydrothermal
ore-depositing systems must be considered. 30 High-salinity
liquid

wt% NaCl

Immiscibility
20
13.2.3 Mineral Solubility in Water and Salt 10
Solutions at High Temperature and Pressure 3.2
0

Quartz solubility (mmol kg-1)


10
The main interests of geochemists investigating hydrothermal
Single-phase fluid Low-salinity
processes, including those of ore formation, are the mecha-
(a) vapor
nisms that control the transport of elements in aqueous solu-
tions and the saturation of the solutions with minerals
50
containing them. At any given set of physicochemical condi- 50 MPa
tions, the concentration of an element in an aqueous solution High-salinity
is limited by the solubility of the least soluble mineral contain- liquid
40
ing this element. For example, in the system Na–Ca–F–H2O,

Immiscibility
the concentration of fluorine is determined by the solubility of
fluorite (CaF2). Consequently, the mineral villiaumite (NaF), wt% NaCl 10
30
which is much more soluble than fluorite, will not form unless

Retrograde
solubility
the buffering capacity of fluorite is exhausted, that is, the con- 3.2
centration of Ca is so low that the solubility product for the ions 20
0
comprising fluorite is less than that for villiaumite. In this
section, a brief review is provided of the data available on the
solubility of the major gangue (non-ore) minerals commonly 10
encountered in hydrothermal ore deposits. Single-phase fluid Low-salinity
vapor
Hydrothermal gangue minerals can be subdivided into two (b)
groups, namely, those that display prograde solubility, that is, 0
their solubility increases with increasing temperature, and 0 100 200 300 400 500
those that display retrograde solubility, that is, decrease in Temperature (⬚C)
solubility with increasing temperature. Most chloride, oxide, Figure 3 Isobaric quartz solubility as a function of temperature in pure
sulfide, and silicate minerals exhibit prograde solubility, H2O, H2O þ 3.2 wt% NaCl, and H2O þ 10 wt% NaCl fluids, (a) at 20 MPa
whereas sulfates, carbonates, and phosphates commonly (bottom) and (b) at 50 MPa; the shaded areas represent regions of
show retrograde solubility. These minerals can be further clas- retrograde solubility. Modified from Steele-MacInnis M, Han L,
sified according to whether their dissolution is congruent or Lowell RP, Rimstidt JD, and Bodnar RJ (2012) The role of fluid phase
incongruent. In the case of the former, all products of the immiscibility in quartz dissolution and precipitation in sub-seafloor
dissolution reaction are transferred stoichiometrically to the hydrothermal systems. Earth and Planetary Science Letters
solution. By contrast, incongruent dissolution involves forma- 321/322: 139–151.
tion of both aqueous species and new solids. Simple com-
pounds such as chlorides and oxides generally display
The solubility of quartz in pure water is controlled by a
congruent dissolution, whereas complex silicates and sulfides
number of hydrated ‘SiO2’ species of which silicic acid,
typically undergo incongruent dissolution.
H4SiO40 (Si(OH)40), and its corresponding anions, H3SiO4
Quartz is without question the gangue mineral most com-
and H2SiO42, are the most important, although H2SiO42 is a
monly encountered in ore-forming hydrothermal systems and
negligible species in hydrothermal fluids in the Earth’s crust.
has been the subject of large numbers of studies by both
Neutrally charged species predominate at pH< 9, and therefore,
experimentalists and theoreticians. In pure water, the solubility
the solubility of quartz is independent of pH, except at condi-
of quartz increases systematically with temperature up to
tions of extreme alkalinity. At these conditions, the formation of
370  C, but at higher temperature, its solubility is retrograde;
ionized species of silicic acid causes quartz solubility to increase
quartz solubility increases with pressure at all temperatures.
with increasing pH and, in salt solutions, this may be further
The presence of alkali salts, such as NaCl and KCl, reduces the
enhanced by the formation of ion pairs such as NaH3SiO40
solubility of quartz (salting out) at low temperature but in-
(Seward, 1974a,b). Aqueous silica is also known to form poly-
creases it at temperatures above 250  C (Figure 3). However,
mers (Chan, 1989; Iler, 1979), as illustrated by the following
an understanding at a molecular level of how the presence of
stepwise reactions (e.g., Newton and Manning, 2010):
an electrolyte salt, such as NaCl, affects quartz solubility over a
wide range of temperature and pressure, as illustrated by the SiO2 þ 4H2 O ¼ SiðOHÞ4ðaqÞ
experiments of Newton and Manning (2000), eludes us at quartz
silica monomer
present. It will necessarily involve a complex interplay of sol-
and
vent interaction with both ionic and molecular species as well
as the effect of temperature and pressure on ion pairing/asso- 2SiðOH4 ÞðaqÞ ¼ Si2 OðOHÞ6ðaqÞ þ H2 O
ciation, salt hydrolysis, and the ionization of silicic acid. silica monomer silica dimer
38 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

An understanding of the nature and kinetics of silica poly- Sherman and Barak, 2000). These studies showed that the
merization as a function of temperature, pressure, pH, and salt CaCO3 and MgCO3 components of dolomite dissolve at con-
concentration (ionic strength) is of considerable importance in siderably different rates. Consequently, evaluations of the
the exploitation of active geothermal systems for energy use solubility of dolomite in aqueous solutions are now based on
(Gunnarsson and Arnorsson, 2005; Rothbaum and Rhode, calorimetric determinations of its enthalpy, entropy, and heat
1979), particularly in managing the reservoir environment capacity (e.g., Hemingway and Robie, 1994; Navrotsky and
when hot, ‘flashed,’ fluids (oversaturated with respect to amor- Capobianco, 1987).
phous silica) are reinjected. However, the chemistry of silica Sulfate minerals, such as anhydrite (CaSO4) and its hy-
polymerization, especially in the presence of aluminum and drated form, gypsum (CaSO4 2H2O), and, to a lesser extent,
salts (e.g., NaCl and KCl), under more extreme hydrothermal barite (BaSO4), are also important in some types of hydrother-
conditions is still poorly known. mal ore deposits, notably porphyry copper and molybdenum
Another very common gangue mineral is calcite. Studies deposits (anhydrite) and VMS deposits (anhydrite and barite).
of its solubility in hydrothermal fluids date back to the The solubility of gypsum, which is significantly lower than that
publication of Ellis (1963) and have continued to the present of calcite, increases to a maximum of 0.015 m at 40  C in
(e.g., Duan and Li, 2008a). In all cases and for most crustal pure water and then decreases with further increases in tem-
conditions, these studies have shown that calcite exhibits ret- perature (Booth and Bidwell, 1950). At 97  C, gypsum dehy-
rograde solubility, that is, its solubility decreases in pure water drates to anhydrite, generally via formation of an intermediate
with increasing temperature. However, its solubility increases hemihydrate phase, CaSO4  0.5H2O; anhydrite displays ret-
with increasing pressure. Calcite solubility also increases with rograde solubility. As is the case for calcite, the solubility of
increasing NaCl concentration up to a molality of  1–1.5 gypsum and anhydrite increases with increasing pressure. Sim-
(Figure 4); however, it decreases at higher NaCl concentration ilarly, their solubility increases with addition of NaCl up to a
(Duan and Li, 2008; Newton and Manning, 2002). Neverthe- concentration of  1–2 m (Figure 4) and decreases at higher
less, at deep crustal conditions (i.e., at very high temperature NaCl concentration (Blount and Dickson, 1973; Raju and
and pressure), the solubility of calcite is prograde, even at Atkinson, 1990). Anhydrite also displays pronounced pro-
relatively high salinity (Newton and Manning, 2002). The grade solubility in saline fluids at high temperature and pres-
aqueous species controlling the solubility of calcite are the sure. The main aqueous species controlling the solubility of
aquated ion and the ion pairs, Ca2þ, CaHCO3þ, and CaCO30 gypsum and anhydrite are Ca2þ, CaSO40, CaHSO4þ, CaClþ,
(the list also includes CaClþ and CaCl20 in chloride-containing CaCl20, HSO4, and SO42.
solutions), and the carbonic acid species, CO2(aq), HCO3, The behavior of barite (BaSO4) is similar to that of gypsum,
and CO32. The hydrolytic species, CaOHþ, may also form, except that it is considerably less soluble. The solubility of
but its contribution to calcite solubility is likely to be barite in pure water increases with temperature to a maximum
insignificant in CO2-saturated solutions. of 0.000 02 m at  150  C and, thereafter, decreases with in-
The other important carbonate mineral in ore-forming hy- creasing temperature. With the addition of NaCl and/or KCl,
drothermal systems, particularly in systems forming MVT de- this maximum shifts to higher temperature. Pressure has a
posits, is dolomite. The behavior of this mineral at elevated similar effect, increasing the solubility of barite by a factor of
temperature, however, has attracted considerably less attention 3 from 100 to 1000 bar (Blount, 1977). The speciation con-
from experimentalists than has calcite. A number of studies trolling barite solubility is effectively the same as that for
were conducted on the solubility of dolomite at ambient tem- gypsum and anhydrite, except that Ca is replaced by Ba.
perature during the 1950s and 1960s (e.g., Garrels et al., 1960) Fluorite (CaF2) is by far the most commonly occurring
but reported values for its solubility product differed by up to fluorine-containing mineral and is an important gangue min-
three orders of magnitude. The probable reason for this has eral in MVT deposits, as well as in rare earth element (REE)
since been revealed by experimental studies of the kinetics of deposits, where it shows a clear spatial association with the
dolomite dissolution (Arvidson and Mackenzie, 1999; REE mineralization (Williams-Jones et al., 2000). In pure
water, fluorite is relatively insoluble, with a maximum solubil-
ity of 0.000 2 m, which is reached at a temperature of 100  C at
6.0
swvp (Strübel, 1965). Thus, like gypsum, its solubility is pro-
300 ⬚C, 500 bar
5.0 grade at low temperature and retrograde at higher temperature;
however, its maximum solubility is only about 10% of that of
Ca (mmol kg-1)

4.0 te
ydri gypsum. With the addition of alkali chlorides, fluorite solubil-
Anh
3.0 ity increases (Figure 4), and in a 2 m NaCl solution, its solu-
bility is prograde at all temperatures. The solubility of fluorite,
2.0 like that of the carbonate and sulfate minerals, also increases
Calcite with pressure (Tropper and Manning, 2007). Species control-
1.0
Fluorite ling the solubility of fluorite comprise Ca2þ, CaClþ, CaCl20,
0.0 HF, and F.
0 0.2 0.4 0.6 0.8 1 From the preceding paragraphs, it is apparent that the
NaCl (mol kg-1) effects of temperature, pressure, and salinity on the solubility
Figure 4 The solubility of anhydrite, calcite, and fluorite as a function of of calcite, gypsum, anhydrite, and fluorite are quite similar,
NaCl molality at 300  C and 500 bar; calculated using thermodynamic although the absolute solubility of these minerals varies con-
data from Holland and Powell (1998) and Shock et al. (1997). siderably at comparable conditions. All these minerals display
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 39

retrograde solubility in dilute aqueous solutions, albeit over sulfur, thiosulfate, or polysulfides, the rate of conversion of
quite different temperature intervals. They also all revert to these intermediate phases to pyrite is extremely fast and may
prograde solubility with the addition of alkali chlorides, and occur within minutes, whereas with only hydrogen sulfide
they all increase in solubility within increasing pressure. The or bisulfide present, the conversion rate is much slower
main reason for their similar behavior is that these minerals (Schoonen and Barnes, 1991a).
only differ in the nature of the ligand that binds with Ca2þ (i.e., One of the important roles of pyrite in hydrothermal sys-
CO32, SO42, or F). These are all hard ligands having high tems is to buffer parameters such as fO2 (or f H2 ) and pH
charge/radius ratios and are strong electron acceptors, and through reactions with minerals like pyrrhotite, magnetite,
therefore, their bonding with Ca2þ is dominantly ionic. More- and hematite. For example, the assemblage pyrite–pyrrhotite
over, as the relative order of hardness for these ligands is F > constrains oxygen fugacity under acidic conditions through the
CO32 > SO42, then Ca2þ should form its strongest bonds reaction
with these ligands in the same order, and consequently, at a
given set of conditions, anhydrite should have the highest ð1  xÞFeS2 þ ð1  xÞH2 O ¼ Fe1x S þ ð1  xÞH2 S
solubility followed, in turn, by calcite and fluorite. This is þ 0:5ð1  xÞO2
what is observed (Figure 4).
and the assemblage pyrite–pyrrhotite–magnetite buffers both
Pyrite (FeS2) is the most important sulfide gangue mineral
f O2 and pH, in both cases for a given total concentration of
and occurs in virtually all major hydrothermal base and pre-
aqueous sulfur (Figure 5).
cious metal mineral deposits, except those subjected to high-
The solubility of most silicate minerals increases with tem-
grade metamorphism, in which pyrrhotite has commonly
perature in aqueous solutions, that is, it is prograde, and de-
replaced pyrite. Furthermore, pyrite is the principal sulfide
pends strongly on pH. In most cases, these minerals dissolve
mineral in most of these deposits, and its concentration is
incongruently, and the nature of the phase formed depends
generally much greater than that of the sulfide minerals con-
strongly on the composition of the solution, including its pH.
taining the ore metals of interest. Numerous studies have
For example, at relatively low pH and/or low activity of Kþ, the
investigated the dissolution of pyrite and its deposition from
dissolution of K-feldspar may involve formation of kaolinite
aqueous solutions (Ohmoto et al., 1994; Schoonen and
(or another aluminosilicate phase) through the reaction
Barnes, 1991a,b). These studies have shown that the solubility
of pyrite is prograde and depends strongly on pH. 2KAlSi3 O8 þ H2 O þ 2Hþ ¼ Al2 Si2 O5 ðOHÞ4
Formally, the dissolution of sulfide minerals can be de- þ 4SiO2ðaqÞ þ 2Kþ
scribed by reactions of the type

MeS ¼ Me2þ þ S2

but as hydrothermal solutions invariably contain insignificant


-10
concentrations of S2 (Ellis and Giggenbach, 1971; Migdisov
250 ⬚C, 1 m NaCl,
et al., 2002), the dissolution of these minerals is more conve- pH 4.5
niently described by reactions involving H2S and HS, the
dominant sulfide species in solution, at acidic and near-neutral
-4
to alkaline conditions, respectively. As mentioned earlier, pK2 -8
for H2S is 17 at 25  C. However, the sulfur in most sulfide
Hematite

minerals is in the 2 state, whereas in pyrite, the sulfur occurs


in both the zerovalent and 2 state as the S0.S2 ion (i.e., as the
simple S22 polysulfide ion). The solubility of pyrite is there- -6
log aH2 aq

fore most appropriately described by a reaction involving the Pyrite


-3
polysulfane, H2S2, but as the stability of polysulfanes and -8
-6
especially their nondissociated forms is poorly known at ele- -4
vated temperature, the dissolution of pyrite is normally -4 Magnetite
described by redox reactions involving H2S and HS, that is,

FeS2 þ 2Hþ þ H2 O ¼ Fe2þ þ 2H2 S þ 0:5O2


-2 -2
or -3
-4
FeS2 þ H2 O ¼ Fe2þ þ 2HS  þ 0:5O2 -2
Pyrrhotite
The dissolution of pyrite is incongruent and, depending on 0
the system, may proceed via complex reactions involving for- -6 -4 -2 0
log aH2S aq
mation of elemental sulfur or iron oxides (e.g., Caldeira et al.,
2010; Rimstidt, 2003). The process of pyrite precipitation is Figure 5 Solubility contours (logm SFe) for pyrite, pyrrhotite,
also complex and, below 300  C (and perhaps at higher magnetite, and hematite as a function of log aH2 S aq and log aH2 aq in a 1 m
temperature), proceeds via intermediate steps involving forma- SCl aqueous solution at pH 4.5, a temperature of 250  C, and vapor-
tion of precursor iron monosulfide, mackinawite (FeS), and saturated water pressure; calculated using thermodynamic data from
the thiospinel, greigite (Fe3S4). In the presence of elemental Holland and Powell (1998) and Shock et al. (1997).
40 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

6 100 ⬚C, swp Table 3 Ligands occurring in hydrothermal ore solutions


5
Hydroxide OH
4 Muscovite
Halide ions F, Cl, Br, I

aH+
aK+
3 K-feldspar Sulfur species HS, Sn, SnS2, SO32, S2O32, SO42
log 2 Ammonia (ammine) NH3
1 Oxyanions CO32, PO43, AsO33, SbO33, MoO42,
0 Kaolinite WO42, SiO44
-1 Thioanions AsS33, SbS33, MoS42, WS42
-2 Carboxylates CH3COO (acetate), C2H5COO (propionate),
-4 -3 -2 -1 0 1 CH2(COO)22 (malonate), (COO)22 (oxalate)
log aSiO2 Miscellaneous HTe, Te22, CN, SCN
ligands of possible
Figure 6 Stability relationships among muscovite, kaolinite, and interest

K-feldspar as a function of log aSiO2 aq and log aKþ =aHþ at 100  C and
vapor-saturated water pressure; calculated using thermodynamic data
from Holland and Powell (1998) and Shock et al. (1997).
stoichiometry of metal complexes with available ligands as a
function of temperature, pressure, pH, and redox potential
By contrast, at higher pH and/or higher activity of Kþ, the disso- comprises the fundamental basis for understanding the pro-
lution reaction leads to the formation of muscovite (or illite): cesses of hydrothermal ore formation.
What ligands are present in hydrothermal ore solutions?
3KAlSi3 O8 þ 2Hþ ¼ KAl3 Si3 O10 ðOHÞ2 þ 6SiO2ðaqÞ þ 2Kþ
The three most important are Cl, HS, and OH, although
Both reactions also depend on the activity of aqueous silica various other electron donors are also present (Table 3) and
(Figure 6). may be important in the complexing and transport of some
Because of their importance as rock-forming minerals, metals and under certain conditions. In natural systems, ligand
many of these silicates play an important role in buffering availability is important. For example, chloride is always pre-
fluid composition, and analysis of hydrothermal equilibria sent at much higher concentrations than, say, iodide, and
involving them may help in evaluating compositional param- hence, in a 1.0 m chloride ore fluid at 500  C and 500 bar,
eters such as pH. As hydrothermal fluids commonly saturate chloride would be the more important complexing ligand for
with respect to quartz, aqueous silica activity is typically buff- gold transport even though the gold(I) iodide complex, AuI2,
ered, and reactions involving K-feldspar, muscovite, and an is more stable than AuCl2.
aluminosilicate mineral can be rewritten as follows: Ligand availability and stability may be affected by various
homogeneous reactions occurring within the hydrothermal
3K-feldspar þ 2Hþ ¼ Muscovite þ 6Quartz þ 2Kþ ore solution itself. The activity of free chloride ion in an aque-
2Muscovite þ 3H2 O þ 2Hþ ¼ 3Kaolinite þ 2Kþ ous NaCl solution is much diminished with increasing tem-
perature due to ion pairing as discussed earlier. A 1.0 m NaCl
Thus, the coexistence of K-feldspar with muscovite in the solution at 400  C and 300 bar will be >99% associated,
presence of quartz or muscovite with kaolinite buffers the Kþ/ which will determine the number of coordinated chlorides in
Hþ activity ratio and, if albite is also present, the Kþ activity can metal complexes. In addition, phase separation and boiling
be deduced from fluid inclusion salinity and the Kþ/Naþ ratio may dramatically decrease the activity of compounds such as
corresponding to equilibrium between K-feldspar and albite, H2S, H2Te, H2Se, NH3, and CO2 in the residual liquid (or
thereby permitting evaluation of pH (see also the discussion in denser supercritical) phase, effectively removing the corre-
Chapter 13.1). At given concentrations of Kþ or Naþ, hydro- sponding weak acid/base ligands. Changes in the redox envi-
lysis of alkali feldspar to muscovite and paragonite, or the latter ronment that accompany boiling and/or mixing with
minerals to an aluminosilicate mineral, buffers pH to a fixed oxygenated fluids in the upper parts of geothermal/hydrother-
value that depends only on temperature and pressure (e.g., mal systems also lead to ligand destruction by oxidation as in
Hemley et al., 1992). the case of H2S/HS, which reacts to sulfate and other inter-
mediate oxidation state species of sulfur. Mineral deposition
(e.g., sulfide precipitation) also causes H2S/HS loss, especially
13.2.4 Ore Metal Transport and Deposition in the ore-depositing environment where a cascade of interre-
lated chemical events leads to the almost complete deposition
13.2.4.1 Ore Fluids with Liquid-Like Densities
of various metals from the ore fluid. In addition, some ligands
13.2.4.1.1 Ligands in hydrothermal ore solutions are not thermally stable. Acetate (CH3COO) is well known to
Ore-forming hydrothermal fluids migrate through the Earth’s decarboxylate at elevated temperatures (Bell and Palmer,
crust transporting metals as complex ions and molecules. 1994), and thiosulfate (S2O32) disproportionates at elevated
Focused deposition of metals to form an ore deposit occurs temperatures Giggenbach, 1974b).
in response to physicochemical changes in the local environ- Finally, it is noted that complex stabilities differ signifi-
ment. In essence then, ore deposition as a manifestation of the cantly because of the differing metal–ligand bonds. The differ-
stability/instability or disproportionation of metal complexes ing complex stabilities can be predicted using hard/soft Lewis
at a given temperature and pressure (or T–P range) may be acid/base theory (Ahrland, 1968; Pearson, 1963) as discussed
considered. A detailed knowledge of the stability and by Seward and Barnes (1997) with reference to hydrothermal
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 41

metal complexing. Thus, a soft electron donor such as HS but, until relatively recently, few studies extended to higher
would be predicted to form very stable complexes with Auþ (a temperatures. Most earlier studies have been discussed/sum-
soft electron acceptor), whereas the fluoride (hard Lewis base/ marized by Seward (1981) and Seward and Barnes (1997).
electron donor) complexes of Auþ are much less stable and the Sverjensky et al. (1997) tried to overcome the paucity of reli-
first isolatable Au(I) fluoride complex has only recently been able, experimentally based thermodynamic data for metal
synthesized (Laitar et al., 2005), although fluoride complexes complex equilibria at extreme conditions with computational
with higher oxidation states of gold are well documented predictions extending to 1000  C and 5000 bar. Studies of
(Mohr, 2004). This approach is simplistic compared to the chloridosilver(I) and chloridolead(II) stepwise complex for-
modern quantum chemical assessment of metal complex ge- mation up to 350  C and 300  C, respectively (Seward, 1976,
ometries and metal–ligand bond energies of complexes of 1984), demonstrated that the higher coordination number
geochemical interest (e.g., Boily and Seward, 2005; Sherman, species, such as AgCl43 and PbCl42, are not observed at
2010) but facilitates an overview of the important metal elevated temperatures (e.g., 350  C) and equilibrium saturated
transport mechanisms in the multicomponent electrolyte vapor pressures, mainly because of the low chloride ion activity
solutions, which comprise the diversity of hydrothermal ore due to NaCl association. Several recent studies (Figure 7) reaf-
solutions. firm earlier observations concerning stepwise complex forma-
tion as well as demonstrating the coordination geometry
13.2.4.1.2 Metal chloride complexing changes that may occur at elevated temperatures, as discussed
Sodium chloride is the ubiquitous salt in hydrothermal ore in some detail by Seward (1981) and Seward et al. (1996). In
solutions, and hence, the stability of chloride complexes of the case of complexing of tin(II) with chloride (Müller and
many elements at high temperatures and pressures plays a Seward, 2001; Figure 7), the distribution of species changes
fundamental role in the transport of these elements by migrat- and the SnCl42 species is not observed in the more concen-
ing ore fluids throughout the Earth’s crust and their subsequent trated chloride solutions up to 3.0 m at 300  C at the equilib-
precipitation as ore minerals in an ore-depositing environment rium saturated vapor pressure.
at a given temperature and pressure. It has long been known Coordination geometry changes attending stepwise com-
that chloride may form a stepwise series of complexes with plex formation and changes in temperature and pressure
many metal cations at ambient temperature as defined by have long been known (Seward, 1981), especially in the case
of cobalt(II) and nickel(II) species in aqueous solution (e.g.,
Mnþ þ mCl ¼ MClm nm Lüdemann and Franck, 1967). Recently, the stepwise formation

100 100
25 ⬚C 300 ⬚C
0 0 3
80 80

60 60
% Sn

% Sn

1 2
2 3 4
40 40
1
20 20

0 0
-3 -2 -1 0 1 -3 -2 -1 0 1
log Cltotal / mol kg-1 log Cltotal / mol kg-1

100 100
35 ⬚C 4 350 ⬚C
80 0 80

60 1 60
% Co
% Co

40 40
2-o 2-t
20 20
1
4
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Cltotal / mol kg-1 Cltotal / mol kg-1

Figure 7 Stepwise formation of chloridotin(II) and chloridocobalt(II) complexes at ambient and elevated temperatures at saturated vapor pressure; the
curves are labeled according to the number of chloride ligands bound to each metal cation (i.e., Sn2þ or Co2þ). In the case of the cobalt(II) chloride
complexes, 2-o refers to octahedrally coordinated CoCl2(H2O)4 and 2-t refers to tetrahedrally coordinated CoCl2(H2O)2. Modified from Müller B and
Seward TM (2001) Spectrophotometric determination of the stability of tin(II) chloride complexes in aqueous solution up to 300 C. Geochimica et
Cosmochimica Acta 65: 4187–4199; Liu W, Borg SJ, Testemale D, Etschmann B, Hazemann J-L, and Brugger J (2011) Speciation and thermodynamic
properties for cobalt chloride complexes in hydrothermal fluids at 35–440 C and 600 bar: An in-situ XAS study. Geochimica et Cosmochimica Acta 75:
1227–1248.
42 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

of cobalt(II) complexes and their octahedral/tetrahedral con- These include studies of the chloride complexes of Ti4þ
figurational changes have been revisited by Liu et al. (2011) (Ryzhenko et al., 2006), Mn2þ (Suleimenov and Seward,
(Figure 7), who demonstrated the predominance of tetrahe- 2000), Fe2þ and Fe3þ (Liu et al., 2007; Stefansson et al.,
drally coordinated moieties, including CoCl42, in expanded 2008; Testemale et al., 2009), Co2þ (Liu et al., 2011; Migdisov
high-temperature water. Of particular interest as well is the et al., 2011a,b), Ni2þ (Tian et al., 2012), Cuþ and Cu2þ
nature of complex geometry and the manner in which the (Etschmann et al., 2010; Fulton et al., 2000; Liu et al., 2002,
solvent molecules interact with metal complex ions/molecules 2008; Sherman, 2007; Xiao et al., 1998), Zn2þ (Anderson et al.,
to affect metal–ligand bonds and complex stability. In the case 2000; Harris et al., 2003; Trevani et al., 2009), Cd2þ (Bazarkina
of stepwise complex formation between Cd2þ and Cl, for et al., 2010; Palmer et al., 2000; Seward and Driesner, 2004;
example, solvent molecules (i.e., H2O) contribute fundamen- Seward et al., 2013), In3þ (Seward et al., 2000), Sn2þ and Sn4þ
tally to the stability of the complex (Seward et al., 2013). As (Kovalenko and Ryzhenko, 1997; Müller and Seward, 2001;
shown in Figure 8, the Cd–O (water) and Cd–Cl distances Sherman et al., 2000a,b; Uchida et al., 2002), Sb3þ (Oelkers
obtained from EXAFS measurements are 2.30 and 2.50 Å, et al., 1998; Pokrovski et al., 2006), Pd2þ (Boily and Seward,
respectively, for the octahedral CdCl(H2O)5þ complex, but 2005; Seward and Driesner, 2004), Auþ and Au3þ (Gammons
these values differ from the quantum chemically calculated et al., 1997; Stefansson and Seward, 2003a,b,c; Usher et al.,
values with only five first-shell water molecules bound to the 2009), Tlþ (Bebié et al., 1998), rare earth(III) elements except
Cd2þ. However, with stepwise addition of solvated waters to the Pm3þ (Gammons et al., 2002; Mayanovic et al., 2002, 2007,
second shell, one observes that the calculated Cd–O distance 2009; Migdisov and Williams-Jones, 2002, 2006; Migdisov
asymptotically approaches the value measured by X-ray absorp- et al., 2008, 2009; Stepanchikova and Kolonin, 2005), Y3þ
tion spectroscopy when n  4. However, at least eight second- (Ragnarsdottir et al., 1998), and U4þ (Kovalenko et al.,
shell waters are necessary to reproduce the measured Cd–Cl 2011). All these studies provide thermodynamic data and/or
distance. These concepts are of considerable importance in molecular insight into complex energetics, geometries, and
understanding the partitioning of metal complex ions and mol- metal–ligand bond lengths, the aim being to understand fun-
ecules to a steam-phase or low-density supercritical fluid in damental aspects of metal transport, metal (complex) parti-
hydrothermal systems. Metal complexes in such lower-density tioning during fluid phase separation, and ore mineral
fluids are also hydrated moieties, which comprise molecular precipitation in ore-forming hydrothermal systems.
aggregates or clusters, as discussed in the succeeding text.
In recent years, further advances in our understanding of 13.2.4.1.3 Complexing with other halide ligands
metal chloride complexing at hydrothermal conditions have With the exception of Gammons and Yu (1997) and Liu et al.’s
come from solubility measurements, UV/Vis and X-ray absorp- (2012) studies of bromide and iodide complexes of silver and
tion spectroscopy, and ab initio and ab initio/MD calculations. zinc at elevated temperatures, there have been few studies of

Stepwise hydration of CdCl(H2O)5+

2.65 Cd–Cl
[CdCl(H2O)5(H2O)n]+
2.60
2.55
EXAFS bulk, Cd–Cl, 298 K
2.50

Cd–Cl (Å)
Cd–O (Å)

2.45
[CdCl(H2O)5(H2O)1]+ [CdCl(H2O)5(H2O)2]+ [CdCl(H2O)5(H2O)3]+
2.40
2.35
Cd–O
2.30 EXAFS bulk, Cd–O, 298 K
MP2/aug-cc-pVDZ(PP), this study
2.25 MP2/aug-cc-pVTZ(PP), this study
HF/MSI*, Butterworth92
2.20
0 1 2 3 4 5 6
n, Solvation number
[CdCl(H2O)5(H2O)4]+ [CdCl(H2O)5(H2O)5]+ [CdCl(H2O)5(H2O)6]+

MP2 optimized geometries


Figure 8 EXAFS and ab initio data for cadmium(II) chloride and oxygen (water) distances for the octahedral monochloridocadmium(II) complex, CdCl
(H2O)5þ; the EXAFS-determined Cd–O (water) and Cd–Cl distances are given by the horizontal dashed lines. The quantum chemically calculated
distances for the optimized geometries resulting from the addition of each second-shell water are also shown. Reproduced from Seward TM, Lemke KH,
Henderson CMB, and Charnock JM (2013) An X-ray absorption spectroscopic and ab initio computational study of the Cd(II) aquated ion and
chlorocadmium(II) complexing in hydrothermal solutions. Chemical Geology (in review).
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 43

metal bromide and iodide complexes under hydrothermal an adequate understanding of the hydrothermal transport and
conditions from which thermodynamic data have been deposition chemistry of uranium is available. The uranyl ion,
obtained. In near-magmatic to epithermal hydrothermal UO22þ, also plays a role in hydrothermal uranium transport in
fluids, bromide and iodide complexes are not considered to environments of intermediate redox potential.
play a significant role in ore transport. Nevertheless, bromide
and iodide complexes may play a role in the volatile metal
complex chemistry in magmatic/volcanic gas systems and tran- 13.2.4.1.4 Metal complexes with hydroxide and other
sition metal transport chemistry. Various sulfide/halide subli- oxygen electron donor ligands
mate minerals containing bromide and iodide occur in Many metal ions hydrolyze in high-temperature aqueous solu-
fumarolic deposits of active volcanoes, such as at Vulcano in tions in the pH range encompassed by natural hydrothermal
Italy (Demartin et al., 2010). Bromide and iodide complexes fluids in the Earth’s crust, and hence, the hydroxide complexes
may play a role in metal transport in sedimentary basinal brines, can play an important role in the chemistry of hydrothermal
but this has been little studied. Bromide interaction with some ore transport. A detailed overview of metal ion hydrolysis
metal cations has been studied at hydrothermal conditions speciation is given by Baes and Mesmer (1976), and
using X-ray absorption spectroscopy (Ni2þ: Hoffmann et al., Wesolowski et al. (2004) provide an erudite summary of
1999; Wallen et al., 1998; Zn2þ: Anderson et al., 2000; Liu metal ion hydrolysis and metal oxide solubility measurements
et al., 2012; Simonet et al., 2002). These latter studies emphasize at elevated temperatures. It is well known that the solubility of
the tendency toward tetrahedrally coordinated species, such as metal oxides in aqueous media over wide ranges of tempera-
ZnBr2(aq) (i.e., ZnBr2(H2O)20) and ZnBr42, with increasing ture and pressure is pH-dependent (Wesolowski et al., 2004),
temperature, as shown also by the high-temperature (up to and as in the case of rutile solubility at 300  C, for example, the
500  C) Raman data of Mibe et al. (2009). minimum solubility in the near-neutral region and the overall
However, fluoride plays an important role in the transport shape of the solubility curve is due to the hydrolysis of Ti4þ
of some metals/elements such as titanium, tin, rare earths, with the formation of various Ti(OH)n4n species (Knauss
zirconium, and uranium in hydrothermal ore solutions. In et al., 2001) according to the reaction
the case of titanium, for example, the solubility of rutile in
high-temperature hydrothermal fluids is much enhanced by TiO2 ðrutileÞ þ ð4  nÞHþ þ ðn  2ÞH2 O ¼ TiðOHÞ4 4n
the presence of fluoride (Rapp et al., 2010) due to the complex-
ing of Ti4þ by F (Ryzhenko et al., 2006) and the formation of The elevated solubility of rutile in supercritical water (up to
mixed hydroxofluoro -complexes, such as TiF(OH)30, 1000  C and 2.3 GPa) in the presence of albite (Audétat and
TiF2(OH)20, and TiF(OH)4. Kovalenko et al. (1992) and Keppler, 2005; Hayden and Manning, 2011; Manning et al.,
Kovalenko and Ryzhenko (1997) have demonstrated the im- 2008) indicates that dissolved Ti(IV) forms some sorts of ad-
portance of tin(II) fluoride complexes, such as SnF(OH)2 and ducts with dissolved aluminum silicate species in solution.
SnFCl0, in hydrothermal fluoride solutions at 500  C and Iron occurs ubiquitously in hydrothermal ore deposits as
1 kbar, but more experimentally based thermodynamic data sulfide, oxide, silicate, and carbonate minerals. The hydrother-
are required in order to adequately understand the hydrother- mal ore transport chemistry of Fe(II) and Fe(III) is dominated
mal transport and deposition chemistry of tin by fluids in the by chloride complexing (see the preceding text) and by hydro-
Earth’s crust. Recent experimental studies (Migdisov et al., lysis equilibria, which have been extensively studied up to
2011a,b; Prisyagina et al., 2008; Ryzhenko et al., 2008) have 300  C (Wesolowski et al., 2004). Recently, the formation
also demonstrated that the zirconium(IV) hydroxofluoride of Fe(OH)2þ has been studied spectrophotometrically by
complexes are important in determining the transport (i.e., Stefansson et al. (2008) up to 300  C, and their thermody-
the mobility) of zirconium (and the solubility of zircon and namic data are in good agreement with the few earlier reported
baddeleyite) in hydrothermal fluids in the Earth’s crust. The data. Unfortunately, there are essentially no experimentally
fluoride complexes of the rare earth(III) elements (except derived thermodynamic data for iron complexing and hydro-
Pm3þ) have been studied using solubility and spectrophoto- lysis at higher temperatures from 350  C to supercritical
metric methods by Migdisov and Williams-Jones (2007) and conditions, which limits our ability to model the transport
Migdisov et al. (2009) up to 300  C at the equilibrium vapor and precipitation chemistry of iron in many ore-depositing
pressure. Previous theoretical predictions had overestimated environments in the Earth’s crust.
the stability at elevated temperatures of species such as NdF2þ Significant gallium enrichments occur in the metal-
and hence its importance in hydrothermal fluids relative to the enriched surface precipitates of some active geothermal sys-
chloride complexes, NdCl2þ and NdCl2þ. The values of the tems (Krupp and Seward, 1987), but the aqueous chemistry
equilibrium constants for the formation of the monofluorido- of gallium (predominantly as Ga3þ) has been little studied at
and monochlorido-complexes of the rare earth(III) elements elevated temperatures and pressures, and the transport chem-
are summarized in Figure 9. More data are required to under- istry by hydrothermal fluids in the Earth’s crust is poorly
stand REE transport and deposition chemistry by hydrother- known. Bénézeth et al. (1997) measured the solubility of
mal fluids at T > 350  C. Kovalenko et al. (2012) have a-GaOOH from 150 to 250  C and estimated the equilibrium
measured the solubility of uraninite (UO2) in HF solutions at formation constants for Ga(OH)n3n from 25 to 300  C.
500  C and 1000 bar and identified the hydroxofluoride spe- The hydrolysis Nd3þ has been studied by solubility and
cies, U(OH)3F0 and U(OH)2F20, thus emphasizing the proba- potentiometric methods up to 290  C by Wood et al. (2000),
ble importance of fluoride complexing of U4þ in hydrothermal but otherwise, there are few experimentally based data available
transport, but more experimental studies are required before on REE hydroxide complexes under hydrothermal conditions.
44 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

9
250 ⬚C

8
200 ⬚C
250 ⬚C
150 ⬚C

Log b F1
7
200 ⬚C
6

5 150 ⬚C

4
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu

3.5
250 ⬚C
3

2.5 200 ⬚C
Log b Cl1

1.5
150 ⬚C
1

0.5
La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Figure 9 The equilibrium cumulative formation constants for the monofluorido- and monochlorido-complexes of the rare earth elements in
aqueous solution up to 250  C and at the saturated vapor pressure. Modified from Migdisov AA, Williams-Jones AE, and Wagner T (2009)
An experimental study of the solubility and speciation of rare earth(III) elements in fluoride and chloride bearing aqueous solutions at temperatures
up to 300 C. Geochimica et Cosmochimica Acta 73: 7087–7109.

Hydrothermal molybdate chemistry has been studied by (i.e., tungstic acid ionization and polynuclear tungstate forma-
Ulrich and Mavrogenes (2008) who measured the solubility tion) up to 300  C. Recently, Minubayeva and Seward (2013)
of Mo and MoO3 in water and KCl solutions from 500 to have also studied (spectrophotometrically) the deprotonation
800  C and to 3000 bar and suggested (with the aid of of simple monomeric H4WO4 up to 300  C and reported
XANES spectra) that solubility was due to the simple molybdic similar results to those of Wesolowski et al. (1984).
acid species as well as to chloromolybdate species (e.g., Arsenic and antimony are fellow travelers in hydrothermal
MoOmCln2) (see also the thioanion discussion in the succeeding ore fluids and are often involved in ore mineral deposition,
text). Minubayeva and Seward (2010) used ultraviolet spectros- and hence, an understanding of their hydrothermal chemistry
copy to obtain the first thermodynamic data for molybdic acid will provide additional insight into ore formation. The equilib-
ionization up to 300  C at the equilibrium saturated vapor rium constants for arsenous and antimonous acid ionization
pressure, that is, have been determined up to 300  C at saturated vapor pressures
by Zakaznova-Herzog et al. (2006) and Zakaznova-Herzog and
H2 MoO4 ¼ HMoO4 þ Hþ and Seward (2006). Their data permit the modeling of arsenic(III)
HMoO4 ¼ MoO4 2 þ Hþ and antimony(III) transport and deposition in hydrothermal/
geothermal fluids having low, reduced sulfur activity. The
and Yan et al. (2011) have confirmed the MoO4 stoichiometry dilemma, of course, is that hydrothermal ore fluids contain
(using EXAFS) in hydrothermal solutions up to 600  C. reduced sulfur and thioarsenite, and thioantimonite species
In addition, Borg et al. (2012) have confirmed molybdate (i.e., H3(As,Sb)S3(As,Sb)(HS)3) will also contribute to the
stoichiometry and geometry up to 385 C and 600 bar using total dissolved As(III) and Sb(III) in the fluids. The thermody-
X-ray absorption spectroscopy and also identified chloromo- namic data for these species are sparse and in poor agreement, as
lybdate species in concentrated HCl solutions. There are few discussed briefly in the succeeding text:
studies of hydrothermal tungstate chemistry, and the earlier
H3 AsO3ðaqÞ ¼ H2 AsO3 þ Hþ and
potentiometric study by Wesolowski et al. (1984) still com-
prises the benchmark work on aqueous tungstate chemistry H3 SbO3ðaqÞ ¼ H2 SbO3 þ Hþ
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 45

Carbonate and bicarbonate are ubiquitous ligands in natu-


ral hydrothermal fluids and form stable complexes with many 100
metals (i.e., with hard/intermediate Lewis acids or class A
acceptors), such as Pb2þ, UO22þ, and REE ions, but there are AuOH0
80
essentially no systematic studies of metal carbonate complexes
up to elevated temperature and pressure and from which ther-
60
modynamic data are available. Such complexes will be pH-

% Au
sensitive, and their stabilities will decrease dramatically during
boiling or with the precipitation of carbonate minerals. 40
Au(OH)2-
Various carboxylic acid and other organic species occur in Au(HS)2-
sedimentary basin brines, the most important of which is AuCl2-
20
acetate (Shock and Koretsky, 1993), and a number of metal AuHS0
acetate complexes have been studied at elevated temperatures 0
up to 300  C. These include potentiometric and spectroscopic
measurements of the acetate complexes of Fe2þ (Giordano and (a)
Drummond, 1991; Palmer and Drummond, 1988), Co2þ
(Bridger et al., 1981), Cuþ (Liu et al., 2001), Zn2þ and Pb2þ 100
(Yang et al., 1989), Cd2þ (Bénézeth and Palmer, 2000), Csþ
and Sr2þ (Ragnarsdottir et al., 2001), and Nd3þ, Eu3þ and Y3þ 80
AuOH0
(Tagirov et al., 2007a,b; Wood et al., 2000; Zotov et al., AuCl2-
2002). Other literature data on metal–organic complexing
60
relevant to ore-depositing systems are discussed by Seward

% Au
and Barnes (1997). Au(OH)2-
40
Au(HS)2-
13.2.4.1.5 Complexing with hydrosulfide/sulfide ligands 20
The hydrosulfide ligand, HS, plays a fundamentally impor- AuHS0
tant role in the transport and deposition chemistry of some 0
elements in hydrothermal ore solutions. Auþ is the archtype (b)
soft Lewis acid (class B) electron acceptor and, as such, forms
very stable complexes with HS. The stoichiometry and stabil- 100 AuCl2-
ity of the AuHS0 and Au(HS)2 species were first reported by AuHS0
Seward (1973), and there have been numerous studies of the 80
interaction of Auþ with reduced sulfur ligands since then (e.g.,
Baranova and Zotov, 1998; Benning and Seward, 1996; Dadze 60 Au(HS)2-
% Au

et al., 2000; Gibert et al., 1998; Loucks and Mavrogenes, 1999;


Pokrovski et al., 2009; Renders and Seward, 1989a; Shenberger
40
and Barnes, 1989; Stefansson and Seward, 2004; Tagirov et al.,
2005, 2006). The study of Stefansson and Seward (2004) pro-
vides thermodynamic data for the formation of the hydrosulfido- 20
gold(I) complexes, AuHS0 and Au(HS)2, up to 500  C and
500 bar, which, together with the data for hydroxide and 0
chloride complexes (Stefansson and Seward, 2003a,b), enable
the modeling of gold transport and deposition chemistry in
(c) 0 2 4 6 8
hydrothermal ore fluids over a wide range of temperature and
pH
pressure (Figure 10). As shown in Figure 11, the solubility of
gold in aqueous sulfide solution at elevated temperature is very Figure 10 The distribution of gold complexes in solutions of different
sensitive to changes in pH and redox potential. In addition, composition at 400  C and 500 bar. (a) H2S þ HS ¼ 0.001 m,
any process that causes a loss of reduced sulfur (i.e., Cl ¼ 0.001 m; (b) H2S þ HS ¼ 0.001 m, Cl ¼ 0.5 m;
H2S þ HS) will lead to complex instability and, hence, the (c) H2S þ HS ¼ 0.5 m, Cl ¼ 0.5 m. Modified from Stefansson A and
precipitation of gold. Boiling a hydrothermal fluid can be Seward TM (2004) Gold(I) complexing in aqueous sulfide solutions to
500 C at 500 bar. Geochimica et Cosmochimica Acta 68: 4121–4143.
catastrophic for the stability of the gold(I) hydrosulfide com-
plexes. Partitioning of volatiles into the vapor phase in epither-
mal environments leads to a loss of H2S (i.e., loss of the HS reduced sulfur from the ore fluid. In supercritical, magmatic
ligand), H2 (i.e., a redox change), and CO2 (i.e., an increase in hydrothermal fluids, the condensation of an iron-rich brine
pH). A change in the redox potential of the residual liquid during phase separation allows the less dense, volatile-rich
phase during boiling and /or mixing with oxygenated meteoric (H2S-containing) phase to then cool, contract, and ascend
waters encountered in the upper parts of geothermal/epither- buoyantly in the Earth’s crust without losing most of its dis-
mal systems removes sulfide sulfur (Seward, 1989). The pre- solved sulfide to pyrite precipitation, as discussed by Heinrich
cipitation of pyrite and other sulfide minerals also extracts et al. (2004).
46 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

-29 or H2S/HS) concentration, as discussed, for example, by


- Stefansson and Seward (2003a,b,c). The stoichiometry and sta-
AuCl2 AuHS0

b
pp
HSO4- bility of zinc(II) hydrosulfide/sulfide complexes have been stud-

10
1 pp SO42- ied by Tagirov et al. (2007a,b) and Tagirov and Seward (2010)
b
-33 up to 250  C. At 300  C and the equilibrium vapor pressure, the
10 ppb Au(HS)2-
two hydrosulfide species, Zn(HS)20 and Zn(HS)3, predominate
1 pp
b in the near-neutral 0.10 m sulfide solutions, but for concentra-
AuHS0
log f O2

tions of Cl >0.1 m, the transport of zinc will be dominated by

10
-37

pp
ZnCln2n complexes (Ruaya and Seward, 1986). The stability of

b
1 ppb

pp
100 ppb
the monohydrosulfido-complex of cobalt(II) has recently been

10

pb
1p
reported by Migdisov et al. (2011a,b), who suggested that the
-41 HS - CoHSþ may play some role in cobalt transport in sulfide-con-
H2S Au(HS)2- taining fluids at T < 200  C. The relevance of the related mono-
hydrosulfide complexes of iron and nickel to ore transport under
hydrothermal conditions is not known. Mercury(II) forms stable
-45
hydrosulfide/sulfide complexes under hydrothermal conditions
2 4 6 8 10
(see Barnes and Seward, 1997, and references therein), and other
(a) pH studies pertaining to ambient/low-temperature conditions using
-2 solubility, ab initio, and EXAFS methods (Bell et al., 2007;
250 ⬚C Au(HS)2- Lennie et al., 2003; Paquette and Helz, 1997; Tossell, 2001a)
emphasize the stability of the mercury(II) hydrosulfide/sulfide
-4
log mAu/ mol kg-1

complexes, but more detailed studies under hydrothermal con-


ditions are required.
-6
13.2.4.1.6 Thioanions
Of particular interest as well are the various thio-substituted
AuHS0
-8 oxyanions of As, Sb, Mo, and W, which have been little studied
at elevated temperatures and pressures. There have been a
AuOH0 number of recent spectroscopic (X-ray absorption and ultravi-
-10 olet) studies of aqueous thioarsenite species (Beak et al., 2008;
Bostick et al., 2005; Zakaznova-Herzog and Seward, 2012, and
0 2 4 6 8 10 12 references therein) that confirm the simple thioarsenous acid
(b) pH
stoichiometry (i.e., H3AsS3 or As(HS)3), but the system is also
Figure 11 (a) The solubility of gold (contour lines in ppb) and complicated by the formation of intermediate oxythio-species
speciation at 250  C and 500 bar as a function of pH and log fO2 in a (i.e., AsOS23 and AsO2S3) and their protonated equivalents,
solution containing 1.0 m NaCl and SS ¼ 0.01 m. The dashed lines define all of whose stabilities are pH-dependent. Thioantimonite and
regions of predominance of the various sulfur species; the regions of thioantimonate species have also been reported at elevated
highest gold solubility are colored yellow (10–100 ppb) and red
temperatures to 350  C (Krupp, 1988; Sherman et al., 2000a,b),
(>100 ppb). Thermodynamic data for the gold(I) hydrosulfide
but there is a desperate need for further systematic, corrob-
complexes are from Stefansson and Seward (2004). Modified from
Williams-Jones AE, Bowell RJ, and Migdisov AA (2009) Gold in solution. orative studies to confirm stoichiometries and provide high-
Elements 5: 281–287. With permission from the Mineralogical quality equilibrium thermodynamic data, especially for
Association of Canada. (b) The solubility (molal) of gold at 250  C and thioantimonite and oxythioantimonite (SbO3mSm3) in hy-
500 bar and showing the species contributing to the solubility curve as a drothermal solutions. As noted earlier, the thermodynamics of
function of pH. Reproduced from Stefansson A and Seward TM (2004) molybdic and tungstic acid ionization up to 300  C (i.e.,
Gold(I) complexing in aqueous sulphide solutions to 500 C at 500 bar. H2(Mo,W)O4) are known, but this is not the case for the
Geochimica et Cosmochimica Acta 68: 4121–4143. sulfur-substituted analogues. Thiomolybdate and thiotung-
state, ((Mo,W)O4mSm2, where m ¼ 1–4), are well known at
ambient temperature (e.g., Erickson and Helz, 2000) and may
In addition to gold, the other group IB (group 11) elements, be extensively polymerized (e.g., Mo4S156, Mo4S132, and
Cu and Ag, also form stable hydrosulfide complexes. Thermody- Mo2S72) (Saxena et al., 1968), depending on the moly-
namic and local structural data for CuHS0 and Cu(HS)2 have bdenum and/or tungsten activity. Zhang et al. (2012) have
been studied under hydrothermal conditions by Mountain measured the solubility of MoS2 (molybdenite) from 600 to
and Seward (1999, 2003) and Etschmann et al. (2010), and 800  C at 200 MPa and suggested that an ion-paired dithiomo-
Stefansson and Seward (2003a,b,c) have studied the stability of lybdate species, NaHMo2S2, may account for the observed sol-
AgHS0 and Ag(HS)2 up to 400  C and at 500 bar. At alkaline ubilities. All of these various thioanions will be potentially
conditions, the Cu2S(HS)22 and Ag2S(HS)22 stoichiometries important transporting moieties for elements such as As, Sb,
have been identified, but these are not considered to play a Mo, and W in sulfide-containing hydrothermal fluids in the
significant role in ore transport. Whether the dominant ore Earth’s crust, but there is a dearth of data at elevated tempera-
transport mechanism is chloride or hydrosulfide complexing tures and pressures. In addition, these thioanions may them-
depends upon the temperature, pressure, pH, and ligand (Cl selves act as complexing ligands as in the case of the Cu(I)
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 47

thioarsenite complex, CuAsS(SH)(OH) (Clarke and Helz, role in metal transport/remobilization in the supergene
2000; Tossell, 2001a,b), but there are no data pertaining to environment, and a number of thiosulfate and sulfite minerals
the stability of such species under hydrothermal conditions. have been identified in oxidized supergene zones of ore deposits
The transport of gold in ore solutions as a gold(I) thioarsenite (e.g., sidpietersite, Pb4(S2O3)O2(OH)2, and scotlandite,
complex has long been hypothesized, but there are no available PbSO3). Hannebachite, CaSO3 H2O, occurs in late-stage,
data pertaining to the stability and stoichiometry of such spe- lower-temperature vesicle fillings of the Eifel volcanics
cies at high temperatures and pressures. (Hentschel et al., 1985). In the case of gold(I), for example,
It should be noted as well that As(V) and Sb(V) thio-species the thiosulfate and sulfite complexes are very stable at 25  C
have been identified in surface hot springs of active geothermal with the equilibrium formation constant, log b2 ¼ 26–29.4 (Au
systems (Planer-Friedrich and Scheinost, 2011; Planer-Friedrich (S2O3)23) and log b2 ¼ 30.1 (Au(SO3)23) (Peshchevitsky
et al., 2010). This raises the possibility of thioarsenate and et al., 1970; Pouradier and Gadet, 1969), and the local, near-
thioantimonate species facilitating the mobilization of As and linear, structure of the thiosulfate moiety has been studied using
Sb in the upper parts of geothermal/hydrothermal systems in EXAFS and ab initio DFT methods (Bryce et al., 2003).
which the redox potential of the liquid phase has become In addition, the mixed ligand ammine/thiosulfate and sulfite/
elevated due to boiling and mixing with circulating, steam- thiosulfate gold(I) species have also been studied at 25  C
heated, oxygenated meteoric waters. (Perera and Senanayake, 2004; Perera et al., 2005). However,
the importance of these complexes in hydrothermal ore trans-
13.2.4.1.7 Complexing with other sulfur-containing port is strongly dependent on the stability of the ligands
ligands themselves, at elevated temperatures and pressures. The poly-
The possible role of intermediate oxidation state sulfur species, thionates, thiosulfate, and the divalent polysulfides all disp-
such as polysulfides (SnS2 and Sn), thiosulfate (S2O32), and roportionate at elevated temperatures (i.e., at T  250  C), and
sulfite (SO32), in the hydrothermal transport of gold has been the kinetics of these disproportionations are also sensitive to pH
discussed by Seward (1982), and this topic has recently been as noted by various workers (Foerster et al., 1923; Giggenbach,
revisited by Pokrovski et al. (2009). The divalent polysulfide 1974a,b; Mizoguchi et al., 1976; Pryor, 1960). Nevertheless,
ions (SnS2) disproportionate to Sn, thiosulfate, and monosul- these intermediate oxidation state, anionic sulfur species may
fide (H2S/HS) at T > 150  C, as noted by Giggenbach (1974a, play some role in the mobilization of some metals such as
b). The SnS2 ligands form complexes with Auþ up to moderate gold in the upper, cooler parts of ore-depositing geothermal
temperatures of  150  C (Berndt et al., 1994; Kakovski and systems.
Tyurin, 1962), and such species may play a role in gold precip- In higher temperature, supercritical hydrothermal systems,
itation and local remobilization in the upper parts of geo- the sulfur redox chemistry is often considered in terms of the
thermal systems where fluids of intermediate redox potential reaction
occur. Polysulfide complexes of other elements, for example,
mercury, are known (e.g., Jay et al., 2000; Paquette and Helz, H2 SðgÞ þ 2H2 OðgÞ ¼ SO2ðgÞ þ 3H2ðgÞ
1997), but there are no studies at hydrothermal conditions from
which thermodynamic data have been obtained. In magmatic and volcanic gases, both H2S and SO2 occur,
However, in terms of hydrothermal complexing of gold and but the thermodynamics of the reaction favors SO2 at high
other ‘soft’ metal cations, the S3 ion might be of more impor- temperatures. However, the stability of solvated SO2, that is,
tance. S3 has long been known to occur in ethanol (a low sulfite as SO32 and HSO3, at T > 300  C, is poorly known,
dielectric protic solvent having some similarities to high- although sulfite (SO32 and HSO3) has been shown to dispro-
temperature water with respect to metal complex equilibria) portionate to elemental sulfur and sulfate at temperatures from
and salt melts at high temperatures to 600  C (Giggenbach, 150 to 200  C, sometimes with the formation of transient thio-
1968, 1971c). This blue polysulfide ion, now considered to be sulfate (Foerster et al., 1923; Mizoguchi et al., 1976), depending
the S3 species (Seel et al., 1977), was also shown to be on pH. Elemental sulfur hydrolyzes to H2S/HS and HSO4/
stabilized in high-temperature water by Giggenbach (1971b), SO42 in high-temperature water (Ellis and Giggenbach, 1971).
suggesting that it could play a role in metal complex formation The reactivity of magmatic sulfur dioxide in the presence of
in some hydrothermal ore solutions of intermediate redox po- supercritical water with decreasing temperature is given by
tential and at temperatures in excess of 300  C, as has also been
noted recently by Pokrovski and Dubrovinsky (2011). Tossell 4SO2 þ 4H2 O ¼ H2 S þ 3HSO4 þ 3Hþ
(2012) has also studied the properties of the S3 anion using
quantum mechanical methods and demonstrated that S3 may and this reaction together with the dissociation of HCl0 ion
form complexes with Cuþ (e.g., Cu(S3)2 and Cu(S3)(H2O)0) pairs generates acidity which is then ‘titrated’ by reaction with
having similar stability to the Cu(I) hydrosulfide moieties. silicate minerals (see Giggenbach, 1992).
The intermediate oxidation state oxyanions of sulfur, thio- Sulfate complexes (ion pairs) of the divalent first-row tran-
sulfate (S2O32), sulfite (SO32), polythionates (SnO62), and, sition element cations, Fe2þ (Rudolph et al., 1997), Ni2þ
of course, sulfate (SO42), are known to occur in active geo- (Madekufamba and Tremaine, 2011), Cu2þ (Mendez de Leo
thermal systems, especially in surface hot springs/pools as well et al., 2005), and Zn2þ (Rudolph et al., 1999), have been
as some well discharges and hot crater lake environments studied under hydrothermal conditions, but none of these
(Kaasalainen and Stefansson, 2011; Takano et al., 1994; simple ion pairs (i.e., MSO40) are considered to play a signif-
Webster, 1987; Wilson, 1941; Xu et al., 2000). The sulfur oxya- icant role in ore metal transport in saline hydrothermal fluids
nions and divalent polysulfides (i.e., SnS2) undoubtedly play a of reduced to intermediate redox potential in the Earth’s crust.
48 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

Migdisov and Williams-Jones (2008) have studied the forma- H2S(aq) þ HS). Henry’s law relates the partial pressure of
tion of Nd(III), Sm(III), and Er(III) sulfate complexes up to the gas species over the liquid to its concentration in the
250  C (i.e., (REE)(SO4)þ and (REE)(SO4)2) and concluded liquid phase and, via the Henry’s law constant, enables the
that these stable complexes would be the principal REE species distribution of the molecular species to be calculated from its
in natural hydrothermal solutions in the absence of elevated thermodynamic properties. Salting-out effects may give rise
concentrations of other complexing ligands, such as fluoride to induced phase separation (e.g., Duan and Li, 2008;
and chloride. Suleimenov and Krupp, 1994), and this calculation requires
that the liquid should be at the same temperature as the gas.
13.2.4.1.8 Other complexing ligands During the past three decades, a large body of experimental
Metal ammine complexes are known at ambient temperature data has been gathered on the solubility of acid gases in aque-
for many elements, but the stability and stoichiometry of metal ous solutions, and these have been used to develop models
ammino-complexes in hydrothermal solutions are essentially that describe the behavior of some of these gases in the two-
unstudied, with the exception of the Cu(II) ammine com- phase region (e.g., CO2, CH4, and H2S) (Dubessy et al., 2005;
plexes, Cu(NH3)n2þ (1  n  4), in Trevani et al. (2001), who Majer et al., 2008). Because of nonideal behavior, accurate
investigated (visible spectroscopy) their formation up to prediction of the partitioning of gas species between vapor
150  C. Amminosilver(I) and amminogold(I) complexes are and liquid necessitates that these models also employ equa-
known at ambient temperature (Perera and Senanayake, 2004; tions of state for the gas phase (e.g., Stryjek and Vera, 1986)
Skibsted and Bjerrum, 1974), but there are no studies of the and activity models for the liquid solutions (e.g., Helgeson
stability and stoichiometry of these potentially important, ore- et al., 1981), in addition to the Henry’s law and ionization
transporting species under hydrothermal conditions. constants referred to earlier.
Deviations from ideal behavior in aqueous vapor are due
13.2.4.1.9 Ore fluids with gas-like density largely to the ability of water molecules to form van der Waals
Hydrothermal fluids, such as those of magmatic hydrothermal and hydrogen bonds, which, as shown experimentally by Liu
systems, are commonly supercritical and may vary in density and Cruzan (1996), leads to the formation of weakly bound
from low (gas-like) to high (liquid-like), depending on the gas-phase clusters, even in pure water vapor. In addition,
level of emplacement of the hydrothermal system. On cooling, solutes in water vapor or supercritical, low-density steam do
these fluids generally separate into vapor and liquid by con- not exist as ‘anhydrous’ molecular moieties but rather as hy-
densation or boiling, although the gas-like fluid may contract drated clusters. For example, calculations by Lemke and
to a liquid and the liquid-like fluid may expand to a vapor Seward (2008a) showed that even at the conditions of the
(Heinrich, 2005). This separation of formerly supercritical hy- Earth’s atmosphere, mixtures of H2O, CO2, and N2O contain
drothermal fluids into vapor and liquid leads to partitioning of ppm levels (molar) of hydrated CO2(H2O)n, N2O(H2O)n, and
the components of the original fluid between these two phases. H2O(H2O)n clusters (n  4). Such relatively modest cluster
Volatile components, such as acidic gases (e.g., CO2, H2S, and formation is also typical of mixtures of water vapor with highly
HCl), partition preferentially into the vapor, whereas compo- volatile compounds. For example, no significant contribution
nents with low volatility, such as NaCl and KCl, partition of hydrated species to arsenic solubility was detected in the
preferentially into the liquid. This does not mean, however, system As(OH)3–H2O (Pokrovski et al., 2002) and to boron
that low-volatility components are absent from the vapor. On solubility in the system B(OH)3–H2O (Kukuljan et al., 1999),
the contrary, salts like NaCl (e.g., Armellini and Tester, 1993; although the well-known solubility of both arsenic and boron
Bischoff et al., 1986) and NH4Cl (Palmer and Simonson, in steam suggests the formation of hydrate clusters.
1993) dissolve in appreciable concentrations in water vapor In contrast to volatile compounds, water clusters contribute
(depending on density), although in much lower concentra- significantly to the solubility of weakly volatile species in water
tions than in liquid water. Nonetheless, there are important vapor. For example, the solubility of NaCl in water vapor has
differences in the nature of the species dissolved in aqueous been shown experimentally to be many orders of magnitude
vapor and liquid. higher than that calculated assuming the absence of hydrated
As discussed earlier, ionic species, including simple and NaCl gas species (e.g., Armellini and Tester, 1993; Bischoff
complex ions, are important in aqueous liquids and com- et al., 1986; Suleimenov et al., 2006). Suleimenov et al.
monly predominate. By contrast, aqueous vapor is dominated (2006) showed that the solubility of NaCl in supercritical
by molecular, uncharged species. For example, the nature of ‘steam’ at 450  C and up 350 bar is due primarily to an assem-
the predominant reduced sulfur species in aqueous liquids blage of the three even-numbered clusters, NaCl(H2O)n, where
depends on pH and can be either H2S(aq) or HS, whereas in n ¼ 4, 6, and 8. Thus, NaCl dissolves in water vapor predomi-
the vapor, the predominant form is molecular H2S. The frac- nantly as NaCl(H2O)n clusters, and the contribution of unhy-
tionation of solution components between aqueous liquids drated NaCl gas species to the NaCl solubility is insignificant.
and vapor therefore varies, even at constant temperature and Based on these experimental data, a set of models describing
pressure, and depends strongly on the properties of the aque- the formation of these clusters and their stability has been
ous liquid, namely, the degree of ionization of the dissolved proposed (Pitzer and Pabalan, 1986; Suleimenov et al.,
components. As a result, the fractionation cannot be described 2006). Similar behavior has been demonstrated for KCl
by a simple partition coefficient. However, it can be predicted (Hovey et al., 1990) and for a set of chlorides and oxides of
by using Henry’s law to determine the distribution of the economically important metals, that is, Cu, Ag, Au, Sn, and Mo
molecular species between gas and liquid, and ionization con- (Figure 12; Archibald et al., 2001, 2002; Migdisov and
stants to determine the proportion of the component of inter- Williams-Jones, 2005; Migdisov et al., 1999; Rempel et al.,
est that is in molecular form in the liquid (e.g., H2S0 relative to 2006; Zezin et al., 2011).
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 49

AgCl–HCl–H2O 100
100 ppb
12.0 r = 0.2 g cm-3
360 ⬚C 75

Au in Vapor (ppm)
10.0
X AgCl (10-9)

50
8.0 AgCl + H2O =
50 ppb AgCl:(H2O)
6.0
25
Ideal solubility
4.0 0
330 ⬚C 400 450 500 550 600 650 700
2.0
10 ppb (a) T (⬚C)
0.0
6.0
0 50 100 150 200
pH2O (bar) 5.0

Au in Vapor (ppm)
Figure 12 Experimentally determined concentration of AgCl in vapor as
a function of pH2O. Ideal solubility refers to the vapor pressure of AgCl 4.0 Au + HCl + H2O =
over the corresponding solid. Modified from Migdisov AA, Williams- AuCl:(H2O) + 0.5H2
Jones AE, and Wagner T (2009) An experimental study of the solubility 3.0
and speciation of rare earth(III) elements in fluoride and chloride bearing r = 0.2 g cm-3
aqueous solutions at temperatures up to 300 C. Geochimica et 2.0 HCl = 0.1 vol%
Cosmochimica Acta 73: 7087–7109. Ni/NiO buffer
1.0
These findings have important implications for theories of
ore formation. It is known, for example, that the hydrothermal 0.0
systems responsible for the formation of porphyry deposits are 400 500 600 700 800
dominated by vapor, rather than liquid (Henley and McNabb, (b) T (⬚C)
1978). There is also compelling evidence that the alteration
associated with the formation of high-sulfidation epithermal Figure 13 The predicted concentration of (a) Ag and (b) Au in aqueous
fluid of density 0.2 g cm3 as a function of temperature. The fluid in
Au–Ag deposits is caused by extremely acidic fluids that prob-
(a) was saturated with AgCl and in (b) with elemental gold; the fO2 of the
ably can only be produced by the condensation of acidic latter system, which contained 0.1 vol% HCl, was buffered by the Ni/NiO
magmatic vapors. Consequently, it is attractive to postulate, buffer. Modified from Migdisov AA and Williams-Jones AE (2013)
as some researchers have, that the ore-forming fluid for these A predictive model for metal transport of silver chloride by aqueous vapor
deposits was a supercritical, low-density fluid that evolved by in ore-forming magmatic-hydrothermal systems. Geochimica et
condensation and/or contraction to a fluid of liquid-like den- Cosmochimica Acta 104: 123–135.
sity (Williams-Jones and Heinrich, 2005). Calculations of the
solubility of Cu and Au in vapor as CuCl(H2O)n and AuCl Wahrenberger et al., 2002) and in high-temperature fumarole
(H2O)n clusters at 360  C and pressures up to 200 bar reach sublimates (e.g., Chaplygin et al., 2005); however, the molecular
levels of 102.8 and 106.5 m, respectively (Archibald et al., basis for its gas-phase transport and volatility is poorly
2001, 2002). Such concentrations are high enough to form a known. Recent FT-ICR/MS (Fourier transform ion cyclotron res-
hypothetical 50 million tonne porphyry copper deposit with a onance mass spectrometry) measurements combined with ab
grade of 0.5% Cu in 20 500–56 000 years (depending on initio quantum chemical calculations have identified an exten-
composition of the fluid) or a 36 tonne Au deposit in sive array of gas-phase cadmium(II) chloride–water clusters
167 000–238 000 years. However, the experimentally deter- (Lemke and Seward, 2013), such as (CdCl)þ(H2O)n (n ¼ 1–4).
mined concentrations are considerably lower than those mea- Various dimers, trimers (with respect to cadmium), and higher-
sured in vapor inclusions, albeit at much higher temperature order polynuclear species, such as (Cd4Cl7)þ(H2O) as well as
and pressure (Ulrich et al., 1999). This apparent disagreement larger, doubly charged clusters, were also identified mass
has recently been resolved as a result of the discovery that spectrometrically.
the solubility of metals in chloride-bearing aqueous fluids in- As noted earlier in the case of volatile Cd chloride–water
creases exponentially with the density of the fluid, due to the clusters, the dominance of molecular solubility in low-density
formation of increasingly larger clusters of water molecules fluids does not exclude the formation of clusters containing
around the metal, and that the free energy of cluster formation ionized species in the vapor phase, especially in the super-
varies linearly with reciprocal temperature (Migdisov and critical region. Indeed, proton–water clusters have been
Williams-Jones, 2013). Using these observations, Migdisov detected over boiling water and in supercritical steam using
and Williams-Jones (2013) have extrapolated the solubility infrared spectroscopy (Carlon, 1979), nuclear magnetic reso-
data for silver and gold and shown that they reach the ppm nance (Matubayasi et al., 1997), and high-pressure mass spec-
levels measured in vapor inclusions at the near-magmatic con- trometry (Likholyot et al., 2007). Clusters containing ionized
ditions of their entrapment (Figure 13). halide ions (Cl, Br, and I) have also been detected experi-
Cadmium exhibits volatility in high-temperature systems mentally at temperatures from 45 to þ175  C (Likholyot
and occurs in volcanic gases (e.g., Symonds et al., 1987; et al., 2005). Recent studies based on ab initio calculations
50 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

have reported standard thermodynamic properties for such transport and ore mineral deposition (i.e., ore formation) is
clusters and predicted them for more complex aggregates, still fraught with inadequacy because of the lack of data per-
such as H3Oþ(H2O)m(H2S)n, NH4þ(H2O)m(H2S)n, and H3Sþ taining to metal complex stoichiometry and stabilities at ele-
(H2O)m(H2S)n, where m  6 and n  4 (Lemke and Seward, vated temperatures and pressures. At temperatures up to about
2008b). 300  C and at moderate pressures not far removed from satu-
Although there is compelling evidence for the transport of rated vapor pressures, we have some considerable knowledge
metals in water vapor, it seems probable that deposition of of complex equilibria for a few metals such as Fe, Cu, Pb, Zn,
these metals occurs after condensation and/or contraction Ag, and Au. For most other metals, the data may pertain to
of the vapor to liquid. Evidence for this is provided by the complexing with only one ligand or derive from a single ex-
blanket-like distribution of gold in high-sulfidation epithermal perimental study at restricted conditions, such as at one ele-
deposits, interpreted to represent a level where the temperature vated temperature and pressure. As summarized in the
of the host rocks is below the dew point of the fluid discussions earlier, information has also been acquired at a
(Chouinard et al., 2005b). In many cases, the mechanisms of molecular level from spectroscopic (X-ray absorption, Raman,
deposition are likely to be the same as those discussed earlier and UV–Vis) and quantum chemical computations. While this
for the liquid. However, as the concentrations of the metals information is fundamental and necessary, high-quality ther-
may be relatively low, it is likely that they may be insufficient to modynamic data with which to model metal transport and
saturate the ore fluid of interest, particularly in relatively low- deposition over the range of fluid compositions, temperature,
pressure environments such as those of some high-sulfidation and pressure occurring in hydrothermal environments
epithermal systems. In these systems, the metals may therefore throughout the Earth’s crust are also required. At present,
concentrate, not by precipitation of an ore mineral of the gold is the only metal for which reliable thermodynamic
metal, but rather by adsorption onto the surface of a gangue data exist over a wide range of temperature and pressure to
mineral like pyrite or other sulfide minerals (Chouinard et al., supercritical conditions(i.e., up to 600  C and 1800 bar) for
2005a), as demonstrated by Renders and Seward (1989), a number of important complexes (i.e., with HS, Cl, and
Cardile et al. (1993), and Widler and Seward (2002). OH). Thus, Heinrich et al. (2004) were able to model gold
Finally, the intriguing but as yet unstudied role of super- transport and deposition associated with phase separation
critical CO2 in the transport of some elements in high- and subsequent magmatic vapor contraction from the mag-
temperature–high-pressure environments in the Earth’s lower matic porphyry to the epithermal environment by combining
crust should be mentioned. Numerous authors have reported the relevant thermodynamic data for the gold(I) complexes
the occurrence of fluid inclusions containing pure CO2 and with phase equilibria in the system NaCl–H2O and fluid
CO2 þ CH4 þ N2 fluids that have been trapped in minerals inclusion data.
formed at granulite and amphibolite facies metamorphic con- A priority must therefore be to acquire high-quality thermo-
ditions (e.g., Andersen et al., 1997; Cuney et al., 2007; dynamic and molecular data on metal complex equilibria per-
Munyanyiwa et al., 1993; Touret, 1971). The question arises tinent to ore transport and deposition processes over wide
as to the role, if any, of supercritical CO2 solvent in element ranges of temperature and pressure. As would be expected, the
transport at high temperatures and pressures in the deep crust. ion pairing of metal complexes with the bulk electrolyte salt
It is well known that molecular (uncharged) metal complexes cations, such as Naþ and Kþ, at near-magmatic supercritical
and metal chelates have appreciable solubility in supercritical conditions will also be of interest, but such species have been
CO2 (e.g., Ashraf-Khorassani et al., 1997; Haruki et al., 2011). little studied. These species (e.g., NaAu(HS)20 or NaHMoO2S20)
What might be the solubility of transition metal carbonyl may play an important role in enhancing metal solubilities and
complexes such as Fe(CO)5 or Ni(CO)4 or even simple ion partitioning into a magmatic volatile-rich phase as noted, for
pairs (e.g., HCl0 and/or NaCl0) in supercritical CO2 at deep example, by Zajacz et al. (2010) and Zhang et al. (2012). This
crustal temperatures and pressures (e.g., 700 to 850  C and up will be a challenging endeavor at high temperatures >350  C
to 15 kbar)? Other molecular components such as TiCl4 and and at pressures up to 5000 bar. The lack of such data comprises
SiCl4 are highly ‘soluble’ in supercritical CO2, and in fact, the major barrier to the further understanding of the formation
the phase equilibria in the binary systems TiCl4–CO2 and of hydrothermal ore deposits as well as to the molecular intri-
SiCl4–CO2 (Suleimenov et al., 2003; Tolley and Tester, 1989) cacies of the chemistry involved.
indicate extensive miscibility between the two end-member
components. For example, in the binary SiCl4–CO2 system at
T > 150  C and P > 125 bar, both components mix/interact Acknowledgments
to give a homogeneous fluid phase throughout the entire
composition range. We thank our colleagues, Chris Heinrich and Steve Scott, for
their helpful suggestions on an earlier version of this chapter.

13.2.5 Epilogue
References
Despite the extensive archive of information on ore deposits,
which includes knowledge of tectonic settings and regional Ahrland S (1968) Thermodynamics of complex formation between hard and soft
acceptors and donors. Structure and Bonding 5: 118–149.
permeabilities, mineralogy, trace element and isotope geo- Anderko A and Pitzer KS (1993) Equation-of-state representation of phase equilibria
chemistry, fluid inclusion compositions, temperatures, and and volumetric properties of the system NaCl-H2O above 573K. Geochimica et
pressures, our understanding of the chemistry of metal Cosmochimica Acta 57: 1657–1680.
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 51

Andersen T, Whitehouse MJ, and Burke EAJ (1997) Fluid inclusions in Scourian Bischoff JL and Rosenbauer R (1988) Liquid–vapor relations in the critical region of the
granulites from the Lewisian complex of NW Scotland: Evidence for CO2-rich fluid system NaCl–H2O from 380 to 415 C: A refined determination of the critical point
in late Archean high grade metamorphism. Lithos 40: 93–104. and two-phase boundary of seawater. Geochimica et Cosmochimica Acta
Anderson AJ, Mayanovic RA, Chou I-M, and Bassett WA (2000) XAFS investigations of 52: 2121–2126.
zinc halide complexes up to supercritical conditions. In: Tremaine PR, Hill PG, Bischoff JL, Rosenbauer R, and Pitzer KS (1986) The system NaCl–H2O: Relations of
Irish DE, and Balakrishnan PV (eds.) Steam, Water and Hydrothermal Systems, vapor–liquid near the critical temperature of water and of vapor–liquid-halite from
pp. 599–606. Ottawa, ON: NRC Research Press. 300 to 500 C. Geochimica et Cosmochimica Acta 50: 1437–1444.
Archibald SM, Migdisov AA, and Williams-Jones AE (2001) The stability of Au-chloride Blount CW (1977) Barite solubilities and thermodynamic quantities up to 300 C and
complexes in water vapor at elevated temperatures and pressures. Geochimica et 1400 bars. American Mineralogist 62: 942–957.
Cosmochimica Acta 65: 4413–4423. Blount CW and Dickson FW (1973) Gypsum–anhydrite equilibria in systems
Archibald SM, Migdisov AA, and Williams-Jones AE (2002) An experimental study of CaSO4–H2O and CaCO3–NaCl–H2O. American Mineralogist 58: 323–331.
the stability of copper chloride complexes in water vapor at elevated temperatures Bodnar RJ, Burnham CW, and Sterner SM (1985) Synthetic fluid inclusions in natural
and pressures. Geochimica et Cosmochimica Acta 66: 1611–1619. quartz. III. Determination of phase equilibrium properties in the system H2O-NaCl to
Armellini FJ and Tester JW (1993) Solubility of sodium chloride and sulfate in sub- and 1000 C and 1500 bars. Geochimica et Cosmochimica Acta 49: 1861–1873.
supercritical water vapor from 450–550 C and 100–250 bar. Fluid Phase Equilibria Boily J-F and Seward TM (2005) Palladium(II) chloride complexation:
84: 123–142. Spectrophotometric investigation in aqueous solutions from 5 to 125 C and
Arnórsson S (1978) Major element chemistry of the geothermal sea-water at Reykjanes theoretical insight into Pd-Cl and Pd-OH2 interactions. Geochimica et
and Svartsengi, Iceland. Mineralogical Magazine 42: 209–220. Cosmochimica Acta 69: 3773–3789.
Arvidson R and Mackenzie FT (1999) The dolomite problem: Control of precipitation Bondarenko GV, Gorbaty YuE, Okhulkov AV, and Kalinichev AG (2006) Structure and
kinetics by temperature and saturation state. American Journal of Science 299: 257–288. hydrogen bonding in liquid and supercritical aqueous NaCl solutions at a pressure
Ashraf-Khorassani M, Combs MT, and Taylor LT (1997) Solubility of metal chelates and of 1000 bar and temperatures up to 500 C: A comprehensive experimental and
their extraction from an aqueous environment via supercritical CO2. Talanta computational study. Journal of Physical Chemistry A 110: 4042–4052.
44: 755–763. Booth HS and Bidwell RM (1950) Solubilities of salts in water at high temperatures.
Audétat A, Gunther D, and Heinrich CA (2000) Causes for large scale metal zonation Journal of the American Chemical Society 72: 2567–2575.
around mineralized plutons: Fluid inclusion LA-ICP-MS evidence from the Mole Borg S, Liu W, Etschmann B, Tian Y, and Brugger J (2012) An XAS study of
Granite, Australia. Economic Geology 95: 1563–1581. molybdenum speciation in hydrothermal chloride solutions from 25–385 C and
Audétat A and Keppler H (2005) Solubility of rutile in subduction zone fluids as 600 bar. Geochimica et Cosmochimica Acta 92: 292–307.
determined by experiments in the hydrothermal diamond anvil cell. Earth and Bostick BC, Fendorf S, and Brown GE Jr. (2005) In-situ analysis of thioarsenite
Planetary Science Letters 232: 393–402. complexes in neutral to alkaline sulfide solutions. Mineralogical Magazine
Audétat A, Pettke T, Heinrich CA, and Bodnar RJ (2008) The composition of 69: 781–795.
magmatic-hydrothermal fluids in barren and mineralized intrusions. Economic Bridger K, Patel RC, and Matijevic E (1981) Temperature dependences of the formation
Geology 103: 877–908. constants of the cobalt(II) acetate complexes. Journal of Inorganic and Nuclear
Baes CF Jr. and Mesmer RE (1976) The Hydrolysis of Cations. New York: Wiley. Chemistry 43: 1011–1016.
Banks DA, Yardley BWD, Campbell AR, and Jarvis KE (1994) REE composition of an Bryce RA, Charnock JM, Patrick RAD, and Lennie AR (2003) EXAFS and density
aqueous magmatic fluid: A fluid inclusion study from the Capitan Pluton, functional study of gold(I) thiosulfate complex in aqueous solution. Journal of
New Mexico, U.S.A. Chemical Geology 113: 259–272. Physical Chemistry A 107: 2516–2523.
Baranova NN and Zotov AV (1998) The stability of gold sulfide species, Au(HS)(aq) and Calabrese S, Aiuppa A, Allard P, et al. (2011) Atmospheric sources and sinks of
Au(HS)2, at 300 and 350 C and 500 bar: Experimental study. Mineralogical volcanogenic elements in a basaltic volcano (Etna, Italy). Geochimica et
Magazine 62A: 116–117. Cosmochimica Acta 75: 7401–7425.
Barnes HL and Seward TM (1997) Geothermal systems and mercury deposits. Caldeira CL, Ciminelli VST, and Osseo-Asare K (2010) The role of carbonate ions
In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 3rd edn., ch.14, in pyrite oxidation in aqueous systems. Geochimica et Cosmochimica Acta
pp. 699–736. New York: Wiley Interscience. 74: 1777–1789.
Bazarkina EF, Pokrovski GS, Zotov AV, and Hazemann J-L (2010) Structure and stability Cardile C, Cashion JD, Renders PJ, and Seward TM (1993) 197Au Mössbauer study of
of cadmium chloride complexes in hydrothermal solutions. Chemical Geology Au2S and gold adsorbed onto As2S3 and Sb2S3 substrates. Geochimica et
276: 1–17. Cosmochimica Acta 57: 2481–2486.
Beak DG, Wilkin RT, Ford RG, and Kelly SD (2008) Examination of arsenic speciation in Carlon H (1979) Ion content and infrared absorption of moist atmospheres. Journal of
sulfidic solutions using X-ray absorption spectroscopy. Environmental Science & the Atmospheric Sciences 36: 832–837.
Technology 42: 1643–1650. Carpenter AB, Trout ML, and Pickett EE (1974) Preliminary report on the origin and
Bebié J, Seward TM, and Hovey JK (1998) Spectrophotometric determination of the chemical evolution of lead- and zinc-rich oil field brines in Central Mississippi.
stability of thallium(I) chloride complexes in aqueous solutions up to 200 C. Economic Geology 69: 1191–1206.
Geochimica et Cosmochimica Acta 62: 1643–1651. Chai L and Navrotsky A (1993) Thermochemistry of carbonate-pyroxene equilibria.
Bell AMT, Charnock JM, Helz GR, et al. (2007) Evidence for dissolved polymeric Contributions to Mineralogy and Petrology 114: 139–147.
mercury(II) sulfur complexes. Chemical Geology 243: 122–127. Chan SH (1989) A review on solubility and polymerization of silica. Geothermics
Bell JLS and Palmer DA (1994) Experimental studies of organic acid decomposition. 18: 49–56.
In: Pittman ED and Lewan MD (eds.) The Role of Organic Acids in Geological Chaplygin IV, Mozgova NN, Magazina LO, et al. (2005) Kudriavite, (Cd,Pb)Bi2S4, a new
Processes, pp. 226–269. Heidelberg: Springer. mineral species from Kudriavy volcano, Iturup Island, Kurile Arc, Russia. The
Bénézeth P, Diakonov II, Pokrovski GS, Dandurand J-L, Schott J, and Khodakovsky IL Canadian Mineralogist 43: 695–701.
(1997) Gallium speciation in aqueous solution. Experimental study and modelling: Chialvo AA and Cummings PT (1994) Hydrogen bonding in supercritical water. Journal
Part 2. Solubility of a-GaOOH in acidic solutions from 150 to 250 C and hydrolysis of Chemical Physics 101: 4466–4469.
constants of gallium(III) to 300 C. Geochimica et Cosmochimica Acta Chialvo AA, Gruszkiewicz MS, Simonson JM, Palmer DA, and Cole DR (2009) Ion pair
61: 1345–1357. association in extreme aqueous environments: Molecular-based and electrical
Bénézeth P and Palmer DA (2000) Potentiometric determination of cadmium-acetate conductance approaches. Journal of Solution Chemistry 38: 827–841.
complexation in aqueous solutions to 250 C. Chemical Geology 167: 11–24. Chou I-M (1987) Phase relations in the system NaCl-KCl-H2O. III: Solubilities of halite
Benning LG and Seward TM (1996) Hydrosulfide complexing of Au(I) in hydrothermal in vapor-saturated liquids above 445 C and redetermination of phase equilibrium
solutions from 150–400 C and 500–1500 bar. Geochimica et Cosmochimica Acta properties in the system NaCl-H2O to 1000 C and 1500 bars. Geochimica et
60: 1849–1871. Cosmochimica Acta 51: 1965–1975.
Bernabei M, Botti A, Bruni F, Ricci MA, and Soper AK (2008) Percolation and Chouinard A, Paquette J, and Williams-Jones AE (2005a) Crystallographic controls on
three-dimensional structure of supercritical water. Physical Review E 78: 021505. trace element incorporation in auriferous pyrite from the Pascua epithermal
Berndt ME, Buttram T, Early D III, and Seyfried WE Jr. (1994) The stability of gold high-sulphidation deposit, Chile–Argentina. The Canadian Mineralogist
polysulfide complexes in aqueous sulfide solutions: 100 to 150 C and 100 bars. 43: 951–963.
Geochimica et Cosmochimica Acta 58: 587–594. Chouinard A, Williams-Jones AE, Leonardson RW, et al. (2005b) Geology and genesis
Beyer RP and Staples BR (1986) Pitzer–Debye–Hückel limiting slopes for water from 0 to of the multistage high-sulfidation epithermal Pascua Au-Ag-Cu deposit, Chile and
350  C and from saturation to 1 kbar. Journal of Solution Chemistry 15: 749–764. Argentina. Economic Geology 100: 463–490.
52 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

Clark JR and Williams-Jones AE (1990) Analogues of epithermal gold-silver deposition Garrels R, Thompson ME, and Siever R (1960) Stability of some carbonates at 25 C and
in geothermal well scales. Nature 346: 644–645. one atmosphere total pressure. American Journal of Science 258: 402–418.
Clarke MB and Helz GR (2000) Metal-thiometallate transport of biologically active trace German CR and Von Damm KL (2003) Hydrothermal processes. In: Holland HD and
elements in sulfidic environments: I. Experimental evidence for copper thioarsenite. Turekian KK (eds.) Treatise on Geochemistry, Vol. 6, pp. 181–222. Oxford:
Environmental Science & Technology 34: 1477–1482. Elsevier.
Cuney M, Coulibaly Y, and Boiron M-C (2007) High density early CO2 fluids in the Gibert F, Pascal M-L, and Pichavant M (1998) Gold solubility and speciation in
ultrahigh-temperature granulites of Ihouhaouene (in Ouzzal, Algeria). Lithos hydrothermal solutions: Experimental study of the stability of hydrosulfide complex
96: 402–414. of gold (AuHS0) at 350 to 450 C and 50 bars. Geochimica et Cosmochimica Acta
Dadze TP, Kaashirtseva GA, and Ryzhenko BN (2000) Gold solubility and species in 62: 2931–2947.
aqueous sulfide solutions at T¼300 C. Geochemistry International 38: 708–712. Giggenbach WF (1968) On the nature of the blue solutions of sulfur. Journal of
Demartin F, Gramaccioli CM, and Campostrini I (2010) Demicheleite-I, a new mineral Inorganic and Nuclear Chemistry 30: 3189–3201.
from La Fossa crater, Vulcano, Aeolian Islands, Italy. Mineralogical Magazine Giggenbach WF (1971a) Optical spectra of highly alkaline sulfide solutions and the
74: 141–145. second dissociation constant of hydrogen sulfide. Inorganic Chemistry
Dong H, Liu W, Doren DJ, and Wood RH (2008) Structure of an accurate ab initio model 10: 1333–1338.
of the aqueous Naþ ion at high temperatures. Journal of Physical Chemistry B Giggenbach WF (1971b) Blue solutions of sulfur in water at elevated temperatures.
112: 13552–13560. Inorganic Chemistry 10: 1306–1308.
Driesner T, Seward TM, and Tironi IG (1998) Molecular dynamics simulation study of Giggenbach WF (1971c) The blue solutions of sulfur in salt melts. Inorganic Chemistry
ionic hydration and ion association in dilute and 1 molal aqueous sodium chloride 10: 1308–1311.
solutions from ambient to supercritical conditions. Geochimica et Cosmochimica Giggenbach WF (1974a) Equilibria involving polysulfide ions in aqueous sulfide
Acta 62: 3095–3107. solutions up to 240 . Inorganic Chemistry 13: 1724–1730.
Duan Z and Li D (2008) Coupled phase and aqueous species equilibrium of the Giggenbach WF (1974b) Kinetics of the polysulfide-thiosulfate disproportionation up to
H2O–CO2–NaCl–CaCO3 system from 0 to 250 C, 1 to 1000 bar with NaCl 240 C. Inorganic Chemistry 13: 1730–1733.
concentrations up to saturation of halite. Geochimica et Cosmochimica Acta Giggenbach WF (1992) Magma degassing and mineral deposition in
72: 5128–5145. hydrothermal systems along convergent plate boundaries. Economic Geology
Duan Z, Møller N, and Weare JH (2003) Equations of state for the NaCl-H2O-CH4 87: 1927–1944.
system and the NaCl-H2O-CO2-CH4 system: Phase equilibria and volumetric Giordano TH and Drummond SE (1991) The potentiometric determination of stability
properties above 573 K. Geochimica et Cosmochimica Acta 67: 671–680. constants for zinc acetate complexes in aqueous solutions to 295 C. Geochimica et
Dubessy J, Tarantola A, and Sterpenich J (2005) Modelling of liquid-vapour equilibria Cosmochimica Acta 55: 359–366.
in the H2O-CO2-NaCl and H2O-H2S-NaCl systems to 270 C. Oil and Gas Science Gorbaty YuE and Kalinichev AG (1995) Hydrogen bonding in supercritical water: 1.
and Technology 60: 339–355. Experimental results. Journal of Physical Chemistry 99: 5336–5340.
Ellis AJ (1963) Solubility of calcite in sodium chloride solutions at high temperatures. Guardia E, Laria D, and Marti J (2006a) Reorientational dynamics of water in aqueous
American Journal of Science 261: 259–267. ionic solutions at supercritical conditions: A computer simulation study. Journal of
Ellis AJ (1979) Explored geothermal systems. In: Barnes HL (ed.) Geochemistry of Molecular Liquids 125: 107–114.
Hydrothermal Ore Deposits, 2nd edn., ch.13, pp. 632–683. New York: Wiley Guardia E, Marti J, and Padro JA (2006b) Ion solvation in aqueous supercritical
Interscience. electrolyte solutions at finite concentrations: A computer simulation study.
Ellis AJ and Giggenbach WF (1971) Hydrogen sulfide ionisation and sulfur hydrolysis Theoretical Chemistry Accounts 15: 161–169.
in high temperature solution. Geochimica et Cosmochimica Acta 35: 247–260. Gunnarsson I and Arnorsson S (2005) Treatment of geothermal waste water to prevent
Ellis AJ and Mahon WAJ (1977) Geochemistry and Geothermal Systems. New York: silica scaling. Proceedings of the World Geothermal Congress, Antalya, Turkey,
Academic Press. 24–29 April 2005.
Erickson BE and Helz GR (2000) Molybdenum(VI) speciation in sulfidic waters: Stability Hall WE and Friedman I (1963) Composition of fluid inclusions, Cave-in-Rock fluorite
and lability of thiomolybdates. Geochimica et Cosmochimica Acta 64: 1149–1158. district, Illinois, and Upper Mississippi Valley zinc-lead district. Economic Geology
Etschmann B, Liu W, Testemale D, et al. (2010) An in situ XAS study of copper(I) 58: 886–911.
transport as hydrosulfide complexes in hydrothermal solutions (25–592 C, Halla VF and van Tassel R (1965) Auflösungserscheinungen bei Erdalkalikarbonaten I.
180–600 bar): Speciation and solubility in vapor and liquid phases. Geochimica et Radex-Rundschau 4: 595–599.
Cosmochimica Acta 74: 4723–4739. Halmer M (2002) The annual volcanic gas input into the atmosphere, in particular into
Fedotova MV (2008) Structural features of concentrated aqueous NaCl solution in the the stratosphere: A global data set for the past 100 years. Journal of Volcanology
sub- and supercritical state at different densities. Journal of Molecular Liquids and Geothermal Research 115: 511–528.
143: 35–41. Hannington MD, de Ronde CEJ, and Petersen S (2005) Sea-floor tectonics and
Fernandez DP, Goodwin ARH, Lemmon EW, Sengers JMHL, and Williams RC (1997) submarine hydrothermal systems. In: Hedenquist JW, Thompson JFH, Goldfarb RJ,
A formulation for the static permittivity of water and steam at temperatures from 238 and Richards JP (eds.) Economic Geology 100th Anniversary Volume,
K to 873 K at pressures up to 1200 MPa, including derivatives and Debye–Hückel pp. 111–141. Littleton, CO: Society of Economic Geologists, Inc.
coefficients. Journal of Physical and Chemical Reference Data 26: 1125–1166. Hardardottir V, Brown KL, Fridriksson T, Hedenquist JW, Hannington MD, and
Foerster F, Lange F, Drossbach O, and Seidel W (1923) Beiträge zur Kenntnis der Thorhallsson S (2009) Metals in deep liquid of the Reykjanes geothermal system,
schwefligen Säure und ihrer Salze. I. Über die Zersetzung der schwefligen southwest Iceland: Implications for the composition of seafloor black smoker fluids.
Säure und ihrer Salze in wäßriger Lösung. Zeitschrift für anorganische und Geology 37: 1103–1106.
allgemeine Chemie 128: 245–342. Harris DJ, Brodholt JP, and Sherman DM (2003) Zinc complexation in hydrothermal
Franck EU (1987) Fluids at high pressures and temperatures. Pure and Applied chloride brines: Results from ab initio molecular dynamics calculations. Journal of
Chemistry 59: 25–34. Physical Chemistry A 107: 1050–1054.
Fulton JL, Hoffmann MM, and Darab JG (2000) An X-ray absorption fine structure study Haruki M, Kobayashi F, Kihara S, and Takishima S (2011) Solubility of b-diketonate
of copper(I) chloride coordination structure in water up to 325 C. Chemical Physics complexes of copper(II) and cobalt(II) in supercritical carbon dioxide. Journal of
Letters 330: 300–308. Chemical and Engineering Data 56: 2230–2235.
Galobardes JF, Hare DR, and Rogers LB (1981) Solubility of sodium chloride in dry Hayden LA and Manning CE (2011) Rutile solubility in supercritical NaAlSi3O8–H2O
steam. Journal of Chemical and Engineering Data 26: 363–366. fluids. Chemical Geology 284: 74–81.
Gammons CH, Wood SA, and Li Y (2002) Complexation of the rare earth elements with Heinrich CA (2005) The physical and chemical evolution of low-salinity magmatic fluids
aqueous chloride at 200 and 300 C and saturated water vapor pressure. at the porphyry to epithermal transition: A thermodynamic study. Mineralium
In: Hellmann R and Wood SA (eds.) Water-Rock Interactions, Ore Deposits and Deposita 39: 864–889.
Environmental Geochemistry. A Tribute to David A. Crerar, Geochemical Society Heinrich CA, Driesner T, Stefansson A, and Seward TM (2004) Magmatic vapor
Special Publication No. 7, pp. 191–207. St. Louis, MO: Geochemical Society. contraction and the transport of gold from the porphyry environment to epithermal
Gammons CH and Yu Y (1997) The stability of silver bromide and iodide complexes at ore deposits. Geology 32: 761–764.
25–300 C: Experiments, theory and geologic applications. Chemical Geology Helgeson HC, Kirkham DH, and Flowers GC (1981) Theoretical prediction of the
137: 155–173. thermodynamic behavior of aqueous electrolytes at high pressures and
Gammons CH, Yu Y, and Williams-Jones AE (1997) The disproportionation of gold(I). temperatures: IV. Calculation of activity coefficients, osmotic coefficients, and
chloride complexes at 25 to 200 C. Geochimica et Cosmochimica Acta apparent molal and standard and relative partial molal properties to 600. American
61: 1971–1983. Journal of Science 281: 1249–1516.
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 53

Hemingway BS and Robie RA (1994) Enthalpy and Gibbs energy of formation of Laitar DS, Müller P, Gray TG, and Sadighi JP (2005) A carbene-stabilized gold(I)
dolomite, CaMg(CO3)2, at 298.15 K from HCl solution calorimetry. US Geological fluoride: Synthesis and theory. Organometallics 24: 4503–4505.
Survey Open-File Report 94–575. Reston, VA: US Geological Survey. Lemke KH and Seward TM (2008a) Ab initio investigation of the structure, stability, and
Hemley JJ, Cygan GL, Fein JB, Robinson GR, and d’Angelo WM (1992) Hydrothermal atmospheric distribution of molecular clusters containing H2O, CO2, and N2O.
ore-forming processes in the light of studies in rock-buffered systems: I. Journal of Geophysical Research 113(D19): 3–4.
Iron-copper-zinc-lead sulfide solubility relations. Economic Geology 87: 1–22. Lemke KH and Seward TM (2008b) Solvation processes in steam: Ab initio calculations
Henley RW and McNabb A (1978) Magmatic vapor plumes and ground-water of ion–solvent structures and clustering equilibria. Geochimica et Cosmochimica
interaction in porphyry copper emplacement. Economic Geology 73: 1–20. Acta 72: 3293–3310.
Hentschel G, Tillmanns E, and Hofmeister W (1985) Hannebachite natural calcium Lemke KH and Seward TM (2013) FT-ICR mass spectrometric and density functional
sulfite hemihydrite CaSO3.½H2O. Neues Jahrbuch für Mineralogie - Monatshefte theory studies of solvated [CdnCl2n–1]þ clusters in low density aqueous flow.
241–250. Chemical Geology (in review).
Hoffmann MM and Conradi MS (1997) Are there hydrogen bonds in supercritical water? Lennie AR, Charnock JM, and Pattrick RAD (2003) Structure of mercury(II)-sulfur
Journal of the American Chemical Society 119: 3811–3817. complexes by EXAFS spectroscopic measurements. Chemical Geology 199: 199–207.
Hoffmann MM, Darab JG, Palmer BJ, and Fulton JL (1999) A transition in the Ni2þ Li D and Duan Z (2007) The speciation equilibrium coupling with phase equilibrium in
complex structure from six- to four-coordinate upon formation of ion pair species in the H2O–CO2–NaCl system from 0 to 250 C, from 0 to 1000 bar, and from 0 to 5
supercritical water: An X-ray absorption fine structure, near-infrared, and molecular molality of NaCl. Chemical Geology 244: 730–751.
dynamics study. Journal of Physical Chemistry A 103: 8471–8482. Likholyot A, Hovey JK, and Seward TM (2005) Experimental and theoretical study of
Holland TBJ and Powell R (1998) An internally consistent data set for phases of hydration of halide ions. Geochimica et Cosmochimica Acta 69: 2949–2958.
petrological interest. Journal of Metamorphic Geology 16: 309–343. Likholyot A, Lemke KH, Hovey JK, and Seward TM (2007) Mass spectrometric and
Hovey JK, Pitzer KS, Tanger JC, Bischoff JL, and Rosenbauer RJ (1990) Vapor–liquid quantum chemical determination of proton water clustering equilibria. Geochimica
phase equilibria of potassium chloride–water mixtures: Equation-of-state et Cosmochimica Acta 71: 2436–2447.
representation for KCl–H2O and NaCl–H2O. Journal of Physical Chemistry Liu K and Cruzan J (1996) Water clusters. Science 271: 929–933.
94: 1175–1179. Liu W, Borg S, Etschmann B, Mei Y, and Brugger J (2012) An XAS study of speciation
Iler RK (1979) The Chemistry of Silica – Solubility, Polymerization, Colloid and Surface and thermodynamic properties of aqueous zinc bromide complexes at 25–150 C.
Properties and Biochemistry. New York: Wiley. Chemical Geology 298–299: 57–69.
Jay JA, Morel FMM, and Hemond HF (2000) Mercury speciation in the presence of Liu W, Borg SJ, Testemale D, Etschmann B, Hazemann J-L, and Brugger J (2011)
polysulfides. Environmental Science & Technology 34: 2196–2200. Speciation and thermodynamic properties for cobalt chloride complexes in
Kaasalainen H and Stefansson A (2011) Sulfur speciation in natural hydrothermal hydrothermal fluids at 35–440 C and 600 bar: An in-situ XAS study. Geochimica et
waters, Iceland. Geochimica et Cosmochimica Acta 75: 2777–2791. Cosmochimica Acta 75: 1227–1248.
Kakovski IA and Tyurin NG (1962) Solubility of gold in polysulfide solutions at elevated Liu W, Brugger J, Etschmann B, Testemale D, and Hazemann J-L (2008) The solubility
temperatures and pressures. Izvestiya Vysshikh Uchebnykh Zavedenii-Tsvetnaya of nantokite (CuCl(s)) and Cu speciation in low-density fluids near the critical
Metallurgiya 2: 104–111. isochore: An in-situ XAS study. Geochimica et Cosmochimica Acta 72: 4094–4106.
Kalinichev AG and Churakov SV (1999) Size and topology of molecular clusters in Liu W, Brugger J, McPhail DC, and Spiccia L (2002) A spectrophotometric study of
supercritical water: A molecular dynamics simulation. Chemical Physics Letters aqueous copper(I) chloride complexes in LiCl solutions between 100 and 250 C.
302: 411–417. Geochimica et Cosmochimica Acta 65: 2937–2948.
Kalinichev AG, Gorbaty Yu E, and Okhulkov AV (1999) Structure and hydrogen bonding Liu W, Etschmann B, Foran G, Shelley M, and Brugger J (2007) Deriving formation
of liquid water at hydrostatic pressures: Monte Carlo NPT-ensemble simulations up constants for aqueous metal complexes from XANES spectra: Zn2þ and Fe2þ
to 10 kilobars. Journal of Molecular Liquids 82: 57–92. chloride complexes in hypersaline solutions. American Mineralogist 92: 761–770.
Kharaka YK, Maest AS, Carothers WW, Law LM, Lamothe PJ, and Fries TA (1987) Liu W, McPhail DC, and Brugger J (2001) An experimental study of copper(I)-chloride
Geochemistry of metal-rich brines from central Mississippi Salt Dome basin, U.S.A. and copper(I)-acetate complexing in hydrothermal solutions between 50 and 250 C
Applied Geochemistry 2: 543–561. and vapour saturated pressure. Geochimica et Cosmochimica Acta 65: 2937–2948.
Klemm LM, Pettke T, Heinrich CA, and Campos E (2007) Hydrothermal evolution of the Lizarralde D, Soule SA, Seewald JS, and Proskurowski G (2010) Carbon release by
El Teniente deposit, Chile: Porphyry Cu-Mo ore deposition from low-salinity off-axis magmatism in a young sedimented spreading centre. Nature Geoscience
magmatic fluids. Economic Geology 102: 1021. 4: 50–54.
Knauss KG, Dibley MJ, Bourcier WL, and Shaw HF (2001) Ti(IV) hydrolysis constants Loucks RR and Mavrogenes JA (1999) Gold solubility in supercritical hydrothermal
derived from rutile solubility measurements made from 100 to 300 C. Applied brines measured in synthetic fluid inclusions. Science 284: 2159–2163.
Geochemistry 16: 1115–1128. Lowell RP, Rona PA, and von Herzen RP (1995) Seafloor hydrothermal systems. Journal
Kovalenko NI and Ryzhenko BN (1997) Stability constants of SnFCl0 at 500 C, of Geophysical Research 100: 327–352.
1kbar and a constant fugacity of H2 (Ni/NiO buffer). Geochemistry International Lüdemann H-D and Franck EU (1967) Absorptionsspektren bei hohen Drucken und
35: 766–769. Temperaturen: I. Wässrige Co(II)- und Ni(II)-halogenide-Lösungen bis zu 500 C
Kovalenko NI, Ryzhenko BN, Dorfeyeva VA, and Bannykh LN (1992) The stability of Sn und 6 kbar. Berichte der Bunsengesellschaft für physikalische Chemie 71: 455–460.
(OH)42-, Sn(OH)2F- and Sn(OH)2Cl- at 500 C and 1 kbar. Geochemistry Madekufamba M and Tremaine PR (2011) Ion association in dilute aqueous magnesium
International 29: 84–94. and nickel sulfate solutions under hydrothermal conditions by flow conductivity
Kovalenko NI, Ryzhenko BN, Prisyagina NI, and Bychkova Ya V (2011) Experimental measurements. Journal of Chemical and Engineering Data 56: 889–898.
study of uraninite solubility in aqueous HCl solutions at 500 C and 1 kbar. Majer V, Sedlbauer J, and Bergin G (2008) Henry’s law constant and related coefficients
Geochemistry International 49: 277–289. for aqueous hydrocarbons, CO2 and H2S over a wide range of temperature and
Kovalenko NI, Ryzhenko BN, Prisyagina NI, and Bychkova Ya V (2012) Experimental pressure. Fluid Phase Equilibria 272: 65–74.
determination of uranium(IV) speciation in HF solutions at 500 C and 1000 bar. Manning CE, Wilke M, Schmidt C, and Cauzid J (2008) Rutile solubility in albite-H2O
Geochemistry International 50: 21–29. and Na2Si3O7-H2O at high temperatures and pressures by in-situ synchrotron
Kramer JR (1959) Correction of some earlier data on calcite and dolomite in sea water. radiation micro-XRF. Earth and Planetary Science Letters 272: 730–737.
Journal of Sedimentary Petrology 29: 465–467. Marshall WL and Franck EU (1981) Ion product of water substance, 0-1000 C,
Kron I, Marshall SL, May PM, Hefter G, and Konigsberger E (1995) The ionic product of 1–10,000 bar. New international formulation and its background. Journal of
water in highly concentrated aqueous electrolyte solutions. Monatshefte für Chemie Physical and Chemical Reference Data 10: 295–304.
126: 819–837. Matubayasi N, Wakai C, and Nakahara M (1997) Structural study of supercritical water:
Krupp RE (1988) Solubility of stibnite in hydrogen sulphide solutions, speciation I. Nuclear magnetic resonance spectroscopy. Journal of Chemical Physics
and equilibrium constants from 25 to 350 C. Geochimica et Cosmochimica Acta 107: 9133.
52: 3005–3015. Mayanovic RA, Anderson AJ, Bassett WA, and Chou I-M (2007) On the formation and
Krupp RE and Seward TM (1987) The Rotokawa geothermal system, New Zealand: An structure of rare earth element complexes in aqueous solutions under hydrothermal
active epithermal gold depositing environment. Economic Geology 82: 1109–1129. conditions with new data on gadolinium aqua and chloro complexes. Chemical
Krupp RE and Seward TM (1990) Transport and deposition of metals in the Rotokawa Geology 239: 266–283.
geothermal system, New Zealand. Mineralium Deposita 25: 73–81. Mayanovic RA, Anderson AJ, Bassett WA, and Chou I-M (2009) The structure and
Kukuljan JA, Alvarez JL, and Fernandez-Prini R (1999) Distribution of B(OH)3 between stability of aqueous rare earth elements in hydrothermal fluids: New results on
water and steam at high temperatures. Journal of Chemical Thermodynamics neodymium(III) aqua and chloroaqua complexes in aqueous solutions to 500 C and
31: 1511–1521. 520 MPa. Chemical Geology 259: 30–38.
54 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

Mayanovic RA, Jayanetti S, Anderson AJ, Bassett WA, and Chou I-M (2002) The Munyanyiwa H, Touret JLR, and Jelsma HA (1993) Thermobarometry and fluid
structure of Yb3þ up to 500 C and 270 bar. Journal of Physical Chemistry A evolution of enderbites within the Magondi Mobile Belt, northern Zimbabwe. Lithos
106: 6591–6599. 29: 163–176.
Mendez de Leo LP, Bianchi HL, and Fernandez-Prini R (2005) Ion pair formation Nahtigal IG and Svishchev IM (2009) Molecular dynamics study of ionic nano-clusters
in copper sulfate aqueous solutions at high temperatures. Journal of Chemical produced from supercritical solutions. Journal of Supercritical Fluids 50: 169–175.
Thermodynamics 37: 499–511. Navrotsky A and Capobianco C (1987) Enthalpies of formation of dolomite and of
Mercado S and Hurtado R (1992) Potash extraction from Cerro Prieto geothermal brine. magnesian calcites. American Mineralogist 72: 782–787.
Geothermics 21: 759–764. Newton RC and Manning CE (2000) Quartz solubility in H2O-NaCl and H2O-CO2
Mibe K, Chou I-M, Anderson AJ, Mayanovic RA, and Bassett WA (2009) The speciation solutions at deep crust-upper mantle pressures and temperatures: 2–15 kbar
of aqueous zinc(II) bromide solutions to 500 C and 900 MPa determined using 500–900 C. Geochimica et Cosmochimica Acta 64: 2993–3005.
Raman spectroscopy. Chemical Geology 259: 48–53. Newton RC and Manning CE (2002) Experimental determination of calcite solubility
Migdisov AA and Williams-Jones AE (2002) A spectrophotometric study of neodymium in H2O-NaCl solutions at deep crust/upper mantle pressures and temperatures:
(III) complexation in chloride solutions. Geochimica et Cosmochimica Acta Implications for metasomatic processes. American Mineralogist 87: 1401–1409.
66: 4311–4323. Newton RC and Manning CE (2010) Role of saline fluids in deep-crustal and upper-mantle
Migdisov AA and Williams-Jones AE (2005) An experimental study of cassiterite metasomatism: Insights from experimental studies. Geofluids 10: 58–72.
solubility in HCl-bearing water vapour at temperatures up to 350 C: Implications for Nilsson A and Petterssen LGM (2011) Perspective on the structure of liquid water.
tin ore formation. Chemical Geology 217: 29–40. Chemical Physics 389: 1–34.
Migdisov AA and Williams-Jones AE (2006) A spectrophotometric study of erbium(III). Oelkers EH, Sherman DM, Ragnarsdottir KV, and Collins C (1998) An EXAFS
speciation in chloride solutions at elevated temperatures. Chemical Geology spectroscopic study of aqueous antimony(III) chloride complexation at temperatures
234: 17–27. from 25 to 250 C. Chemical Geology 151: 21–27.
Migdisov AA and Williams-Jones AE (2007) An experimental study of the solubility and Ohmoto H, Hayashi K-I, and Kajisa Y (1994) Experimental study of the solubilities
speciation of neodymium(III) fluoride in F-bearing aqueous solutions. Geochimica of pyrite in NaCl-bearing aqueous solutions at 250–350 C. Geochimica et
et Cosmochimica Acta 71: 3056–3069. Cosmochimica Acta 58: 2169–2185.
Migdisov AA and Williams-Jones AE (2008) A spectrophotometric study of Nd(III), Sm Palmer DA, Corti HR, Grotewold A, and Hyde KE (2000) Potentiometric measurements
(III) and Er(III) complexation in sulfate-bearing solutions at elevated temperatures. of the thermodynamics of cadmium(II) chloride complexation to higher
Geochimica et Cosmochimica Acta 72: 5291–5303. temperatures. In: Tremaine PR, Hill PG, Irish DE, and Balakrishnan PV (eds.) Steam,
Migdisov AA and Williams-Jones AE (2013) A predictive model for metal transport of Water and Hydrothermal Systems, pp. 736–743. Ottawa, ON: NRC Press.
silver chloride by aqueous vapor in ore-forming magmatic-hydrothermal systems. Palmer DA and Drummond SE (1988) Potentiometric formation of the molal formation
Geochimica et Cosmochimica Acta 104: 123–135. constants of ferrous acetate complexes in aqueous solutions to high temperatures.
Migdisov AA, Williams-Jones AE, Lakshtanov L, and Alekhin YV (2002) Estimates of the Journal of Physical Chemistry 92: 6795–6800.
second dissociation constant of H2S from the surface sulfidation of crystalline Palmer DA and Simonson JM (1993) Volatility of ammonium chloride over aqueous
sulfur. Geochimica et Cosmochimica Acta 66: 1713–1725. solutions to high temperatures. Journal of Chemical and Engineering Data 38(3):
Migdisov AA Williams-Jones AE Norman C, and Wood SA (2008) A spectrophotometric 465–474.
study of samarium(III) in chloride solutions at elevated temperatures. Geochimica et Palmer DA, Simonson JM, and Jensen JP (2004) Partitioning of electrolytes to steam
Cosmochimica Acta 72: 1611–1623. and their solubilities in steam. In: Palmer DA, Fernandez-Prini R, and Harvey AH
Migdisov AA, Williams-Jones AE, and Suleimenov OM (1999) Solubility of (eds.) Aqueous Systems at Elevated Temperatures and Pressures: Physical
chlorargyrite (AgCl) in water vapor at elevated temperatures and pressures. Chemistry in Water, Steam and Hydrothermal Solutions, ch. 12, pp. 409–439.
Geochimica et Cosmochimica Acta 63: 3817–3827. Amsterdam: Elsevier.
Migdisov AA, Williams-Jones AE, van Hinsberg V, and Salvi S (2011a) An experimental Paquette K and Helz GR (1997) Inorganic speciation of mercury in sulfidic waters: The
study of the solubility of baddeleyite (ZrO2) in fluoride-bearing solutions at elevated importance of zero valent sulphur. Environmental Science & Technology
temperature. Geochimica et Cosmochimica Acta 75: 7426–7434. 31: 2148–2153.
Migdisov AA, Williams-Jones AE, and Wagner T (2009) An experimental study of the Patel N (2001) Calculation of vapor–liquid equilibria for a 10-component system:
solubility and speciation of the rare earth elements(III) in fluoride- and chloride- Comparison of EOS, EOS–GE and GE–Henry’s law models. Fluid Phase Equilibria
bearing aqueous solutions at temperatures up to 300 C. Geochimica et 185: 397–405.
Cosmochimica Acta 73: 7087–7109. Pearson RG (1963) Hard and soft acids and bases. Journal of the American Chemical
Migdisov AA, Zezin D, and Williams-Jones AE (2011b) An experimental study of cobalt Society 85: 3533–3539.
(II) complexation in Cl and H2S-bearing hydrothermal solutions. Geochimica et Perera WN and Senanayake G (2004) The ammine, thiosulfato and mixed ammine/
Cosmochimica Acta 75: 4065–4079. thiosulfato complexes of silver(I) and gold(I). Inorganic Chemistry 43: 3048–3056.
Minubayeva Z and Seward TM (2010) Molybdic acid ionisation under hydrothermal Perera WN, Senanayake G, and Nicol MJ (2005) Interaction of gold(I) with thiosulfate-
conditions to 300 C. Geochimica et Cosmochimica Acta 74: 4365–4374. sulfite mixed ligand systems. Inorganica Chimica Acta 358: 2183–2190.
Minubayeva Z and Seward TM (2013) A UV spectrophotometric study of tungstic acid Peshchevitsky BI, Yerenburg AM, Belavantsev BI, and Kazakov VP (1970) Stability of
deprotonation to 300 C at saturated vapour pressures. Chemical Geology gold complexes in aqueous solutions. Izvestiya Sibirskogo Otdeleniya Akademii
(in review). Nauk Seriya Khimicheskikh 4: 75–83.
Mizoguchi T, Takei Y, and Okabe T (1976) The chemical behaviour of low valence sulfur Pitzer KS and Pabalan R (1986) Thermodynamics of NaCl in steam. Geochimica et
compounds. X. Disproportionation of thiosulfate, trithionate, tetrathionate and Cosmochimica Acta 50: 1445–1454.
sulfite under acidic conditions. Bulletin of the Chemical Society of Japan 49: 70–75. Planer-Friedrich B and Scheinost AC (2011) Formation and structural characterization
Mohr F (2004) The chemistry of gold-fluoro-compounds: A continuing challenge for of thioantimony species and their natural occurrence in geothermal waters.
gold chemists. Gold Bulletin 37: 164–169. Environmental Science & Technology 45: 6855–6863.
Møller N (1988) The prediction of mineral solubilities in natural waters: A chemical Planer-Friedrich B, Suess E, Scheinost AG, and Wallschläger D (2010) Arsenic
equilibrium model for the Na-Ca-Cl-SO4-H2O system, to high temperature and speciation in sulfidic waters: Reconciling contradictory spectroscopic and
concentration. Geochimica et Cosmochimica Acta 52: 821–837. chromatographic evidence. Analytical Chemistry 82: 10228–10235.
Monnin C (1990) The influence of pressure on the activity coefficients of the solutes and Pokrovski GS, Borisova AYu, Roux J, et al. (2006) Antimony speciation in saline
on the solubility of minerals in the system Na-Ca-Cl-SO4-H2O to 200 C and 1 kbar hydrothermal fluids: A combined X-ray absorption fine structure spectroscopy and
and to high NaCl concentration. Geochimica et Cosmochimica Acta 54: 3265–3282. solubility study. Geochimica et Cosmochimica Acta 70: 4196–4214.
Mountain BW and Seward TM (1999) The hydrosulphide/sulphide complexes of Pokrovski GS and Dubrovinsky LS (2011) The S3 ion is stable in geological fluids at
copper(I): Experimental determination of stoichiometry and stability at 22 C and elevated temperatures and pressures. Science 331: 1052–1054.
reassessment of high temperature data. Geochimica et Cosmochimica Acta 63: 11–29. Pokrovski GS, Tagirov BR, Schott J, Hazemann J-L, and Proux O (2009) A new view of
Mountain BW and Seward TM (2003) The hydrosulfide/sulfide complexes of copper gold speciation in sulfur-bearing hydrothermal fluids from in-situ X-ray absorption
(I): Experimental confirmation of the stoichiometry and stability of Cu(HS)2 to spectroscopy and quantum chemical modelling. Geochimica et Cosmochimica Acta
elevated temperatures. Geochimica et Cosmochimica Acta 67: 3005–3014. 73: 5406–5427.
Müller B and Seward TM (2001) Spectrophotometric determination of the stability of tin Pokrovski G, Zakirov I, and Roux J (2002) Experimental study of arsenic speciation in
(II) chloride complexes in aqueous solution up to 300 C. Geochimica et vapor phase to 500 C: Implications for As transport and fractionation in low-density
Cosmochimica Acta 65: 4187–4199. crustal fluids and volcanic gases. Geochimica et Cosmochimica Acta 66: 3453–3480.
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 55

Pouradier J and Gadet MC (1969) Electrochemistry of gold salts: VII. Thiosulfate. Scott SD (1997) Seafloor hydrothermal systems and deposits. In: Barnes HL (ed.)
Journal de Chimie Physique et de Physico-Chimie Biologique 66: 109–112. Geochemistry of Hydrothermal Ore Deposits, 3rd edn., ch. 16, pp. 797–874.
Prisyagina NI, Kovalenko NI, Ryzhenko BN, and Starshinova NP (2008) Experimental New York: Wiley Interscience.
determination of ZrO2 solubility in alkali metal fluoride solutions at 500 C and 1000 Seel F, Güttler H-J, Simon G, and Wieckowski A (1977) Colored sulfur species in
bar. Geokhimiya 9: 1234–1237. EPD-solvents. Pure and Applied Chemistry 49: 45–54.
Pryor WA (1960) The kinetics of the disproportionation of sodium thiosulfate to sodium Seward TM (1973) Thiocomplexes of gold and the transport of gold in hydrothermal ore
sulfide and sulfate. Journal of the American Chemical Society 82: 4794–4797. solutions. Geochimica et Cosmochimica Acta 37: 379–399.
Quist AS and Marshall WL (1968) Electrical conductances of aqueous sodium chloride Seward TM (1974a) Determination of the first ionisation constant of silicic acid from
solutions from 0 to 800 degrees and at pressures to 4000 bars. Journal of Physical quartz solubility in borate buffer solutions to 350 C. Geochimica et Cosmochimica
Chemistry 72: 684–704. Acta 38: 1651–1664.
Ragnarsdottir KV, Fournier P, Oelkers EH, and Harrichoury J-C (2001) Experimental Seward TM (1974b) Equilibrium and oxidation potential in geothermal waters at
determination of the complexation of strontium and cesium with acetate in high Broadlands, New Zealand. American Journal of Science 274: 190–192.
temperature aqueous solutions. Geochimica et Cosmochimica Acta Seward TM (1976) The stability of chloride complexes of silver in hydrothermal
107: 3955–3964. solutions up to 350 C. Geochimica et Cosmochimica Acta 40: 1329–1341.
Ragnarsdottir KV, Oelkers EH, Sherman DM, and Collins CR (1998) Aqueous speciation Seward TM (1981) Metal complex formation in aqueous solutions at elevated
of yttrium at temperatures from 25 to 340 C at P-sat: An in-situ EXAFS study. temperatures and pressures. In: Rickard DT and Wickman FE (eds.) Chemistry and
Chemical Geology 151: 29–39. Geochemistry of Solutions at High Temperatures and Pressures, Physics and
Raju KUG and Atkinson G (1990) The thermodynamics of ‘scale’ mineral solubilities: 3. Chemistry of the Earth, vol. 13 and 14, pp. 113–132. New York: Pergamon Press.
Calcium sulfate in aqueous sodium chloride. Journal of Chemical and Engineering Seward TM (1982) The transport and deposition of gold in hydrothermal systems.
Data 35: 361–367. In: Foster RP (ed.) Gold 82: The Geology, Geochemistry and Genesis of Gold
Rapp JF, Klemme S, Butler IB, and Harley SL (2010) Extremely high solubility of rutile Deposits, pp. 165–181. Amsterdam: Balkema.
in chloride and fluoride-bearing metamorphic fluids: An experimental investigation. Seward TM (1984) The formation of lead(II) chloride complexes to 300 C: A
Geology 38: 323–326. spectrophotometric study. Geochimica et Cosmochimica Acta 48: 121–134.
Reeves EP, Seewald JS, Saccocia P, et al. (2011) Geochemistry of hydrothermal fluids Seward TM (1989) The hydrothermal chemistry of gold and its implication for ore
from the PACMANUS, Northeast Pual and Vienna Woods hydrothermal fields, formation: Boiling and conductive cooling as examples. Economic Geology
Manus Basin, Papua New Guinea. Geochimica et Cosmochimica Acta Monographs 6: 398–404.
75: 1088–1123. Seward TM and Barnes HL (1997) Metal transport in hydrothermal ore fluids.
Rempel KU, Migdisov AA, and Williams-Jones AE (2006) The solubility and In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 3rd edn., ch. 9,
speciation of molybdenum in water vapour at elevated temperatures and pp. 435–486. New York: Wiley Interscience.
pressures: Implications for ore genesis. Geochimica et Cosmochimica Acta Seward TM and Driesner T (2004) Hydrothermal solution structure: Experiments and
70: 687–696. computer simulations. In: Palmer DA, Fernandez-Prini R, and Harvey AH (eds.)
Renders PJ and Seward TM (1989a) The stability of hydrosulphido- and sulphido- Aqueous Systems at Elevated Temperatures and Pressures: Physical Chemistry in
complexes of Au(I) and Ag(I) at 25 C. Geochimica et Cosmochimica Acta Water, Steam and Hydrothermal Solutions, ch. 5, pp. 149–182. Amsterdam:
53: 245–253. Elsevier.
Renders PJ and Seward TM (1989b) The adsorption of thio gold(I) complexes by Seward TM, Henderson CMB, and Charnock JM (2000) In(III) chloride complexing and
amorphous As2S3 and Sb2S3 at 25 C and 90 C. Geochimica et Cosmochimica Acta solvation of In3þ in hydrothermal solutions to 350 C: An EXAFS study. Chemical
53: 255–267. Geology 167: 117–127.
Rimstidt J (2003) Pyrite oxidation: A state-of-the-art assessment of the reaction Seward TM, Henderson CMB, Charnock JM, and Dobson BR (1996) An X-ray
mechanism. Geochimica et Cosmochimica Acta 67: 873–880. absorption. (EXAFS) spectroscopic study of aquated Agþ in hydrothermal solutions
Rothbaum HP and Rhode AG (1979) Kinetics of silica polymerization and deposition to 350 C. Geochimica et Cosmochimica Acta 60: 2273–2382.
from dilute solutions between 5 and 180 C. Journal of Colloid and Interface Science Seward TM, Henderson CMB, Charnock JM, and Driesner T (1999) An EXAFS study of
71: 533–559. solvation and ion pairing in strontium chloride solutions to 330 C. Geochimica et
Ruaya JR and Seward TM (1986) The stability of chloro-zinc(II) complexes in Cosmochimica Acta 63: 2409–2418.
hydrothermal solutions up to 350 C. Geochimica et Cosmochimica Acta Seward TM, Lemke KH, Henderson CMB, and Charnock JM (2013) An X-ray absorption
50: 651–662. spectroscopic and ab initio computational study of the Cd(II) aquated ion and
Rudolph WW, Brooker MH, and Tremaine PR (1997) Raman spectroscopic chlorocadmium(II) complexing in hydrothermal solutions. Chemical Geology
investigations of aqueous FeSO4 in neutral and acidic solutions from 25 to 303 C: (in review).
Inner and outer sphere complexes. Journal of Solution Chemistry 26: 757–777. Seyfried WE (1987) Experimental and theoretical constraints on hydrothermal alteration
Rudolph WW, Brooker MH, and Tremaine PR (1999) Raman spectroscopy of aqueous processes at mid-ocean ridges. Annual Review of Earth and Planetary Sciences
under hydrothermal conditions: Solubility, hydrolysis and sulphate ion pairing. 15: 317–335.
Journal of Solution Chemistry 28: 621–630. Shanks WC (2001) Stable isotopes in seafloor hydrothermal systems: Vent fluids,
Ryzhenko BN, Kovalenko NI, and Prisyagina NI (2006) Titanium complexation in hydrothermal deposits, hydrothermal alteration, and microbial processes. Reviews
hydrothermal systems. Geochemistry International 44: 879–895. in Mineralogy and Geochemistry 43: 469–525.
Ryzhenko BN, Kovalenko NI, Prisyagina NI, Starshinova NP, and Krupskaya VV (2008) Shenberger DM and Barnes HL (1989) Solubility of gold in aqueous sulfide solutions
Experimental determination of zirconium speciation in hydrothermal solutions. from 150 to 350 C. Geochimica et Cosmochimica Acta 53: 269–278.
Geochemistry International 46: 328–339. Sherman DM (2007) Complexation of Cuþ in hydrothermal NaCl brines: Ab initio
Samson IM, Williams-Jones AE, Ault KM, Gagnon JE, and Fryer BJ (2008) Source of molecular dynamics and energetics. Geochimica et Cosmochimica Acta
fluids forming distal Zn-Pb-Ag skarns: Evidence from laser ablation–inductively 71: 714–722.
coupled plasma–mass spectrometry analysis of fluid inclusions from El Mochito, Sherman DM (2010) Metal complexation and ion association in hydrothermal fluids:
Honduras. Geology 36: 947. Insights from quantum chemistry and molecular dynamics. Geofluids 10: 41–57.
Saunders J and Swann C (1990) Trace metal content of Mississippi oil field brines. Sherman DM and Collings MD (2002) Ion association in concentrated NaCl brines from
Journal of Geochemical Exploration 37: 171–183. ambient to supercritical conditions: Results from classical molecular dynamics.
Saxena RS, Jain MC, and Mittal ML (1968) Electrometric investigations of an acid- Geochemical Transactions 3: 102–107.
thiomolybdate system and the formation of polyanions. Australian Journal of Sherman DM, Ragnarsdottir KV, and Oelkers EH (2000a) Antimony transport in
Chemistry 21: 91–96. hydrothermal solutions: An EXAFS study of antimony(V) complexation in alkaline
Schoonen MAA and Barnes HL (1988) An approximation of the second dissociation sulfide and sulfide–chloride brines at temperatures from 25 C to 300 C at Psat.
constant for H2S. Geochimica et Cosmochimica Acta 52: 649–654. Chemical Geology 167: 161–167.
Schoonen MAA and Barnes H (1991a) Mechanisms of pyrite and marcasite formation Sherman DM, Ragnarsdottir KV, Oelkers EH, and Collins CR (2000b) Speciation of tin
from solution: III. Hydrothermal processes. Geochimica et Cosmochimica Acta (Sn2þ and Sn4þ) in aqueous Cl solutions from 25 C to 350 C: An in situ EXAFS
55: 3491–3504. study. Chemical Geology 167: 169–176.
Schoonen MAA and Barnes HL (1991b) Reactions forming pyrite and marcasite from Sherman LA and Barak P (2000) Solubility and dissolution kinetics of dolomite in
solution: I. Nucleation of FeS2 below 100 C. Geochimica et Cosmochimica Acta Ca–Mg–HCO3/CO3 solutions at 25  C and 0.1 MPa carbon dioxide. Soil Science
55: 1495–1504. Society of America Journal 64: 1959–1968.
56 The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids

Shmulovich KI, Yardley BWD, and Graham CM (2006) The solubility of quartz in crustal gases at Merapi volcano, Indonesia. Geochimica et Cosmochimica Acta
fluids: Experiments and general equations for salt solutions and H2O-CO2 mixtures 51: 2083–2101.
at 400–800 C and 0.1–0.9 GPa. Geofluids 6: 154–167. Tagirov BR, Baranova NN, Zotov AV, Schott J, and Bannykh LN (2006) Experimental
Shock EL and Koretsky CM (1993) Metal-organic complexes in geochemical processes: determination of the stabilities of Au2S(cr) at 25 C and Au(HS)2 at 25–250 C.
Calculation of standard partial molal thermodynamic properties of aqueous acetate. Geochimica et Cosmochimica Acta 70: 3689–3701.
Geochimica et Cosmochimica Acta 57: 4899–4922. Tagirov BR, Salvi S, Schott J, and Baranova NN (2005) Experimental study of
Shock EL, Sassani DC, Willis M, and Sverjensky DA (1997) Inorganic species in gold-hydrosulfide complexing in aqueous solutions at 350–500 C and 1000 bars
geologic fluids: Correlations among standard molal thermodynamic properties of using mineral buffers. Geochimica et Cosmochimica Acta 69: 2119–2132.
aqueous ions and hydroxide complexes. Geochimica et Cosmochimica Acta Tagirov BR and Seward TM (2010) Hydrosulfide/sulfide complexes of zinc to 250 C
61: 907–950. and the thermodynamic properties of sphalerite. Chemical Geology 269: 301–311.
Simmons SF and Brown KL (2006) Gold in magmatic hydrothermal solutions and the Tagirov BR, Suleimenov OM, and Seward TM (2007a) Zinc complexation in aqueous
rapid formation of a giant ore deposit. Science 314: 288–291. sulfide solution: Determination of the stoichiometry and stability of complexes via
Simmons SF and Brown KL (2007) The flux of gold and related metals through a ZnS(cr) solubility measurements at 100 C and 150 bar. Geochimica et
volcanic arc, Taupo Volcanic Zone, New Zealand. Geology 35: 1099. Cosmochimica Acta 71: 4942–4953.
Simonet V, Calzavara V, Hazemann J-L, Argoud R, Geaymond O, and Raoux D (2002) Tagirov BR, Zotov A, Schott J, Suleimenov OM, and Koroleva L (2007b) A
X-ray absorption spectroscopy studies of ionic association in aqueous solutions of potentiometric study of the stability of aqueous yttrium acetate complexes from 25 to
zinc bromide from normal to critical conditions. Journal of Chemical Physics 175 C and 1–1000 bar. Geochimica et Cosmochimica Acta 71: 1689–1708.
117: 2771–2781. Takano B, Ohsawa S, and Glover RB (1994) Surveillance of Ruapehu crater lake,
Skibsted LH and Bjerrum J (1974) Gold complexes: II. Equilibrium between gold(I) and New Zealand, by aqueous polythionates. Journal of Volcanology and Geothermal
gold(III) in the ammonia system and the standard potentials of the couples involving Research 60: 29–57.
gold diamminegold(I) and tetramminegold(III). Acta Chemica Scandinavica Testemale D, Brugger J, Liu W, Etschmann B, and Hazemann J-L (2009) In-situ X-ray
A28: 764–770. absorption study of iron(II) speciation in brines up to supercritical conditions.
Soper AK (2000) Radial distribution functions of water and ice from 220 to 673K and at Chemical Geology 264: 295–310.
pressures up 4400 MPa. Chemical Physics 258: 121–137. Tian Y, Etschmann B, Liu W, et al. (2012) Speciation of nickel(II) chloride complexes in
Steele-MacInnis M, Han L, Lowell RP, Rimstidt JD, and Bodnar RJ (2012) The role of hydrothermal fluids: In-situ XAS study. Chemical Geology 334: 345–363.
fluid phase immiscibility in quartz dissolution and precipitation in sub-seafloor Tivey M (2005) Generation of seafloor hydrothermal vent fluids and associated mineral
hydrothermal systems. Earth and Planetary Science Letters 321(322): 139–151. deposits. Oceanography 20: 50–65.
Stefansson A, Lemke KH, and Seward TM (2008) Iron(III) complexation in hydrothermal Tolley WK and Tester LS (1989) Supercritical carbon dioxide solubility of titanium
solutions: An experimental and theoretical study. In: Proceedings of ICPWS XV tetrachloride. US Bureau of Mines Report of Investigation 9216.
(15th International Conference on the Properties of Water and Steam), Berlin, 8–11 Tossell JA (2001a) Calculation of the structures, stabilities and properties of mercury
September 2008, geo 01–09. sulphide species in aqueous solution. Journal of Physical Chemistry
Stefansson A and Seward TM (2003a) Experimental determination of the stability and A105: 935–941.
stoichiometry of sulphide complexes of silver(I) in hydrothermal solutions to Tossell JA (2001b) Computing the properties of the copper thioarsenite complex,
400 C. Geochimica et Cosmochimica Acta 67: 1395–1413. CuAsS(SH)(OH). Inorganic Chemistry 40: 6487–6492.
Stefansson A and Seward TM (2003b) The hydrolysis of gold(I) in aqueous solutions to Tossell JA (2012) Calculation of the properties of the S3-radical anion and its
600 C and 1500 bar. Geochimica et Cosmochimica Acta 67: 1677–1688. complexes with Cuþ in aqueous solution. Geochimica et Cosmochimica Acta
Stefansson A and Seward TM (2003c) The stability of chloridogold(I) complexes in 95: 79–92.
aqueous solutions from 300 to 600 C and from 500 to 1800 bar. Geochimica et Touret JLR (1971) Le faciès granulite en Norvège méridionale: II. Les inclusions fluides.
Cosmochimica Acta 67: 4459–4576. Lithos 4: 423–436.
Stefansson A and Seward TM (2004) Gold(I) complexing in aqueous sulphide solutions Trevani L, Ehlerova J, Sedlbauer J, and Tremaine PR (2009) Complexation in the
to 500 C at 500 bar. Geochimica et Cosmochimica Acta 68: 4121–4143. Cu(II)-LiCl-H2O system at temperatures to 423 K by UV-visible spectroscopy.
Stepanchikova SA and Kolonin GR (2005) Spectrophotometric study of Nd, Sm and Ho International Journal of Hydrogen Energy 35: 10905–10926.
complexation in chloride solutions at 100–250 C. Russian Journal of Coordination Trevani LN, Roberts JC, and Tremaine PR (2001) Copper(II) ammonia complexation at
Chemistry 31: 193–202. temperatures from 30 to 250 C by visible spectroscopy. Journal of Solution
Stillinger FH (1980) Water revisited. Science 209: 451–457. Chemistry 30: 585–622.
Stoffell B, Wilkinson JJ, and Jeffries TE (2004) Metal transport and deposition in Tropper P and Manning CE (2007) The solubility of fluorite in H2O and H2O–NaCl at
hydrothermal veins revealed by 213 nm UV laser ablation microanalyses of single high pressure and temperature. Chemical Geology 242: 299–306.
fluid inclusions. American Journal of Science 304: 533–557. Uchida E, Sakamori T, and Matsunaga J (2002) Aqueous speciation of lead and tin
Strübel G (1965) Quantitative Untersuchungen über die hydrothermale Löslichkeit von chlorides in supercritical hydrothermal solutions. Geochemical Journal 36: 61–72.
Flussspat. Geologische Rundschau 3: 83–95. Uematsu M and Franck EU (1980) Static dielectric constant of water and steam. Journal
Stryjek R and Vera JH (1986) PRSV: An improved Peng-Robinson equation of state for of Physical and Chemical Reference Data 9: 1291–1306.
pure compounds and mixtures. Canadian Journal of Chemical Engineering Ulrich T, Günther D, and Heinrich CA (1999) Gold concentrations of magmatic brines
64: 323–333. and the metal budget of porphyry copper deposits. Nature 399: 676–679.
Suleimenov OM, Bastrakov E, Shvarov YV, and Seward TM (2006) Sodium Ulrich T, Gunther D, and Heinrich CA (2002) The evolution of a Cu-Au deposit based on
chloride-water clusters in steam: Classical, quantum chemical and LA-ICP-MS analysis of fluid inclusions: Bajo de la Alumbrera, Argentina. Economic
thermodynamic computational approach. Geochimica et Cosmochimica Acta Geology 97: 1889–1920.
70(supplement): 623. Ulrich T and Mavrogenes J (2008) An experimental study of the solubility of
Suleimenov OM and Krupp RE (1994) Solubility of hydrogen sulfide in pure water and molybdenum in H2O and KCl-H2O solutions from 500 C to 800 C and 150 to 300
in NaCl solutions from 20 to 320 C and at saturation pressures. Geochimica et MPa. Geochimica et Cosmochimica Acta 72: 2316–2330.
Cosmochimica Acta 58: 2433–2444. Usher A, McPhail DC, and Brugger J (2009) A spectrophotometric study of Au(III)
Suleimenov OM, Panagiotopoulos AZ, and Seward TM (2003) Grand canonical Monte halide–hydroxide complexes at 25–80 C. Geochimica et Cosmochimica Acta
Carlo simulations of phase equilibria of pure silicon tetrachloride and its binary 77: 3359–3380.
mixture with carbon dioxide. Molecular Physics 101: 3213–3221. Wahrenberger C, Seward TM, and Dietrich V (2002) Volatile trace element transport in
Suleimenov OM and Seward TM (1997) The ionisation of hydrogen sulphide in high temperature gases from Kudriavy volcano (Iturup, Kuril Island, Russia).
aqueous solution to 350 C. Geochimica et Cosmochimica Acta 61: 5187–5198. In: Hellmann R and Wood SA (eds.) Water-Rock Interaction, Ore Deposits and
Suleimenov OM and Seward TM (2000) Spectrophotometric measurements of metal Environmental Geochemistry: A Tribute to David A. Crerar, Geochemical Society
complex formation at high temperatures: The stability of Mn(II) chloride species. Special Publication No. 7, pp. 307–327. St. Louis, MO: Geochemical Society.
Chemical Geology 167: 177–192. Wallen SL, Palmer BJ, and Fulton JL (1998) The ion pairing and hydration structure of
Sverjensky DA, Shock EL, and Helgeson HC (1997) Prediction of thermodynamic Ni2þ in supercritical water at 425 C determined by x-ray absorption fine structure
properties of aqueous metal complexes to 1000 C and 5 kbar. Geochimica et and molecular dynamics studies. Journal of Chemical Physics 108: 4039–4046.
Cosmochimica Acta 61: 1359–1412. Walrafen GE, Chu YC, and Piermarini GJ (1996) Low frequency Raman scattering from
Symonds RB, Rose WI, Reed MH, Lichte FE, and Finnegan DL (1987) Volatilization, water at high pressures and high temperatures. Journal of Physical Chemistry
transport and sublimation of metallic and non-metallic elements in high temperature 100: 10363–10372.
The Chemistry of Metal Transport and Deposition by Ore-Forming Hydrothermal Fluids 57

Webster JG (1987) Thiosulfate in surficial geothermal waters, North Island, Wood SA, Wesolowski DJ, and Palmer DA (2000) The aqueous geochemistry of
New Zealand. Applied Geochemistry 2: 579–584. the rare earth elements: IX. A potentiometric study of Nd3þ complexation with
Weiss RF, Lonsdale P, Lupton J, Bainbridge A, and Craig H (1977) Hydrothermal acetate in 0.1 molal NaCl solution from 25 to 250 C. Chemical Geology
plumes in the Galapagos Rift. Nature 267: 600–603. 167: 231–253.
Weissberg BG, Browne PRL, and Seward TM (1979) Ore metals in geothermal systems. Xiao Z, Gammons CH, and Williams-Jones AE (1998) Experimental study of copper(I)
In: Barnes HL (ed.) Geochemistry of Hydrothermal Ore Deposits, 2nd edn., ch. 15, chloride complexing in hydrothermal solutions at 40 to 300 C and saturated water
pp. 738–780. New York: Wiley Interscience. vapour pressure. Geochimica et Cosmochimica Acta 62: 2949–2964.
Wernet Ph, Testemale D, Hazemann J-L, et al. (2005) Spectroscopic characterization of Xu Y, Schoonen MAA, Nordstrom DK, Cunningham KM, and Ball JW (2000) Sulfur
microscopic hydrogen-bonding disparities in supercritical water. Journal of geochemistry of hydrothermal waters in Yellowstone National Park, Wyoming, USA.
Chemical Physics 123: 154503. II. Formation and decomposition of thiosulfate and polythionate in Cinder Pool.
Wesolowski D, Drummond SE, Mesmer RE, and Ohmoto H (1984) Hydrolysis equilibria Journal of Volcanology and Geothermal Research 97: 407–423.
of tungsten(VI) in aqueous sodium chloride solutions to 300 C. Inorganic Yan H, Mayanovic RA, Anderson AJ, and Meredith PR (2011) An in-situ X-ray
Chemistry 23: 1120–1132. spectroscopic study of Mo6þ speciation in supercritical aqueous solutions. Nuclear
Wesolowski DJ, Ziemniak SE, Anovitz LM, Machesky ML, Bénézeth P, and Palmer DA Instruments and Methods in Physics Research, Section A 649: 207–209.
(2004) Solubility and surface adsorption characteristics of metal oxides. Yanat’eva OK (1952) Solubility of dolomite in water-salt solutions. Izvestiya Sektora
In: Palmer DA, Fernandez-Prini R, and Harvey AH (eds.) Aqueous Systems at Fiziko-Kimcheskogo Analiza Akademii Nauk SSSR 20: 252–268.
Elevated Temperatures and Pressures: Physical Chemistry in Water, Steam and Yang MM, Crerar DA, and Irish DE (1989) A Raman spectroscopic study of lead and
Hydrothermal Solutions, ch. 14, pp. 493–595. Amsterdam: Elsevier. zinc acetate complexes in hydrothermal solutions. Geochimica et Cosmochimica
Widler A and Seward TM (2002) The adsorption of hydrosulphide complexes of gold(I) Acta 53: 319–326.
onto iron sulphide mineral substrates. Geochimica et Cosmochimica Acta Yang K and Scott SD (1996) Possible contribution of a metal-rich magmatic fluid to a
66: 383–402. sea-floor hydrothermal system. Nature 383: 420–423.
Wiersma C and Rimstidt J (1984) Rates of reaction of pyrite and marcasite with ferric Yang K and Scott SD (2006) Magmatic fluids as a source of metals in seafloor
iron at pH 2. Geochimica et Cosmochimica Acta 48: 85–92. hydrothermal systems. In: Christie DM, Fisher CR, Lee S-M, and Givens S (eds.)
Wilkinson JJ, Stoffell B, Wilkinson CC, Jeffries TE, and Appold MS (2009) Anomalously Back-Arc Spreading Systems: Geological, Biological, Chemical, and Physical
metal-rich fluids from hydrothermal ore deposits. Science 323: 764–767. Interactions. Geophysical Monograph Series, vol. 166, pp. 163–184. Washington,
Williams AE and McKibben M (1989) A brine interface in the Salton Sea Geothermal DC: American Geophysical Union.
System, California: Fluid geochemical and isotopic characteristics. Geochimica et Yui K, Sakuma M, and Fanazukuri T (2010) Molecular dynamics simulation on ion-pair
Cosmochimica Acta 53: 1905–1920. association of NaCl from ambient to supercritical water. Fluid Phase Equilibria
Williams-Jones AE, Bowell RJ, and Migdisov AA (2009) Gold in solution. Elements 297: 227–235.
5: 281–287. Zajacz Z, Seo JH, Candela PA, Piccoli PM, Heinrich CA, and Guillong M (2010) Alkali
Williams-Jones AE and Heinrich CA (2005) 100th Anniversary Special Paper: Vapor metals control the release of gold from volatile-rich magmas. Earth and Planetary
transport of metals and the formation of magmatic-hydrothermal ore deposits. Science Letters 297: 50–56.
Economic Geology 100: 1287–1312. Zakaznova-Herzog VP and Seward TM (2006) Antimonous acid protonation/
Williams-Jones AE, Migdisov AA, Archibald SM, and Xiao Z (2002) Vapor-transport of deprotonation equilibria in hydrothermal solutions to 300 C. Geochimica et
ore metals. In: Hellmann R and Wood SA (eds.) Water-Rock Interactions, Ore Cosmochimica Acta 70: 2298–2310.
Deposits and Environmental Geochemistry. A Tribute to David A. Crerar, Zakaznova-Herzog VP and Seward TM (2012) A spectrophotometric study of the
Geochemical Society Special Publication No. 7, pp. 279–305. St. Louis, MO: formation and deprotonation of thioarsenite species in aqueous solution at 22 C.
Geochemical Society. Geochimica et Cosmochimica Acta 83: 48–60.
Williams-Jones AE, Samson IM, and Olivo GR (2000) The genesis of hydrothermal Zakaznova-Herzog VP, Seward TM, and Suleimenov OM (2006) Arsenous acid
fluorite-REE deposits in the Gallinas Mountains, New Mexico. Economic Geology ionization in aqueous solutions from 25 to 300 C. Geochimica et Cosmochimica
95: 327–341. Acta 70: 1928–1938.
Williamson M and Rimstidt JD (1994) The kinetics and electrochemical rate-determining Zezin DY, Migdisov AA, and Williams-Jones AE (2011) PVTx properties of H2O–H2S
step of aqueous pyrite oxidation. Geochimica et Cosmochimica Acta 58: 5443–5454. fluid mixtures at elevated temperature and pressure based on new experimental data.
Wilson SH (1941) Natural occurrence of polythionic acids. Nature 148: 502–503. Geochimica et Cosmochimica Acta 75: 5483–5495.
Wilson T and Long D (1993) Geochemistry and isotope chemistry of Michigan Basin Zhang L, Audétat A, and Dolejs D (2012) Solubility of molybdenite (MoS2) in aqueous
brines: Devonian formations. Applied Geochemistry 8: 81–100. fluids at 600–800 C, 200MPa: A synthetic fluid inclusion study. Geochimica et
Wood SA, Palmer DA, Wesolowski DJ, and Bénézeth P (2002) The aqueous Cosmochimica Acta 77: 175–185.
geochemistry of the rare earth elements and yttrium. Part XI. The solubility of Zorigtkhuu O-E, Tsunogae T, and Dash B (2012) Carbonic inclusions in amphibolite
Nd(OH)3 and hydrolysis of Nd3þ from 30 to 290 C at saturated water vapor facies pelitic schists from Bodonch area, western Mongolian Altai. Journal of
pressure with in-situ pHm measurement. In: Hellmann R and Wood SA (eds.) Water- Mineralogical and Petrological Sciences 107: 44–49.
Rock Interactions, Ore Deposits and Environmental Geochemistry. A Tribute to Zotov AV, Tagirov BR, Diakonov II, and Ragnarsdottir KV (2002) A potentiometric study
David A. Crerar, Geochemical Society Special Publication No. 7, pp. 229–256. of Eu3þ complexation with acetate ligand from 25 to 170 C at Psat. Geochimica et
St. Louis, MO: Geochemical Society. Cosmochimica Acta 66: 3599–3613.

Das könnte Ihnen auch gefallen