You are on page 1of 21

Geoscience Frontiers 14 (2023) 101624

Contents lists available at ScienceDirect

Geoscience Frontiers
journal homepage: www.elsevier.com/locate/gsf

Research Paper

Tin transport and cassiterite precipitation from hydrothermal fluids


Xiangchong Liu a,b,c, Pingping Yu a, Changhao Xiao a,⇑
a
Institute of Geomechanics, Chinese Academy of Geological Sciences, Beijing 100081, China
b
Key Laboratory of Paleomagnetism and Tectonic Reconstruction, Ministry of Natural Resources, Beijing 100081, China
c
The Laboratory of Dynamic Diagenesis and Metallogenesis, Institute of Geomechanics, CAGS, Beijing 100081, China

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

Article history: Cassiterite (SnO2) is the main ore mineral of tin in magmatic–hydrothermal tin deposits, but tin transport
Received 8 November 2022 and precipitation mechanisms from hydrothermal fluids remain poorly understood. We critically evalu-
Revised 24 March 2023 ated aqueous tin speciation in hydrothermal fluids from extensive experimental data and thermody-
Accepted 22 April 2023
namic modeling. Sn(II) chloride complexes in hydrothermal fluids exist mainly as SnCl+, SnCl2(aq), and
Available online 27 April 2023
Handling Editor: Kristoffer Szilas
SnCl3 . The revised Helgeson–Kirkham–Flowers model parameters of these three tin species and two
tin ions (Sn4+ and Sn2+) were derived from the correlation algorithms among these parameters, and
the standard molar properties of cassiterite were optimized to be internally consistent with the available
Keywords:
thermodynamic dataset. These thermodynamic parameters, together with the available equilibrium con-
Cassiterite stant equation of Sn(IV) chloride complexes, could reproduce the available solubility data of cassiterite in
Aqueous tin species acidic solutions at 400–700 °C under oxygen fugacity (f O2 ) levels buffered by hematite–magnetite (HM)
Precipitation mechanisms or nickel–nickel oxide (NNO). These comparisons allow modeling chemical systems of SnO2–NaCl–HCl–
Tin deposits H2O (liquid phase) to examine tin transport and cassiterite precipitation mechanisms under
Thermodynamic modeling tin-mineralizing conditions: 300–500 °C, 50–150 MPa, 2 molal NaCl, and f O2 levels from QFM
(quartz–fayalite–magnetite) to HM. Sn(II) chloride complexes are commonly interpreted to dominate
in aqueous tin speciation under f O2 = NNO, but our modeling results indicate that considerable contents
of Sn(IV) chloride complexes also exist in those reduced fluids with high HCl contents, consistent with
recent in situ high-temperature experiments and molecular dynamic simulations. The Sn(II)/Sn(IV) ratios
in fluids depends on f O2 , temperature, and HCl contents. A considerable amount of Sn(IV) possibly exist in
an early mineralization stage even under f O2 = NNO; if so, redox reactions are unnecessary to precipitate
cassiterite from these mineralizing fluids. We find that even if the f O2 levels are constant, simple cooling
can alter mineralizing fluids to be more oxidized (e.g., from QFM to HM) and cause cassiterite
precipitation, indicating that oxidizing agents are not necessary as previously thought. This explains
why cassiterite can precipitate in host rocks (e.g., sandstone or quartzite) that do not provide oxidizing
agents. A simple rise in f O2 levels and pH neutralization (e.g., greisenization) also cause cassiterite
precipitation. Cassiterite solubility in oxidized acidic hydrothermal fluids (NNO < f O2  HM) is high
enough to account for the tin contents of fluid inclusions from typical tin deposits, but the mineralization
potential of oxdized fluids is inferior to reduced fluids (f O2  NNO) under the same conditions.
Ó 2023 China University of Geosciences (Beijing) and Peking University. Published by Elsevier B.V. on
behalf of China University of Geosciences (Beijing). This is an open access article under the CC BY-NC-ND
license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction evolved granites emplaced in the upper crust (Lehmann, 1990;


Mao et al., 2019). Recent studies also identified the major roles
Tin (Sn) is considered a critical and strategic metal (Kamilli of metamorphic fluids in mobilizing tin from source rocks
et al., 2017). By far, the predominant ore mineral of tin is cassiterite (Romer et al., 2022).
(SnO2), and cassiterite occurs in magmatic–hydrothermal deposits Tin-associated granites are generally crystallized from reduced
and placer deposits, which account for approximately 20% and 80% (between quartz–fayalite–magnetite and nickel–nickel oxide buf-
of the world’s tin resources, respectively (Kamilli et al., 2017). fers) metaluminous to peraluminous granitic melts (Dubessy
These two types of tin deposits are genetically related to highly et al., 1987; Mlynarczyk et al., 2003; Chen et al., 2013; Kinnaird
et al., 2016), in which Sn(II) dominates over Sn(IV) (e.g., Linnen
⇑ Corresponding author. et al., 1995; Linnen et al., 1996; Bhalla et al., 2005; Farges et al.,
E-mail addresses: xiaochanghao1986@126.com, xcliu@cags.ac.cn (C. Xiao). 2006). Tin diffusivity in hydrous silicate melts was quantified by

https://doi.org/10.1016/j.gsf.2023.101624
1674-9871/Ó 2023 China University of Geosciences (Beijing) and Peking University. Published by Elsevier B.V. on behalf of China University of Geosciences (Beijing).
This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Yang et al. (2016), and the partitioning of tin from granitic melts salinities <10 wt.% NaCl equivalent (Naumov et al., 2011), but the
into aqueous phase was re-examined by Zhao et al. (2022). Follow- fluid salinities at earlier stages may reach higher levels (e.g.,
ing the dominance of Sn(II) at magmatic stage, extensive experi- Myint et al., 2018; Korges et al., 2020). The mineralization pres-
mental data under reduced hydrothermal conditions have been sures of most fluid inclusion data lie in the range of 50–150 MPa
derived for tin speciation and cassiterite solubility in the last sev- (Naumov et al., 2011).
eral decades (e.g., Kovalenko et al., 1986; Wilson and Eugster, Tin-mineralizing fluids are generally acidic and leave typical
1990; Dorofeeva et al., 1994; Duc-Tin et al., 2007), and tin in min- mineral assemblages of quartz–muscovite–fluorite–topaz (grei-
eralizing fluids is generally interpreted to be transported mainly as sens) after flowing through granitic rocks (Štemprok, 1987;
Sn(II) chloride complexes (Jackson and Helgeson, 1985a, 1985b; Pirajno, 1992; Meinert et al., 2005). Constraints from common
Heinrich, 1990; Halter et al., 1998). Thus, oxidation (H2 consump- mineral assemblages and the mineral–solution equilibrium in tin
tion) is required for cassiterite precipitation in previous formation deposits suggest that the mineralizing fluids have pH levels of 4–
models (cf. Heinrich, 1990; Lehmann, 2021): 6 (Patterson et al., 1981; Jackson and Helgeson, 1985b; Polya,

1989). At their sources, the mineralizing fluids may be more acidic
þ 2H2 O ¼ SnO2 ðcassiteriteÞ þ 2Hþ þ xCl þ 0:5H2
2x
SnClx than those aforementioned pH levels. Values of oxygen fugacity
ð1Þ (see Fig. 1) are estimated from gas compositions (CO2 and CH4)
of fluid inclusions and from mineral equilibrium and mostly lie
However, the dominance of Sn(II) chloride complexes was
between those of the mineral assemblages of quartz–fayalite–mag
derived by regressing the solubility data after quenching. In con-
netite (QFM) and nickel–nickel oxide (NNO) buffers (Patterson
trast, recent in situ high-temperature experiments and molecular
et al., 1981; Jackson and Helgeson, 1985b; Polya, 1989; Xiang
dynamic simulations identified notable contents of Sn(IV) chloride
et al., 2018; Schmidt et al., 2021).
complexes in reduced solutions (Schmidt, 2018; Chou et al., 2021;
Published in situ compositions of fluid inclusions from Sn–W
Wang et al., 2021), challenging the earlier view of the dominance
and W–Sn deposits were compiled in this study to show the distri-
of Sn(II) in mineralizing fluids. Therefore, it is necessary to evaluate
bution of tin contents in mineralizing fluids and place a quantita-
the available experimental data in the literature and re-examined
tive constraint on subsequent thermodynamic modeling. These
tin speciation under hydrothermal conditions.
content data come from typical tin deposits of four world-class
Reliable thermodynamic data of major aqueous tin species are
tin provinces, including the Sn–W deposits around the Mole Gran-
fundamental to modeling tin transport and cassiterite solubility
ite, Australia (Audétat et al., 2000); the Sn–W deposits in the Erzge-
in hydrothermal fluids. The standard molar properties and revised
birge, Germany and the Czech Republic (Korges et al., 2017); the
Helgeson–Kirkham–Flowers (HKF) model parameters of five aque-
Huanuni tin deposit, Bolivia (Müller et al., 2001); and the Piaotang
ous tin species are derived in this study by evaluating extensive
and Maoping W–Sn deposits in the Nanling Range, China (Legros
experimental data and using the correlation algorithms among
et al., 2019). The boxplot in Fig. 2 shows that the tin contents of
these parameters for aqueous metal ions and complexes proposed
mineralizing fluids range from a few to 104 parts per million
by Shock et al. (1997) and Sverjensky et al. (1997). The estimated
(ppm), with an average of 529 ppm. Tin contents between 10
thermodynamic parameters of Sn(II) chloride complexes can
and 1000 ppm account for 83% of the total data, so this range of
reproduce the available experimental solubility data. This paper
tin concentrations can be used to place a lower limit on cassiterite
is organized as follows. First, the geochemical characteristics of
solubility in hydrothermal fluids under tin-mineralizing
tin-mineralizing fluids are summarized to obtain the geochemical
conditions.
characteristics and the tin contents in the mineralizing fluids. Sec-
ond, the methods used for estimating the thermodynamic proper-
3. Methods
ties of aqueous tin species and thermodynamic modeling are
described. Third, extensive experimental data of tin speciation
Chemical mass transfer modeling has been used to decipher
and cassiterite solubility under hydrothermal conditions are criti-
tin transport and cassiterite deposition from mineralizing fluids
cally evaluated to derive major tin species in hydrothermal fluids.
(Jackson and Helgeson, 1985a, 1985b; Heinrich, 1990; Tornos,
Fourth, how tin is transported and the main variables affecting cas-
1997; Halter et al., 1998). Heinrich (1990) evaluated several
siterite solubility under tin-mineralizing conditions are analyzed
key chemical and physical parameters that may cause cassi-
from modeling results.
terite precipitation using fluid-buffered and rock-buffered mass
transfer modeling, in which the thermodynamic data of Sn(II)
2. Geochemical characteristics of tin-mineralizing fluids chloride complexes came from Pabalan (1986). As shown later,
the high contents of SnCl2 4 interpreted by Pabalan (1986) in
Tin-mineralizing fluids belong to the NaCl–H2O ± CO2 systems their experiments is problematic. Halter et al. (1998) modeled
(Naumov et al., 2011). CO2-bearing fluid inclusions in gangue fluid–rock interactions during greisenization in a tin deposit
minerals such as quartz and topaz are commonly found among using the equilibrium constants of SnCl+ fitted by Wilson and
many magmatic-hydrothermal tin deposits (e.g., Mangas and Eugster (1990). The chemical systems of cassitierite in
Arribas, 1987; Halter and Williams-Jones, 1995; Borges et al., hydrothermal fluids were modeled and the concentrations of
2009; Jia et al., 2019) but occasionally in cassiterite (e.g., aqueous tin species and the solubility of cassiterite were calcu-
Linnen, 1998; Schmidt et al., 2021). These aqueous–carbonic lated by the method of equilibrium constants. This method
inclusions were interpreted to have a magmatic or metamorphic assumes that a chemical system is at equilibrium and that all
origin (e.g., Linnen, 1998; Dewaele et al., 2010; Van Daele et al., species in the chemical system are stable. The concentrations
2018). and activities of all species can be derived by solving nonlinear
The dataset of Naumov et al. (2011), which includes 320 Sn and equations.
Sn–W deposits, suggests that most fluid inclusions in cassiterite Each chemical reaction in the chemical system at equilibrium
have homogenization temperatures ranging from 300 to 500 °C. produces a nonlinear equation. For example, the nonlinear equa-
This temperature range is consistent with earlier compiled data- tion for the association of Na+ and Cl into NaCl(aq) is:
sets (e.g., Lehmann, 1990) and numerous later studies (e.g.,
cNaClðaqÞ mNaClðaqÞ
Korges et al., 2017; Myint et al., 2018; Jia et al., 2019; Liu et al., K¼ ð2Þ
2020a; Han et al., 2023). Tin-mineralizing fluids generally have cNaþ mNaþ cCl mCl
2
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Fig. 1. The oxygen fugacity of tin-mineralizing fluids available in the literature. The curves of mineral–buffered oxygen fugacity were calculated at a fixed pressure of
200 MPa using thermodynamic data from Helgeson (1978), Robie et al. (1978), and Berman (1988). The points come from Jackson and Helgeson (1985b) (the tin deposits from
the Southeast Asian tin belt), Patterson et al. (1981) (the Renison Bell tin deposit, Australia), Polya (1989) (the Panasqueira tungsten-tin deposit, Portugal), and Xiang et al.
(2018) (the Jiumao tin deposits, China). The data of Schmidt et al. (2021) were calculated at a fixed pressure of 50 MPa based on average compositions of primary fluid
inclusions in cassiterite from seventeen tin deposits in the world. HM, hematite-magnetite buffer; NNO, nickel-nickel oxide buffer; QFM, quartz-fayalite-magnetite buffer;
FM, ferrous oxide-magnetite; FI, ferrous oxide-iron.

Fig. 2. The in situ Sn contents of fluid inclusions from typical Sn–W and W–Sn deposits. The tin contents (unit in ppm) have been logarithmically transformed with respect to
base 10. On each box, the central mark indicates the median, and the bottom and top edges of the box indicate the 25th and 75th percentiles, respectively. N represents the
sampling number of each dataset. The average of the 325 data is 529 ppm.

in which K is the equilibrium constant; and mNaClðaqÞ and cNaClðaqÞ are solutions, respectively. The activity coefficients of electrically
the concentration and the activity coefficient of NaCl(aq) in aqueous charged species in the models were calculated by an extended

3
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Debye–Hückel equation (Helgeson et al., 1981), and those of neutral from previous experimental data. Despite these available datasets
species were derived from the Setchénow equation (see Eq. (3) in of cassiterite, its thermodynamic properties were updated for the
Miron et al., 2016). The activity of H2O is generally assumed to be following two reasons. The first and most important reason is that
unity at room temperature in geochemical modelling (Anderson, each dataset comes from several sources and no one has tested its
2005). However, the high-temperature experimental data of consistency with the available internally consistent thermody-
Aranovich and Newton (1996) suggest that the activity of H2O namic dataset of minerals and gas species (Holland and Powell,
approximately equal its mole fraction at 620–640 °C and 200 MPa 2011). The second reason is that the standard molar entropy and
in concentrated NaCl solutions, and H2O activity rapidly decreases heat capacity measured by Gurevich et al. (2004), whose experi-
below the mole fraction with increasing pressure. The tin- mental data are more reliable than those of Zhogin et al. (1980)
mineralizing conditions and most experimental data in the litera- (see Gamsjäger et al., 2012, pp. 128–133), are 5.8% and 3.8% higher
ture do not exceed the temperature and pressure conditions of than the corresponding values recommended by Robie and
Aranovich and Newton (1996), so H2O activity in this study was Hemingway (1995) (see Table 1).
assumed to equal its mole fraction in solutions. These equations To make the thermodynamic properties of cassiterite consistent
of activity coefficients were calculated using the R language free with the dataset of Holland and Powell (2011), we first updated the
package CHNOSZ (Dick, 2019). CHNOSZ is an open–source R lan- heat capacity function from the available experimental data. Then,
guage package that uses the revised HKF electrostatic equations of we optimized the standard molar Gibbs energy of cassiterite from
state (Helgeson et al., 1981) and the thermodynamic data of miner- the available thermodynamic equilibrium data that are associated
als and aqueous species in SUPCRT (Johnson et al., 1992). Charge with cassiterite, gas species (e.g., O2(g) and H2(g)), and other oxides
and mass balance relations are also required to calculate the con- (e.g., Cu2O and NiO).
centrations of the species in the models. The nonlinear equations Kapustinskii et al. (1936) reported the most reliable high-
were solved using another free package in R language, rootSolve temperature heat capacity of cassiterite until now, and their data
(cf. Soetaert and Herman, 2008), and the solving process was similar were used to fit the heat capacity equation by Robie et al. (1978)
to that presented in the appendix of our previous paper (Liu and and Robie and Hemingway (1995). Their high-temperature data,
Xiao, 2020). The concentration of a species is expressed as molal together with the low-temperature heat capacity data of
(mol kg1 H2O) throughout this article. Gurevich et al. (2004), were employed to fit the Hass–Fisher heat
The revised HKF electrostatic equations of state (Helgeson et al., capacity function against temperature (Haas and Fisher, 1976):
1981; Tanger and Helgeson, 1988) is the most accepted model in
geochemical applications and is conventionally restricted to C oP ¼ a þ bT þ cT 2 þ dT 0:5 þ eT 2 ð3Þ
1000 °C and 500 MPa at liquid-like fluid densities (>0.35 g cm3)
Fig. 3 shows that the fitted coefficients reproduce the experi-
(Shock et al., 1997; Dolejš, 2013). Most thermodynamic data of
mental data with an absolute error no >0.6 J mol1 K1. To be con-
minerals and aqueous species (e.g., H+, OH, and Cl) used in the
sistent with the heat capacity data of Gurevich et al. (2004), the
models come from the SUPCRT database (Johnson et al., 1992).
standard molar entropy of 51.82 J mol1 was used in this study.
The HKF model parameters of Na+, NaCl(aq), and NaOH(aq)
updated by Miron et al. (2016) were used, and those of cassiterite For a solid phase, its molar Gibbs free energy of formation Df G0PT
were updated in the present study. The standard molar properties can be calculated from the equation
and revised HKF model parameters of five aqueous tin species Z Z
(Sn4+, Sn2+, SnCl+, SnCl2(aq), and SnCl
T T
C 0P
3 ) were calculated from the Df G0PT ¼ Df G0T r ;Pr  ðT  T r ÞS0T r ;Pr þ C 0P dT  T dT
correlation algorithms proposed by Shock et al. (1997) and Tr Tr T
Sverjensky et al. (1997). The Akinfiev-Diamond model was used 0

for aqueous nonelectrolytes such as O2(g) and H2(g) (Akinfiev þ V T r ;Pr ðP  Pr Þ ð4Þ
and Diamond, 2003).
in which T r = 298.15 K; P r = 0.1 MPa; and Df GoT r ;Pr is the standard
molar Gibbs free energy of formation, the standard molar entropy
4. The thermodynamic properties of cassiterite
SoT r ;Pr , the heat capacity C op , and the standard molar volume V oT r ;Pr .
To our knowledge, at least three datasets for the thermody- The Df G0T r ;Pr values of cassiterite were calculated from the ther-
namic properties of cassiterite are available in the literature (see modynamic equilibrium data of the following chemical systems:
Table 1). The thermodynamic parameters of cassiterite in the Sn–O (Bannister, 1986; Li-Zi et al., 1994; Mallika et al., 2001),
SUPCRT database come from Robie et al. (1978). Later, these data Sn–Cu–O (Schaefer, 1984), Sn–H–O (Atarashiya et al., 1960), and
were updated by Robie and Hemingway (1995) from the experi- Sn–Ni–O (Belford and Alcock, 1965; Petot-Ervas et al., 1975). These
mental data of Zhogin et al. (1980) and Bannister (1986) and the experimental studies reported the molar Gibbs energy changes of
dataset reported by CODATA (Cox et al., 1989). Gamsjäger et al. the chemical reactions shown in Table 2; thus, the Df G0T r ;Pr of cassi-
(2012) also calculated the standard molar properties of cassiterite terite is the only variable for each experimental point (see Eq. (4)).
The thermodynamic data of cuprite (Cu2O) and bunsenite (NiO) in
Holland and Powell (2011) were used. A total of 138 experimental
points produce an average Df GoT r ;Pr ¼ 514.62 kJ mol1 of cassi-
Table 1
The available standard molar properties of cassiterite in the literature. terite (Fig. 4), slightly greater than those of Robie and
Hemingway (1995) (515.8 kJ mol1) and Gamsjäger et al.
Sources Df GoPr ;T r SoPr ;T r C op
(2012) (516.64 kJ mol1). Note that the standard deviation of this
(kJ mol1) (J mol1) (J mol1 K1)
study is significantly larger than those reported in the other two
SUPCRT1 519.90 ± 0.75 52.30 ± 1.25 52.59 studies. Such a large standard deviation is possibly caused by mult-
Robie and Hemingway (1995) 515.8 ± 0.2 49.0 ± 0.2 53.22
Gamsjäger et al. (2012) 516.64 ± 0.21 51.77 ± 0.14 55.26 ± 0.09 ple sources of experimental erros. For example, the Df G0T r ;Pr values
Gurevich et al. (2004) – 51.82 ± 0.07 55.24 ± 0.03 of cassiterite derived from the data of Bannister (1986) at <700 °C
This study 514.62 ± 2.83 51.82 55.24 deviate from the average value by up to 7 kJ mol1, while those
P r = 0.1 MPa, T r = 298.15 K. over 700 °C are mostly within 1–2 kJ mol1. Overall, the updated
1
, data from Robie et al. (1978). thermodynamic properties of cassiterite are internally consistent
4
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Fig. 3. The heat capacity of cassiterite fitted from the available experimental data. (a) The empirical equation fitted by this study matches the low-temperature and high-
temperature experimental data, while the values reproduced from the equation of Robie and Hemingway (1995) are significantly lower than the low-temperature
experimental data. The heat capacities of Robie and Hemingway (1995) under hydrothermal conditions are close to the empirical equation fitted by this study because both
two equations employed the high-temperature data of Kapustinskii et al. (1936). Two samples were used by Gurevich et al. (2004), Russian pure tin (IV) oxide (RP) and
commercial tin (IV) oxide product of Alfa Aesar (AA). (b) The absolute residuals of the fitted heat capacity do not exceed 0.6 J mol1 K1.

Table 2 with those of cuprite (Cu2O) and bunsenite (NiO) in Holland and
The chemical reactions for optimizing the Df GoPr ;T r value of cassiterite. Powell (2011).
Numbers Chemical reactions References
1 Sn(l) + O2(g) = SnO2(s) 1, 2, 3 5. Tin species and chemical reactions under hydrothermal
2 Sn(s) + 2Cu2O = SnO2(s) + 4Cu 4 conditions
3 Sn(l) + 2H2O = SnO2(s) + 2H2(g) 5
4 Sn(s) + 2NiO = SnO2(s) + 2Ni 6, 7
Tin species in hydrothermal fluids exist mainly as tin chloride
References: 1, Bannister (1986); 2, Li-Zi et al. (1994); 3, Mallika et al. (2001); 4, and fluoride complexes (Wilson and Eugster, 1990; Taylor and
Schaefer (1984); 5, Atarashiya et al. (1960); 6, Belford and Alcock (1965); 7, Petot-
Wall, 1993; Sherman et al., 2000; Müller and Seward, 2001; Duc-
Ervas et al. (1975).
Tin et al., 2007). Although water vapor can carry an appreciable

Fig. 4. The optimized Df G0T r ;Pr value of cassiterite from the available equilibrium data. A total of 138 experimental data points give an average of Df G0T r ;Pr = 514.62 KJ mol1
with a standard deviation of 2.83 KJ mol1. The shaded area (514.62 ± 5.66 kJ mol1) covers most of the experimental determinations.

5
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

amount of tin (Migdisov and Williams-Jones, 2005), the typical tin- (e.g., Klintsova and Barsukov, 1973; Dadze et al., 1981; Sorokin
mineralizing conditions (see Section 2) indicate that the aqueous and Dadze, 1994; Sherman et al., 2000; Liu et al., 2020b). For sim-
liquid phase plays the main role in ore forming processes. There- plicity, only aqueous Sn(IV) species were used to analyze the
fore, we focused on the thermodynamic properties of aqueous Sn experimental data under unconstrained f O2 , and both Sn(IV) and
(II) and Sn(IV) ions, hydroxides, and chloride complexes. Tin fluo- Sn(II) species were considered to interpret the experimental data
ride complexes are not considered in this study. if the f O2 values were known.
There are several reviews about inorganic Sn(II) and Sn(IV) spe-
cies from extensive experimental data available in the literature 5.2. Sn4+
(Lothenbach et al., 1999; Séby et al., 2001; Cigala et al., 2012;
Gamsjäger et al., 2012). Séby et al. (2001) published a review of The oxidation state of tin in cassiterite is stannic (Sn4+). The
Sn(II) hydrolytic and complex species at room temperature from available standard molar Gibbs free energies of formation Df GoT r ;Pr
the literature. Later, Cigala et al. (2012) critically reviewed Sn(II)
of Sn4+ are highly controversial. Wagman et al. (1968) calculated
species with the most important natural inorganic ligands (OH,
the Df GoT r ;Pr of Sn4+ to be 2.51 kJ mol1, which was recommended
Cl, F, CO2 2 3
3 , SO4 , and PO4 ) and derived their formation con-
stants and the formation enthalpy changes at 25 °C. Gamsjäger by National Bureau of Standards/National Institute of Standards
et al. (2012) also evaluated aqueous tin species in solutions mostly and Technology (NBS/NIST) (Wagman et al., 1982; Reed, 2020).
at room temperature. These previous reviews provide a basis for This value was also used by Jackson and Helgeson (1985a) to pre-
interpreting aqueous tin species in hydrothermal fluids at temper- dict the molar Gibbs free energies of formation of Sn4+ under
atures over 300 °C. hydrothermal conditions. In contrast, the Df GoT r ;Pr = 19.318 kJ mol1
We evaluated the available experimental data (Table 3) and derived by Ryzhenko et al. (1997) is much lower than that of
estimated the thermodynamic parameters of five aqueous tin spe- Wagman et al. (1968). Note that the procedure by which
cies (Table 4) from the correlation algorithms shown in Supple- Ryzhenko et al. (1997) derived this value is unknown. Gamsjäger
mentary Data 1. All the cited experimental solubility data are et al. (2012) derived a much higher Df GoT r ;Pr ¼ 46:7  3:9 kJ mol1
listed in Supplementary Data 2. Here, we follow the convention from the standard electrode potential of the Sn4+/Sn2+ redox couple
adopted by Wagman et al. (1982) and Shock et al. (1997) that in Gajda et al. (2009). Rai et al. (2011) derived logK =  64.39 ±
H2O is subtracted from the stoichiometry of some aqueous species 0.32 of the following chemical reaction from their experimental
(e.g., hydroxide complexes). For example, the uncharged Sn(II) data at 23 ± 2 °C and 1 bar:
hydroxide Sn(OH)2(aq) is replaced by SnO(aq), and Sn(OH)4(aq) is
replaced by SnO2(aq). The reason for this simplicity is because SnO2 ðsÞ þ 2H2 O ¼ Sn4þ þ 4OH ð5Þ
the formulas with or without the number of molecules of water However, this logK value is possibly problematic. First, Rai et al.
are generally thermodynamically equivalent (cf, Wagman et al., (2011) neglected the formation of Sn(IV) chloride complexes in
1982, pp. 2–22). This pragmatic decision ignores the molecular 0.0003–1.0 molal HCl solutions, whereas the experimental data
structure of aqueous species studied. Note that no one has tested of Gajda et al. (2009) suggest the dominance of Sn(IV) chloride
whether the hydroxides and chloride complexes of Sn(II) and Sn complexes at >3  105 molal Cl. Second, Sn4+ is a minor species
(IV) with a few number of water molecules are thermodynamically in the solutions of Rai et al. (2011) and calculating the logK of Reac-
equivalent to those without water molecules. The following abbre- tion (5) is easily affected by ignoring Sn(IV) chloride complexes.
viations of mineral assemblage–buffered f O2 levels are used for Therefore, the logK of Reaction (5) derived by Rai et al. (2011) can-
simplicity: NNO, nickel–nickel oxide; MM, MnO2–Mn2O3; HM, not be used to test the above Df GoT r ;Pr values of Sn4+. As a result, a
hematite–magnetite; PPM: pyrite–pyrrhotite–magnetite. recommended Df GoT r ;Pr value of Sn4+ is not generated here. From
later calculations, even if a lower Df GoT r ;Pr value of Sn4+ is used, its
5.1. The f O2 boundary between the Sn(II) and Sn(IV) dominance fields concentrations in solutions are extremely low and can be ignored
for geologic fluids.
Tin in hydrothermal fluids has two oxidation states, +2 and +4. The standard molar entropy SoT r ;Pr of Sn4+ in the dataset compiled
The dominant tin species in solutions under f O2  NNO is com- by Wagman et al. (1968) is 117.15 J mol1 K1, but the source is
monly interpreted to be stannous (e.g., Kovalenko et al., 1986; unknown. Sassani and Shock (1992) derived a value of SoT r ;Pr = 472.
Kovalenko et al., 1989; Wilson and Eugster, 1990; Kovalenko 37 J mol1 K1 from their correlations between the SoT r ;Pr of aqueous
et al., 1992; Taylor and Wall, 1993; Dorofeeva et al., 1994; Duc- ions and their effective electrostatic radii, which is close to the one
Tin et al., 2007). However, this interpretation is derived by regress- predicted by Gamsjäger et al. (2012) using the modified Powell–
ing the solubility data after quenching and is challenged by recent Latimer correlation (468 ± 33 J mol1 K1). To maintain consis-
in situ experiments. Schmidt (2018) proposed from his in situ tency with the revised HKF model parameters of other aqueous
Raman spectroscopic data that the f O2 boundary between Sn(II)
tin species, SoT r ;Pr = 472.37 J mol1 K1 derived by Sassani and
and Sn(IV) dominance fields lies between f O2 ¼ HM and a few log
Shock (1992) was used in this study. The standard heat capacity
units below f O2 ¼ NNO. Later, Schmidt et al. (2021) constrained C op , the standard molar volume V oT r ;Pr , and the revisd HKF model
the boundary to be between f O2 near NNO and the rhenium- parameters of Sn4+ were derived using the correlation algorithms
rhenium oxide buffered f O2 levels (close to HM), and Chou et al. in Supplementary Data 1.
(2021) identified the in situ Raman spectra of Sn(IV) chloride com-
plexes in solutions at 300 °C and f O2 ¼ NNO (see their Fig. 3a). 5.3. Sn2+
These recent experimental data suggest that the amount of aque-
ous Sn(IV) species in those solubility experiments under The structure of hydrated Sn(II) ions in aqueous solutions was
f O2 ¼ NNO cannot be neglected. recently identified by Persson et al. (2016). We found that the C op
The tin species in solutions under f O2  HM is generally inter- of Sn2+ in the SUPCRT database was not the value calculated from
preted to be stannic (Wilson and Eugster, 1990). The experiments Eq. (18) of Shock et al. (1997) (see Eqs. A1–A2 in Supplementary
under unconstrained oxygen fugacities should be at higher f O2 Data 1) and the SoT r ;Pr recommended by Cox et al. (1989), and this
levels, and the predominant aqueous tin species was also stannic incorrect value leads to incorrect estimates of its revised HKF
6
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Table 3
Summary of hydrothermal experiments for cassiterite solubility or tin speciation.

Authors Methods T (°C) P (MPa) Aqueous solution f O2 Dominant tin species

Kovalenko et al. Solubility 500 100 NaOH (0.0001–0.5 molal) NNO Sn(OH)2
4
(1992)
Kovalenko et al. Solubility 500 100 NaOH (0.18–0.5 molal)–NaCl NNO Sn(OH)2Cl
(1992) (0.5–4 molal)
Chen (1986) Solubility 300–450 0.1–45 HCl (0.001–0.1 molal) charcoal* Sn(II)-Cl complexes
Kovalenko et al. Solubility 500–700 100 HCl–NaCl–KCl NNO SnCl2(aq), SnOHCl (netural pH)
(1986)
Pabalan (1986) Solubility 200–350 Psat HCl (?) IM SnCl2n
n (1  n  4)
Li (1988) Solubility 200–500 100 HCl NNO SnCl2(aq), SnOHCl
+
Wilson and Solubility 400–600 150 HCl (0.2–4.45 molal) NNO SnCl , SnCl2(aq)
Eugster (1990)
Taylor and Wall Solubility 700–800 200 HCl + NaCl + KCl [QFM  0.7, SnCl2(aq), SnOHCl, alkali-chloride
(1993) QFM + 1.5] complexes (e.g. KSnOHCI2)
Sherman et al. EXAFS spectra 25–350 0.1 SnCl2–HCl (0.5–2.5 molal) unknown** SnCl2
4
(2000)
Müller and Seward spectrophotometry 25–300 Psat HCl (0.01 molal)–NaCl (0.01– Ar* SnCl2n
n (1  n  4)
(2001) 2.936 molal)
Duc-Tin et al. Synthetic fluid 700 140 HCl (0.5–4.4 molal)–NaCl (5– NNO SnCl2(aq), Sn(OH)Cl
(2007) inclusions 35 wt.%)
Duc-Tin et al. Synthetic fluid 700 140 NaCl (5–35 wt.%) NNO Sn(OH)Cl
(2007) inclusions
Wang et al. (2021) Molecular 57–700 0.1–150 SnCl2n
n (n = 0–4) unknown** SnCl
3 (<300 °C), SnCl2(aq) (500–700
dynamics °C)
simulation
Klintsova and Solubility 25–200 0.1 Water unknown** SnO2(aq)
Barsukov
(1973)
Dadze et al. (1981) Solubility 200–400 101.3 Water unknown** SnO2(aq)
Liu and Chen Barnes volumetry 25–350 0.1–60 Water, NaCl (0.01–1 molal) unknown** SnO2(aq)
(1986)
Sorokin and Dadze Solubility 200–400 1.6–150 water unknown** SnO2(aq)
(1994)
Klintsova and Solubility 25–200 0.1 NaOH unknown** Sn(OH) 2
5 ,Sn(OH)6
Barsukov
(1973)
Liu and Xiao Raman 400–850 300–850 Li2CO3, Na2CO3 unknown** Sn(OH)2
6
(2020) spectroscopy
Klintsova et al. Solubility 25–200 0.1 HNO3 + NaF unknown** SnO2(aq)
(1975)
Dadze et al. (1981) Solubility 200–400 101.3 HCl (0.0025–0.04 molal) unknown** SnO2(aq), Sn(OH)+3
Dadze et al. (1981) Solubility 200–400 101.3 HNO3 (0.01–0.5 molal) unknown** SnO2(aq)
Kovalenko et al. Solubility 500 100 HCl–NaCl–KCl MM SnOHCl03, SnO2(aq)
(1986)
Li, 1988) Solubility 200–500 100 HCl (0.01–1.0 molal) unknown** SnO2(aq), Sn(OH)3Cl0
Wilson and Solubility 500–600 150 HCl (0.174–4.64 molal) HM SnCl+3
Eugster, 1990
Sorokin and Dadze Solubility 200–400 101.3 HCl (0.001–0.5 unknown** SnO2(aq), Sn(OH)+3, Sn(OH)3Cl
(1994) molal) ± HNO3 ± KCl

Sherman et al. EXAFS spectra 25–350 0.1 SnCl4 (0.01 molal) + HCl (0.11 unknown** SnCl2
6
(2000) molal) + NaCl (2 molal)
Chou et al. (2021) Raman 30 Psat 0.5 molal SnCl4 + 0.5 molal HCl unknown** SnCl 0 +
5 , SnCl4(aq), SnCl3
spectroscopy
Kokh et al. (2016) hydrothermal 349–351 13.9–23.2 CO2–NaCl–KCl HM SnCl4(aq)
reactor
Kokh et al. (2017) hydrothermal 450 57–68 CO2–KCl PPM SnO2(aq)
reactor
Schmidt (2018) Raman 500–600 194–1330 HCl (1.63–4.26 molal) Ar ± H2 ± CH4* SnCl4(aq), SnCl+3, SnCl
5
spectroscopy

The oxygen fugacity f O2 : IM, iron–magnetite; NNO, nickel-nickel oxide; MM, MnO2–Mn2O3; HM, hematite–magnetite; PPM: pyrite–pyrrhotite–magnetite. EXAFS, extended
X-ray absorption fine structure. One asterisk (*) above unknown f O2 represents that the oxygen fugacity was controlled by the material before the asterisk, but the precise f O2
is unknown; two asterisks (**) above unknown f O2 represent that the oxygen fugacity was unconstrained. Under the mineral assemblage of MnO2–Mn2O3 at 500 °C and
100 MPa, logf O2 ¼ –0.22 bar (1 bar = 0.1 MPa).

model parameters (see Eqs. A10–A11 in Supplementary Data 1). 5.4. Sn(II) hydroxide complexes
The C op and revised HKF model parameters of Sn2+ were recalcu-
lated from the corresponding correlation algorithms. Fig. 5 shows Sn(II) has a strong tendency to hydrolyze and forms several
that the calculated molar Gibbs free energy of formation of Sn2+ aqueous hydroxide complexes at room temperature (Cigala et al.,
at 100–500 °C and 100 MPa differs from those by Shock et al. 2012; Gamsjäger et al., 2012). At low tin levels (<0.02 molal =
(1997) by no more than 3.3 kJ mol1. 2374 ppm), Pettine et al. (1981) identified only three mononuclear

7
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Table 4
The standard state properties and revised HKF model parameters of aqueous tin species and Th4+.

Species Df GoPr ;T r (cal HoPr ;T r (cal K1 SoPr ;T r (cal K1 C op (cal K V oPr ;T r (cm3 a1*10 a2*102 a3 a4*104 c1 c2*104 xPr ;T r *105
mol1) mol1) mol1) mol1) mol1) (cal K1 (cal (cal*K mol1 (cal*K (cal K1 (cal*K (cal mol1)
mol1) mol1) bar1) mol1) mol1) mol1)
Sn4+ 600a 18,083.8x 112.9b 56.12x 61.76x 5.3731x 20.9004x 13.9648x 1.915x 9.0585x 14.4665x 3.884x
Th4+ 168,500 101c 183,800 53.54d 60.6d 5.2739x 20.6582x 13.8696x 1.9250x 8.9637x 13.9401x 3.7093x
Sn2+ 6570 2249.42 3.99 11.14x 15.5584 0.0142 7.7466 8.7948 2.4588 9.9465x 5.3042x 1.1190
SnCl+ 39,962b 41,670x 17.5442x 8.7913x 9.2717x 3.1277x 0.1446 x 5.8069x 2.7730x 13.8796 x 1.2438x 0.278x
SnCl2(aq) 72,232b 81,852x 32.2825x 13.9082x 37.0637x 6.8239 x 8.8804x 2.2598x 3.1461x 13.9676x 0.2015x 0.038x
SnCl3 103,547b 121,668x 45.0481 x 4.2091x 68.0292x 11.3814 x 20.0081x 2.1139x 3.6061x 17.0106x 2.1772x 0.9093x

The data with a superscript come from the following references: a, Wagman et al. (1968); b, Sassani and Shock (1992); c, Morss and McCue (1976); d, Hovey (1997); x, this
study. Those parameters without a superscript are from Shock et al. (1997). Using calorie in the above parameters is a tradition in the thermodynamic database SUPCRT.
1 cal = 4.184 J. P r = 0.1 MPa, T r = 298.15 K.

Fig. 5. The molar Gibbs free energy of formation of Sn2+ calculated by the HKF parameters given by Shock et al. (1997) and those updated by this study. The Gibbs free
energies of formation calculated by the updated parameters are smaller than those by Shock et al. (1997) and their difference increases with temperature.

complexes SnOH+, SnO(aq), and HSnO 2 at 20 °C and pH = 2–11. solutions at 25 °C (e.g., Bajnóczi et al., 2014). To our knowledge,
From potentiometric titrations, Raman, Mössbauer and X-ray the data of Kovalenko et al. (1992) are the only solubility data of
absorption spectroscopic data, Bajnóczi et al. (2014) found that cassiterite in alkaline solutions under reducing hydrothermal con-
Sn(OH) 
3 (here replaced by its dehydrated form HSnO2 ) was the ditions. Therefore, more hydrothermal solubility experiments
predominant Sn(II) species in hyper-alkaline (pH > 13) aqueous should be conducted to confirm or reject the high contents of
solutions at room temperature. The local structure of tin in their high-order Sn(II) hydrolytic species.
experiments was independent of the NaOH (2–12 molal) and Sn
(II) (0.05–0.2 molal) contents. Therefore, it is safe to ignore SnO2
2 5.5. Sn(IV) hydroxide complexes
and polynuclear complexes (e.g., Sn2(OH)2+ 2 ) even in hyperalkaline
solutions at room temperature. 5.5.1. The dominance of SnO2(aq) in weakly pH–acidic and neutral
The experimental data of Sn(II) hydrolytic species under solutions
hydrothermal conditions are extremely limited (Kovalenko et al., Sn(IV) also tends to be hydrolyzed in solutions and the available
1992). Kovalenko et al. (1992) measured cassiterite solubility in experimental data of Sn(IV) hydrolytic species at room tempera-
NaOH solutions (0.001–0.5 molal) at 500 °C, 100 MPa, f O2 ¼ NNO. ture are consistent (Hummel et al., 2002; Gamsjäger et al., 2012).
They speculated the dominance of SnO2 2 to explain the high con- From the independent relationship between tin concentrations
centrations of tin in alkaline solutions, but significant contents of and pH = 1.9–8 at 25 °C, the uncharged hydrolytic species SnO2(aq)
this species are inconsistent with recent findings in alkaline was interpreted to be the predominant aqueous Sn(IV) species
8
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

(Amaya et al., 1997; Gamsjäger et al., 2012). As pH increases to groups of data were conducted under unconstrained f O2 conditions
alkaline levels, tin concentrations in solutions significantly (Klintsova and Barsukov, 1973; Liu and Chen, 1986; Sorokin and
increase and HSnO 3 and SnO3
2
were interpreted to exceed Dadze, 1994; Amaya et al., 1997). It should be noted that the solu-
SnO2(aq) (Amaya et al., 1997). Rai et al. (2011) derived similar bility data in Sorokin and Dadze (1994) contain those of Dadze
results from their solubility experiments at 23 ± 2 °C. et al. (1981), so the latter is not shown here.
The influence of pH on the dominant Sn(IV) hydroxide com- The comparisons suggest that several discrepancies exist
plexes extends to hydrothermal conditions. SnO2(aq) was inter- among these experimental data (see Fig. 6). First, Liu and Chen
preted by earlier experimental studies to be the dominant Sn(IV) (1986) substantially underestimated cassiterite solubility at room
species in water at 200–400 °C (Dadze et al., 1981; Liu and Chen, temperature and higher temperatures for unknown reasons. Sec-
1986; Sorokin and Dadze, 1994) and in weakly acidic solutions at ond, the negative correlation between tin contents and tempera-
200–450 °C (Dadze et al., 1981; Sorokin and Dadze, 1994). This is ture in Klintsova and Barsukov (1973) is problematic because of
similar to the dominance of the neutral hydrolytic species GeO2(- a positive correlation suggested by most other data. Third, the
aq) of germanium (another element in Group 14 of the periodic influences of f O2 on tin contents are contradictory, with an inde-
table) in water and acidic solutions (Pokrovski and Schott, 1998; pendent relation suggested by the data of Kovalenko et al. (1986)
Pokrovski et al., 2005). From the absence of a positive correlation and a positive correlation suggested by the difference between
between tin concentrations and Cl concentrations in their experi- the data of Sorokin and Dadze (1994) and Kokh et al. (2017).
ments at 350–450 °C, Kokh et al. (2017) also raised the possibility Despite these inconsistencies, the total contents of both Sn(II)
that SnO2(aq) was the dominant tin species in low-concentration and Sn(IV) hydroxides under tin-mineralizing conditions do not
CO2 and CO2-free solutions. In alkaline solutions, Klintsova and exceed 1 ppm; therefore, aqueous tin hydroxides can be safely
Barsukov (1973) employed SnO2 3 and SnO2(aq) to interpret their ignored.
solubility data at 200 °C. Recently, Liu et al. (2020b) identified
the Raman spectrum of SnO2 3 in SnO2–NaOH–H2O experiments
5.6. Sn(II) chloride and mixed hydrolytic complexes
conducted in a hydrothermal diamond-anvil cell. The above analy-
sis suggests that SnO2(aq) is the main Sn(IV) hydrolytic species in
Both SnCl+ and SnCl2(aq) are stable in aqueous solutions at
pH neutral and acidic hydrothermal fluids.
25 °C (Séby et al., 2001; Cigala et al., 2012). Séby et al. (2001)
and Gamsjäger et al. (2012) derived their formation constants at
5.5.2. Cassiterite solubility in water 25 °C, 0.1 MPa, and infinite dilution from available experimental
The experimental data of cassiterite solubility in water are com- data using specific interaction theory (SIT):
piled in Fig. 6. Kovalenko et al. (1986) derived the tin contents in
 2n
water under two different f O2 levels, one buffered by f O2 = NNO Sn2þ þ nCl  ¼ SnCln ðn ¼ 1  4Þ ð6Þ
and the other buffered by f O2 = MM. The latter f O2 gave a high oxy- Their results are close to the voltammetric measurements con-
gen fugacity of logf O2 = 0.22 bar at 500 °C and 100 MPa, close to ducted by Cigala et al. (2012). The presence of SnCl+ and SnCl2(aq)
that under unconstrained f O2 conditions. Kokh et al. (2017) con- under hydrothermal conditions is also supported by different ana-
ducted solubility experiments under f O2 = PPM. The remaining four lytical techniques (e.g., solubility experiments, spectrophotometry,

Fig. 6. The available solubility data of cassiterite in water under hydrothermal conditions. Although several major discrepancies exist in these seven groups of experimental
data (see the text), cassiterite solubility in water at 100–500 °C does not exceed 1 ppm, indicating that either aqueous Sn(II) or Sn(IV) hydroxides play a minor role in
transporting tin in mineralizing fluids.

9
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

and molecular dynamics simulations) (Pabalan, 1986; Wilson and logK values of Reaction (6) (n = 1, 2, and 3) were calculated from
Eugster, 1990; Müller and Seward, 2001). Recent molecular their HKF model parameters and those of Sn2+ and Cl.
dynamics simulations of Wang et al. (2021) also indicate that The experimental logK values derived by Müller and Seward
SnCl2(aq) is stable at 500–700 °C and 150 MPa. (2001) are 1–4 orders of magnitude lower than those calculated
SnCl3 is another major Sn(II) chloride complex at room temper- in the present study and Barsukov et al. (1990) from their revised
ature (Cigala et al., 2012; Séby et al., 2001). The existence of SnCl3 Ryzhenko–Bryzgalin parameters, and their differences increase
in solutions at 350 °C is supported by different analytical tech- with temperature (Fig. 7). The logK values of Müller and Seward
niques (e.g., EXAFS, solubility, spectrophotometry, Raman spectra, (2001) are possibly problematic for the following reasons. First,
and molecular dynamic simulations) (Pabalan, 1986; Sherman the RB parameters of Barsukov et al. (1990) could reproduce the
et al., 2000; Müller and Seward, 2001; Chou et al., 2021; Wang solubility data of Wilson and Eugster (1990) under f O2 = NNO (cf.
et al., 2021), while its presence at higher temperatures (400– Ryzhenko et al., 1997), so using the logK values of Müller and
700 °C) is not supported by solubility experiments (Wilson and Seward (2001) would not reproduce the data of Wilson and
Eugster, 1990; Duc-Tin et al., 2007). Recent molecular dynamic Eugster (1990). Second, Müller and Seward (2001) proposed from
simulations also suggest that the dominant Sn(II) chloride complex thermodynamic calculations that SnCl 3 was the dominant Sn(II)
at 500–700 °C is SnCl2(aq) rather than SnCl 3 (Wang et al., 2021). It chloride complex at 400 °C, 50 MPa, 4.2 molal NaCl, and
should be noted that the in situ Raman spectra of Sn(II)-Cl at 500– pH = 2.4. This is inconsistent with the dominance of SnCl2(aq)
600 °C identified by Schmidt (2018) (see his Figs. 4 and 5) are sim- found by Wang et al. (2021). Third, to support the validity of their
ilar to those of SnCl3 identified by Chou et al. (2021) at 300 °C (see logK values, Müller and Seward (2001) calculated the tin content at
their Fig. 3), indicating the presence of SnCl3 in the experiments of 400 °C, 50 MPa, 4.2 molal NaCl, and pH = 2.4 to be 450 ppm, close
both studies. From the above analysis, SnCl 3 should be considered to that (381 ppm) of fluid inclusions from the Huanuni tin deposit,
under tin-mineralizing conditions. Bolivia (Müller et al., 2001). However, the pH value used by Müller
The presence of SnCl2 4 in aqueous solutions under room and and Seward (2001) is considerably more acidic than that of the
hydrothermal conditions is more controversial than other Sn(II) fluid inclusions (see Section 2). As our thermodynamic modeling
chloride complexes. Gamsjäger et al. (2012) argued that SnCl2 4 is results show later, the tin solubility in such acidic fluids is approx-
likely present at high chloride concentrations, whereas Cigala imately two orders of magnitude higher than the value derived by
et al. (2012) did not find SnCl2 4 in solutions with 2.44 molal Cl Müller and Seward (2001).
at 25 °C. Pabalan (1986) and Sherman et al. (2000) identified SnCl2- 4 The systematic difference in logK values of SnCl+ between this

in aqueous saline solutions at temperatures reaching 350 °C from study and Barsukov et al. (1990) does not exceed one order of mag-
solubility experiments and EXAFS spectroscopic data, respectively, nitude, but the differences for SnCl2(aq) and SnCl 3 at temperatures
while Wang et al. (2021) proposed from molecular dynamic simu- over 200 °C reach 1–2 orders of magnitude. The reasons for these
lations that SnCl24 has a low stability at 55 °C and is unstable at large differences will be discussed later.
300 °C. The solubility data at higher temperatures (400–700 °C)
of Wilson and Eugster (1990) and Duc-Tin et al. (2007) denied
the presence of substantial concentrations of SnCl2 5.7. Sn(IV) chloride complexes and mixed hydrolytic species
4 , and Müller
and Seward’s (2001) spectrophotometric data did not detect SnCl2 4
in concentrated chloride solutions (reaching 3 molal) above 150 °C. Sn(IV) chloride complexes SnCl4n n (n = 1–6) in aqueous solu-
Müller and Seward (2001) inferred that Pabalan (1986) data may tions are stable at room temperature (Taylor and Coddington,
have been affected by hydrolyzed Sn(II) species and that the Sn 1992; Gamsjäger et al., 2012). Taylor and Coddington (1992)
(II) solutions used by Sherman et al. (2000) may have been partly reported the NMR and Raman spectra data of SnCl4n n (n = 1–6)
oxidized. From the analysis above, SnCl2 at 20–60 °C, building a basis for the Raman peak assignments in
4 is negligible under tin-
mineralizing temperatures. later studies. The spectrophotometric data of Gajda et al. (2009)
Hydrolyzed Sn(II) can also form mixed hydroxy–chloride com- suggest that the main Sn(IV) chloride complexes in solutions >0.1

plexes (e.g., SnOHCl) at 25 °C when Cl is available (see references molal Cl at 25 °C are SnCl2 6 , SnCl5 , and SnCl4(aq). The dominance

in Müller and Seward, 2001; Séby et al., 2001; Cigala et al., of SnCl26 is supported by the EXAFS spectroscopic data of Sherman
2012). Only neutral SnOHCl was reported to exist at 25 °C (see et al. (2000) at 25–270 °C. In contrast, Chou et al. (2021) did not
Table 1 in Cigala et al., 2012). SnOHCl can account for 60% of the identify SnCl26 from in situ Raman spectra of SnCl4 + HCl solutions

total Sn(II) content at pH  4.5 but is less important than Sn(II) at 30 °C and the corresponding Sn(IV)–Cl band is composed of 55%
chloride complexes at more acidic pH levels (cf. Cigala et al., SnCl +
5 , 26% SnCl4(aq), and 19% SnCl3 (see their supplementary

2012). SnOHCl was interpreted to be a major tin-bearing complex information of S-5). Chou et al. (2021) used 0.5 molal SnCl4 and
under neutral pH levels at higher temperatures (Kovalenko et al., 0.5 molal HCl in their solutions, which are more acidic than those
1986; Li, 1988; Taylor and Wall, 1993). Duc-Tin et al. (2007) used (0.01 molal SnCl4, 0.11 molal HCl, and 2 molal NaCl) of Sherman
SnOHCl as one of two potential tin-bearing complexes in their HCl- et al. (2000). Therefore, the availability of Cl in these two groups
bearing experiments at 700 °C. At alkaline pH levels, the charged of experiments cannot explain their great discrepancy on the main
species Sn(OH)2Cl was reported to exist by Kovalenko et al. Sn(IV) chloride complexes and more reliable experimental data are
(1992), while Taylor and Wall (1993) argued that alkali chloride needed in future.
complexes (e.g., KSnOHCI2) may be important in fluids with high Five groups of solubility data of cassiterite under oxidized
concentrations of alkali chloride. Until now, there has been no hydrothermal conditions are available in the literature. Dadze
in situ experimental evidence for mixed hydroxy–chloride com- et al. (1981) and Li (1988) conducted their experiments under
plexes in acidic hydrothermal fluids (e.g., Chou et al., 2021; unconstrained f O2 conditions. The f O2 ¼ MM used by Kovalenko
Wang et al., 2021). et al. (1986) gave an oxygen fugacity of logf O2 = 0.22 bar at
The above analysis suggests that Sn(II) chloride complexes in 500 °C and 100 MPa, while the solutions used by Wilson and
tin-mineralizing fluids exist mainly as SnCl+, SnCl2(aq), and SnCl 3.
Eugster (1990) were buffered by the hematite–magnitite mineral
The standard molar properties and revised HKF model parameters assemblage. Schmidt (2018) did not used mineral buffers, thus
of these three complexes were estimated from the correlation algo- the precise f O2 levels in his experiments are unknown (probably
rithms in Supplementary Data 1. To enable comparisons with the not far from NNO in the experiments in which Ar + CH4 gas was
data of Müller and Seward (2001) and Barsukov et al. (1990), the used as flushing gas (Schmidt et al., 2021)).
10
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

The cassiterite solubility data in the same group increase with


temperature and HCl concentrations in solutions, but the solubil-
ity data among different groups under the same conditions differ
from each other (Fig. 8). For example, the solubility data in 0.1
molal HCl solutions of Dadze et al. (1981) exceed those of Li
(1988) except at 200 °C (Fig. 8). In contrast, the data of Dadze
et al. (1981) in 0.01 molal HCl solutions are approximately two
orders of magnitude lower than those of Li (1988). The reasons
for their discrepancies are unknown and at least one group of
experimental data are problematic. Fig. 9a shows that the data
of Kovalenko et al. (1986) at 500 °C are systematically lower than
those of Li (1988) and Wilson and Eugster (1990), the latter of
which are consistent with Schmidt (2018) if the influences of
fluid pressure are assumed to be minor (Fig. 9c). Note that most
data of Kovalenko et al. (1986) are lower than the detection limit.
Therefore, only the data of Wilson and Eugster (1990) and
Schmidt (2018) can be trusted.
The dominant Sn(IV) chloride complexes in high-temperature
acidic solutions are controversial. SnCl+3 was interpreted by
Wilson and Eugster (1990) to be the dominant tin species in HCl
solutions at 500–600 °C, whereas Schmidt (2018) proposed that
the most abundant complex at the same temperatures was SnCl4(-
aq) with additional SnCl+3. The HCl molalities (0.17–4.64 molal)
used by Wilson and Eugster (1990) are slightly wider than those
(1.63–4.16 molal) of Schmidt (2018). The Raman spectra of SnClo4(-
aq) and SnCl+3 were identified by Schmidt (2018), while the coordi-
nation number of Cl around Sn4+ was calculated by Wilson and
Eugster (1990) using nonlinear least squares regressions. From
the fitting results of Wilson and Eugster (1990) (see their Fig. 6),
the variation for fitting SnCl+3 to the experimental data is slightly
lower than that for SnCl4(aq). Accepting the in situ evidence of
Schmidt (2018), we interpreted that a great amount of SnCl4(aq)
existed in the solutions of Wilson and Eugster (1990), producing
the average coordination number of Cl around Sn4+ between 3
and 4. The existence of Sn(II) chloride complexes (e.g., SnCl2(aq))
in the solutions of Wilson and Eugster (1990) also lowered the

Fig. 7. Comparisons between experimental and calculated logK values of SnCl+,


SnCl2(aq), and SnCl
3 . The experimental logK values derived by Müller and Seward Fig. 8. Major discrepancies of cassiterite solubility in acidic solutions under
(2001) are 1–4 orders of magnitude lower than those calculated by this study and unconstrained f O2 conditions. The solubility data in 0.1 molal HCl solutions of
Barsukov et al. (1990) from their revised Ryzhenko-Bryzgalin parameters. The Dadze et al. (1981) exceed those of Li (1988) except the point at 200 °C, while the
reasons for these major discrepancies are possibly because the spectrophotometric data of Dadze et al. (1981) in 0.01 molal HCl solutions are approximately two orders
data of Müller and Seward (2001) underestimated the logK values of Sn(II) chloride of magnitude lower than those of Li (1988). These discrepancies suggest that at
complexes. least one group of experimental data are problematic.

11
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Fig. 9. The experimental and calculated cassiterite solubility data in HCl solutions under f O2 ¼ HM. (a, c) the calculated solubility data (model 1a) at 500 and 600 °C are
consistent with the experimental data of Wilson and Eugster (1990); (b, d) the dominant tin species in most cases is SnCl+3, followed by SnCl2(aq).

average coordination number. The above analysis suggests that the 6. Thermodynamic modeling results for cassiterite solubility
major Sn(IV) chloride complexes under hydrothermal acidic condi-
tions are possibly SnCl4(aq) and SnCl+3, and the empirical logK Cassiterite solubility under hydrothermal conditions was mod-
equation of SnCl+3 fitted by Wilson and Eugster (1990) possibly eled using a tin speciation model consisting of SnCl+, SnCl2(aq),
integrates the contributions of these two Sn(IV) chloride SnCl + 2+ 4+
3 , SnCl3, Sn , and Sn . Despite the dispute on Sn(IV) chloride
complexes. complexes (see Section 5.7), the empirical equation for the logK of
Mixed hydroxy–chloride complexes (e.g., Sn(OH)3Cl and SnCl+3 fitted by Wilson and Eugster (1990) was provisionally used
SnOHCl3) were also interpreted to be major Sn(IV)-bearing species to represent the contribution of Sn(IV) chloride complexes to cas-
in HCl solutions under oxidized hydrothermal conditions siterite solubility, especially under oxidized conditions. Table 5
(Kovalenko et al., 1986; Li, 1988; Sorokin and Dadze, 1994). lists all the chemical reactions in the models and Table 6 shows
Kovalenko et al. (1986) tried to explain their solubility data using the charge and mass balance equations. The modeling results are
SnOHCl3, but their least squares fitting was not satisfactory. Li listed in Supplementary Data 3.
(1988) proposed from his solubility data that mixed hydroxy–chlo-
ride complexes (e.g., Sn(OH)3Cl) were the dominant Sn(IV)-bearing 6.1. Cassiterite solubility in HCl solutions and f O2 = HM (Model 1a)
complex in solutions of HCl < 0.1 molal, while the dominant com-
plex in solutions of >0.1 molal HCl became SnCl4(aq). Sorokin and Cassiterite solubility in HCl solutions at 500–600 °C, 150 MPa,
Dadze (1994) also proposed that Sn(OH)3Cl may be important in and f O2 = HM was first modeled to compare with the corresponding
HCl solutions. However, those mixed complexes were identified solubility data of Wilson and Eugster (1990) and test the thermo-
neither by the EXAFS spectroscopic data of Sherman et al. (2000) dynamic parameters of aqueous tin species. Their solubility exper-
nor by the in situ Raman spectra of Chou et al. (2021). According iments correspond to the SnO2–HCl–H2O chemical system (Model
to Pearson’s hard/soft acid/base principle (Pearson, 1963), Sn4+ (a 1a; see Tables 5 and 6). The calculated tin contents at 500 °C are
hard acid) bonds preferentially with OH (a hard base) compared consistent with the experimental data of Wilson and Eugster
to Cl (a borderline base). The Cl concentrations of the solutions (1990) (Fig. 9a), and the SnCl+3 contents exceed Sn(II) chloride com-
used by Li (1988) and Sorokin and Dadze (1994) did not exceed plexes by a few orders of magnitude (Fig. 9b), except in 0.01 molal
1 molal, lower than those of Sherman et al. (2000) and Chou HCl solutions. The calculated tin contents at 600 °C are slightly
et al. (2021). Therefore, the ligand numbers of OH and Cl of greater than the experimental data of Wilson and Eugster (1990)
stable Sn(IV) hydroxy–chloride complexes may depend on the (Fig. 9c), and the dominant tin species in >0.1 molal HCl solutions
availability of Cl and further experiments are needed to support remains SnCl+3 (Fig. 9d). Note that the concentrations of Sn4+ in all
or reject their existence. models are zero.
12
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Table 5
Chemical reactions in the thermodynamic models.

Number Chemical reactions Model numbers Source of logK


+ 
1 H2O = H + OH 1a, 1b, 2a, 2b, 2c a
2 H+ + Cl = HCl(aq) 1a, 1b, 2a, 2b, 2c a
3 SnO2(s) + 2H+ = Sn2+ + H2O + 0.5O2(g) 1a, 1b, 2a, 2b, 2c a, b
4 Sn2+ + 2H+ + 0.5O2(g) = Sn4+ + H2O 1a, 1b, 2a, 2b, 2c a, b
5 SnO2(s) + 4HCl(aq) = SnCl+3 + Cl + 2H2O 1a, 1b, 2a, 2b, 2c a, b
6 Sn2+ + Cl = SnCl+ 1a, 1b, 2a, 2b, 2c a, b
7 Sn2+ + 2Cl = SnCl2(aq) 1a, 1b, 2a, 2b, 2c a, b
8 Sn2+ + 3Cl = SnCl3 1a, 1b, 2a, 2b, 2c a, b
9 Na+ + OH = NaOH(aq) 2a, 2b, 2c a, c
+ 
10 Na + Cl = NaCl(aq) 2a, 2b, 2c a, c
11 4Fe3O4(magnetite) + O2(g) = 6Fe2O3(hematite) 1a a
12 Ni(s) + 0.5O2(g) = NiO(s) 1b, 2b a
13 3SiO2(quartz) + 2Fe3O4(magnetite) = 3Fe2SiO4(fayalite) + O2(g) 2a a

a, SUPCRT; b, this study; c, Miron et al. (2016). The letters in the brackets: s, solid phase; aq, aqueous phase; g, gaseous phase.

Table 6 6.3. Cassiterite solubility in the NaCl–HCl–H2O chemical system


The charge and mass balance equations used in the models.

Number Charge and mass balance Equations Model


The cassiterite solubility under tin-mineralizing conditions was
relations numbers modeled in this section. To better compare the in situ tin contents
P P of fluid inclusions from typical Sn–W and W–Sn deposits (Fig. 2),
1 Charge balance ½cations ¼ ½anions 1a, 1b, 2a,
2b, 2c cassiterite solubility in the SnO2–NaCl–HCl–H2O chemical system
2 A given total molality of – 1a, 1b, 2a, was calculated under the following conditions: 300–500 °C, 50–
Cl 2b, 2c 150 MPa, 2 molal NaCl (1.7 molal NaCl = 10 wt.%), and pH = 3–4.
3 A fixed pH mHþ =10pH =cHþ 2a, 2b
Two mineral assemblage–buffered f O2 levels were considered in
4 An initial HCl content – 2c
5 A fixed oxygen fugacity – 2c two models, f O2 = QFM in Model 2a and f O2 = NNO in Model 2b.
In addition to aqueous tin species, other species in these two mod-
mHþ represents the molality of Hþ and cHþ is the activity coefficient of Hþ .
els include H+, Cl, HCl(aq), Na+, NaOH(aq), and NaCl(aq).
The tin contents at 400 °C, 50 MPa, 4.2 molal NaCl, and pH = 2.4
were calculated for comparison with the calculations of Müller and
6.2. Cassiterite solubility in HCl solutions and f O2 = NNO (Model 1b) Seward (2001) (see the second last paragraph in section 5.6). Under
these conditions, the tin contents were calculated to be
Cassiterite solubility in HCl solutions at 400–700 °C and 31,479 ppm under f O2 = NNO and 54,511 ppm under f O2 = QFM,
f O2 = NNO was modeled to compare with the solubility data of approximately two orders of magnitude higher than the value
Kovalenko et al. (1986), Wilson and Eugster (1990) and Duc-Tin (450 ppm) derived by Müller and Seward (2001) and the tin con-
et al. (2007). The only difference between this model and the pre- tent (381 ppm) of fluid inclusions from the Huanuni tin deposit,
Bolivia (Müller et al., 2001).
vious one is the mineral assemblage–buffered f O2 .
The contour maps under f O2 = NNO and QFM (logarithmically
The calculated solubilities at 400 °C are slightly smaller than the
corresponding experimental data of Wilson and Eugster (1990) transformed to base 10) in Fig. 11 show that cassiterite solubility
(Fig. 10a). Fig. 10b shows that SnCl2(aq) is the dominant aqueous under tin-mineralizing conditions increases with increasing tem-
species, and the contents of other Sn(II) and Sn(IV) chloride com- perature and decreasing pH. Cassiterite solubility is slightly nega-
plexes in >1 molal HCl solutions are also appreciable. The calcu- tively correlated with fluid pressure, and the negative influence
lated solubilities at 500 °C are close to the data of Wilson and of fluid pressure is stronger at higher temperatures. The tin content
Eugster (1990) and Kovalenko et al. (1986) (Fig. 10c). The domi- range (10–1000 ppm) of fluid inclusions is also shown in Fig. 11
nant tin species depends on the initial HCl contents in the solu- (see the dotted lines) for better comparisons. The calculated solu-
tions, SnCl2(aq) in <0.1 molal HCl solutions and SnCl+3 in >0.1 bility data at a fixed pH = 4 exceed 10 ppm only at temperatures
molal HCl solutions (Fig. 10d). over 350–380 °C, depending on the f O2 levels (Fig. 11a and c).
The calculated solubility data at 600 °C and 150 MPa are also When the fluid acidity decreases to pH = 3, the fluids at 300 °C
slightly smaller than the corresponding experimental data of can dissolve 10 ppm tin (Fig. 11b and d).
Wilson and Eugster (1990) except the data point in 0.302 molal The f O2 levels buffered by a mineral assemblage increase with
HCl solution (Fig. 10e). This experimental point was interpreted increasing temperature (Fig. 1), so this variable is temperature-
to be abnormal because it exceeds the tin concentrations both in dependent. Neutral pH levels also depend on temperature, and fix-
0.614 molal and 1.11 molal HCl solutions. The dominant tin species ing pH (i.e. the activity of H+) but increasing temperature would
is SnCl2(aq) in 0.1 molal HCl solutions and changes to SnCl+3 in elevate the activity of OH. To avoid these problems, we used an
>0.15 molal HCl solutions (Fig. 10f). initial HCl content in Model 2c and the f O2 levels vary from QFM
Fig. 10g compares the experimental and calculated solubility to HM. Using HCl contents also enables comparisons between
data at 700 °C and 140 MPa. The calculated values in >2 molal our modelling results and the available experimental data. The
HCl solutions are close to the experimental data of Duc-Tin et al. HCl contents of magmatic–hydrothermal fluids are poorly con-
(2007). However, the experimental data in 1 molal HCl solutions strained, especially for tin-mineralizing fluids. Piccoli et al.
are systematically higher than our calculated values, and their dif- (1999) used aplite and apatite chemistry to constrain the HCl con-
ference in 0.1–0.3 molal HCl solutions exceeds one order of magni- tents to be approximately 0.02 molal. Note that no mineralization
tude. The dominant tin species is SnCl2(aq) in 3 molal HCl occurs around the two granites studied by Piccoli et al. (1999) (cf.
solutions and changes to SnCl+3 in >3 molal HCl olutions (Fig. 10h). Kistler and Swanson, 1981), so tin-mineralizing fluids possibly
13
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Fig. 10. Comparisons between calculated and experimentally determined cassiterite solubilities in HCl solutions under 400–700 °C and f O2 ¼ NNO. The calculated solubility
data (model 1b) are consistent with most experimental data of Kovalenko et al. (1986), Wilson and Eugster (1990), and Duc-Tin et al. (2007). The dominant aqueous tin
species is SnCl2(aq), but the contents of tin(IV) chloride complexes are also appreciable and can exceed SnCl2(aq) in high HCl solutions.

14
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Fig. 11. The contour maps of cassiterite solubility (ppm) at a fixed salinity and pH. (a) pH = 4 and f O2 ¼ Q FM (model 2a); (b) pH = 3 and f O2 ¼ Q FM (model 2a); (c) pH = 4 and
f O2 ¼ NNO (model 2b); (b) pH = 3 and f O2 ¼ NNO (model 2b). The solubility data have been logarithmically transformed (base 10). The area circled by dotted lines represents
the tin content range (10–1000 ppm, see Fig. 2) of fluid inclusion from typical tin-tungsten deposits. Approximately half of the cassiterite solubilities at pH = 4 are
considerably lower than the in situ tin concentrations of fluid inclusions, while those at pH = 3 are generally higher than tin concentrations from tin-mineralizing fluids. The
solubilities at the right and lower corners are not calculated because the HKF model is not applicable in regions of low water density (<0.35 g cm3).

have higher HCl contents. HCl contents of 0.1 molal and 0.5 molal solutions occurs at lower temperatures (>380 °C) and lower f O2
are used in the Model 2c. levels (from slightly higher than NNO to HM).
The contour maps of Fig. 12a and b illustrate that the fluid
pH levels mostly lie 1.6–3.4 in 0.1 molal HCl solutions at 300–
500 °C and f O2 = QFM–HM, and they decrease to 1.0–3.0 in 0.5 7. Discussion
molal HCl solutions. These pH levels are 2–3 log units lower
than the values (pH = 4–6 in section 2) constrained by fluid- 7.1. Consistency and uncertainties of thermodynamic data of aqueous
mineral equilibrium. Consequently, cassiterite solubility in 0.1 Sn(II) and Sn(IV) species
molal HCl solutions under f O2 = QFM–NNO reach thousands of
ppm (Fig. 12c), while the solubilities under higher f O2 levels The internal consistency of thermodynamic data is vital for con-
fident calculations of aqueous geochemical processes (Tutolo et al.,
(HM) decrease to hundreds of ppm. This negative correlation
2014; Wolery and Jové Colón, 2017). Most thermodynamic data in
between cassiterite solubility and f O2 levels also exist in 0.5
the models come from the SUPCRT database, while the empirical
molal HCl solutions. From Fig. 12c and d, cassiterite solubility
logK equation of SnCl+3 in the models was fitted by Wilson and
increases with temperature and HCl contents and decreases with
Eugster (1990) from their solubility data. This compromise was
f O2 levels. The molar ratios of Sn(II) against Sn(IV) were calcu-
made because of the lack of more reliable thermodynamic data
lated to show their relative importance in transporting tin in for Sn(IV) chloride complexes. The internal consistency between
hydrothermal fluids (Fig. 12e and f). The Sn(II)/Sn(IV) molar cassiterite and aqueous tin species should be examined after reli-
ratios are negatively correlated to f O2 levels, temperature, and able thermodynamic parameters of major Sn(IV) chloride com-
HCl contents. Sn(IV) in 0.1 molal HCl solutions exceeds Sn(II) plexes are available.
only at >420 °C and f O2 levels 2–3 log units greater than NNO, The uncertainties of our modeling results stem from the corre-
while the dominance of Sn(IV) over Sn(II) in 0.5 molal HCl lation algorithms and the empirical equation of Wilson and Eugster
15
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Fig. 12. The contour maps of pH (a, b), cassiterite solubility (c, d), and Sn(II)/Sn(IV) molar ratios (e, f) under 300–500 °C and f O2 levels from QFM to NNO. The fluid pressure
and the concentrations of NaCl are fixed, and the differences between the left and right maps lie in the HCl contents (0.1 molal and 0.5 molal). The three dotted lines in each
contour map represent f O2 ¼ HM, NNO, and QFM from the top down. The solubility data and Sn(II)/Sn(IV) molar ratios have been logarithmically transformed (base 10).

(1990). Sverjensky et al. (1997) identified three major sources of Gamsjäger et al., 2012), and the remaining two sources may lead
uncertainty caused by the correlation algorithms, the logK value to an uncertainty for their logK values at 500 °C by no more than
at 25 °C and 1 bar, the equation for the revised HKF model param- one log unit (see Section 9 in Sverjensky et al., 1997). The compar-
eters (Eqs. A5–A12 in Supplementary Data 1), and the equations isons made by Sverjensky et al. (1997) showed that the discrepan-
among the standard molar properties (Eqs. A13–A20 in Supple- cies between the experimental and calculated logK values for
mentary Data 1). For Sn(II) chloride complexes, logK values at chloride complexes of bivalent cations are well within the esti-
25 °C and 1 bar are reliable (Séby et al., 2001; Cigala et al., 2012; mated uncertainties.

16
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Wilson and Eugster (1990) derived the empirical equation for Sn(II) to Sn(IV) would decrease by approximately one order of
the logK of Sn(IV) chloride complexes from the experimental data magnitude as the fluids at 400–500 °C cool by 100 °C (Fig. 12e
at 500 and 600 °C. Whether their empirical equation remains valid and f). HCl contents also alter the roles of Sn(II) and Sn(IV) in trans-
at other temperatures is unknown, but the comparisons shown in porting tin. Our modeling results indicate that Sn(IV) chloride com-
Figs. 9 and 10 suggest their empirical equation, together with the plexes are also major species in high HCl solutions (Figs. 9–12). The
revised HKF model parameters of Sn(II) chloride complexes, can influence of HCl contents on the valence of major tin species can be
reproduce most experimental solubility data at 400–700 °C. The interpreted by the chemical reaction:
following comparisons suggest that extending the empirical equa-
tion for SnCl+3 to 300 °C possibly underestimates the contents of Sn SnCl2 (aq) + 0.5O2 (g) + 2HCl(aq) = SnCl4 (aq) + H2 O ð7Þ
(IV) chlorides. From in situ Raman spectra of 0.5 molal SnCl4 + 0.5 HCl is less dissociated under hydrothermal conditions than at
molal HCl solution at 300 °C and f O2 = QFM (i.e. P H2 = 1.3 bar), Chou room temperature, so it is predominantly associated HCl(aq) in
et al. (2021) identified that almost all the Sn(IV) species were con- the model conditions (see Supplementary Data 3). According to
verted to Sn(II) within 48 hours, consistent with the dominance of Le Châtelier’s Principle, a higher content of HCl in fluids would
Sn(IV) in our modelling results under similar conditions (see make the equilibrium shift to the right side of Reaction (7), increas-
Fig. 12f). However, they identified Raman spectra of Sn(IV) at ing the contents of Sn(IV) chloride complexes. Therefore, a major
300 °C and f O2 = NNO (i.e. P H2 = 0.2 bar), indicating notable con- part of tin is possibly transported as Sn(IV) chloride complexes at
tents of Sn(IV) chloride complexes. In contrast, the Sn(IV) content an early mineralization stage, supporting the proposal of Schmidt
in our modelling results is approximately seven orders of magni- (2018). This is also a major supplement to the earlier view of the
tude lower than the Sn(II) content, leaving a minor role of Sn(IV) dominance of Sn(II) chloride complexes in tin-mineralizing fluids.
in the total tin content. Chou et al. (2021) did not detect cassiterite Wang et al.’s (2021) simulation results can be interpreted by two
precipitation after the solutions having 0.5 molal SnCl4 were possible cases. The first case is that the cassiterite they used for
heated to 300 °C, which suggests that their solutions having 0.5 Sn isotope analysis was precipitated from high-temperature and
molal Sn at 300 °C were not saturated with cassiterite. Under the high-HCl fluids although the cassiterite formation temperatures
experimental conditions of Chou et al. (2021), we calculated the are unknown. The second case is that their cassiterite precipitated
cassiterite solubility to be 0.13 molal at f O2 = NNO and 0.19 molal from oxidized fluids. For hydrothermal experiments of tin specia-
at f O2 = QFM, accounting for 26% and 30% of the tin contents in tion and cassiterite solubility, the influences of HCl contents on
solutions used by Chou et al. (2021). Therefore, more cassiterite the valence of major tin species should be considered because
solubility data under <500 °C and oxidized conditions are needed experimental HCl contents are commonly considerably greater
to refine the empirical equation of Wilson and Eugster (1990). than those in natural hydrothermal systems (cf. Piccoli et al.,
Ryzhenko et al. (1997) reported revised Ryzhenko–Bryzgalin 1999).
model parameters of aqueous Sn(II) and Sn(IV) chlorides and Most solubility experiments of cassiterite under f O2 ¼ NNO
hydroxides, but their RB model parameters failed to reproduce were intepreted using Sn(II) chloride complexes (Kovalenko
the data of Wilson and Eugster (1990) under f O2 = HM. One of pos- et al., 1986; Li, 1988; Wilson and Eugster, 1990; Duc-Tin et al.,
sible reasons for the failure is that their RB model parameters of 2007), but our modeling results under the same conditions suggest
aqueous Sn(IV) chlorides were estimated from the experimental that the main aqueous tin species would change to be Sn(IV) chlo-
data of Kovalenko et al. (1986). From Fig. 9a, most data of ride complexes in high-HCl solutions (Fig. 10). Consequently,
Kovalenko et al. (1986) are lower than the detection limit for ignoring Sn(IV) chloride complexes in the regression of solubility
unknown reasons, and their data are systematically lower than experiments under f O2 ¼ NNO (the method used by Barsukov
those of Wilson and Eugster (1990). et al. (1990)) would overestimate the logK values of Sn(II) chloride
complexes. This interprets why the logK values calculated by
Barsukov et al. (1990) are systematically greater than our calcu-
7.2. The roles of Sn(II) and Sn(IV) in hydrothermal fluids lated values (see Fig. 7 and the last paragraph of section 5.6).

It has long been accepted that tin in mineralizing fluids is trans-


7.3. Cassiterite precipitation mechanisms from hydrothermal fluids
ported mainly as Sn(II) chloride complexes (Lehmann, 1990, 2021).
However, this earlier point of view is challenged by recent in situ
Redox reactions are required for cassiterite precipitation in pre-
high-temperature experiments and molecular dynamic simula-
vious formation models (cf. Lehmann, 1990, 2021). From our mod-
tions (Schmidt, 2018; Chou et al., 2021; Wang et al., 2021). Wang
eling results (Fig. 12e and f), the contents of Sn(IV) chloride
et al. (2021) argued from Sn isotope variations in their molecular
complexes are possibly appreciable when the mineralizing fluids
dynamic simulations and natural cassiterite that tin in mineraliz-
contain high contents of HCl and their temperatures are high.
ing fluids was transported mainly as Sn(IV) species, opposing the
These conditions are more easily satisfied at an early mineraliza-
earlier view of Sn(II) dominance. Schmidt (2018) proposed from
tion stage. In this case, redox reactions are not necessary for pre-
in situ Raman data that Sn(IV) chloride complexes are capable of
cipitating cassiterite. Thus, our modeling results support Schmidt
transporting considerable amounts of tin and are possibly preva-
et al. (2018, 2021), who proposed that redox reactions play a smal-
lent tin–transporting species. The in situ Raman data of Chou
ler role than assumed in previous formation models. Also, even if
et al. (2021) suggest that notable contents of both Sn(II) and Sn
the f O2 levels are kept to be constant, simple cooling moves fluids
(IV) exist in solutions under f O2 ¼ NNO. Our modeling results sug-
to more oxidized levels (e.g., from QFM to HM) and can decrease
gest that the roles of Sn(II) and Sn(IV) in transporting tin in acidic
cassiterite solubility by up to one order of magnitude. Therefore,
hydrothermal fluids depend on f O2 , temperature, and HCl contents oxidizing agents are not necessary for precipitating cassiterite from
(Figs. 9 and 10). Sn(II)-dominated hydrothermal fluids. This can explain why
The f O2 levels are one of the key variables controlling the cassiterite-bearing quartz veins in many tin deposits are hosted
valence of tin in acidic saline fluids. As the f O2 rises from QFM to by sandstone or quartzite that do not provide oxidizing agents
HM at a constant temperature, the molar ratios of total Sn(II) to (Rozendaal et al., 1995; Liu et al., 2021; Li et al., 2022). Besides, cas-
Sn(IV) can decrease by approximately three orders of magnitude siterite precipitation from acid hydrothermal fluids is also caused
(Fig. 12e and f). Under a fixed f O2 level, the molar ratios of total by pH neutralization (e.g., greisens) (Fig. 11) and an increase in
17
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

f O2 through redox reactions, while a drop in fluid pressure cannot reduced fluids under f O2 ¼ QFM–NNO. These strict condi-
precipitate cassiterite from liquid fluids. tions decrease the tin mineralization potential of oxidized
hydrothermal fluids.
7.4. Incomplete buffering of mineralizing fluids by granitic rocks (5) The calculated solubilities under the pH levels constrained
by mineral–solution equilibrium are considerably lower
Previous studies indicate that tin-mineralizing fluids have pH than the tin contents in fluid inclusions, indicating that
levels of 4–6 (Patterson et al., 1981; Jackson and Helgeson, tin-mineralizing fluids should be more acidic (pH < 4).
1985b; Polya, 1989), which correspond to the pH levels of fluids Incomplete chemical buffering of mineralizing fluids by
in equilibrium with the mineral assemblages of quartz–muscovite. granitic rocks is necessary to carry high concentrations of
However, the calculated cassiterite solubilities at pH = 4 and tin from the source to deposition sites.
f O2 = QFM or NNO are substantially lower than the in situ tin con-
centrations of fluid inclusions from typical tin–tungsten deposits CRediT authorship contribution statement
(Fig. 11a and c). This discrepancy can be reconciled by the interpre-
tation that the pH range (4–6) constrained by mineral–solution Xiangchong Liu: Conceptualization, Investigation, Methodol-
equilibrium may represent the postmineralization pH levels and ogy, Data curation, Visualization, Funding acquisition, Software,
that the actual pH levels of mineralizing fluids should be more Validation, Writing – original draft, Writing – review & editing.
acidic (pH < 4). This comparison between our modeling results Pingping Yu: Visualization, Investigation, Resources. Changhao
and tin contents in fluid inclusions also suggests that the mineral- Xiao: Conceptualization, Funding acquisition, Writing – review &
izing fluids forming these typical tin–tungsten deposits are incom- editing.
pletely buffered by granitic rocks. This interpretation is consistent
with Heinrich (1990), who proposed from his thermodynamic Declaration of Competing Interest
modeling that incomplete chemical buffering of mineralizing fluids
by granitic rocks is a prerequisite for transporting high concentra- The authors declare that they have no known competing finan-
tions of tin from the source to the deposition sites. Although the Sn cial interests or personal relationships that could have appeared
(II) chloride complexes (e.g., SnCl2
4 ) used by Heinrich (1990) are
to influence the work reported in this paper.
problematic (see section 4.7), the conclusion of Heinrich (1990)
remains reasonable. Therefore, fluid focusing through a confined Acknowledgements
channel (such as a vein), if mineralizing fluids are available, is
vitally important for producing economic tin mineralization. The work was financially funded by CGS Research Fund
(DZLXJK202103, DZLXJK202206, DZLXJK202203), a China Geologi-
8. Conclusions cal Survey project (DD20230344), a Guizhou Provincial Science and
Technology Project (Qiankehezhicheng [2021]408), and a major
The standard molar properties and revised HKF model parame- project of Guizhou Bureau of Geology and Mineral Resources
ters of five aqueous tin species were derived from the correlation Exploration and Development (Qiandikuangkehe [2021]1). The
algorithms among these parameters, and the standard molar prop- authors thanked Jeffrey Dick for his help in using the R package
erties of cassiterite were optimized to internally consistent with CHNOSZ and Christoph Heinrich for sharing his database fitted
the available dataset. These estimated thermodynamic parameters from the experimental data of Pabalan (1986). David Wesolowski
reproduced the experimental solubility data in acidic fluids under is thanked for his patient explanation of extrapolating logK at zero
f O2 = NNO and higher f O2 conditions. Then, cassiterite solubility in ionic strength. The staff from National Geological Library of China,
especially Liu Hong, is appreciated for their help in finding the Rus-
the SnO2–NaCl–HCl–H2O chemical system was calculated under
sian publications used in this study. Christian Schmidt, Wang Xin-
typical tin–mineralizing conditions. Our modeling results provide
song, and two anonymous reviewers are thanked for their
the following implications for tin transport and cassiterite precip-
constructive comments that significantly improved the original
itation mechanisms from hydrothermal fluids:
manuscript.
(1) The total contents of Sn (II and IV) hydroxide complexes do
not exceed 1 ppm under hydrothermal conditions, consider- Appendix A. Supplementary data
ably lower than the tin contents (10–1000 ppm) of fluid
inclusions from typical tin–tungsten deposits in the world. Supplementary data to this article can be found online at
(2) The roles of Sn(II) and Sn(IV) in transporting tin in acidic https://doi.org/10.1016/j.gsf.2023.101624.
hydrothermal fluids depend on f O2 , temperature, and HCl
contents. Sn(IV) chloride complexes can be major aqueous
References
tin species in high temperature and high-HCl fluids even
under reduced conditions (e.g., f O2 ¼ NNO), so a great Akinfiev, N.N., Diamond, L.W., 2003. Thermodynamic description of aqueous
amount of Sn(IV) chloride complexes possibly exist in an nonelectrolytes at infinite dilution over a wide range of state parameters.
Geochim. Cosmochim. Acta 67 (4), 613–629.
early mineralization. In this case, redox reactions are unnec- Amaya, T., Chiba, T., Suzuki, K., Oda, C., Yoshikawa, H., Yui, M., 1997. Solubility of Sn
essary for precipitating cassiterite. (IV) oxide in dilute NaClO4 solution at ambient temperature. MRS Proceedings
(3) Even if the f O2 levels are constant, simple cooling can move 465, 751–758.
Anderson, G.M., 2005. Thermodynamics of Natural Systems. Cambridge University
mineralizing fluids to more oxidized levels and cause cassi- Press, Cambridge.
terite precipitation; therefore, oxidizing agents are not nec- Aranovich, L.Y., Newton, R.C., 1996. H2O activity in concentrated NaCl solutions at
essary for cassiterite precipitation. A simple rise in oxygen high pressures and temperatures measured by the brucite-periclase
equilibrium. Contrib. Mineral. Petrol. 125 (2), 200–212.
fugacity and acid neutralization also precipitate cassiterite Atarashiya, K., Uta, M., Shimoji, M., Niwa, K., 1960. The equilibrium between the Sn-
from acidic fluids, while a drop in fluid pressure does not. Pb liquid solution and the H2–H2O gaseous mixture. Bulletin of the Georgian
(4) Oxidized hydrothermal fluids under f O2 ¼ HM can dissolve Academy of Sciences. Chem. Soc. Japan 33 (5), 706–710.
Audétat, A., Günther, D., Heinrich, C.A., 2000. Causes for large-scale metal zonation
as much tin as fluid inclusions of typical tin-tungsten depos- around mineralized plutons: fluid inclusion LA-ICP-MS evidence from the Mole
its, but it requires fluids to contain more HCl contents than granite, Australia. Econ. Geol. 95 (8), 1563–1581.

18
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Bajnóczi, É.G., Czeglédi, E., Kuzmann, E., Homonnay, Z., Bálint, S., Dombi, G., Forgo, Helgeson, H.C., Kirkham, D.H., Flowers, G.C., 1981. Theoretical prediction of the
P., Berkesi, O., Pálinkó, I., Peintler, G., Sipos, P., Persson, I., 2014. Speciation and thermodynamic behavior of aqueous electrolytes by high pressures and
structure of tin(ii) in hyper-alkaline aqueous solution. Dalton Trans. 43 (48), temperatures: IV, Calculation of activity coefficients, osmotic coefficients, and
17971–17979. apparent molal and standard and relative partial molal properties to 600 °C and
Bannister, M.J., 1986. Oxygen concentration-cell determination of the standard 5kb. Am. J. Sci. 281 (10), 1249–1516.
molar Gibbs free energy of formation of SnO2. J. Chem. Thermodyn. 18 (5), 455– Holland, T.J.B., Powell, R., 2011. An improved and extended internally consistent
463. thermodynamic dataset for phases of petrological interest, involving a new
Barsukov, V.L., Volosov, A.G., Ryzhenko, B.N., 1990. Cassiterite stability under equation of state for solids. J. Metamorph. Geol. 29 (3), 333–383.
hydrothermal conditions: thermodynamic properties of tin compounds. Hovey, J.K., 1997. Thermodynamics of hydration of a 4+ aqueous ion: partial molar
Geokhimiya 12, 16671676 (in Russia). heat capacities and volumes of aqueous thorium(IV) from 10 to 55 °C. The J.
Belford, T.N., Alcock, C.B., 1965. Thermodynamics and solubility of oxygen in liquid Phys. Chem. B 101 (21), 4321–4334.
metals from e.m.f. measurements involving solid electrolytes. Part 2.—Tin. Hummel, W., Berner, U., Curti, E., Pearson, F.J., Thoenen, T., 2002. Nagra/PSI chemical
Transactions of the Faraday Society 61, 443–453. thermodynamic data base 01/01. Radiochim. Acta 90 (9–11), 805–813.
Berman, R.G., 1988. Internally-consistent thermodynamic data for minerals in the Jackson, K.J., Helgeson, H.C., 1985a. Chemical and thermodynamic constraints on
system Na2O-K2O-CaO-MgO-FeO-Fe2O3-Al2O3-SiO2-TiO2-H2O-CO2. J. Petrol. 29 the hydrothermal transport and deposition of tin: I. Calculation of the solubility
(2), 445–522. of cassiterite at high pressures and temperatures. Geochim. Cosmochim. Acta
Bhalla, P., Holtz, F., Linnen, R.L., Behrens, H., 2005. Solubility of cassiterite in evolved 49 (1), 1–22.
granitic melts: effect of T, fO2, and additional volatiles. Lithos 80 (1), 387–400. Jackson, K.J., Helgeson, H.C., 1985b. Chemical and thermodynamic constraints on
Borges, R.M.K., Villas, R.N.N., Fuzikawa, K., Dall’Agnol, R., Pimenta, M.A., 2009. Phase the hydrothermal transport and deposition of tin; II, Interpretation of phase
separation, fluid mixing, and origin of the greisens and potassic episyenite relations in the Southeast Asian tin belt. Econ. Geol. 80 (5), 1365–1378.
associated with the Água Boa pluton, Pitinga tin province, Amazonian Craton, Jia, H.X., Pang, Z.S., Chen, R.Y., Xue, J.-L., Chen, H., Lin, L.-J., 2019. Genesis and
Brazil. J. S. Am. Earth. Sci. 27 (2), 161–183. hydrothermal evolution of the Tiantangshan tin-polymetallic deposit, south-
Chen, J., 1986. Experiment on solubility of cassiterite in the presence of charcoal. eastern Nanling Range, South China. Geol. J. 54 (6), 3958–3979.
Geological Review 32(3), 287-294 (in Chinese with English abstract). Johnson, J.W., Oelkers, E.H., Helgeson, H.C., 1992. SUPCRT92: a software package for
Chen, J., Wang, R., Zhu, J., Lu, J., Ma, D., 2013. Multiple-aged granitoids and related calculating the standard molal thermodynamic properties of minerals, gases,
tungsten-tin mineralization in the Nanling Range, South China. Sci. China Earth aqueous species, and reactions from 1 to 5000 bar and 0 to 1000°C. Comput.
Sciences 56 (12), 2045–2055. Geosci. 18 (7), 899–947.
Chou, I.M., Wang, R., Fang, J., 2021. In situ redox control and Raman spectroscopic Kamilli, R.J., Kimball, B.E., Carlin Jr, J.F., 2017. Tin. In: Schulz, K.J., DeYoung, J.J.H.,
characterisation of solutions below 300 °C. Geochem. Perspect. Lett. 20, 1–5. Seal, R.R., Bradley, D.C. (Eds.), Critical mineral resources of the United States—
Cigala, R.M., Crea, F., De Stefano, C., Lando, G., Milea, D., Sammartano, S., 2012. The Economic and environmental geology and prospects for future supply. U.S.
inorganic speciation of tin(II) in aqueous solution. Geochim. Cosmochim. Acta Geological Survey Professional Paper 1802, 797 p.
87, 1–20. Kapustinskii, A., Zil’berman, A., Veselovskii, B., 1936. Chemical equilibria in
Cox, J.D., Wagman, D.D., Medvedev, V.A., 1989. CODATA key values for inorganic systems. IV. System: tin-carbon-oxygen, Tr. Vses, Nauchno-Issled.
thermodynamics. Hemisphere Publishing Corp, New York, p. 271. Inst. miner. Syr’ya 109 (1936), 68–94.
Dadze, T.P., Sorokin, V.I., Nekrasov, I.Y., 1981. Solubility of SnO2 in water and in Kinnaird, J.A., Nex, P.A., Milani, L., 2016. Tin in Africa. Episodes 39 (2), 361–380.
aqueous solutions of HCl, HCl+ KCl, and HNO3 at 200–400 oC and 1013 bar. Kistler, R.W., Swanson, S.E., 1981. Petrology and geochronology of metamorphosed
Geochem. Int. 18 (5), 142–152. volcanic rocks and a Middle Cretaceous volcanic neck in the east-central Sierra
Dewaele, S., De Clerq, F., Muchez, P., Scheider, J., Burgess, R., Boyce, A., Alonso, Nevada, California. J. Geophys. Res. (Solid Earth) 86 (B11), 10489–10501.
M.F., 2010. Geology of the cassiterite mineralisation in the Rutongo area, Klintsova, A.P., Barsukov, V.L., 1973. Solubility of cassiterite in water and in aqueous
Rwanda (Central Africa); current state of knowledge. Geol. Belg. 13 (1–2), NaOH solutions at elevated temperatures. Geochem. Int. 10 (5), 540–546.
91–111. Klintsova, A.P., Barsukov, V.L., Shemarykina, T.P., Khodakovskiy, I.L., 1975.
Dick, J.M., 2019. CHNOSZ: thermodynamic calculations and diagrams for Measurment of the stability constants of Sn(IV) hydroxofluoride complexes.
geochemistry. Front. Earth Sci. 7 (180), 1–18. Geochemistry Int. 12, 207–215.
Dolejš, D., 2013. Thermodynamics of aqueous species at high temperatures and Kokh, M.A., Akinfiev, N.N., Pokrovski, G.S., Salvi, S., Guillaume, D., 2017. The role of
pressures: Equations of state and transport theory. Rev. Mineral. Geochem. 76 carbon dioxide in the transport and fractionation of metals by geological fluids.
(1), 35–79. Geochim. Cosmochim. Acta 197(Supplement C), 433–466.
Dorofeeva, V.A., Kovalenko, N.I., Ryzhenko, B.N., 1994. The SnO2-HF-NaF-H2O Kokh, M.A., Lopez, M., Gisquet, P., Lanzanova, A., Candaudap, F., Besson, P.,
system at 500 °C, 1 kbar, and hydrogen fugacity of the buffer Ni/NiO: Pokrovski, G.S., 2016. Combined effect of carbon dioxide and sulfur on vapor–
refinement of the stability constant of Sn(II) fluorine complexes and HF 2. liquid partitioning of metals in hydrothermal systems. Geochim. Cosmochim.
Geochem. Int. 5, 755–759 (in Russia). Acta 187(Supplement C), 311–333.
Dubessy, J., Ramboz, C., Nguyen-Trung, C., Cathelineau, M., Charoy, B., Cuney, M., Korges, M., Weis, P., Lüders, V., Laurent, O., 2017. Depressurization and boiling of a
Leroy, J., Poty, B., Weisbrod, A., 1987. Physical and chemical controls (fO2, T, pH) single magmatic fluid as a mechanism for tin-tungsten deposit formation.
of the opposite behaviour of U and Sn-W as examplified by hydrothermal Geology 46 (1), 75–78.
deposits in France and Great-Britain, and solubility data. Bull. Georg. Acad. Sci. Korges, M., Weis, P., Lüders, V., Laurent, O., 2020. Sequential evolution of Sn–Zn–In
Mineral. 110 (2), 261–281. mineralization at the skarn-hosted Hämmerlein deposit, Erzgebirge, Germany,
Duc-Tin, Q., Audétat, A., Keppler, H., 2007. Solubility of tin in (Cl, F)-bearing aqueous from fluid inclusions in ore and gangue minerals. Mineral. Deposita 55, 937–
fluids at 700°C, 140MPa: A LA-ICP-MS study on synthetic fluid inclusions. 952.
Geochim. Cosmochim. Acta 71 (13), 3323–3335. Kovalenko, N.I., Dorofeeva, V.A., Velyukhanova, T.R., 1989. Thermodynamic
Farges, F., Linnen, R.L., Brown Jr., G.E., 2006. Redox and speciation of tin in hydrous properties of Sn(ll) complexes in supercritical acidic fluoride solutions. Sci.
silicate glasses: a comparison with Nb, Ta, Mo and W. Can. Mineral. 44 (3), 795– Géol. Bull. Mémoires 42 (4), 343–350.
810. Kovalenko, Ryzhenko, B.N., Barsukov, V.L., Klintsova, A.P., Velyukhanova, T.K., 1986.
Gajda, T., Sipos, P., Gamsjäger, H., 2009. The standard electrode potential of the Sn4+/ Experimental study of cassiterite solubility in solutions of HCl and HCl+NaCl
Sn2+ couple revisited. Monatshefte für Chemie - Chemical Monthly 140 (11), (KCl) solutions at 500 °C and 1000 atm under fixed redox conditions.
1293–1303. Geokhimiya 23(7), 1-9.
Gamsjäger, H., Gajda, T., Sangster, J., Saxena, S.K., Voigt, W., 2012. Chemical Kovalenko, N.I., Ryzhenko, B.N., Dorofeeva, V.A., Bannvkh, L.N., 1992. The stability of
 
Thermodynamics of Tin. North Holland Elsevier Science Publishers, Amsterdam, Sn(OH)24 , Sn(OH)2F , Sn(OH)2Cl complexes at 500 °C and 1 kbar. Geochem. Int
p. 646. 29 (8), 84–94.
Gurevich, V.M., Gavrichev, K.S., Polyakov, V.B., Clayton, R.N., Mineev, S.D., Hu, G., Legros, H., Richard, A., Tarantola, A., Kouzmanov, K., Mercadier, J., Vennemann, T.,
Gorbunov, V.E., Golushina, L.N., 2004. Low-temperature heat capacity of tin Marignac, C., Cuney, M., Wang, R.C., Charles, N., Bailly, L., Lespinasse, M.Y., 2019.
dioxide: new standard data on thermodynamic functions. Thermochim. Acta Multiple fluids involved in granite-related W-Sn deposits from the world-class
421 (1), 179–184. Jiangxi province (China). Chem Geol. 508, 92–115.
Haas, J.L., Fisher, J.R., 1976. Simultaneous evaluation and correlation of Lehmann, P.D.B., 1990. Metallogeny of Tin. Springer Verlag, Berlin, p. 211.
thermodynamic data. Am. J. Sci. 276 (4), 525–545. Lehmann, B., 2021. Formation of tin ore deposits: a reassessment. Lithos 402–403
Halter, W.E., Williams-Jones, A.E., 1995. Origin and evolution of the greisenizing (15), 105756.
fluid at the East Kemptville tin deposit, Nova Scotia, Canada. Econ. Geol. 93 (7), Li, T.J., 1988. Experimental studies of the solubility of cassiterite and the extraction
1026–1051. of tin from grantic melt. Geochimica 17 (2), 150–160 (in Chinese with English
Halter, W.E., Williams-Jones, A.E., Kontak, D.J., 1998. Modeling fluid–rock abstract).
interaction during greisenization at the East Kemptville tin deposit: Li, Y., Zhang, R.-Q., He, S., Chiaradia, M., Li, X.-H., 2022. Pulsed exsolution of
implications for mineralization. Chem. Geol. 150 (1–2), 1–17. magmatic ore-forming fluids in tin-tungsten systems: a SIMS cassiterite oxygen
Han, L., Pan, J.-Y., Ni, P., Chen, H., 2023. Cassiterite deposition induced by cooling of isotope record. Mineral. Deposita 57, 343–352.
a single-phase magmatic fluid: Evidence from SEM-CL and fluid inclusion LA- Linnen, R.L., 1998. Depth of emplacement, fluid provenance and metallogeny in
ICP-MS analysis. Geochim. Cosmochim. Acta 342, 108–127. granitic terranes: a comparison of western Thailand with other tin belts.
Heinrich, C.A., 1990. The chemistry of hydrothermal tin(-tungsten) ore deposition. Mineral. Deposita 33 (5), 461–476.
Econ. Geol. 85 (3), 457–481. Linnen, R.L., Pichavant, M., Holtz, F., Burgess, S., 1995. The effect of f O2 on the
Helgeson, H.C., 1978. Summary and critique of the thermodynamic properties of solubility, diffusion, and speciation of tin in haplogranitic melt at 850°C and 2
rock-forming minerals. Am. J. Sci. 278, 1–229. kbar. Geochim. Cosmochim. Acta 59 (8), 1579–1588.

19
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Linnen, R.L., Pichavant, M., Holtz, F., 1996. The combined effects of fO2 and melt Polya, D.A., 1989. Chemistry of the main-stage ore-forming fluids of the Panasqueira
composition on SnO2 solubility and tin diffusivity in haplogranitic melts. W-Cu(Ag)-Sn deposit, Portugal; implications for models of ore genesis. Econ.
Geochim. Cosmochim. Acta 60 (24), 4965–4976. Geol. 84 (5), 1134–1152.
Liu, Y.S., Chen, S.Q., 1986. An experimental study on cassiterite solubility and tin Rai, D., Yui, M., Schaef, H.T., Kitamura, A., 2011. Thermodynamic model for SnO2(cr)
transport during mineralization. Acta Geol. Sin. 60 (1), 78–88 (in Chinese with and SnO2(am) solubility in the aqueous Na+–H+–OH–Cl–H2O system. J.
English abstract). Solution Chem. 40 (7), 1155.
Liu, Y.C., Li, J.K., Chou, I.M., 2020b. Cassiterite crystallization experiments in alkali Reed, J.J., 2020. Digitizing ‘‘The NBS tables of chemical thermodynamic properties:
carbonate aqueous solutions using a hydrothermal diamond-anvil cell. Am. selected values for inorganic and C1 and C2 organic substances in SI units”. J.
Mineral. 105 (5), 664–673. Res. Natl. Inst. Stand. Technol. 125, 125007.
Liu, P., Mao, J., Jian, W., Mathur, R., 2020a. Fluid mixing leads to main-stage Robie, R.A., Hemingway, B.S., 1995. Thermodynamic properties of minerals and
cassiterite precipitation at the Xiling Sn polymetallic deposit, SE China: related substances at 298.15 K and 1 bar (105 Pascals) pressure and at higher
evidence from fluid inclusions and multiple stable isotopes (H–O–S). Mineral. temperatures. US Geol. Surv. Bull. 2131, 461.
Deposita 55, 1233–1246. https://doi.org/10.1007/s00126-019-00933-0. Robie, R.A., Hemingway, B.S., Fisher, J.R., 1978. Thermodynamic properties of
Liu, X., Wang, W., Zhang, D., 2021. The mechanisms forming the five–floor zonation minerals and related substances at 298.15K and 1 bar pressure and at higher
of quartz veins: A case study in the Piaotang tungsten–tin deposit, southern temperatures. U.S. Geol. Surv. Bull. 1452, 456.
China. Minerals 11 (8), 883. Romer, R.L., Kroner, U., Schmidt, C., Legler, C., 2022. Mobilization of tin during
Liu, X.C., Xiao, C.H., 2020. Wolframite solubility and precipitation in hydrothermal continental subduction-accretion processes. Geology 50 (12), 1361–1365.
fluids: Insight from thermodynamic modeling. Ore Geol. Rev. 117, 103289. Rozendaal, A., Misiewicz, J.E., Scheepers, R., 1995. The tin zone: sediment-hosted
Li-Zi, Y., Zhi-Tong, S., Chan-Zheng, W., 1994. A thermodynamic study of tin oxides hydrothermal tin mineralization at Rooiberg, South Africa. Mineral. Deposita 30
by coulometric titration. J. Solid State Chem. 113 (2), 221–224. (2), 178–187.
Lothenbach, B., Ochs, M., Wanner, H., Yui, M., 1999. Thermodynamic Date for the Ryzhenko, B.N., Shvarov, Y.V., Kovalenko, N.I., 1997. The Sn-Cl-F-C-S-H-O-Na
Speciation and Solubility of Pd, Pb, Sn, Sb, Nb, and Bi in Aqueous Solution. JNC, system: Thermodynamic properties of components within the conditions of
340. the earth’s crust. Geochem. Int. 35, 1016–1020.
Mallika, C., Edwin Suresh Raj, A.M., Nagaraja, K.S., Sreedharan, O.M., 2001. Use Sassani, D.C., Shock, E.L., 1992. Estimation of standard partial molal entropies of
of SnO for the determination of standard Gibbs energy of formation of SnO2 aqueous ions at 25°C and 1 bar. Geochim. Cosmochim. Acta 56 (11), 3895–3908.
by oxide electrolyte e.m.f. measurements. Thermochimica Acta 371(1), Schaefer, S.C., 1984. Electrochemical determination of thermodynamic properties of
95-101. bismuth sesquioxide and stannic oxide. US Department of the Interior, Bureau
Mangas, J., Arribas, A., 1987. Fluid inclusion study in different types of tin deposits of Mines Report of Investigations.
associated with the Hercynian granites of western Spain. Chem. Geol. 61 (1), Schmidt, C., 2018. Formation of hydrothermal tin deposits: Raman spectroscopic
193–208. evidence for an important role of aqueous Sn(IV) species. Geochim. Cosmochim.
Mao, J.W., Ouyang, H., Song, S.W., Santosh, M., 2019. Geology and metallogeny of Acta 220 (Suppl. C), 499–511.
tungsten and tin deposits in China. Rev. Econ. Geol. 22, 411–482. Schmidt, C., Gottschalk, M., Zhang, R., Lu, J., 2021. Oxygen fugacity during tin ore
Meinert, L.D., Dipple, G.M., Nicolescu, S., 2005. World Skarn Deposits. In: deposition from primary fluid inclusions in cassiterite. Ore Geol. Rev. 139,
Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richards, J.P. (Eds.), One 104451.
Hundredth Anniversary Volume. Society of Economic Geologists, pp. 299–336. Séby, F., Potin-Gautier, M., Giffaut, E., Donard, O.F.X., 2001. A critical review of
Migdisov, A.A., Williams-Jones, A.E., 2005. An experimental study of cassiterite thermodynamic data for inorganic tin species. Geochim. Cosmochim. Acta 65
solubility in HCl-bearing water vapour at temperatures up to 350 °C. (18), 3041–3053.
Implications for tin ore formation. Chem. Geol. 217 (1–2), 29–40. Sherman, D.M., Ragnarsdottir, K.V., Oelkers, E.H., Collins, C.R., 2000. Speciation of tin
Miron, G.D., Wagner, T., Kulik, D.A., Heinrich, C.A., 2016. Internally consistent (Sn2+ and Sn4+) in aqueous Cl solutions from 25°C to 350°C: an in situ EXAFS
thermodynamic data for aqueous species in the system Na–K–Al–Si–O–H–Cl. study. Chem. Geol. 167 (1), 169–176.
Geochim. Cosmochim. Acta 187, 41–78. Shock, E.L., Sassani, D.C., Willis, M., Sverjensky, D.A., 1997. Inorganic species in
Mlynarczyk, M.J., Sherlock, R., Williams-Jones, A., 2003. San Rafael, Peru: geology geologic fluids: Correlations among standard molal thermodynamic properties
and structure of the worlds richest tin lode. Mineral. Deposita 38 (5), 555–567. of aqueous ions and hydroxide complexes. Geochim. Cosmochim. Acta 61 (5),
Morss, L.R., McCue, M.C., 1976. Partial molal entropy and heat capacity of the 907–950.
aqueous thorium(IV) ion. Thermochemistry of thorium nitrate pentahydrate. J. Soetaert, K., Herman, P.M.J., 2008. A Practical Guide to Ecological Modelling - Using
Chem. Eng. Data 21 (3), 337–341. R as a Simulation Platform. Springer, Dordrecht, p. 372.
Müller, B., Frischknecht, R., Seward, T., Heinrich, C., Camargo Gallegos, W., 2001. A Sorokin, V.I., Dadze, T.y.P., 1994. Solubility and complex formation in the systems
fluid inclusion reconnaissance study of the Huanuni tin deposit (Bolivia), using Hg-H2O, S-H2O, SiO2-H2O and SnO2-H2O. In: Shmulovich, K.I., Yardley, B.W.D.,
LA-ICP-MS micro-analysis. Mineral. Deposita 36 (7), 680–688. Gonchar, G.G. (Eds.), Fluids in the Crust. Springer, pp. 57-93.
Müller, B., Seward, T.M., 2001. Spectrophotometric determination of the stability of Štemprok, M., 1987. Greisenization (a review). Geol. Rundsch. 76 (1), 169–175.
tin(II) chloride complexes in aqueous solution up to 300°C. Geochim. Sverjensky, D.A., Shock, E.L., Helgeson, H.C., 1997. Prediction of the thermodynamic
Cosmochim. Acta 65 (22), 4187–4199. properties of aqueous metal complexes to 1000°C and 5 kb. Geochim.
Myint, A.Z., Yonezu, K., Boyce, A.J., Selby, D., Scherstén, A., Tindell, T., Watanabe, K., Cosmochim. Acta 61 (7), 1359–1412.
Swe, Y.M., 2018. Stable isotope and geochronological study of the Mawchi Sn-W Tanger, J.C., Helgeson, H.C., 1988. Calculation of the thermodynamic and transport
deposit, Myanmar: Implications for timing of mineralization and ore genesis. properties of aqueous species at high pressures and temperatures; revised
Ore Geol. Rev. 95, 663–679. equations of state for the standard partial molal properties of ions and
Naumov, V.B., Dorofeev, V.A., Mironova, O.F., 2011. Physicochemical parameters of electrolytes. Am. J. Sci. 288 (1), 19–98.
the formation of hydrothermal deposits: a fluid inclusion study. I. tin and Taylor, M.J., Coddington, J.M., 1992. The constitution of aqueous tin(IV) chloride and
tungsten deposits. Geochem. Int. 49 (10), 1002–1021. bromide solutions and solvent extracts studied by 119Sn NMR and vibrational
Pabalan, R.T., 1986. Solubility of cassiterite (SnO2) in NaCl solutions from 200 °C- spectroscopy. Polyhedron 11 (12), 1531–1544.
350 °C, with geologic applications. Pennsylvania State University. Taylor, J.R., Wall, V.J., 1993. Cassiterite solubility, tin speciation, and transport in a
Patterson, D.J., Ohmoto, H., Solomon, M., 1981. Geologic setting and genesis of magmatic aqueous phase. Econ. Geol. 88 (2), 437–460.
cassiterite-sulfide mineralization at Renison Bell, western Tasmania. Econ. Geol. Tornos, F., 1997. A numerical approach for the formation of chlorite and sulphide-
76 (2), 393–438. beating greisen: a study based on the Navalcubilla system (Spanish Central
Pearson, R.G., 1963. Hard and soft acids and bases. J. Am. Chem. Soc. 85 (22), 3533– System). Mineral. Mag. 61 (408), 639–654.
3539. Tutolo, B.M., Kong, X.-Z., Seyfried, W.E., Saar, M.O., 2014. Internal consistency in
Persson, I., D’Angelo, P., Lundberg, D., 2016. Hydrated and solvated tin(ii) ions in aqueous geochemical data revisited: Applications to the aluminum system.
solution and the solid state, and a coordination chemistry overview of the d10s2 Geochim. Cosmochim. Acta 133, 216–234.
metal ions. Chemistry – A European Journal 22 (51), 18583–18592. Van Daele, J., Hulsbosch, N., Dewaele, S., Boiron, M.C., Piessens, K., Boyce, A.,
Petot-Ervas, G., Farhi, R., Petot, C., 1975. Standard Gibbs free energy of formation of Muchez, P., 2018. Mixing of magmatic-hydrothermal and metamorphic fluids
SnO2 from high-temperature e.m.f. measurements. J. Chem. Thermodyn. 7 (12), and the origin of peribatholitic Sn vein-type deposits in Rwanda. Ore Geol. Rev.
1131–1136. 101, 481–501.
Pettine, M., Millero, F.J., Macchi, G., 1981. Hydrolysis of tin(II) in aqueous solutions. Wagman, D.D., Evans, W.H., Parker, V.B., Halow, I., Bailey, S.M., Schumm, R.H., 1968.
Anal. Chem. 53 (7), 1039–1043. Selected values of chemical thermodynamic properties, tables for the first
Piccoli, P.M., Candela, P.A., Williams, T.J., 1999. Estimation of aqueous HCl and Cl thirty-four elements in the standard order of arrangement. National Bureau of
concentrations in felsic systems. Lithos 46 (3), 591–604. Standards Technical Note(270–3), p. 264.
Pirajno, F., 1992. Greisen systems. In: Pirajno, F. (Ed.), Hydrothermal Mineral Wagman, D.D., Evans, W.H., Parker, V.B., Schumm, R.H., Halow, I., Bailey, S.M.,
Deposits: Principles and Fundamental Concepts for the Exploration Geologist. Churney, K.L., Nuttall, R.L., 1982. The NBS tables of chemical thermodynamic
Springer, Berlin Heidelberg, Berlin, Heidelberg, pp. 280–324. properties. Selected values for inorganic and C1 and C2 organic substances in SI
Pokrovski, G.S., Roux, J., Hazemann, J.-L., Testemale, D., 2005. An X-ray absorption units. J. Phys. Chem. Ref. Data 11 (Suppl. 2), 1–407.
spectroscopy study of argutite solubility and aqueous Ge(IV) speciation in Wang, T.H., She, J.X., Yin, K., Wang, K., Zhang, Y.J., Lu, X.C., Liu, X.D., Li, W.Q., 2021. Sn
hydrothermal fluids to 500 °C and 400 bar. Chem. Geol. 217(1), 127-145. (II) chloride speciation and equilibrium Sn isotope fractionation under
Pokrovski, G.S., Schott, J., 1998. Thermodynamic properties of aqueous Ge(IV) hydrothermal conditions: A first principles study. Geochim. Cosmochim. Acta
hydroxide complexes from 25 to 350°C: implications for the behavior of 300, 25–43.
germanium and the Ge/Si ratio in hydrothermal fluids. Geochim. Cosmochim. Wilson, G.A., Eugster, H.P., 1990. Cassiterite solubility and tin speciation in super
Acta 62 (9), 1631–1642. critical chloride solutions. Geochem. Soc. Spec. Publ. 2, 179–195.

20
X. Liu, P. Yu and C. Xiao Geoscience Frontiers 14 (2023) 101624

Wolery, T.J., Jové Colón, C.F., 2017. Chemical thermodynamic data. 1. The concept of Yang, Y., Zhang, Y., Simon, A., Ni, P., 2016. Cassiterite dissolution and Sn diffusion in
links to the chemical elements and the historical development of key silicate melts of variable water content. Chem. Geol. 441, 162–176.
thermodynamic data. Geochim. Cosmochim. Acta 213, 635–676. Zhao, P., Zajacz, Z., Tsay, A., Yuan, S., 2022. Magmatic-hydrothermal tin deposits
Xiang, L., Wang, R.C., Erdmann, S., Sizaret, S., Lu, J.J., Zhang, W.L., Xie, L., Che, X.D., form in response to efficient tin extraction upon magma degassing. Geochim.
Zhang, R.Q., 2018. Neoproterozoic mineralization in a hydrothermal cassiterite- Cosmochim. Acta 316, 331–346.
sulfide deposit at Jiumao, northern Guangxi, South China: Mineral-scale Zhogin, D.Y., Kosarukina, E.A., Kolesov, V.P., 1980. Heat capacity in the 10–300 K
constraints on metal origins and ore-forming processes. Ore Geol. Rev. 94, range and thermodynamic functions of tetragonal tin dioxide. Zhurnal
172–192. Fizicheskoi Khimii 54 (4), 916–920.

21

You might also like