Вы находитесь на странице: 1из 9

Author's personal copy

Journal of Geochemical Exploration 124 (2013) 92100

Contents lists available at SciVerse ScienceDirect

Journal of Geochemical Exploration


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

Oxygen isotope equilibrium in sulfatewater systems: A revision of geothermometric


applications in low-enthalpy systems
Tiziano Boschetti
Department of Earth Sciences, University of Parma, viale G.P. Usberti 157/a, I-43100 Parma, Italy

a r t i c l e

i n f o

Article history:
Received 16 February 2012
Accepted 11 August 2012
Available online 21 August 2012
Keywords:
Low-enthalpy resources
Isotope
Fractionation
Sulfatewater equilibrium
Geothermometers
Aqueous speciation

a b s t r a c t
Since approximately forty years ago, the equation relating the oxygen isotope fractionation between bisulfate
and water has been extensively used as a geothermometer for acidic and high-enthalpy hydrothermal systems (T > 150 C). Its application to near-neutral or basic, low-enthalpy (T b 150 C) systems was considered
to be unreliable. Recently, empirical and theoretical geothermometric equations were formulated for the
(SO42 H2O) oxygen isotope system, opening new perspectives on isotope geochemistry applied to geothermal research. In the present paper, these geothermometers are applied while revisiting a variety of hydrological case studies. In addition to the equilibrium fractionation, isotopic effects on geothermometric estimates
due to mixingdilution, salinity, sulde oxidation, sulfate reduction, standardization of the analytical results
and aqueous speciation are also discussed.
2012 Elsevier B.V. All rights reserved.

1. Introduction
The geothermal resources of the earth are huge, with low-enthalpy
(Tb 150 C) resources having a much larger potential and a much wider
regional distribution than high-enthalpy resources. Continuous development of innovative drilling and power generation technologies makes
low-enthalpy geothermal resources the best option available to meet future electricity demands (e.g., Chandrasekharam and Bundschuh, 2008).
Estimating reservoir temperatures before beginning of any drilling activity is an important task in geothermal exploration. The most accurate
temperature data are obtained by static, bottom-hole temperature
surveys in exploratory or production wells. However, this method of temperature surveying, although often inevitable, is the most expensive due
to the need for costly investment in exploratory drilling (e.g., Porowski
and Dowgiao, 2009). The application of chemical and isotopic groundwater temperature indicators, the so-called geothermometers, allows a
considerable reduction in the cost of geothermal exploration. Chemical
and isotopic geothermometers are equations or models based on
temperature-dependent chemical reactions or isotope equilibrium
fractionation relations from which the equilibrium temperatures
of these reactions can be calculated. The majority of known
geothermometers were developed and calibrated for high-enthalpy
geothermal systems, so their application to low-enthalpy geothermal
systems always raises controversies, especially for sulfate isotope
geothermometers that take a long time to reach equilibrium with
water molecules at low temperatures and a basic pH.
This paper shows the results obtained through the application of
recently published sulfatewater fractionation equations, here applied
E-mail address: work@tizianoboschetti.com.
0375-6742/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.gexplo.2012.08.011

as isotopic geothermometers to various low-enthalpy hydrogeothermal


systems of the world.
2. Isotope fractionation factors
Sulfatewater oxygen isotope thermometry is based on the equilibrium attained through an exchange reaction between the two phases. From
a theoretical point of view and prescinding from the abundance of
isotopologues, the equilibrium equation between SO42 and H2O can be
written as:
1

.
4

16

18

S O4 H2 O 1

.
4

18

16

S O4 H2 O:

Similar forms of relations are valid also for other dissolved sulfate
species as HSO4 and CaSO40. At equilibrium, the isotopic relation between sulfate and water in terms of 18O values, expressed in per
mil relative to a V-SMOW standard, is described by the fractionation
factor :

i h

 h
i
2
3
18
2
3
18
= 10 OH2 O
SO4 H2 O 10 O SO4

and the general thermometric equation is:




3
2
2
10 ln SO4 H2 O A=T B

where temperature T is in Kelvin, and A and B are constants


depending on the experimental, empirical or theoretical type of equation (e.g., Hoefs, 2009; Seal et al., 2000).

Author's personal copy


T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

Experimental fractionation factors for the HSO4H2O system in the


70350 C range were determined by Lloyd (1968) and Mizutani and
Rafter (1969) and then combined by Seal et al. (2000) in the equation
(N= 12):

10 ln HSO4 H2 Oexperimental 3:2610 =T 5:81:

Over the last forty years, HSO4H2O isotopic fractionation was extensively employed as a geothermometer in acidic, high-enthalpy geothermal systems (T>150 C; e.g., Cortecci, 1974), sometimes with minor
adjustments (Giggenbach et al., 1983; McKenzie and Truesdell, 1977).
This geothermometer was used also in low-enthalpy (Tb 150 C),
near-neutral or basic water (e.g., Sonney and Vuataz, 2010; Zhonge,
2001) in which SO42 was the dominant sulfur species in solution. Often
in basic texts of applied geothermometry, sulfate (SO42) and bisulfate
(HSO4) are not distinguished (e.g., D'Amore and Arnrsson, 2000;
Nicholson, 1993) possibly because at 300 C>T>250 C, bisulfate prevails also at near-neutral pH (Fig. 1) as hypothesized by McKenzie and
Truesdell (1977). However, it should be stressed that HSO4 is restricted
to low pH and oxidizing solutions (pHb 2 and pHb 4 at 25 C and
150 C, respectively; Fig. 1), and, therefore, unreliably estimated temperatures can be obtained out of these conditions for low-temperature
systems.
The problem connected with the shortcomings of the SO42H2O
fractionation equation was partially overcome by Sakai (1977), who
recommended the use of the BaSO4(s)H2O equation (Kusakabe and
Robinson, 1977) in near-neutral or basic, high-enthalpy geothermal
systems to mimic the isotopic effect on dissolved SO42:
3

10 ln BaSO4 H2 Oexperimental 3:0110 =T 7:30:

Successively, experimental studies (Chiba and Sakai, 1985; Hoering


and Kennedy, 1957; Lloyd, 1968; Seal et al., 2000) showed that the
rate of the oxygen isotope exchange between water and sulfate species

S = 10

25

HSO4

93

i) increases rapidly with increasing temperature and acidity and ii) is


controlled by the HSO4H2O exchange also at near-neutral pH. This latter control on exchange conicts with the ndings of Sakai (1977) on
Japanese geothermal waters with near-neutral pH and temperatures of
200 to 300 C, in which the role of the SO42H2O pair in the isotopic exchange is competitive with the HSO4H2O one.
As far as the time required for sulfatewater isotopic equilibrium,
Chiba and Sakai (1985) suggested that 97% equilibrium in ambient seawater (pH =8, T= 4 C) and deep-sea interstitial water should require
107109 years, which is much higher than the 105 years calculated by
Lloyd (1968). Biological processes such as sulfate reduction, disproportionation of elemental sulfur and thiosulfate, and sulde oxidation can
enhance the isotopic exchange through the formation of sulte as an intermediate. Unlike sulfate, sulte readily exchanges oxygen isotopes
with water (e.g., van Stempvoort and Krouse, 1994), and then its oxidation will produce sulfate with an oxygen isotope signature largely
inherited from sulte (Brunner et al., 2006; Turchyn et al., 2010).
Apart from the sultewater exchange, the residence time of SO42 in
sedimentary groundwaters (i.e., connate or formation brines) can be
long enough to attain equilibrium with water at temperatures even
below 150 C. Halas and Pluta (2000) investigated this possibility by
analyzing salt waters from different sites; the correlation between the
sulfatewater fractionation and the measured temperature (T b 50 C)
was good (R2 =0.85, N=17) and was described by the equation:


3
2
10 ln SO4 H2 O

empirical

2:4110 =T 5:77:

This equation appears to be comparable with the theoretical relationship recently formulated by Zeebe (2010) within the temperature
range of 0 to 150 C (the intersection between the two equations occurs at 114 C, Fig. 2):


3
2
10 ln SO4 H2 O

theoretical

2:6810 =T 7:45:

-3

15

0C

HSO4

25

Eh (Volts)

.5
30

S0

SO4

2-

H2S(aq)

.5

HS
HS

10

12

2-

14

pH
Fig. 1. Pourbaix plot for aqueous species in the system SO2H2O at 25 C and 1.013 bar
(surface conditions; black lines) and 150 C and 4.665 bar (vapor pressure of water; gray
lines). The HSO4SO42 boundaries (dotted gray lines) at 250 C and 300 C are also
shown for comparison. Log activity of total sulfur was set at log[S]=3 to show the S0
eld at 25 C. Note that SO42 is substituted by CaSO40 for log[Ca]2.3 (T=25 C) and
by CaSO4(s) for log[Ca]3.2 (T=150 C). Thermodynamic database: thermo.dat
(www.geology.illinois.edu/Hydrogeology/hydro_thermo.htm).

Fig. 2. Oxygen isotope fractionation vs. temperature for different SOH2O systems
(solid lines). Circles: data from Halas and Pluta (2000); diamonds: data from Table 1
in Cortecci and Dowgiao (1975); triangles: data from Table 2 in Porowski and
Dowgiao (2009). Black symbols refer to the original data; the white and gray ones
refer to the original data and data that were adjusted by standardization and by
chemical geothermometry, respectively (see inset and text for explanations).
HSO4SO42 line is from Eq. (4), empirical and theoretical SO42 H2O lines are
from Eqs. (6) and (7), respectively; see text for details and references.

Author's personal copy


94

T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

Zeebe (2010) also showed that the microbially mediated oxygen


isotope fractionation between SO42 and H2O during sulfate reduction
in porewater is consistent with the inorganic equilibrium between
the two phases. In Fig. 2, the theoretical and empirical fractionation
factors within the range of 0150 C are plotted along with the original empirical data from Halas and Pluta (2000), the data points of
waters from low-enthalpy groundwater (Table 1 in Cortecci and
Dowgiao, 1975; Table 2 in Porowski and Dowgiao, 2009) and the
data points of interstitial waters from the Baltic Sea (Halas et al.,
1993). Samples suspected to be diluted by recent meteoric water
based on their 3H and 14C contents were excluded from the plot.
While many samples t the SO42H2O equilibrium area, some are
shifted from it (e.g., the 710 samples from Table 2 in Porowski and
Dowgiao, 2009). This shift may be due to isotopic effects on sulfate related to sultewater exchange followed by oxidation (see Fig. 2) or/and
to the decoupling of decreasing temperatures and sulfatewater isotopic
re-equilibration in the uid during its ascent (conductive cooling). In
this latter case, the effect of water cooling can be quantied by means
of estimation of temperature at depth using the local geothermal gradient and/or the chemical geothermometers as shown in Fig. 2 (chemical
geothermometric estimation from Table 3 in Porowski and Dowgiao,
2009). Shifts towards the HSO4H2O equilibrium line after correction,
especially at relatively high temperatures (>100 C), could denote a signicant presence of HSO4 together with SO42 in the uids. Along with
the cooling effect, an additional so called standard effect should be considered. In older publications researchers commonly accepted +9.5 for
the present-day seawater sulfate (NBS-127 standard), whereas with advent of continuous ow, on-line method this value was lowered to
+8.6 relative to VSMOW. The time boundary of this shift is about
2000 therefore, in order to ll the gap between off-line and the on-line
techniques, published 18O (SO42) data before this time should be decreased of 0.9, or depending by the obtained value on seawater sulfate
declared in the analytic method (Boschetti and Iacumin, 2005; Boschetti
et al., 2011a).
The departure from the SO42H2O equilibrium line shown in
Fig. 2 by the interstitial (glacial) waters from the Baltic Sea was ascribed to salt effects (Halas and Pluta, 2000). This does not seem to
be a satisfactory explanation because the salt effect can affect the
18O of brine with TDS > 100 g/L (e.g., Horita et al., 1993), while the
waters in question are brackish to saline (1 g/L b TDS b 100 g/L;
Halas et al., 1993). Moreover, it should be noted that SO42H2O isotopic equilibrium was not attained in these waters probably due to
their residence time, estimated to be approximately 2 10 4 years
(Halas et al., 1993), which is lower than the time of 6 10 4 necessary
to reach a 97% equilibrium as calculated at the stated T = 10 C and
pH =8 following the approach of Cortecci (1981). However, according
to Horita et al. (1994), the salt effect on isotope equilibrium should be
considered in isotope geothermometric estimation between brines and
dissolved species as sulfate. More about calculation of this effect will
be explained in the next section.
3. Case studies
It is still unexplained why the HSO4H2O isotopic fractionations
gave coherent results in many low-enthalpy geothermalsedimentary
systems where SO42 was the dominant sulfate form (e.g., Dogger aquifer, Paris Basin, Fouillac et al., 1990; Frio Formation, Texas Gulf Coast,
Dworkin and Land, 1996; Salsomaggiore, Po Basin, Italy, Boschetti et
al., 2011b).
In the following review, the 18O values of water and dissolved
sulfate and chemical data from a variety of case studies will be
revisited to validate the new sulfatewater oxygen isotope fractionation equations as a geothermometric tool considering the
abundance of the aqueous sulfur species and, when possible,
correcting isotope effects at the equilibrium from the above mentioned secondary effects.

3.1. Thermodynamic calculation: saturation index, activity of the dissolved


sulfur species and isotope salt effect
Saturation index and activity of sulfur-bearing mineral and aqueous
species, respectively, were computed from published data using complete chemical analyses from the literature along with the EQ3/6
software, supported by the Lawrence Livermore National Laboratory
thermodynamic database, its modication for pressure, and the Davies
or B-Dot equation to calculate the activity of waters with ionic strength
(IS) of 0.0 b IS b 0.3 and 0.3b ISb 0.5 m, respectively (Wolery and Jarek,
2003). In brine in equilibrium with gypsum and with NaCl >0.5 m,
SO42 is the dominant species according to the virial activity model of
HarvieMllerWeare (e.g. Bethke, 2008). However, in the present
study, the activity of sulfur-bearing species in IS> 0.5 brines was also
calculated by means of BrnstedGuggenheimScatchard (SIT model,
e.g., Grenthe et al., 1997) or/and by the BromleyMeissner activity
coefcient model for cationanion interaction and the Pitzer activity
coefcient model for ionneutral molecule and moleculemolecule interactions, all coupled with the HelgesonKirkhamFlowersTanger
equation of state that consider TP deviation from the standard state
and an extensive database of aqueous species (Rafal et al., 1994). Before
speciation, carbonate alkalinity was rst calculated from the measured
total alkalinity along with the complete physico-chemical analysis by
PHREEQCI software using LLNL database (Parkhurst and Appelo,
1999). The values of activity and saturation index S.I. = log (Q / K),
where K is the equilibrium constant for the reaction of dissolution
of a pure solid phase, and Q the activity product referred to the
dissolved species obtained by computation are only approximate
because of analytical and thermodynamic uncertainty; thus, saturation at equilibrium have been assumed conventionally in the
range 0.2 b S.I. b 0.2. The results of the calculations are reported
in the Appendix.
It has long been recognized that dissolved salts in water can
change oxygen isotope partitioning between water and other phases
(vapor, minerals, and dissolved species) due to the hydration of ions
upon the dissolution of salts in water, therefore the so called salt effect
must be taken into account for isotopic studies of geothermal systems
(Horita et al., 1994). Salt effect on brines can be satisfactorily described as an additive property of the experimental salt effects of individual single salts (Horita et al., 1993):
3

10 ln

18


mixed salt solution




3
18
10 ln
O
 

b
mi ai i
T

single salt solution

where is the salt effect, i is the ith single salt with m molality in
mixed salt solutions, ai and bi are experimentally determined coefcients, and T (K). In this study, mi in brine samples was computed
by means of complete chemical composition, computer software
and thermodynamic database using the virial-coefcients approach
(e.g. Boschetti et al., 2011b). Afterwards, salt effect on oxygen isotope
fractionation over the temperature range of 0150 C was obtained
from:
3

10 ln sulfatebrine 10 ln

18


3
O 10 ln sulfatewater :

3.2. Lake Vanda (Antarctica)


Lake Vanda is situated in the Wright Valley, approximately 50 km
west of the Victoria Land coast, on the west side of McMurdo Sound,
Antarctica. It occupies the lowest part of the valley at an elevation of
95 m above sea level. The lake is perennially covered by 4 m thick ice.
During summer, a moat of meltwater a few meters wide develops at
the lake margin. The lake receives glacial meltwater from the east

Author's personal copy


T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

via the Onyx River, which drains mainly the Wright Lower glacier, has
no outow and losses water only by evaporation (Nakai et al., 1975).
It is a meromictic lake characterized by the following features:
a) In spite of the mean annual air temperature of approximately
20 C, the water temperature of the lake increases and pH decreases gradually with depth, reaching values of 26 C and 6.05.6
in the bottom layer at 68 m depth, respectively (Angino et al., 1965;
Nakai et al., 1975).
b) The lake water is permanently density-stratied by salt concentration. The lowest salinity is observed just underneath the surface ice
sheet. The chemocline coincides approximately with the thermocline
and begins between 40 and 50 m of depth. The salt content reaches
100 g/L in the bottom brine and 200 g/L in the interstitial brine,
with a corresponding IS of 2.4 and 5.4 m, respectively.
c) A high concentration of H2S (up to 76.8 mg/L) occurs in the anoxic
bottom layer. The gas diffuses upwards, undergoing a nearly continuous oxidation starting from approximately 10 m above the bottom
up to the surface. The bottom CaCl brine was formed approximately
3000 years ago by seawater evaporation during valley uplift, and it
was then diluted at the surface by glacial meltwaters (Friedman et
al., 1995).
Published data on the oxygen isotope composition of water and
dissolved sulfate are shown in Fig. 3. All published 18O(SO42 ) data
were produced in the Rafter's Laboratory, Lower Hutt (New Zealand),
where a mean oxygen sulfate value of + 9.9 was obtained for seawater around New Zealand (Rafter and Mizutani, 1967). Considering
the real seawater value of 18O(SO42 ) = +8.6, a correction for
standard effect of 1.3 was applied to these data. Some clusters
are easily determined from Fig. 3: brines at the chemocline t with
the SO42H2O equilibrium, in keeping with the dominance of SO42
in brine with IS higher than 2 (Appendix A.1), the age of brines
(up to 3000 years), and the relatively high temperature and low
pH that enhance equilibration of sulfate with water. A time interval
of 3000 years should be enough for a 97% isotopic exchange in
these bottom water (e.g., Cortecci, 1981). Samples with the lowest

Fig. 3. Sulfatewater oxygen isotope fractionation vs. temperature in Lake Vanda waters, Antarctica (Friedman et al., 1995; Nakai et al., 1975; Rafter and Mizutani, 1967).
Isotopic data were corrected for standardization effect, while theoretical SO42H2O
(Eq. 7) and CaSO4H2O (Eq. 8) fractionations were corrected for salt effect as specied
in the text; therefore, thick lines represents SO42interstitial brine and CaSO4bottom
brine, respectively.

95

18O (SO42 ) values determine, with their 18O (H2O) values, the
lowest 10 3 ln fractionation, probably because part of the sulfate
derives from H2S oxidation, whereas the trend to lower temperatures is shown in samples from shallow depths. In this latter cluster, the scattering from the SO42 H2O equilibrium line is due to
mixing with recent ice melt-water. Bottom brines show a trend
towards high 10 3 ln values, possibly due bacterial dissimilatory
sulfate reduction. However, it's worthy of note that brines samples
with high fractionation cluster near the anhydritewater equilibrium
line of isotopic equation Zheng (1999):1
3

10 ln CaSO4 H2 Otheoretical 3:9410 =T 5:5410 =T1:87:

This agrees with the brines supersaturation and presence of


authigenic anhydrite in the sediments (Walter and Bauld, 1983).
The deep warm water of the lake up to 26 C has prompted a number
of papers suggesting possible heating mechanisms. Explorative drilling
registered that bottom sediment is about 22 C, that is less than the bottom water at the deepest part of the lake (Bydder and Holdsworth,
1977). The line of best t from chemocline to bottom brines intercept
the anhydritewater isotopic equation corrected for salt effect at 26 C
(Fig. 3), in agreement with the measured values. Interstitial brines are
anhydrite supersaturated too but show a decrease of 103 ln from 29
to 22 along with the deepening of sediments (Friedman et al., 1995).
Also in this case, the temperature of 22 C obtained using the salt
corrected SO42H2O equilibrium equation agrees with the measured
ones (Fig. 3). Finally, the measured and estimated low temperatures at
lake bottom entail an absence of geothermal heating, whereas the temperature prole support the thermohaline circulation solar radiation
as combined heating mechanism (Torii et al., 1981).
3.3. The Dogger and Rhaetian aquifers (Paris Basin)
In the Paris Basin, the saline waters from the Dogger and Rhaetian
aquifers have a common origin and are related to Triassic brines and
salts. In the Dogger layers, downhole temperature and salinity proles are 55 C to 85 C and 1 to 35 g/L, respectively (Matray and
Fontes, 1990). A combination of chemical and isotopic tracers showed
that this water is a mixture of two brine components, both of meteoric origin, and were formed by dissolution of Triassic evaporites, which
in turn were diluted by present-day meteoric groundwater (Matray
and Chery, 1998; Matray and Fontes, 1990). The isotopic temperatures of twelve wells using the HSO4H2O fractionation are in good
agreement with temperatures measured at depth (Fouillac et al.,
1990), and the attainment of isotopic equilibrium should have been
catalyzed by the bacterial sulfate reduction that was active in the
Dogger aquifer (Fouillac et al., 1990; Fritz et al., 1989).
However, some evidences cast doubt on the applicability of the
HSO4H2O fractionation to the studied system. Firstly, speciation of
the most saline brine using the complete physicochemical data set
from the literature (sample GAY1; Azaroual et al., 1997; Rojas et al.,
1989; Appendix A.2) reveals that SO42 is the predominant form,
while HSO4 is the minor species. Secondly, it should be considered
that sulfate in solution derives from the mixing of two different
sources (Fig. 4A), one with 18O (SO42) = + 22, and Cl = 0 g/L,
and the second with 18O (SO42) = +10 and Cl = 20 g/L (Rojas et
al., 1989). This latter saline source has a 18O (H2O) of 3.5
(Matray and Chery, 1998) and a 18O (SO42) of +10 that corresponds

1
Other equations were published for the CaSO4(s)H2O system. In this paper, Zheng's
(1999) equation was chosen because its wide temperature range (01200 C) covers the
low temperature range considered in this study (Tb 150 C). A critical review of previously
published equations for that system is beyond the scope of this work, but the reader can
nd a good review in Zheng (1999) or use the web-isotope fractionation calculator of
Beaudoin and Therrien (2009).

Author's personal copy


96

T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

determining the isotopic sulfate composition of the Rhaetian brines


(Fontes and Matray, 1993). As in the Dogger aquifer, the SO42 specie
is predominant over HSO4 and CaSO40 in these brines. The estimated
chemical (LiMg method; Fontes and Matray, 1993) and the isotopic
(SO42H2O equation) temperatures t along a dilution trend (Fig. 4B).
The isotopic and chemical thermometric estimates are nearly identical
(ca. 117 C) for samples with the highest TDS value; this brine probably
represented the reservoir term as supported by a study on uid inclusion geothermometry (Guilhaumou, 1993, Fig. 4B).

3.4. Soultz-sous-Forts (Rhine Graben)

Since 1987, Soultz-sous-Forts has been the site of the European


Hot Dry Rock project. Until now, much geochemical work has been
carried out to characterize the chemical and isotopic composition of
uids in the local granite complex and especially in the GPK-1 borehole
(Sanjuan et al., 2010). Chemical, gas and isotope geothermometry were
applied to estimate the temperature at depth (Pauwels et al., 1993;
Sanjuan et al., 2010). A temperature of 160165 C was measured at a
depth of 3500 m during the sampling of the GPK-1 brine in the granite
core, well below the estimations of 230240 C from chemical and gas
geothermometers. This discrepancy was interpreted as being due to the
interaction of the geothermal brine with the sedimentary rocks of the
Trias rather than with the granite (Sanjuan et al., 2010). The HSO4H2O
isotope geothermometer provided intermediate estimates of approximately 199 C using the Seal et al. (2000) equation. Several interpretations were invoked to explain these lower temperature estimates,
including dilution and barite scaling, but sulfate speciation in solution
was ignored. This analysis, in fact, reveals a tendency to precipitate
anhydrite and barite at depth as well as the dominance of CaSO40 and
SO42 in the uid (Appendix A.3); therefore, the use of a HSO4H2O
geothermometer is not appropriate. Considering the oversaturation
relative to anhydrite, the CaSO4H2O fractionations should be used,
and in fact the application of Eq. (8) to the GPK-1 brine, 18O (H2O)=
2.9 and 18O (SO42) = +5.9 (Aquilina et al., 1997) gives a temperature of 233 C, which is in perfect agreement with the chemical
geothermometric results.

3.5. Alkaline water in granites (Central Pyrenees, Spain)

Fig. 4. A: 18O (SO42) vs. 34O (SO42) relationship for aqueous sulfate in the Dogger
aquifer brines (Fontes and Matray, 1993; Rojas et al., 1989). Curves depict the
evolution of residual sulfate at slow to fast specic bacterial sulfate reduction rates
(see Brunner et al., 2005). Mixing processes (see text) are indicated by arrow. B: a
comparison between temperatures calculated by the MgLi chemical thermometer
(Kharaka and Mariner, 1989) and the SO42H2O oxygen isotope thermometer
(Zeebe, 2010) for the Rhaetian brines (Fontes and Matray, 1993). Circle diameter is
proportional to brine salinity. The shaded box refers to temperature estimates from a
uid inclusion study (Guilhaumou, 1993).

to the pristine sulfate before bacterial reduction (Fig. 4A; Fouillac et al.,
1990). Using these data and the SO42H2O theoretical fractionation
equation (Eq. 7), we estimate an equilibrium temperature of 85 C
that agrees with the maximum temperature observed. The residence
time of the saline groundwater in the aquifer (between 4 My and
1 My; Matray and Chery, 1998) should have been sufcient for isotopic equilibration at this temperature and stated pH.
In the Rhaetian formation, the salinity of groundwater ranges from
5 g/L to 220 g/L (Matray and Fontes, 1990). The aquifer is a mixture
of two water sources: a meteoric water source and a seawater source
that dissolved evaporites (mainly halite). In addition, a small amount
of a primary brine (i.e., the mother liquor of halite) was considered
to account for the observed chemical compositions; however, this
contribution was small (~ 1.4%) and likely did not play a role in

Alkaline thermal waters spout from faults in granitic rocks and gneiss
along the Pyrenees Range. Equilibrium temperatures evaluated by
chemical geothermometers (110 20 C; Auqu et al., 1996; Criaud
and Vuataz, 1984) were much higher that the highest temperatures of
6570 C measured in a borehole (Forage 1, Luchon, at 114 m; Criaud
and Vuataz, 1984). Approximately 30 thermal zones, extending over
200 km from the Mediterranean Sea from the east to the west,
half-way from the Atlantic Ocean, display very similar geochemical
features: TDS b 500 mg/L, a pH of approximately 9 (a little higher in
colder waters), 90% sodium relative to total cations, and a uniform concentration of anions with none of them prevailing (Vuataz et al., 1985).
In spite of the dominance of sulde at depth (Criaud and Vuataz, 1984),
its oxidation during ascent and dilution with shallow meteoric water,
the HSO4H2O isotopic geothermometer was applied to the
springs (Auqu et al., 1996). As expected, the estimated temperature of 86 C for the hottest and less diluted spring at Forage 1
is notably lower than the chemical estimates of 110 C. However,
the resulting SO42 H2O isotopic temperatures for the other
samples match with the measured values at the spring, possibly
reecting a rapid isotopic exchange between SO32 and H2O during oxidation to sulfate. These results agree with the position of
the samples in the sulte oxidation band in the plot of Fig. 5.
Therefore, the use of sulfatewater oxygen isotope thermometers should be avoided in these sulde-rich waters.

Author's personal copy


T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

97

Fig. 5. 18O (SO42) vs. 18O (H2O) of alkaline thermal waters from the Central Pyrenees (diamonds: data from Auqu et al., 1996). Broken and solid lines delimit areas of possible
18O values for aqueous sulfate obtained by inorganic oxidation of reduced sulfur involving or not involving sulte as intermediate form (van Stempvoort and Krouse, 1994).

3.6. Salsomaggiore (Northern Apennine Foredeep, Italy)


The waters of Salsomaggiore are typical CaCl brines (up to TDS =
190 g/L) from a hydrocarbon-bearing sedimentary basin. Sulfate content in the brine is quite low, likely due to bacterial or thermochemical
reduction. In any case, sulfur and oxygen isotopes indicate a Miocene
origin for dissolved sulfate (Fig. 6; Boschetti et al., 2011b). The brines
are not affected by mixing with present-day meteoric water as revealed
by their very low tritium content (0.006 0.019 UT, unpublished
results) that probably implies that SO42 H2O isotopic equilibrium
was reached.
In the aquifers, dolomitization occurred at approximately 150 C
(i.e., 57 km of burial) as inferred from the MgLi and NaLi chemical
geothermometers and was accompanied by a water 18O-shift of up
to + 11 (Boschetti et al., 2011b). The chemical temperature was
conrmed by the HSO4H2O isotopic fractionations (Eq. 4, T=145
23 C, N=2 samples; Boschetti et al., 2011b) even if the brine speciation
reveals that SO42 is the dominant form at depth (84.7 mol%). Therefore,
the SO42H2O isotopic fractionations should be used, thus obtaining
lower temperature estimates of 9026 C (Eq. 7). However, extrapolation of the Salsomaggiore brine 18O34S correlation towards the local
Messinian sulfate value (34S=+24.9) provides a 18O (SO42)=
+18 pre-reduction value in the brine (Fig. 6). Coupling this value for
SO42 and the +11 value for H2O, an isotopic temperature of 146 C
is calculated (Eq. 7). The agreement between results obtained by this
late estimation and the chemical geothermometers, the long residence
time and the 34S18O relation shown in Fig. 6 for the Salsomaggiore
brine indicates that the SO42H2O system achieved an abiotic oxygen isotope equilibrium before sulfate reduction. The high consistency between
pre-reduction SO42H2O and post-reduction HSO4H2O isotopic temperatures is most likely due to the important role played by HSO4 during
the thermochemical reduction of sulfate (Ma et al., 2008).

volcanics range from basalt to rhyolite, overlying and interbedding the


sedimentary rocks.
As the application of chemical geothermometers to these waters
produces contradictory temperature estimates (Mariner et al.,
1993); Mariner et al. (1980) switched to use the bisulfatewater
isotope geothermometer. Higher temperature estimates were obtained
that completely disagreed with temperatures given by the chemical
geothermometers, and the matter was closed. However, using chemical
modeling, it was noted that the HSO4H2O temperatures in the NaCa
Cl and NaCl waters facies were rather consistent with re-equilibration
of solutions at the anhydrite saturation temperature, probably accounting for the reactions (Mariner et al., 1993):
NaCl water volcanic glass NaCaCl water calcite anhydrite
sodic zeolites
NaCl water plagioclase silica NaCaCl water calcite
anhydrite sodic zeolites

9
10

whereas in NaSO4 waters the temperatures obtained by using chemical


geothermometers (silica, Na/K) are concordant between them but not
with the HSO4H2O (Mariner et al., 1993). In accord with that chemical
modeling, many NaCl and NaCaCl waters plot close to or between
the CaSO4H2O and HSO4H2O isotopic lines (Fig. 7). Probably, alteration of primary rock occurred via volcanogenic sulfuric acid according
to the reaction:
CaAl2 Si2 O8 H2 SO4 H2 O AL2 Si2 O5 OH4 CaSO4 :
anorthite

kaolinite

11

anhydrite

On the other hand, most NaSO4 waters (Mariner et al., 1993) t


the theoretical SO42H2O isotopic line (Eq. 7) and are in chemical
equilibrium with quartz (concordant temperatures), probably accounting for the reaction:

3.7. Thermal waters from the western United States


The fresh to saline thermal waters from the Cascade Range (Oregon)
and the Modoc Plateau (California) emerge after passing through a thick
sequence of volcanic and sedimentary rocks of Cenozoic age. The

2NaAlSi3 O8 H4 SiO4 5H2 O CaSO4 CaAl2 Si7 O18 7H2 O


albite

anhydrite

2Na SO4 :

stilbite

12

Author's personal copy


98

T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

l
-

l
l
l

l
l

Fig. 6. 18O (SO42) vs. 34O (SO42) plot for Salsomaggiore brines compared to other relevant systems and Messinian evaporites, both local and Mediterranean (modied from
Boschetti et al., 2011b). Note that the Salsomaggiore brines (circles) and pore waters interacting with Messinian deposits are arranged along the slow sulfate reduction (bacterial
or thermochemical?) curve (insert, see also Fig. 4A). Back extrapolation of the Salsomaggiore brines allows to dene that 18O (SO42) = +18 of the parental evaporite, which is
considered to be the pristine isotope composition of dissolved sulfate (see text for details).

Some NaSO4 waters deviate towards the HSO4H2O systems, as


one can expect from the following controlling reaction:

2NaAlSi3 O8 H2 SO4 9H2 O 4H4 SiO4 Al2 Si2 O5 OH4 2Na


albite

kaolinite
2

SO4 :

13

Finally, a number of water samples with intermediate chemical


composition plot between the NaSO4 and NaCaCl groups.
The thermal waters from the western United States represent an emblematic case of successful application to low- and high-enthalpy hydrothermal systems of the oxygen isotope equilibrium equations between
speciation-recognized aqueous sulfate chemical forms and water.
4. Conclusions
Empirical and theoretical thermometric equations based on the oxygen isotope fractionation between SO42 and H2O provide comparable
results in the 0150 C range. Coherent geothermometric results are
obtained in aquifers with pH, temperature and residence time appropriate for oxygen isotope exchange between phases (e.g., in sedimentary
fossil brines). In the examined cases, measured residence times are

comparable with those calculable from Cortecci (1981) rather than


those calculable from Chiba and Sakai's (1985) approach.
Speciation of the sulfate forms and calculation of the sulfatemineral
saturation indexes of the solutions plays a key role in applying the appropriate isotopic fractionation with water. It's noteworthy that CaSO4(s)
H2O isotopic equation gives high-quality thermometric estimates in
anhydrite-saturated brines or in fresh to saline solutions where CaSO40
is the dominant aqueous species. This is not an unexpected effect
considering that thermodynamic equilibrium constant K is directly
related to fractionation factor a (e.g. Sharp, 2007). However, more
efforts are necessary to conrm this hypothesis. For examples, the
similarity between CaSO40H2O and CaSO4(s)H2O isotope fractionation
should be checked using, for example, the quantum-chemistry approach.
Furthermore, someone might pinpoint that anhydrite precipitation effect
should be considered, but should be noted that this effect is self-included
when considering the distance between SO42H2O and CaSO4(s)H2O
lines. In fact:




3
3
2
3
2
10 ln CaSO4 H2 O10 ln SO4 H2 O 10 ln CaSO4 SO4 :
Therefore, the new theoretical and experimental sulfatewater
equations may better quantify also the never claried effect due to
anhydrite precipitation (Holser et al., 1979).

Author's personal copy


T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

99

comments to the rst version and to S. Halas, one anonymous reviewer


and Associate Editor D. Huston for their critical comments to this work.
Appendix A. Supplementary data
Supplementary data to this article can be found online at http://
dx.doi.org/10.1016/j.gexplo.2012.08.011.
References

Fig. 7. Sulfatewater and sultewater oxygen isotope fractionation vs. temperature for
thermal springs in the Western United States. Gray diamonds are temperatures measured
at the outlet (Mariner et al., 1980, 1993); black diamonds are the temperatures at depth
calculated using chemical modeling (anhydrite equilibrium in NaCaCl springs) or
chemical geothermometers (NaSO4 waters).

In a 10 3 ln vs. 10 6/T 2 diagram, the shifts from the oxygen isotope


SO42 H2O line towards equilibria between water and other
sulfur-bearing species can help to recognize isotopic effects not directly
linked to (inorganic) isotopic exchange between phases, such as the salt
effect, cooling (related to the difference between temperature at the
issue and at depth), mixing and/or dilution, sulde oxidation and sulfate
reduction effects. In addition, sulfate derived from sulte oxidation can
be effectively identied in the same diagram and conrmed on a 18O
(SO42) vs. 18O (H2O) plot.
Finally, it may be concluded that the application of a SO42H2O
oxygen isotope geothermometer can be useful to validate chemical
geothermometric results, particularly in low-enthalpy systems
where a reservoir temperature is difcult to reliably assess.

Acknowledgments
This work is dedicated to my colleague and former tutor
Gianni Cortecci, who in 1974 published a pioneering work on
the geothermometric application of the HSO4H2O oxygen isotope
fractionations to hydrothermal systems (see reference list). During his
scientic career, Gianni Cortecci has become highly regarded worldwide
as a stable isotope geochemist. His main area of interest was (and partially still is) sulfur and oxygen isotope geochemistry applied to waters,
rocks and mineral deposits. His rst studies in isotope geochemistry
date back to the 1960s at the Laboratorio di Geologia Nucleare of the
University of Pisa where, among other things, he improved the sulfur
and oxygen isotope analytical methods on suldes and sulfates. From
1994 to 2004, he was a full professor of Geochemistry at the University
of Bologna, and then went back to Pisa as a pensioner and scientic consultant at the Institute of Geosciences and Georesources of the CNR of
Italy. He continued teaching and research activities until a denitive retirement in December 2010. However, the leopard never changes his
spots/the wolf changes the hair but not vice, and, accordingly, Gianni
Cortecci is still keen on and concerned about science, and I would like
to thank him once more for his critical review and useful comments to
this manuscript. I wish to express my gratitude also to C. Pierre for

Angino, E.E., Armitage, K.B., Tosh, J.C., 1965. A chemical and limnological study of Lake
Vanda, Victoria Land, Antarctica. The University of Kansas Science Bulletin 45,
10971118.
Aquilina, L., Pauwels, H., Genter, A., Fouillac, C., 1997. Waterrock interaction processes in
the Triassic sandstone and the granitic basement of the Rhine Graben: geochemical investigation of a geothermal reservoir. Geochimica et Cosmochimica Acta 42814295.
Auqu, L.F., Mandado, J., Lpez, P.L., Gimeno, M.J., 1996. Aplicacin del geotermmetro
isotpico sulfato-agua (18O SO4218O H2O) en algunos sistemas termales alcalinos.
del Pirineo Revista de la Academia de Ciencias de Zaragoza, 51, pp. 195207.
Azaroual, M., Fouillac, C., Matray, J.M., 1997. Solubility of silica polymorphs in electrolyte
solutions, II. Activity of aqueous silica and solid silica polymorphs in deep solutions
from the sedimentary Paris Basin. Chemical Geology 140, 167179.
Beaudoin, G., Therrien, P., 2009. The updated web stable isotope fractionation calculator. In: De Groot, P.A. (Ed.), Handbook of Stable Isotope Analytical Techniques,
Volume-II. Elsevier, pp. 11201122.
Bethke, C.M., 2008. Geochemical and Biogeochemical Reaction Modeling, Second edition.
Cambridge University Press.
Boschetti, T., Cortecci, G., Toscani, L., Iacumin, P., 2011a. Sulfur and oxygen isotope compositions of Upper Triassic sulfates from Northern Apennines (Italy): palaeogeographic
and hydrogeochemical implications. Geologica Acta 9, 129147.
Boschetti, T., Iacumin, P., 2005. Continuous ow 18O measurements: new approach to
standardization, high temperature thermodynamics and sulfate analysis. Rapid
Communication in Mass Spectrometry 19, 30073014.
Boschetti, T., Toscani, L., Shouakar-Stash, O., Iacumin, P., Venturelli, G., Mucchino, C., Frape,
S.K., 2011b. Salt waters of the Northern Apennine Foredeep Basin (Italy): origin and
evolution. Aquatic Geochemistry 17, 71108.
Brunner, B., Bernasconi, S., Kleikemper, J., Schroth, M.H., 2005. A model for oxygen and
sulphur isotope fractionation in sulfate during bacterial sulfate reduction processes.
Geochimica et Cosmochimica Acta 20, 47334785.
Brunner, B., Mielke, R.E., Coleman, M., 2006. Abiotic oxygen isotope equilibrium fractionation between sulte and water. Eos Transactions, American Geophysical
Union AGU.
Bydder, E.L., Holdsworth, R., 1977. Lake Vanda (Antarctica) revisited. New Zealand
Journal of Geology and Geophysics 20, 10271032.
Chandrasekharam, D., Bundschuh, J., 2008. Low-Enthalpy Geothermal Resources for
Power Generation. CRC Press, Taylor & Francis Group.
Chiba, H., Sakai, H., 1985. Oxygen isotope exchange rate between dissolved sulfate and
water at hydrothermal temperatures. Geochimica et Cosmochimica Acta 49,
9931000 Geochim. Cosmochim. Acta 49.
Cortecci, G., 1974. Oxygen isotopic ratios of sulfate ionswater pairs as a possible
geothermometer. Geothermics 3, 6064.
Cortecci, G., 1981. Prospezione e studio di sistemi idrotermali mediante geotermometria
isotopica SO4H2O e SO4H2S. PEG Editrice, Roma, pp. 2335.
Cortecci, G., Dowgiao, J., 1975. Oxygen and sulfur isotopic composition of the sulfate ions
from mineral and thermal groundwaters of Poland. Journal of Hydrology 24, 271282.
Criaud, A., Vuataz, D., 1984. Etude gochimique et gothermique des eaux sulfures
sodiques de Luchon, Pyrnes. Rapport du BRGM 84 SGN 384 IRG.
D'Amore, F., Arnrsson, S., 2000. Geothermometry. In: Arnrsson, S. (Ed.), Isotopic and
Chemical Techniques in Geothermal Exploration, Development and Use. IAEA,
Vienna, pp. 152199.
Dworkin, S.I., Land, L.S., 1996. The origin of aqueous sulfate in Frio pore uids and its
implication for the origin of oil eld brines. Applied Geochemistry 11, 403408.
Fontes, J.C., Matray, J.M., 1993. Geochemistry and origin of formation brines from the
Paris Basin. Part 2: saline solutions associated with oil elds. Chemical Geology
109, 177200.
Fouillac, C., Fouillac, A.M., Criaud, A., 1990. Sulphur and oxygen isotopes of dissolved
sulphur species in formation waters from the Dogger geothermal aquifer, Paris
Basin, France. Applied Geochemistry 5, 415427.
Friedman, I., Rafter, A., Smith, G.I., 1995. A thermal, isotopic and chemical study of Lake
Vanda and Don Juan Pond, Antarctica. In: Elliot, D.H., Blaisdell, G.L. (Eds.),
Contributions to Antarctic Research IV, pp. 4774.
Fritz, P., Basharmal, G.M., Drimmie, R.J., Ibsen, J., Qureshi, R.M., 1989. Oxygen isotope
exchange between sulphate and water during bacterial reduction of sulphate.
Chemical Geology: Isotope Geoscience Section 79, 99105.
Giggenbach, W.F., Gonantini, R., Jangi, B.L., Truesdell, A.H., 1983. Isotopic and chemical
composition of Parbati Valley geothermal discharges, NW Himalaya, India.
Geothermics 12, 199222.
Grenthe, I., Plyasunov, A.V., Spahiu, K., 1997. Estimation of medium effects on thermodynamic data. In: Grenthe, I., Puigdomenech, I. (Eds.), Modeling in Aquatic Chemistry. OECD Nuclear Energy Agency, Issy-les-Moulineaux, France, pp. 325426.
Guilhaumou, N., 1993. Palaeotemperatures inferred from uid inclusions in diagenetic
cements; implications for the thermal history of the Paris Basin. European Journal
of Mineralogy 5, 12171226.

Author's personal copy


100

T. Boschetti / Journal of Geochemical Exploration 124 (2013) 92100

Halas, S., Pluta, I., 2000. Empirical calibration of isotope thermometer 18O
(SO42)18O (H2O) for low temperature brines, V Isotope Workshop. European
Society for Isotope Research, Krakw, Poland, pp. 6871.
Halas, S., Trembaczowski, A., Soltyk, W., Valendziak, J., 1993. Sulphur and oxygen in natural waters: (2) deep-waters from horizons below Baltic Sea oor. Isotopenpraxis 28,
229235.
Hoefs, J., 2009. Stable Isotope Geochemistry, Sixth edition. Springer.
Hoering, T.C., Kennedy, J.W., 1957. The exchange of oxygen between sulfuric acid and
water. Journal of the American Chemical Society 79, 5660.
Holser, W.T., Kaplan, I.R., Sakai, H., Zak, I., 1979. Isotope geochemistry of oxygen in the
sedimentary sulfate cycle. Chemical Geology 25, 117.
Horita, J., Cole, D.R., Wesolowski, D.J., 1993. The activitycomposition relationship of
oxygen and hydrogen isotopes in aqueous salt solutions: II. Vaporliquid water
equilibration of mixed salt solutions from 50 to 100 C and geochemical implications. Geochimica et Cosmochimica Acta 57, 47034711.
Horita, J., Cole, D.R., Wesolowski, D.J., 1994. Salt Effects on Stable Isotope Partitioning
and their Geochemical Implications for Geothermal Brines. Nineteenth Workshop
on Geothermal Reservoir Engineering, Stanford University, Stanford, California.
Kharaka, Y.K., Mariner, R.H., 1989. Chemical geothermometers and their application to
formation waters from sedimentary basins. In: Naeser, N.D., McCollin, T.H. (Eds.),
Thermal History of Sedimentary Basins. Springer-Verlag, New York, pp. 99117.
Kusakabe, M., Robinson, B.W., 1977. Oxygen and sulfur isotope equilibria in the system from
110 to 350 C and applications. Geochimica et Cosmochimica Acta 41, 10331040.
Lloyd, R.M., 1968. Oxygen isotope behaviour in the sulfatewater system. Journal of
Geophysical Research 73, 60996110.
Ma, Q., Ellis, G.S., Amrani, A., Zhang, T., Tang, Y., 2008. Theoretical study on the reactivity of
sulfate species with hydrocarbons. Geochimica et Cosmochimica Acta 72, 45654576.
Mariner, R.H., Presser, T.S., Evans, W.C., 1993. Geothermometry and waterrock interaction in selected thermal systems in the Cascade Range and Modoc Plateau, western
United States. Geothermics 22, 115.
Mariner, R.H., Swanson, J.R., Orris, G.L., Presser, T.S., Evans, W.C., 1980. Chemical and
Isotopic Data for Water from Thermal Springs and Wells of Oregon. U.S. Geological
Survey Open-File Report, p. 50.
Matray, J.M., Chery, L., 1998. Origin and age of deep waters of the Paris Basin. In:
Causse, C., Gasse, F. (Eds.), Hydrologie et gochimie isotopique. Orstrom, Paris,
pp. 117133.
Matray, J.M., Fontes, J.C., 1990. Origin of the oil-eld brines in the Paris Basin. Geology
18, 501504.
McKenzie, W.F., Truesdell, A.H., 1977. Geothermal reservoir temperatures estimated
from the oxygen isotope composition of dissolved sulfate and water from hot
springs and shallow drillholes. Geothermics 5, 5161.
Mizutani, Y., Rafter, T.A., 1969. Oxygen isotopic composition of sulphates Part 3.
Oxygen isotopic fractionation in the bisulphate ionwater system. New Zealand
Journal of Science 12, 5459.
Nakai, N., Wada, H., Kiyosu, Y., Takimoto, M., 1975. Stable isotope studies on the origin
and geological history of water and salts in the Lake Vanda area, Antarctica.
Geochemical Journal 9, 724.
Nicholson, K.N., 1993. Geothermal Fluids: Chemistry and Exploration Techniques.
Springer-Verlag, Berlin.
Parkhurst, D.L., Appelo, C.A.J., 1999. User's guide to PHREEQC (version 2) a computer
program for speciation, batch reaction, one-dimensional transport and inverse
geochemical calculations. US Geological Survey 312.
Pauwels, H., Fouillac, C., A.M., F., 1993. Chemistry and isotopes of deep geothermal saline uids in the Upper Rhine Graben: origin of compounds and waterrock interactions. Geochimica et Cosmochimica Acta 57, 27372749.

Porowski, A., Dowgiao, J., 2009. Application of selected geothermometers to exploration of low-enthalpy thermal water: the Sudetic Geothermal Region in Poland. Environmental Geology 58, 16291638.
Rafal, M., Berthold, J.W., Scrivner, N.C., Grise, S.L., 1994. Models for electrolyte solutions. In: Sandler, S.I. (Ed.), Models for Thermodynamic and Phase Equilibria Calculations. Marcel Dekker, New York, pp. 601670.
Rafter, T.A., Mizutani, Y., 1967. Oxygen isotopic composition of sulphates Part 2. Preliminary results on oxygen isotopic variation in sulphates and the relationship to their
environment and their 34S values. New Zealand Journal of Science 10, 816840.
Rojas, J., Giot, D., LeNindre, Y.M., Criaud, A., Fouillac, C., Brach, M., Menjoz, A., Martin, J.C.,
Lambert, M., 1989. Caractrisation et modlisation du rservoir gothermique du
Dogger, Bassin Parisien, France. Rapport nal CCE, EN 3G-0046-F(CD), BRGM R 30 IRG
SGN 89.
Sakai, H., 1977. Sulfatewater isotope thermometry applied to geothermal systems.
Geothermics 5, 6774.
Sanjuan, B., Millot, R., Dezayes, C., Brach, M., 2010. Main characteristics of the deep geothermal brine (5 km) at Soultz-sous-Forts (France) determined using geochemical and tracer test data. Comptes Rendus Geoscience 342, 546559.
Seal, R.R.I., Alpers, C.N., Rye, R.O., 2000. Stable isotope systematics of sulfate minerals.
In: Alpers, C.N., Jambor, J.L., Nordstrom, D.K. (Eds.), Sulfate Minerals Crystallography: Geochemistry and Environmental Signicance, pp. 541602.
Sharp, Z., 2007. Principles of Stable Isotope Geochemistry. PearsonPrentice Hall,
Upper Saddle River, New Jersey.
Sonney, R., Vuataz, F.D., 2010. Validation of Chemical and Isotopic Geothermometers
from Low Temperature Deep Fluids of Northern Switzerland. Proceeding World
Geothermal Congress 2010, Bali, Indonesia, p. 12.
Torii, T., Murata, S., Yamagata, N., 1981. Geochemistry of the Dry Valley lakes. Journal of
the Royal Society of New Zealand 11, 387399.
Turchyn, A.V., Brchert, V., Lyons, T.W., Engel, G.S., N., B., Schrag, D.P., Brunner, B., 2010.
Kinetic oxygen isotope effects during dissimilatory sulfate reduction: a combined
theoretical and experimental approach. Geochimica et Cosmochimica Acta 74,
20112024.
van Stempvoort, D.R., Krouse, H.R., 1994. Controls of 18O in sulfate: review of experimental data application to specic environments. In: Alpers, C.A., Blowes, D.W.
(Eds.), Environmental Geochemistry of Sulde Oxidation. American Chemical Society, Washington DC, pp. 446480.
Vuataz, F.D., Criaud, A., Fouillac, C., 1985. Detailed geochemical study of alkaline
thermal waters; a geothermal evaluation in the Pyrenees range, southern France.
Geothermal Resources Council Transactions 9, 375380.
Walter, M.R., Bauld, J., 1983. The association of sulphate evaporites, stromatolitic carbonates and glacial sediments: examples from the Proterozoic of Australia and
the Cainozoic of Antarctica. Precambrian Research 21, 129148.
Wolery, T.W., Jarek, R.L., 2003. EQ3/6, Version 8.0 Software User's Manual. Civilian
Radioactive Waste Management System, Management & Operating Contractor.
Sandia National Laboratories, Albuquerque, New Mexico.
Zeebe, R.E., 2010. A new value for the stable oxygen isotope fractionation between
dissolved sulfate ion and water. Geochimica et Cosmochimica Acta 74, 818828.
Zheng, Y.F., 1999. Oxygen isotope fractionation in carbonate and sulfate minerals.
Geochemical Journal 33, 109126.
Zhonge, P., 2001. Isotope and chemical geothermometry and its applications. Science in
China Series E: Technological Sciences 44, 1620 supplement.

Вам также может понравиться