Vous êtes sur la page 1sur 23

Chemical Geology 153 1999.

187209

A thermodynamic model for the solubility of barite and celestite in electrolyte solutions and seawater to 2008C and to 1 kbar
Christophe Monnin
) CNRSr Uniersite Paul Sabatier, Laboratoire de Geochimie, 38 rue des Trente-Six Ponts, 31400 Toulouse, France Received 24 December 1997; accepted 17 September 1998

Abstract This paper describes a model for barite and celestite solubilities in the NaKCaMgBaSrClSO4 H 2 O system to 2008C and to 1 kbar. It is based on Pitzers ion interaction model for the thermodynamic properties of the aqueous phase and on values of the solubility products of the solids revised in this work. It is shown how barite and celestite solubilities in electrolyte solutions can be accurately predicted as a function of temperature and pressure from previously determined Pitzers parameters. The equilibrium constant for the BaSO4aq. ion pair dissociation reaction is calculated from recently reported barite solubility in Na 2 SO4 solutions from 0 to 808C. Pressure corrections are evaluated through partial molal volume calculations and are partially validated by comparing model predictions to measured barite and celestite solubilities in pure water to 1 kbar and in NaCl solutions to 500 bars. The model is then used to investigate the tendency of ion pairing of Ca, Sr and Ba with sulfate in seawater. Finally, the activity coefficient of aqueous barium sulfate in seawater is calculated for temperature, pressure and salinity values found in the ocean and compared to published values. q 1999 Elsevier Science B.V. All rights reserved.
Keywords: Aqueous solutions; Barium sulfate; Strontium sulfate; Activity coefficients; Seawater; Thermodynamic properties

1. Introduction The question of an eventual control of the barium concentration in seawater by an equilibrium with solid barium sulfate barite. can be addressed through the calculation of the activity coefficient of aqueous BaSO4 as well as the barite solubility product at oceanic conditions. This has been done by Church and Wolgemuth 1972. who used an aqueous solution model based on the ion pairing phenomenology along with stability constants of aqueous species available at that time. This waterrock interaction
)

Fax: q33-561520544; E-mail:monnin@lucid.ups-tlse.fr

problem can be reconsidered in light of the recent development of thermodynamic models of aqueous electrolyte solutions based on Pitzers ion interaction approach, along with the wealth of data on Ba distribution in the oceans. Pitzers formalism has allowed the construction of accurate solubility models for systems including the major species of natural waters see Weare, 1987; Pitzer, 1991, for reviews.. After the landmark paper of Harvie et al. 1984. for the NaKCaMgHClSO4 HCO 3 CO 3 CO 2 H 2 O system at 258C and 1 bar, efforts were made at extending such models to elevated temperatures Pabalan and Pitzer, 1987; Greenberg and Moller, 1989, among others. and pressures Monnin, 1989,

0009-2541r99r$ - see front matter q 1999 Elsevier Science B.V. All rights reserved. PII: S 0 0 0 9 - 2 5 4 1 9 8 . 0 0 1 7 1 - 5

188

C. Monninr Chemical Geology 153 (1999) 187209

1990. and to other species of geochemical significance see Pitzer, 1991.. Pitzers ion interaction approach is semi-empirical, as are most aqueous solution models of practical interest. It allows the mathematical representation of the properties of simple systems binary and ternary with a common ion solutions. up to high concentrations through the evaluation of empirical parameters from experimental data. From this representation the thermodynamic properties of highly complex solutions like seawater can be predicted. The accuracy of the calculations is then established by extensive comparisons between model predictions and independent experimental data Weare, 1987.. The present paper follows the procedure outlined above to show that barite and celestite solubilities in chloride solutions can be accurately predicted in large temperature and composition ranges using available Pitzers parameters describing interactions between aqueous Na, K, Ca, Mg, Ba, Sr and chloride, and those for interactions between aqueous Na, K, Ca, Mg and sulfate. Interactions between aqueous barium and sulfate are accounted for by an explicit 0 equilibrium between the free ions and the BaSO4 aq. ion pair. The enthalpy and entropy of the dissociation reaction of this aqueous complex is calculated from the data of Jiang 1996. on barite solubility in sodium sulfate solutions. No such data for celestite solubility in sodium sulfate solutions is available outside 258C. An evaluation of the ion pairing tendency for alkaline earth sulfates in sulfate rich solutions like seawater nevertheless allows one to esti0 mate the contribution of the SrSO4 aq. ion pair to celestite solubility at temperatures other than 258C. Pressure effects are evaluated following the methods published by Monnin 1989, 1990.. Finally, the predicted BaSO4 aq. activity coefficient in seawater is compared to values given by Church and Wolgemuth 1972. and by Falkner Kenisson et al. 1993..

Instead, in order to derive BaSO4 aq. and SrSO4aq. thermodynamic properties, one has to use barite and celestite solubility measurements in more complex systems, the simplest of which are ternary common ion systems. Barite and celestite solubility measurements at 258C were compiled and critically evaluated by Monnin and Galinier 1988.. Experimental data at other temperatures and pressures for NaCl, KCl, CaCl 2 , and MgCl 2 binary aqueous solutions and some of their mixtures Templeton, 1960; Lucchesi and Whitney, 1962; Uchameyshvili et al., 1966; Strubel, 1966; Puchelt, 1967; Strubel, 1967; Mac Donald and North, 1974; Blount, 1977; Jacques and Bourland, 1983; Vetter et al., 1983; Reardon and Amstrong, 1987; Schulien, 1987. have allowed the development of several high temperaturehigh pressure solubility models Blount, 1977; Jacques and Bourland, 1983; Langmuir and Melchior, 1984; Reardon and Amstrong, 1987; Moller, 1988; Raju and Atkinson, 1988, 1989; Yuan and Todd, 1991.. All of these models are fits of measured solubilities, based either on empirical expressions Jacques and Bourland, 1983., the DebyeHuckel equation for aqueous species activity coefficients Blount, 1977., or, more recently, Pitzers formalism Langmuir and Melchior, 1984; Reardon and Amstrong, 1987; Moller, 1988; Raju and Atkinson, 1988, 1989; Yuan and Todd, 1991.. Several of these models are limited to NaCl solutions Moller, 1988; Raju and Atkinson, 1988, 1989; Yuan and Todd, 1991.. On the contrary, the present study takes into account the chlorides and sulfates of the main seawater cations and validates the model by thorough comparison between model predictions and measured solubilities. 2.1. BaSO4 and SrSO4 interactions Felmy et al. 1990. studied the NaBaSO4 H 2 O and NaSrSO4 H 2 O systems at 258C up to 0.01 mol Na 2 SO4rkg H 2 O. These authors found that celestite solubility in sodium sulfate solutions can be accurately calculated to moderate sulfate concentrations 0.01 molrkg H 2 O. within the standard Pitzer approach complete dissociation of the aqueous solutes. with SrSO4 interaction parameters assumed equal to those for CaSO4 . Conversely, in the Na BaSO4 H 2 O system, they showed that a solution

2. Thermodynamic properties of the NaKCa MgBaSrClSO4 H 2 O system to 2008C and 1 kbar Due to the low solubilities of barite and celestite, BaSO4 H 2 O and SrSO4 H 2 O binary solutions cannot be studied as a function of Ba or Sr molality.

C. Monninr Chemical Geology 153 (1999) 187209


0 model which includes the BaSO4 aq. ion pair is somewhat superior to a strong electrolyte model in reproducing experimental barite solubilities to 0.01 mol Na 2 SO4rkg H 2 O Felmy et al., 1990.. 0 0 The effect of BaSO4 aq. or SrSO4 aq. ion pair formation on barite or celestite solubilities can be illustrated as follows. The dissolution of a solid alkaline earth sulfate can be written as:

189

any complex or ion pair wsee Johnson and Pytckowicz 1979., among other authors, for a full description of this approachx. Equilibrium between an ion pair and an aqueous cation and sulfate can be written as:
0 2y MSO4 aq . z M 2q aq . q SO4 aq .

4.

MSO4 s . z M

2q

aq .

2y q SO4

aq .

1.

where K ip stands for the equilibrium constant for reaction 4.: K ip s


2y 2y m M 2q aq .,F P m SO 4 aq .,F P g M 2q aq .,F P g SO 4 aq .,F 0 0 m MS O 4 aq . P g MSO 4 aq .

The equilibrium constant of reaction 1. is the solubility product of the solid. If the model used to calculate aqueous species activity coefficients is a strong electrolyte model i.e., based on the hypothesis of complete dissociation of the aqueous solutes., then the solubility product is given by: 2y 2y K sp s m M 2q aq . P m SO 4 aq . P g M 2q aq . P g SO 4 aq . 2 . where m is the molality and g the activity coefficient of the designated aqueous species. In a strong electrolyte model, activity coefficients can be generated from numerous models, including the classic DebyeHuckel theory or Pitzers ion interaction ap proach in its original formulation Pitzer, 1991.. For aqueous solutes exhibiting strong association, the complete dissociation hypothesis is no longer valid. The explicit definition of one or more aqueous complexes is more efficient for calculating the thermodynamic properties of associating electrolytes. Pitzer 1991. p. 93. and Weare 1987. p. 148 et seq.. provide discussions of the limit between strong and weak electrolytes. When expressed within the ion pairing or weak electrolyte. formalism, Eq. 2. becomes: 2y 2y K sp s m M 2q aq .,F P m SO 4 aq .,F P g M 2q aq .,F P g SO 4 aq .,F

5.
Combining Eqs. 3. and 5. leads to:
0 m MS O 4 aq . s

K sp K ip

1 P
0 g MSO 4 aq .

6.

3.
where the subscript F designates free ions, i.e., the part of the total solute concentration not involved in

Eq. 6. demonstrates that the ion pair concentration is proportional to the ratio of the mineral solubility product to the ion pair dissociation constant. Assuming that activity coefficients for neutral species are close to unity, this ratio can be taken as an estimate of the ion pair concentration in saturated solutions. Table 1 compares this ratio for the Ca, Ba and Sr sulfates, to the solubility of their anhydrous salts anhydrite, celestite and barite. either in pure water or in dilute sodium sulfate solution. Aqueous complex stability constants were taken from the literature and are given in Table 1. It must be emphasized that, in the present approach, interaction of the element M in solution with sulfate is accounted for by the 0 MSO4 aq. ion pair. Interactions with other aqueous anions are accounted for by interaction parameters in the activity coefficient expressions in Pitzers formalism. Such an approach has been termed a hybrid model by Whitfield 1975..

Table 1 An estimate of the cationsulfate ion pair concentration for Ca, Sr and Ba. See text Salt Log K sp Log K ip K sprK ip Solubility in pure water molrkg H 2 O. 0.015 a. 0.0012 b. 1.0 = 10y5 b. Solubility in 0.01 M Na 2 SO4 0.010 a. 2.0 = 10y4 c. 1.0 = 10y7 c.

CaSO4 anhydrite SrSO4 celestite BaSO4 barite

y4.32 a. y6.62 b. y10.05 b.

y1.65 a. y1.86 c. y2.72 c.

0.002 1.7 = 10y5 4.7 = 10y8

a. Moller 1988.; b. Monnin and Galinier 1988.; c. Felmy et al., 1990.

190

C. Monninr Chemical Geology 153 (1999) 187209

The data listed in Table 1 indicate that the less soluble the salt, the stronger the common ion effect. The addition of 0.01 mol of sulfate to the solution has almost no effect on anhydrite solubility compared to pure water, but it lowers barite solubility by two orders of magnitude. Estimates of calcium sulfate ion pair formation indicates that this complex accounts for less than 10% of total dissolved calcium in pure water or sodium sulfate solutions at equilibrium with anhydrite. Consequently, it is possible to treat this salt at 258C, either as fully dissociated Harvie et al., 1984. or as a weak electrolyte Moller, 1988., with comparable accuracy. At higher temperatures where ion pairing is enhanced, the inclusion of 0 the CaSO4 aq. complex is necessary Moller, 1988.. Strontium sulfate ion pair formation is negligible in pure water at equilibrium with celestite, but it accounts for about 10% of total strontium in celestite saturated Na 2 SO4 solutions Table 1.. As shown by Felmy et al. 1990., the thermodynamic properties of strontium sulfate can be equally well calculated by assuming total dissociation ion interaction approach. or assuming aqueous complex formation ion pairing phenomenology.. Reardon and Amstrong 1987. and Monnin and Galinier 1988. also calculated celestite solubilities in seawater assuming total dissociation which are in close agreement with the corresponding experimental values reported by Culberson et al. 1978.. Moreover, the Reardon and Amstrong 1987. model unambiguously shows that celestite crystals observed in deep-sea carbonate sediments of DSDP Leg 90 are at equilibrium with pore waters Baker and Bloomer, 1988.. Barium sulfate ion pair concentration is negligible in barite saturated pure water Table 1.. In 0.01 M sodium sulfate, however, about 50% of the total 0 barium in solution is present as the BaSO4 aq. aqueous complex, when Felmy et al. 1990. value of the ion pair stability constant is retained. In this case, ionic association must be explicitly taken into account and aqueous barium sulfate has to be treated as a weak electrolyte. As can be seen on solubility diagrams reported by Harvie et al. 1984. or Monnin and Galinier 1988., there is a marked increase in anhydrite, barite and celestite solubility with increasing aqueous chloride concentration salting-in effect. at 258C. As shown above, ion pair formation is negligible in barite and

celestite saturated pure water. This is also true in chloride bearing solutions. Consequently, as noted by Monnin and Galinier 1988. and Reardon and Amstrong 1987. at 258C, calculated barite or celestite solubilities in chloride solutions are insensitive to the BaSO4 or SrSO4 Pitzer interaction parameters. In models based on the ion pairing phenomenology, calculated barite or celestite solubilities in chloride solutions are insensitive to the ion pair 0 0 stability constant values for BaSO4 aq. or SrSO4 aq.. For example, Moller 1988. calculated barite solubility in NaCl solutions to 2508C at Psat 1 by assuming 0 that the BaSO4 aq. stability constant is equal to that 0 . for CaSO4 aq at these temperatures. Moller 1988. showed that the calculated barite solubilities are sensitive to the NaBa interaction parameter. Conversely, because calculated barite and celestite solubilities in chloride solutions are insensitive to Pitzers parameters for Ba or Sr interactions with aqueous 0 0 sulfate wor equivalently to BaSO4 aq. or SrSO4 aq. stability constantsx, these parameters cannot be unambiguously obtained from fits of solubility data in chloride solutions. 2.2. Binary interaction parameters for aqueous Na, K, Ca and Mg chlorides and sulfates at high temperature The full high temperature model of the NaK CaClSO4 H 2 O system of Greenberg and Moller 1989. has been adopted in this work. These authors provide empirical expressions for the temperature variation of binary interaction parameters for the aqueous sodium, potassium, and calcium sulfates and chlorides. Binary interaction parameters for MgCl 2aq. and MgSO4aq. were taken from Pabalan and Pitzer 1987.. Those for SrCl 2 aq. were taken from Phutela et al. 1987. and those for BaCl 2 aq. were taken from Monnin 1995.. Alternate ion interaction models exist for some of these salts. For example, Holmes et al. 1994. proposed a description of CaCl 2 aq. thermodynamic properties valid to 526 K and 400 bars based on a regression of calorimetric

1 Psat refers to a pressure of 1 bar if temperature is below 1008C, and to pressures corresponding to the liquid vaporequilibrium of H 2 O at higher temperatures.

C. Monninr Chemical Geology 153 (1999) 187209

191

and volumetric data using Pitzers equations. Such new expressions do extend the range of application of ion interaction models to higher temperatures, pressures andror molalities. In the present work, the internally consistent model of Greenberg and Moller 1989. for the NaKCaClSO4 H 2 O system and the MgCl 2 model of Pabalan and Pitzer 1987. were sufficient to study the available barite and celestite solubility data. 2.3. Ternary interaction parameters Ternary interaction parameters for the NaK CaClSO4 H 2 O system are from Greenberg and Moller 1989.. Note that the PabalanPitzer model for the NaKMgClSO4 H 2 O system Pabalan and Pitzer, 1987. is not compatible with the GreenbergMoller model for the NaKCaClSO4 H 2 O system. It follows that the Greenberg and Moller 1989. model cannot be directly merged with the Pabalan and Pitzer 1987. model to build a high temperature model for concentrated solutions in the six-ion NaK CaMgClSO 4 H 2 O system. Among their differences, they use different expressions to describe the temperature dependence of NaCl and NaSO4 interaction parameters. Although these two expressions are of comparable accuracies in calculating the thermodynamic properties and mineral solubilities in the Na 2 SO4 H 2 O and NaClH 2 O systems to high temperature, they lead to different values of ternary ClSO4 and NaClSO4 interaction parameters in the NaClSO4 H 2 O system. Inconsistent parameter sets can lead to a significant accuracy deterioration in solubility calculations at high concentrations see, for example, Fig. 1 in the paper of Greenberg and Moller 1989. or Fig. 2 in the paper of Monnin, 1995.. Reardon and Amstrong 1987. calculated ternary mixing parameters u Na,Sr and c Na,Sr,Cl from emf data Lanier, 1965. for the NaSrClH 2 O system at 258C. Monnin 1995. studied the solubility of solid BaCl 2 P 2H 2 O in NaCl, KCl, and CaCl 2 solutions, to generate the corresponding Pitzers ternary mixing parameters involving the interaction of aqueous Ba with either Na, K or Ca. A large number of ternary mixing parameters are still unknown. Calculated activity coefficients are likely insensitive to mixing parameters representing

MBaSO4 , BaSO4 X, MSrSO4 or SrSO4 X interactions M being a cation and X an anion. due to the limitation of barium or strontium concentrations to low values. Missing parameters are set to zero in the present work, although in some instances see, for example, BaSO4 in KCl solutions. they can be adjusted to the data in order to correct for the discrepancies between calculated and experimental solubilities. Such corrections are sensitive only for concentrations above 1 M or so and, in many cases, they lead to meaningless parameter values that cannot be recommended. As such the present model is limited to moderately concentrated solutions and is suited for seawater, but should be used with caution for brines. 2.4. Pressure effects The effect of pressure on aqueous activity coefficients can be obtained through calculation of partial molal volumes within Pitzers formalism Monnin, 1989, 1990.. Expressions for the partial molal volumes and compressibilities of aqueous solutes in complex mixtures are given by Monnin 1989, 1990. along with expressions of the corresponding Debye Huckel slopes. Compressibility has only a second order effect which is negligible at moderate pressures Millero, 1979; Monnin, 1990.. Because no compressibility data is available for BaCl 2 solutions, no estimate of compressibility effects have been included in this work. One can refer to Monnin 1990. for a discussion of the contribution of compressibility to the thermodynamic properties of aqueous electrolytes. First derivatives of Pitzers interaction parameters with respect to pressure are given by Phutela et al. 1987. for SrCl 2 aq. and by Monnin 1990. 2 for NaClaq., Na 2 SO4aq. and CaCl 2 aq.. Puchalska and Atkinson 1991. give the temperature variation of the standard molal volume of BaCl 2 aq. while temperature and pressure dependent expressions of the BaCl 2 interaction parameters are reported by Manohar et al. 1994..

Note that the a1 parameter in Table 1 in the paper of Monnin 1990. should be 8.525003=10q2 instead of 8.525003=10y2 .

192

C. Monninr Chemical Geology 153 (1999) 187209

3. Barite and celestite solubility products The variation of the solubility product with pressure is given by: ln K sp T , P . s ln K sp T , P0 . y

with the same accuracy using the three-parameter Eq. 8. ln K sp T , P0 . s 224.069 y 35.9422 lnT y 10302.32rT 10 . suggesting that the original interpretation Reardon and Amstrong, 1987. was an overfit. Such overfits may lead to poor values to the solubility products when extrapolated outside the range of the fitted data. Eqs. 9. and 10. lead to Dr S 0 s y138.5 " 0.5 0 Jrmol, Dr H 0 s y3.44 " 0.12 kJrmol and Dr Cp s y1 y298.8 " 0.3 Jrmol K . The standard molal volume of the celestite dissolution reaction is:

Dr V
RT

P y P0 . 7.

where Dr V 0 stands for the standard molal volume of the dissolution reaction. The solubility product, K sp T, P0 . at temperature T and reference pressure P0 can be obtained from the standard heat capacity of the dissolution reaction. When this heat capacity is constant over the investigated temperature range, then K sp T, P0 . is given by: ln K sp T , P0 . s A q B lnT q with As Bs C T

Dr V 0 s V 0 SrSO4 ,aq . y V 0 SrSO4 ,s .


0

11 .

8.

Dr S 0
R 0 Dr C p

0 Dr C p

w 1 q lnT0 x
9.

R 0 Dr H 0 T0 D r C p Csy q R R

The standard volume of celestite V SrSO4 ,s. is 46.25 cm3rmol at 258C and 1 bar Robie et al., 1979. and is considered to be constant over the temperature and pressure ranges investigated in this study. The standard molal volume of aqueous strontium sulfate is calculated at T and Psat from those of sodium sulfate and sodium chloride Monnin, 1990. and of strontium chloride Phutela et al., 1987. at the same temperature and pressure using the additivity rule which is expressed by: V 0 SrSO4 ,aq . s V 0 SrCl 2 ,aq . q V 0 Na 2 SO4 ,aq . y 2V 0 NaCl,aq .
0

12 .

0 where Dr S 0 , Dr H 0 and Dr Cp respectively refer to the standard entropy, enthalpy and heat capacity of the dissolution reaction. R is the gas constant and T0 the reference temperature 298.15 K..

where V is the standard molal volume of the designated aqueous electrolyte. 3.2. Barite The barite solubility product at Psat is calculated from pure water barite solubilities measured by Blount 1977. to 2508C and fit to the following equation: 15806.30 ln K sp T , Psat . s 275.053 y 43.014 lnT y T 13 . from which the following thermodynamic quantities are retrieved: Dr S 0 s y108.5 " 1.5 Jrmol, Dr H 0 0 s 24.755 " 0.35 kJrmol and Dr Cp s y357.2 " 0.8 y1 Jrmol K . The uncertainties attached to these values are based on an uncertainty on K sp of 1% at 298.15 K and of 10% at other temperatures.

3.1. Celestite The celestite solubility product was determined by Reardon and Amstrong 1987. from their own pure water celestite solubility measurements from 10 to 908C. Their 258C value is in close agreement with those of Monnin and Galinier 1988. and Felmy et al. 1990.. Reardon and Amstrong 1987. represented K sp values as a function of temperature using four- and five-parameter expressions which both imply a temperature dependent heat capacity of the dissolution reaction. Their K sp values can be fitted

C. Monninr Chemical Geology 153 (1999) 187209

193

The standard volume of the barite dissolution reaction can be calculated from: DVr 0 s V 0 BaCl 2 ,aq . q V 0 Na 2 SO4 ,aq . y 2V 0 NaCl,aq . y V 0 BaSO4 ,s .
0 3

14 .

where V BaSO4 ,s. s 52.1 cm rmol Robie et al., 1979.. The standard molal volume of aqueous barium chloride is given by Puchalska and Atkinson 1991..

4. Calculated barite and celestite solubilities in electrolyte solutions versus experimental data 4.1. Barite and celestite solubilities in pure water at high pressure Pure water celestite solubilities were measured at 2, 22 and 358C to 1000 bars by MacDonald and North 1974.. Blount 1977. reported pure water barite solubility to 2508C and to 1 kbar. A close agreement between the predicted and measured values can be seen in Fig. 1. It must be emphasized that the only data used in model construction is pure water barite solubility at Psat to get the barite solubility product. The curves in Fig. 1 are model predictions. Keeping in mind that no BaSO4 or SrSO4 interaction parameters or association constants are included in the model, Pitzers expression for the activity coefficient reduces to a simple Debye Huckel term. The close correspondence between the symbols and the curves in Fig. 1 demonstrates that this simple DebyeHuckel term is sufficient to accu rately calculate pure water barite and celestite solubilities. These results confirm the analysis that was carried out above from the results reported in Table 1. 4.2. The Na2 SO4 BaSO4 H2 O system Barite solubility in sodium sulfate solutions to 0.008 mol SO4rkg H 2 O is reported by Felmy et al. 1990. at 258C and Jiang 1996. from 0 to 808C. Unfortunately, Felmy et al. 1990. reported their data only as plots. Fig. 2 compares measured barite solubilities black squares. of Jiang 1996. at 0, 40 and 808C, to those calculated when only NaSO4 interactions are taken into account dashed curves., i.e., when BaSO4 interactions are neglected. The calculated solubilities are lower that the measured values, indicating that the calculated BaSO4aq. activity coefficient is overestimated. As ion association tends to 0 lower activity coefficients, a BaSO4 aq. ion pair may account for the difference between calculated and measured barite solubilities. The difference be-

Fig. 1. The solubility of barite and celestite vs. pressure at various temperatures. The experimental data for celestite filled circles. is from MacDonald and North 1974.. The calculated and experimental celestite solubilities at 22 and 358C are vertically offset by 0.001 molrkg H 2 O for clarity. The experimental data for barite filled squares. is from Blount 1977.. The experimental barite solubility at 248C is represented by open squares and the calculated solubility at this temperature by a dashed curve to avoid overlap with the results at 1898C.

194

C. Monninr Chemical Geology 153 (1999) 187209

Fig. 3. Difference between measured barite solubility in sodium sulfate solutions at 808C and solubility calculated without the 0 BaSO4 aq. ion pair vs. the sodium molality.

Fig. 2. Barite solubility in sodium sulfate solutions. The experimental data filled squares. is from Jiang 1996.. The dashed curve is calculated without the barium sulfate ion pair and the plain curve includes this aqueous complex.

tween calculated and measured barium concentration in sodium sulfate solutions is reported in Fig. 3 for a temperature of 808C. One can see in Fig. 3 that, in accordance with Eq. 6., this difference is constant for sodium concentrations above 0.0025 molrkg H 2 O. Therefore, it provides an estimate of the 0 0 BaSO4 aq. ion pair concentration. If the BaSO4 aq. 0 activity coefficient is taken equal to unity, BaSO4 aq. ion pair dissociation constant values can be calculated from Eq. 6. and the experimental data of Jiang 1996.. These values are reported in Fig. 4 and fit to:
0 ln K BaSO 4 s y6.5032 q

which falls within the range of literature values compiled by Felmy et al. 1990.. Felmy et al.s value 0 for the BaSO4 aq. stability constant lead to a degree of BaSO4 association of 50% at 258C Table 1. while it is only about 15% with the value determined in the present work. It only slightly increases to 0 about 20% at 808C. Adjusting the BaSO4 aq. aqueous complex stability constant to the data is enough to bring calculated barite solubilities in good agreement with the measurements plain curves in Fig. 2. of Jiang 1996.. Note that, with this new value of the 0 BaSO4 aq. stability constant, the logarithms of the

383.677 T

15 .

0 The enthalpy and entropy of the BaSO4 ion pair 0 dissociation reaction are Dr H s y3.190 " 0.039 kJrmol and Dr S 0 s y54.07 " 0.13 Jrmol. The 0 decimal logarithm of the BaSO4 aq. stability con15. is y2.26 at 258C, stant calculated from Eq.

0 Fig. 4. The decimal logarithm of the BaSO4 aq. ion pair dissociation constant vs. temperature.

C. Monninr Chemical Geology 153 (1999) 187209

195

Fig. 5. pK values for the dissociation constants of MSO4 aq. ion pairs with M sCa, Ba, Sr. vs. the solubility products of the corresponding anhydrous solids i.e., anhydrite, barite and celestite.. Plotted values are those given in Table 1 except that for 0 BaSO4 aq.. The value determined in this work has been used.

barite, celestite and anhydrite solubility products at 258C are linearly correlated to the logarithms of the 0 0 0 CaSO4 aq., SrSO4 aq. and BaSO4 aq. dissociation constants Fig. 5.. This may reflect the fact that the more stable the aqueous complex, the more soluble the mineral. The correlation exhibited in Fig. 5 does not hold 0 if the values for the SrSO4 aq. dissociation constant of Reardon and Amstrong 1987. are retained. These values have been determined in pure SrSO4 solutions by Reardon 1983. using conductimetric and ion 0 exchange techniques. His pKwSrSO4 aq.x value at 258C is 2.29, compared to the value of 1.86 of Felmy 0 et al. 1990.. As said above, the SrSO4 aq. dissociation constant would be best determined from celestite solubility in sodium sulfate solutions at various temperatures. For now, in light of the discrepancy between various sources, it is impossible to 0 recommend any value of the SrSO4 aq. dissociation constant. An estimate can be obtained using the value of Felmy et al. 1990. at 258C and the enthalpy

Fig. 6. The solubility of celestite in sodium chloride solutions at 80 and 1208C and at 1 and 400 bars. Plain curves: predicted values; filled squares: experimental data from Schulien 1987.; open squares: experimental data from Strubel 1966..

196

C. Monninr Chemical Geology 153 (1999) 187209

0 of Reardon 1983. for the SrSO4 aq. dissociation reaction 8.7 kJrmol at 258C..

4.3. The NaClSrSO4 H2 O system There is a large body of data for this system. Reardon and Amstrong 1987. studied celestite solubility in sodium chloride solutions from 10 to 908C using their own experimental data and those of Strubel 1966.. The measurements of Jacques and Bourland 1983. are consistent with these data in the region of overlap. Celestite solubilities predicted by the present model are in close agreement with the experimental values, as well as with the calculations of Reardon and Amstrong 1987.. We here illustrate results for some data not used by Reardon and Amstrong 1987.. Celestite solubility measurements reported by Schulien 1987. in solutions up to 2 M NaCl at 80 and 1208 and at 1 and 400 bars are shown in Fig. 6. It is assumed in this study that Schulien reported his data using the molality scale. Adoption of this assumption renders these data consistent with the data of Strubel 1966. in the region of overlap 808C; Fig. 6.. Our predicted solubilities are in good agreement with the measured solubilities of Strubel 1966. and Schulien 1987. Fig. 6. as well as with other data, reported by Jacques and Bourland 1983. at pressures up to 200 bars and temperatures to 1498C that are not included in Fig. 6. Vetter et al. 1983. reported a large body of celestite solubility data in NaCl, CaCl 2 and MgCl 2 solutions and their mixtures at 25, 75, 95 and 1258C. Solubility was measured using a radioactive tracer technique using 90 Sr. Some of the results were cross-checked using 35 S instead of 90 Sr, but the two methods lead to solubility values differing by as much as 30%. Our calculations are in agreement with the 35 S measurements, which is consistent with the results of Monnin and Galinier 1988. at 258C. Lucchesi and Whitney 1962. reported solubility data at 08C. Similar to the 258C results of Monnin and Galinier 1988., calculated solubilities at 08C are higher than the measured values by a factor of about three for the highest NaCl concentration 5.6 molrkg H 2 O.. Despite the fact that measurements below 258C are of great importance for oceanographic stud-

ies, these data are not given further consideration in this work. Note that celestite solubility can be accurately calculated assuming that mixing parameters involving NaSr and NaSrCl interactions at 258C Reardon and Amstrong, 1987. do not vary with temperature and can be used up to 1508C. 4.4. The MgCl 2 SrSO4 H2 O system Data for this system is reported by Vetter et al. 1983. at 75, 95 and 1258C at Psat and magnesium chloride concentrations up to 0.5 M Fig. 7.. Note the discrepancy between the celestite solubility measured with 35 S and that obtained with 90 Sr.

Fig. 7. The solubility of celestite in magnesium chloride solutions at 75, 95 and 1258C and 1 bar or water vapor saturation pressure. Open squares: measured values using 35 S Vetter et al., 1983.; filled squares: measured values using 90 Sr Vetter et al., 1983.; plain curves: solubility calculated using u Cl ,SO 4 and c Mg ,Cl,SO 4 from Pabalan and Pitzer 1987.; dashed curves: solubility calculated using u Cl ,SO 4 and c Mg ,Cl,SO 4 from Moller 1988..

C. Monninr Chemical Geology 153 (1999) 187209

197

The aqueous solution model for this system is built from the binary interaction parameters of Phutela et al. 1987. for SrCl 2 and of Pabalan and Pitzer 1987. for MgCl 2 and MgSO4 . To maintain model consistency, Pitzers mixing parameters involving ClSO4 and MgClSO4 interactions u ClSO and c MgClSO . are also taken from Pabalan 4 4 and Pitzer 1987.. Ternary interaction parameters involving Sr are set to zero. As seen in Fig. 7, predicted solubility agrees closely with the 35 S experimental data at 758C, but departs increasingly from the measured values as temperature increases. Solubility calculated using u ClSO 4 and c MgClSO 4 taken from Moller 1988. are also plotted in Fig. 7. These results deviate from those obtained with the PabalanPitzer ternary parameters; the difference increases with increasing magnesium chloride concentration. This difference is as large as the data scatter at 0.5 M MgCl 2 . This result demonstrates the need to maintain model consistency. As emphasized above it is not possible to build a high temperature NaK CaMgClSO4 H 2 O model for concentrated solutions by combining Mollers and Pitzer and Pabalans models without refitting some parameters. Celestite solubility in moderately concentrated magnesium chloride solutions can nevertheless be satisfactorily calculated up to 1008C by setting u Mg ,Sr and c Mg,Sr,Cl equal to zero. 4.5. The CaCl 2 SrSO4 H2 O system Vetter et al. 1983. give data for this system at 75, 95 and 1258C at Psat . It can be seen in Fig. 8 that celestite solubility is satisfactorily predicted by setting the mixing parameters u Ca,Sr and c Ca,Sr,Cl equal to zero. When the calcium concentration of celestite saturated solutions increases, the solid calcium sulfate stability field is reached. At 258C, gypsum can form in solutions having a calcium molality above 0.75 molrkg H 2 O Monnin and Galinier, 1988.. At higher temperatures, anhydrite is the stable solid calcium sulfate phase. Its solubility decreases with temperature. We find that anhydrite is saturated in celestite saturated solutions when the calcium concentration is equal to 0.5 molrkg H 2 O at 758C, 0.24 molrkg H 2 O at 958C, and 0.06 molrkg H 2 O at 1258C Fig. 8..

Fig. 8. The solubility of celestite expressed as strontium molality. in calcium chloride solutions at 75, 95 and 1258C and 1 bar or water vapor saturation pressure. Open squares: measured values using 35 S Vetter et al., 1983.; filled squares: measured values using 90 Sr Vetter et al., 1983.; plain curves: celestite solubility calculated taking into account anhydrite precipitation; dashed curves: celestite solubility calculated for the sole equilibrium with celestite.

When anhydrite precipitates, the calcium and the sulfate concentrations are lowered, which increases the strontium content of the solution. It can be seen in Fig. 8 that calculated solubilities closely match the experimental data when anhydrite formation is not taken into account. This suggests that anhydrite did not precipitate in the experiments of Vetter et al. 1983., although its stability field was reached. 4.6. The NaClMgCl 2 SrSO4 H2 O, NaClCaCl 2 SrSO4 H 2 O, CaCl 2 MgCl 2 SrSO4 H 2 O and NaClCaCl 2 MgCl 2 SrSO4 H2 O systems Data for these complex systems are from Vetter et al. 1983.. In general our calculated celestite solubili-

198

C. Monninr Chemical Geology 153 (1999) 187209

to 2 M NaCl at 80, 100, and 1208C and at pressures of 1, 200, and 400 bars. Blount 1977. reports data up to 2508C, 500 bars and 4 M NaCl. Uchameyshvili et al. 1966. give barite solubilities up to 3708C at Psat and sodium chloride concentrations up to 2 M. Strubel 1967. reports data for a single 2 M NaCl solution, but to 6008C. It can be seen in Fig. 9 that the data of Templeton 1960., Puchelt 1967. and Strubel 1967. at 508C are consistent, and predicted barite solubilities are within experimental uncertainty. Model predictions are in full agreement with Templetons data at 808C, but are up to 10% higher than the experimental values of Blount 1977. and Strubel 1967. in the 1 to 3 M NaCl concentration range at 1008C. It can

Fig. 9. Barite solubility in sodium chloride solutions at 50, 80 and 1008C. The curves represent calculated values. Experimental data: open squares: Templeton 1960.; filled squares: Schulien 1987.; filled circles: Strubel 1967.; open diamonds: Blount 1977.: open circles: Uchameyshvili et al. 1966.; plusses: Puchelt 1967..

ties agree with the 35 S data, but are about 25% higher than the 90 Sr data, as found for other simpler systems. In addition, data for calcium containing systems were found to be supersaturated with respect to anhydrite at 1258C. 4.7. The NaclBaSO4 H2 O system Numerous experimental solubility data are available in this system. Barite solubility has been measured at 1 bar by Strubel 1967. at 108C intervals from 20 to 1008C in solutions containing up to 2 M NaCl. Puchelt 1967. reports barite solubilities at 508C as a function of NaCl concentration up to halite saturation. Templeton 1960. reports smoothed solubility data to 958C and to 5 M NaCl. Schulien 1987. measured barite solubility in solutions containing up

Fig. 10. Barite solubility in sodium chloride solutions at 150, 200 and 2508C and water vapor saturation pressure. The dashed curves represent solubility values calculated with u Na ,Ba s 0.0. The plain curves are calculated with u Na ,Ba equal to 0.02 at 1508C, 0.05 at 2008C and 0.12 at 2508C. Experimental data: filled circles: Strubel 1967.; open diamonds: Blount 1977.: filled squares: Uchameyshvili et al. 1966..

C. Monninr Chemical Geology 153 (1999) 187209

199

also be seen in Fig. 9 that the data of Schulien 1987. is inconsistent with Templetons results at 808C and with other reported solubilities at 1008C. We also found that calculated barite solubilities are systematically higher than Schuliens measurements at 80, 100 and 1208C and at 1, 200 and 400 bars. Experimental data reported for temperatures above 1008C and at Psat are shown in Fig. 10. Data reported by Blount 1977. and by Uchameyshvili et al. 1966. are in agreement in the region of overlap. Strubels datum agrees with these data at 1508C, but not at 2008C and 2508C. Barite solubilities calculated with u Na,Ba equal to zero and c Na,Ba,Cl equal to 0.0128 Monnin, 1995., which are depicted by the dashed curve in Fig. 10, are higher than the experimental data, with a discrepancy increasing with temperature. As shown by Moller 1988., calculated solubilities can be brought into agreement with the experimental data by adjusting only u Na,Ba . This leads to u Na,Ba values equal to 0.02 at 1508C, to 0.05 at 2008C and 0.12 at 2508C. These values are large

relative to characteristic Pitzers mixing parameter values describing interaction between univalent and divalent cations Pitzer, 1991.. Consequently they should only be used for interpolating between these experimental data. Note that the model with u Na,Ba s 0 and c Na,Ba,Cl s 0.0128 Monnin, 1995. can accurately predict barite solubility to moderate NaCl concentrations even at 2508C, despite the fact that these parameters are based on regression of low temperature data. Barite solubilities in NaCl bearing solutions at 500 bars up to 2508C reported by Blount 1977. are depicted in Fig. 11. Barite solubilities at 1508C and 500 bars calculated using a 1 bar barium sulfate activity coefficient are substantially above Blounts data. The pressure effect on BaSO4aq. activity coefficient can be calculated by integration of its partial molal volume that can be obtained by additivity from the NaClaq., Na 2 SO4 aq., and BaCl 2 aq. partial molal volumes. All the needed parameters first derivatives of Pitzers parameters with respect to

Fig. 11. Barite solubility in sodium chloride solutions at 150, 200 and 2508C and 500 bars. The dashed curve labeled A is calculated with the barium sulfate activity coefficient calculated at water vapor saturation pressure instead of 500 bars. The dashed curve labeled C is calculated with Pitzers parameters for the volumetric properties of BaCl 2 given by Manohar et al. 1994.. The dashed curves labeled B are calculated with Pitzers parameters for the volumetric properties of BaCl 2 equal to those for CaCl 2 Monnin, 1990. at the same temperature. The plain curves are calculated with the values of u Na ,Ba adjusted to the experimental data at Psat see Fig. 10 and text.. The filled squares are the experimental data of Blount 1977..

200

C. Monninr Chemical Geology 153 (1999) 187209

pressure. are taken from Monnin 1990. except those for BaCl 2 which are from Manohar et al. 1994.. The barite solubility calculated assuming u Na,Ba s 0 is in close agreement with the experimental data up to 3 M NaCl, but deviates at higher NaCl molalities. It must be kept in mind that the BaCl 2 volumetric parameters of Manohar et al. 1994. have been determined from density measurements up to 1408C and that the results depicted in Fig. 11 already represent extrapolations outside the range of validity of these parameters. When the same calculation is carried out using CaCl 2 volume parameters Monnin, 1990. in place of those for BaCl 2 , the discrepancy at high NaCl molalities disappears and barite solubility is predicted at 1508 and 500 bars with the same accuracy as at 1508C and Psat . When u Na,Ba is set to 0.02, the agreement between predicted and measured values is very good. It thus appears that the pressure effect on barium sulfate activity coefficients at elevated temperatures above 1508C. can be accounted for by using CaCl 2 volumetric properties as a proxy for those of BaCl 2 . The plain curves at 200 and 2508C and 500 bars in Fig. 11 are calculated assuming this pressure correction and by adopting u Na,Ba values determined from the Psat data. It can again be seen that barite solubility is accurately predicted at 2508C and 500 bars up to 3 M NaCl.

Fig. 12. Barite solubility in potassium chloride solutions at 508C. The filled squares represent measurements of Puchelt 1967.. The dashed curve is calculated with u K,Ba equal to zero and the plain curve with u K,Ba s 0.04.

4.9. The MgCl 2 BaSO4 H2 O system Barite solubilities in MgCl 2 solutions were measured by Puchelt 1967. at 508C. Similar results are found for 508C as were found for 258C by Monnin and Galinier 1988.: the calculated value is in close agreement with the experimental data up to 0.5 mol MgCl 2rkg H 2 O, but lower at higher concentrations. Data reported by Uchameyshvili et al. 1966. for a 0.083 mol MgCl 2rkg H 2 O solution up to 2308C show that barite solubility increases with temperature. The present model predicts this tendency, but the agreement between measured and calculated values is only rough. 4.10. The CaCl 2 BaSO4 H2 O system Barite solubilities calculated at 508C are in close agreement with the data of Puchelt 1967. at concentrations up to about 1 molrkg H 2 O, but depart for experimental data at higher concentrations, similarly to what was observed for KCl solutions see Fig. 11.. Uchameyshvili et al. 1966. measured barite solubility in solutions containing 0.165 and 0.5 mol CaCl 2rkg H 2 O to 2408C. They observed anhydrite formation at temperatures above 2308C. As can be

4.8. The KClBaSO4 H2 O system Barite solubilities in this system are reported by Puchelt 1967. at 508C up to 5 mol KClrkg H 2 O. It can be seen in Fig. 12 that these data are consistent with a u K,Ba parameter value of 0.04, but use of u K,Ba s 0 leads to a discrepancy no greater than about 15% at the highest KCl concentration. Fitting u K,Ba to the barite solubility data lead to meaningless parameters, especially if one keeps in mind that it was found Monnin, 1995. that the solubility of BaCl 2 P 2H 2 O in potassium chloride solutions below 1008C can be accurately calculated with u K,Ba s 0. Moreover, barite solubilities calculated assuming u K,Ba s 0 have been found in rough agreement with the scattered data reported by Uchameyshvili et al. 1966. for a solution containing 0.25 mol KClrkg H 2 O between 100 and 2508C.

C. Monninr Chemical Geology 153 (1999) 187209

201

seen in Fig. 13, calculated barite solubilities are higher than measured values by a factor of 2 at 1008C and by about 25% at 2008C. The anhydrite stability field is reached when temperature increases at constant calcium molality Fig. 13. or when the calcium content of the solution is increased at constant temperature. It can be seen in Fig. 13 that the aqueous phase becomes anhydrite-saturated at about 2108C for a solution containing 0.165 mol CaCl 2rkg H 2 O and at 1658C for a solution containing 0.5 mol CaCl 2rkg H 2 O. At 2308C when anhydrite precipitates from barite saturated solutions, aqueous sulfate content decreases and barium concentration increases. This has been experimentally observed by Uchameyshvili et al. 1966.. The dashed curves in Fig. 13 represent calculated barium molality for equilibrium with barite alone anhydrite is prevented

from precipitating., while the upper part of the plain curves is calculated for a solution at equilibrium with both barite and anhydrite. Note that the calculations at 2308C require extrapolation of some model parameters, including the interaction parameters for barium chloride beyond their 2008C validity limit Monnin, 1995.. It is nevertheless evident in Fig. 13 that calculated values are in better agreement with experimental data if the absence of anhydrite is assumed. This type of agreement was also found for celestite see Fig. 10.. 5. Discussion The experimental data reported in Figs. 113 show that the solubility of barite and celestite in aqueous solutions containing alkali and alkaline earth chlorides increases with the chloride content of the solution, at least to 2508C. This salting-in effect is due to a lowering of the aqueous barium or strontium activity coefficient when the ionic strength of the solution increases. BaSO4aq. and SrSO4 aq. activity coefficients in chloride solutions can be calculated by taking into account the sole binary interactions between aqueous barium or strontium, chloride and the other cations. The interactions between aqueous strontium or barium. and aqueous sulfate which are usually taken into account either as interaction parameters in Pitzers formalism or as ion pair stability constants. can be neglected in this case. In the present work, BaSO4 interactions have been evaluated from barite solubility data in sodium sulfate solutions by determining the aqueous barium sulfate ion pair stability constant as a function of temperature. The determination of the variation of this stability constant with respect to pressure requires new experimental data. It is nevertheless very likely that this variation, which is neglected in the present model, has only a minor effect on calculated barite solubilities at high pressure in sulfate-rich systems. The present model treats aqueous barium sulfate as a weakly associated electrolyte in solution and its degree of association less than 20% up to 808C., only slightly varies with temperature. The results reported in Table 1 indicate that aqueous strontium sulfate is an even stronger electrolyte than barium sulfate. Due to the lack of celestite solubility data in sodium sulfate solutions at various temperatures, the stability

Fig. 13. Barite solubility in calcium chloride solutions vs. temperature. The experimental data filled squares. is from Uchameyshvili et al. 1966.. The plain curves are calculated allowing anhydrite to form. The dashed curves are calculated for equilibrium with barite alone anhydrite remains supersaturated..

202

Table 2 Calculated mean activity coefficients of aqueous barium sulfate in seawater at various salinities and the barite solubility product as a function of temperature and pressure Pressure bars. 1 10 20 1 10 20 1 100 200 1 10 20 1 22 22 22 22 22 22 22 22 22 22 22 22 18 102 105 108 114 118 121 128 162 201 226 231 236 204 Falkner et al. 1992. 0.156 9 0.158 3 0.1578 0.1572 0.156 3 0.156 2 0.1572 0.156 7 0.156 1 0.150 6 0.150 7 0.150 8 0.191 5 This work 0.1650 0.1656 0.1659 0.1653 0.1656 0.1659 0.1652 0.168 9 0.171 3 0.163 3 0.1636 0.1638 0.179 0 Salinity . Ba nmolr kg H 2 O.

Temperature 8C.

g BaSO4 .

Log K sp Barite.

C. Monninr Chemical Geology 153 (1999) 187209

5 5 5 7 7 7 9 9 9 20 20 20 25

Falkner et al. 1992. y10.36 y10.34 y10.33 y10.31 y10.30 y10.29 y10.26 y10.16 y10.07 y10.05 y10.04 y10.03 y9.96

This work y10.38 6 y10.377 y10.36 6 y10.34 1 y10.332 y10.32 1 y10.29 7 y10.19 9 y10.09 9 y10.08 6 y10.078 y10.06 9 y10.00 8

25 1 1

1 1 500

35 35 35

260 180 360

Church and Wolgemuth 1972. 0.121 6 0.1031 0.121 8

This work 0.1271 0.128 0 0.144 2

Church and Wolgemuth 1972. y9.96 y10.27 y9.82

This work y10.00 8 y10.482 y9.957

C. Monninr Chemical Geology 153 (1999) 187209


0 constant of the SrSO4 ion pair cannot be determined 0 as accurately as that of BaSO4 aq.. Provisional values can be obtained using the value of Felmy et al. 1990. at 258C and the enthalpy of Reardon 1983. 0 for the SrSO4 aq. dissociation reaction. Only a few Pitzers mixing parameters involving interactions between aqueous barium or strontium and other cations are available for now. Missing parameters are best determined from data on ternary systems that do not exist at present. As such the present model is limited to moderately concentrated solutions, although it is difficult to give a quantitative meaning to moderately. For example, barite solubility in NaCl solutions can be accurately predicted to high concentrations at temperatures below 1008C Fig. 9., but not above without fitting additional parameters Figs. 10 and 11.. Pressure effects on the thermodynamic properties of aqueous strontium and barium can be calculated using the equations for the volumetric properties of SrCl 2 aq. Phutela et al., 1987. and of BaCl 2 aq. Manohar et al., 1994. along with those for other salts Monnin, 1990. in order to compute partial molal volumes. The derived pressure effect brings calculated celestite and barite solubilities either in pure water to 1 kbar Fig. 1. or in NaCl solutions to 500 bars Figs. 6 and 11. in good agreement with measured values. In the present model, due to the fact that most of the mixing parameters are taken equal to zero, the barite and celestite solubilities are almost exclusively calculated from Pitzers binary parameters for simple salts like NaCl, BaCl 2 , SrCl 2 , etc. Expressions giving these binary parameters as a function of temperature have been shown to be valid to 08C, allowing calculations to be carried out for oceanic conditions. These expressions are summarized in Appendix A. We can compare predicted activity coefficients of BaSO4 aq. in seawater to the results of Church and Wolgemuth 1972. and of Falkner Kenisson et al. 1993.. In Table 2 we report the total or stochiometric. mean activity coefficient of aqueous barium sulfate calculated from the distribution of species and the activity coefficients of the free ions with the following relationship: 2 2y m Ba2q aq .,t P m SO 4 aq .,t P g BaSO 4 aq .,t 2 2y s m Ba2q aq .,F P m SO 4 aq .,F P g BaSO 4 aq .,F

203

The subscript t indicates the total concentration or activity coefficient. of the designated species. One can see Table 2. that our results for the barite solubility product are close to those of Falkner Kenisson et al. 1993., but our values of the BaSO4aq. activity coefficient are higher by about 6% than those of Falkner et al. In agreement with Falkner et al., we find that the variation of the barium sulfate activity coefficient in seawater in the moderate temperature and pressure range considered here is slight, but that a salinity change from 18 to 22 has a noticeable effect. Although our value at 258C and 1 bar is close to that of Church and Wolgemuth 1972., it is markedly different for 18C, 1 bar and for 18C, 500 bars. We also find that the barite solubility product determined in this work is in close agreement with values given by Falkner Kenisson et al. 1993. at all temperatures and pressures where comparison is possible. It is within 10% of Church and Wolgemuth value at 258C and 1 bar, but our value is smaller by almost a factor 2 at 18C, 1 bar and by about 40% at 18C, 500 bars. We have used the present model to calculate the barite saturation state of the worlds ocean. The results are presented elsewhere Monnin et al., submitted..

Acknowledgements I am grateful to Roberto Pabalan and an anonymous reviewer for their careful and insightful reviews of the present paper.

Appendix A The MX activity coefficient and its partial molal volume at T and Psat are obtained by calculating Pitzers interaction parameters and the Debye Huckel slopes at these conditions. The expressions of the various thermodynamic properties of aqueous solutions activity coefficients, partial molal volumes, activity of water, etc.. for simple systems and for multicomponent mixtures within Pitzers formalism are given in many papers see Harvie et al.,

204

C. Monninr Chemical Geology 153 (1999) 187209

Table 3 Values of the fitting constants Eq. A1.. for the binary interaction parameters for aqueous electrolytes a1 a6 AF a. NaCl a. 3.36901531Ey 01 1.92118597Ey 06 a2 a7 y6.32100430Ey 04 4.52586464Eq 01 5.60767406Ey 03 4.43854508Eq 00 1.40677479Ey 03 0.0 y1.80529413Ey 05 6.83040995Ey 02 3.01104957Ey 02 0.0 5.77453682Ey 01 1.46772243Eq 03 y3.54521126Ey 02 y2.42272049Eq 00 1.00721050Ey 02 0.0 0.0 0.0 y1.29807848Ey 03 0.0 8.26906675Ey 03 0.0 2.35793239Ey 02 0.0 0.0 0.0 y4.04750026Ey 02 y1.39082000Eq 00 y1.54170000Ey 02 0.0 9.77090932Ey 03 9.91113465Eq 00 0.0 0.0 0.0 0.0 4.00431027Ey 01 0.0 0.0 0.0 y9.31654000Ey 04 0.0 y1.09438000Ey 02 0.0 y2.49949000Ey 04 0.0 a3 a8 9.14252359E00 0.0 y4.22185236Eq 02 y1.70502337Eq 00 1.19311989Eq 02 y4.23433299Eq 00 8.61185543Eq 00 2.93922611Ey 01 y2.32193726Eq 03 0.0 y2.18434467Eq 04 0.0 2.02438830Eq 03 0.0 y7.58485453Eq 02 0.0 3.22892989Eq 02 y5.94578140Eq 00 9.12712100Eq 01 0.0 y1.41842998Eq 03 0.0 2.06712594Eq 03 0.0 0.0 0.0 2.34550368Eq 03 0.0 0.0 0.0 y4.28383748Eq 02 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 a4 a9 y1.35143986Ey 02 0.0 y2.51226677Eq 00 0.0 0.0 0.0 1.24880954Ey 02 0.0 y1.43780207Eq 01 0.0 y1.89110656Eq 02 0.0 1.46197730Eq 01 0.0 y4.70624175Eq 00 0.0 1.16438557Eq 00 0.0 5.86450181Ey 01 0.0 y6.74728848Eq 00 0.0 0.0 0.0 0.0 0.0 1.70912300Eq 01 0.0 0.0 0.0 y3.57996343Eq 00 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 a5 a10 2.26089488Ey 03 0.0 0.0 0.0 0.0 0.0 0.0 0.0 y6.66496111Ey 01 0.0 y2.03550488Ey 01 0.0 y9.16974740Ey 02 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 y9.22885841Ey 01 0.0 0.0 0.0 8.82068538Ey 02 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

b 0. b 1.
CF

1.43783204Eq 01 y2.61718135Ey 06 y4.83060685Ey 01 0.0 y1.00588714Ey 01 3.41172108Ey 08 8.16920027Eq 01 y1.03923656Ey 05 1.00463018Eq 03 y3.23949532Ey 04 y8.07816886Eq 01 1.43946005Ey 05 2.67375563Eq 01 y3.75994338Ey 06 y7.41559626Eq 00 0.0 y3.30531334Eq 00 4.95713573Ey 07 4.07908797Eq 01 0.0 y1.31669651Eq 01 0.0 y1.88000000Ey 02 0.0 y9.41895832Eq 01 1.51488122Ey 05 3.47870000Eq 00 3.17910000Ey 05 1.93056024Eq 01 y4.62270238Ey 06 0.15 f. 0.0 3.00000000Eq 00 0.0 y1.29399287Eq 02 0.0 0.0 0.0 5.76066000Ey 01 5.93915000Ey 07 2.60135000Eq 00 2.60169000Ey 05 5.95320000Ey 02 2.41831000Ey 07

Na 2 SO4 a.

b 0. b 1.
CF

KCl b.

b 0. b 1.
CF

K 2 SO4 b.

b 0. b 1.
CF

CaCl 2 a.

b 0. b 1.
CF

CaSO4 a.

b 0. b 1. b 2.
CF

MgCl 2 c.

b 0. b 1.
CF

C. Monninr Chemical Geology 153 (1999) 187209 Table 3 continued. a1 a6 MgSO4 c. a2 a7 y5.14100000Ey 01 0.0 y1.47980000Ey 01 0.0 y6.88200000Eq 00 0.0 2.10820004Ey 01 0.0 y1.09557500Ey 02 0.0 y3.36065000Ey 02 0.0 0.0 0.0 6.37500000Ey 04 0.0 3.22500000Ey 03 0.0 y1.53700000Ey 04 0.0 a3 a8 y6.84801984Eq 03 0.0 y5.78048532Eq 03 0.0 y6.78768888Eq 04 0.0 2.62610734Eq 03 0.0 y5.12215000Eq 02 0.0 y1.06459300Eq 03 0.0 4.51223000Eq 00 0.0 y1.33653000Eq 03 0.0 4.37411000Eq 03 0.0 7.87197800Eq 02 0.0 a4 a9 0.0 y1.94722500Ey 06 0.0 0.0 0.0 y2.30350000Ey 05 0.0 8.36666686Ey 07 0.0 0.0 0.0 0.0 0.0 0.0 y5.30213100Eq 00 0.0 1.58751700Eq 01 0.0 3.90395300Eq 00 8.76150800Ey 09 a5 a10

205

b 0. b 1. b 2.
CF

9.39251515Eq 01 1.41316667Ey 03 5.28624842Eq 02 1.57606667Ey 04 1.06150006Eq 03 2.02016667Ey 02 y3.71761334Eq 01 y5.95440002Ey 04 4.42847700Eq 00 9.44275000Ey 06 1.14437700Eq 01 4.24357000Ey 05 y1.54490900Ey 02 0.0 3.43831400Eq 01 4.60872500Ey 06 y1.04230500Eq 02 y6.77403000Ey 06 y2.41201200Eq 01 y1.10262000Ey 05

0.0 1.07875000Ey 09 0.0 0.0 0.0 0.0 0.0 y4.6872000Ey 10 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

SrCl 2 d.

b 0. b 1.
CF

BaCl 2 e.

b 0. b 1.
CF

a. Moller 1988.; b. Greenberg and Moller 1989.; c. Pabalan and Pitzer 1987.; d. Phutela et al. 1987.; e. Monnin 1995.; f. note that the b 0. CaSO4 . parameter is reported as 0.015 by Greenberg and Moller 1989.. This seems to be an error although these authors refer to the paper of Moller 1988.. When used instead of 0.15, this value 0.015. leads to noticeable discrepancies in the calculation of the thermodynamic properties of aqueous calcium ion.

1984; Pabalan and Pitzer, 1987; Weare, 1987; Moller, 1988; Greenberg and Moller, 1989; Monnin, 1989, 1990; Pitzer, 1991, among others.. Pabalan and Pitzer 1987., Phutela et al. 1987., Moller 1988., Greenberg and Moller 1989. and Monnin 1995. have used different empirical functions describing the variation of Pitzers parameters with respect to temperature. For computational ease, these functions can be recast into the following ten parameter expression using simple algebraic transformations: X T . s a1 q a 2 T q q a7 680 y T a3 T q q a 4 lnT q a8 T y 227 a5 T y 263 q a6 T 2

Q v or C v. The a i constants in Eq. A1. are given in Table 3 for Pitzers interaction parameters for pure electrolytes and in Table 4 for the ternary parameters. These constants for the standard partial molal volumes of the aqueous solutes and for the interaction parameters for their volumetric properties are reported in Table 5. Moller 1988. established the following expression for the variation with temperature of AF , the DebyeHuckel slope for the activity coefficient:
AF s 3.36901531 = 10y1 y 6.32100430 = 10y4 T q 9.14252359rTy 1.35143986 = 10y2 lnT q 2.26089488 = 10y3 r T y 263 . q 1.92118597 = 10y6 T 2 q 4.52586464 = 10 1r 680 y T .

q a9 T 3 q a10 T 4

A1.
with X T . being either Pitzers parameters b 0., b 1., b 2., CF , Q or C or their first derivatives with respect to pressure noted b 0.,v , b 1.,v , b 2.,v, CF ,v,

206

Table 4 Fitting constants Eq. A1.. for Pitzers ternary interaction parameters a1 a2 a3 a4 a5 a6

C. Monninr Chemical Geology 153 (1999) 187209

u Cl,SO4 . 25150 a. u Cl,SO4 . 150250 a. c Na,Cl,SO4 . 25150 a. c Na,Cl,SO4 . 150250 a. c K,Cl,SO4 . a. c Ca,Cl,SO4 . a. c Mg,Cl,SO4 . b. u Na,K. a. c Na,K,Cl. a. c Na,K,SO4 . 0150C a. c Na,K,SO4 . 150250C a. u Na,Ca. a. c Na,Ca,Cl. a. c Na,Ca,SO4 . a. u Na,Mg. b. c Na,Mg,Cl. b. u Na,Sr. c. c Na,Sr,Cl. c. c Na,Ba,Cl. d. u K,Ca. a. c K,Ca,Cl. a. c K,Ca,SO4 . a. u K,Mg. b. c K,Mg,Cl. b.
0.07 5.67983244Eq1 y0.009 y3.29811409Eq2 y2.12481475Ey1 y0.018 y1.174Ey1 y5.02312111Ey2 1.34211308Ey2 3.4811517Ey2 6.56482122y2 0.05 y0.003 y0.012 0.07 1.99Ey2 0.051 y0.0021 0.0128 0.1156 4.76278977Ey2 0.0 0.0 2.586Ey2 0.0 y1.63021206Ey1 0.0 y4.42410302Ey2 2.4869833Ey4 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 y1.8747982Eq4 0.0 1.62957351Eq4 3.7561961Eq1 0.0 32.63 1.40213141Eq1 y5.10212917 y8.21656777 y2.12621122Eq1 0.0 0.0 0.0 0.0 y9.51 0.0 0.0 0.0 0.0 y2.70770507Eq1 0.0 0.0 y14.27 0.0 5.70511185 0.0 5.16258079Eq1 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

0.0 8.900099309Eq2 0.0 y3.53341751Eq2 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

0.0 9.2144343Ey5 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

a. Greenberg and Moller 1989.; b. Pabalan and Pitzer 1987.; c. Reardon and Amstrong 1987.; d. Monnin 1995.. All other parameters are equal to zero.

C. Monninr Chemical Geology 153 (1999) 187209

207

Table 5 Fitting coefficients for the interaction parameters for the partial molal volumes and for the standard partial molal volumes of aqueous electrolytes Eq. A1.. a1 a6 NaCl a. V0 8.520003E y 02 7.234513E y 03 5.3699517E y 05 8.6255554E y 10 3.200188E y 3 y7.1016610Ey 9 1.145144E y 05 4.34633E y 11 1.281259E q 03 4.267199E y 03 5.3088E y 05 y3.262413Ey 09 3.200188E y 03 y7.1016610Ey 09 y2.7186409Ey 04 y7.10166104Ey 09 4.175000E q 01 y2.841949Ey 4 y9.949027Ey 05 y3.008955Ey 10 2.34138E y 5 2.97171E y 10 y1.798135Eq 03 y5.421483Ey 02 y5.4648E y 04 y3.2424E y 09 2.2968216E y 05 1.8013521E y 07 y1.22812E q 03 y4.80302E y 03 y2.14925E y 04 0.0 y7.54375E y 05 0.0 0.0 0.0 a2 a7 y3.581619 0.0 y2.6538013Ey 07 y2.682931Ey 02 y1.092875Ey 06 0.0 y4.527545Ey 08 2.595239E y 04 y3.292342 y9.132116Eq 04 4.33707E y 6 0.0 y1.092875Ey 06 0.0 1.84544081Ey 6 0.0 4.121094E y 01 y4.35200E q 04 4.549863E y 7 y1.201352Ey 2 y1.405386Ey 7 0.0 1.719652E q 01 0.0 2.6088E y 06 0.0 7.121304E y 080.0 0.0 4.2668 0.0 0.0 0.0 0.0 0.0 0.0 0.0 6.3836E y 01 0.0 y4.977946Ey 7 0.0 0.0 0.0 2.433938E y 7 0.0 a3 a8 y7.515469Eq 04 y3.007338Eq 02 0.0 2.2020163E y 03 y8.373935Ey 01 4.901041E y 02 0.0 y2.165713Ey 04 y1.231424Eq 05 1.067946E q 3 1.42266E y 01 4.51986E y 03 y8.373935Ey 01 4.901041E y 02 7.1598902E y 02 y4.1886382Ey 03 0.0 y5.54500E q 02 0.0 2.501298E y 3 0.0 y1.7478264Ey 4 0.0 0.0 0.0 5.116E y 03 0.0 y2.131572Ey 04 1.19602E q 5 0.0 7.65753E y 02 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 2.154935E y 3 0.0 0.0 0.0 7.142877E y 4 a4 a9 0.0 y5.839699Ey 06 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 y2.70953E y 04 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 y2.146539Ey 13 0.0 5.6747E y 05 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 a5 a10 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

b 0.,v b 1.,v
CF ,v Na 2 SO4 b. V0

b 0.,v b 1.,v
CF ,v CaCl 2 b. V0

b 0.,v
CF ,v MgCl 2 c. V0

b 0.,v
CF ,v SrCl 2 d. V0

b 0.,v b 1.,v
CF ,v BaCl 2 V 0 e.

y1.553467Eq 02 y1.7000E y 02 b 0.,v f. 8.670239E y 5 5.583802E y 10 b 1.,v f. 0.0 0.0 CF ,v f. y8.240557Ey 5 0.0

a. Rogers and Pitzer 1982.; b. Monnin 1990.; c. Monnin 1989.; d. Phutela et al. 1987.; e. Puchalska and Atkinson 1991.; f. Manohar et al. 1994.. Note that the CaCl 2 parameters can be used in place of these. See text.

208

C. Monninr Chemical Geology 153 (1999) 187209 Ed.., Activity Coefficients in Electrolyte Solutions, Vol. II. CRC Press, pp. 63152. Lanier, R.D., 1965. Activity coefficients of sodium chloride in aqueous three-component solutions by cation-sensitive electrodes. J. Phys. Chem. 69, 39923998. Langmuir, D., Melchior, D., 1984. The geochemistry of Ca, Sr, Ba and Ra sulfates in some deep brines from the Palo Duro Basin, Texas. Geochim. Cosmochim. Acta 49, 24232432. Lucchesi, P.J., Whitney, E.D., 1962. Solubility of strontium sulfate in water and aqueous solutions of hydrogen chloride, sodium chloride, sulfuric acid and sodium sulfate by the radiotracer method. J. Appl. Chem 12, 277279. MacDonald, R.W., North, N.A., 1974. The effect of pressure on the solubility of CaCO 3 , CaF2 and SrSO4 in water. Can. J. Chem. 52, 31813186. Manohar, S., Puchalska, D., Atkinson, G., 1994. Pressure volumetemperature properties of aqueous mixed electrolyte solutions: NaClqBaCl 2 from 25 to 1408C. J. Chem. Eng. Data 39, 150154. Millero, F.J., 1979. Effects of temperature and pressure on activity coefficients. In: Pytckowicz, R.M. Ed.., Activity Coefficients in Electrolyte Solutions, Vol. II. CRC Press, pp. 63152. Moller, N., 1988. The prediction of mineral solubilities in natural waters: a chemical equilibrium model for the system to high temperature and concentration. Geochim. Cosmochim. Acta 52, 821837. Monnin, C., 1989. An ion interaction model for the volumetric properties of natural waters: density of the solution and partial molal volumes of electrolytes to high concentration at 258C. Geochim. Cosmochim. Acta 53, 11771188. Monnin, C., 1990. The influence of pressure on the activity coefficients of the solutes and on the solubility of minerals in the system NaCaClSO4 H 2 O to 2008C and 1 kbar, and to high NaCl concentration. Geochim. Cosmochim. Acta 54, 32653282. Monnin, C., 1994. Density calculation and concentration scale conversions for natural waters. Comp. Geosci. 20 10., 1435 1445. Monnin, C., 1995. Thermodynamic properties of the NaKCa BaClH 2 O system to 473.15 K and solubility of barium chloride hydrates. J. Chem. Eng. Data 40, 828832. Monnin, C., Galinier, C., 1988. The solubility of celestite and barite in electrolyte solutions and natural waters at 258C: a thermodynamic study. Chem. Geol. 71, 283296. Monnin, C., Jeandel, C., Dehairs, F., Cattaldo, T., submitted. The barite saturation state of the worlds ocean. Mar. Chem. Pabalan, Pitzer, 1987. Thermodynamics of concentrated electrolyte mixtures and the prediction of mineral solubilities to high temperature for mixtures in the system NaKMgCl SO4 H 2 O. Geochim. Cosmochim. Acta 51, 24292443. Pitzer, K.S., 1991. Ion interaction approach: theory and data correlation. In: Pitzer, K.S. Ed.., Activity coefficients in Electrolyte Solutions, 2nd edn. CRC Press, pp. 75155. Phutela, R.C., Pitzer, K.S., Saluja, P.P.S., 1987. Thermodynamics of aqueous magnesium chloride, calcium chloride and strontium chloride at elevated temperatures. J. Chem. Eng. Data 32, 7680.

The DebyeHuckel slope for the partial molal volume, Av, is given by Monnin 1990. as a function of temperature: from 0 to 1008C, Av s 8.106377 y 1.256008 = 10 1 T q 7.760276 = 10y4 T 2 y 2.098163 = 10y6 T 3 q 2.25777 = 10y9 T 4 from 100 to 3008C: Av s 3.849971 = 10 2 y 6.982754T q 3.877068 = 10y2 T 2 y 1.11381 = 10y4 T 3 q 1.589736 = 10y7 T 4 y 9.395266 = 10y1 1 T 5 q 6.469241 = 10q4 r 680 y T . When necessary, concentration scale conversions have been carried out using the VOPO program Monnin, 1994.. References
Baker, P.A., Bloomer, S.H., 1988. The origin of celestite in deep-sea carbonate sediments. Geochim. Cosmochim. Acta 52, 335339. Blount, C.W., 1977. Barite solubilities and thermodynamic quantities up to 3008C and 1400 bars. Am. Min. 62, 942957. Church, T.M., Wolgemuth, K., 1972. Marine barite saturation. Earth Planet. Sci. Lett. 15, 3544. Culberson, C.H., Latham, G., Bates, R.G., 1978. Solubilities and activity coefficients of calcium and strontium sulfates in synthetic seawater at 0.5 and 258C. J. Phys. Chem. 82, 26932699. Falkner Kenisson, K., Klinkhammer, G.P., Bowers, T.S., Todd, J.F., Lewis, B.L., Landing, W.M., Edmond, J.M., 1993. The behavior of barium in anoxic marine waters. Geochim. Cosmochim. Acta 57, 537554. Felmy, A.R., Rai, D., Amonette, J.E., 1990. The solubility of barite and celestite in sodium sulfate: evaluation of thermodynamic data. J. Sol. Chem. 19 2., 175185. Greenberg, J.P., Moller, N., 1989. The prediction of mineral solubilities in natural waters: a chemical equilibrium model for the system to high concentration from 0 to 2508C. Geochim. Cosmochim. Acta 53, 25032518. Harvie, C., Moller, N., Weare, J.H., 1984. The prediction of mineral solubilities in natural waters: the system to high ionic strengths at 258C. Geochim. Cosmochim. Acta 48, 723751. Holmes, H.F., Busey, R.H., Simonson, J.M., Mesmer, R.E., 1994. CaCl 2 aq. at elevated temperatures. Enthalpies of dilution, isopiestic molalities and thermodynamic properties. J. Chem. Therm. 26, 271298. Jacques, D.F., Bourland, B.I., 1983. A study of the solubility of strontium sulfate. Soc. Am. Pet. Eng. J. 4, 292300. Jiang, C., 1996. Solubility and solubility constant of barium sulfate in aqueous sodium sulfate solutions between 0 and 808C. J. Sol. Chem. 25 1., 105111. Johnson, K.S., Pytckowicz, R.M., 1979. Ion association and activity coefficients in electrolyte solutions. In: Pytckowicz, R.M.

C. Monninr Chemical Geology 153 (1999) 187209 Puchalska, D., Atkinson, G., 1991. Densities and apparent molal volumes of aqueous BaCl 2 solutions from 15 to 1408C and from 1 to 200 bars. J. Chem. Eng. Data 36, 449452. Puchelt, H., 1967. Zur Geochemie des Bariums in Exogenem Zyklus. Springer, Heidelberg, 205 pp. Raju, K.U.G., Atkinson, G., 1988. Thermodynamics of scale mineral solubilities: 1. BaSO4 s. in H 2 O and aqueous NaCl. J. Chem. Eng. Data 33 s., 490495. Raju, K.U.G., Atkinson, G., 1989. Thermodynamics of scale mineral solubilities: 2. SrSO4 s. in aqueous NaCl. J. Chem. Eng. Data 34 s., 361364. 0 Reardon, E.J., 1983. Determination of SrSO4 ion pair using conductimetric and ion exchange techniques. Geochim. Cosmochim. Acta 47, 19171922. Reardon, E.J., Amstrong, D.K., 1987. Celestite SrSO4s . . solubility in water, seawater and NaCl solutions. Geochim. Cosmochim. Acta 51, 6372. Robie, R.A., Hemingway, B.S., Fisher, J.R., 1979. Thermodynamic properties of minerals and related substances and 298.15 K and 1 bar pressure and at higher temperatures. USGS Bulletin 1452. Schulien, S., 1987. High temperaturerhigh pressure solubility measurements in the systems BaSO4 NaClH 2 O and SrSO4 NaClH 2 O in connection with scale studies. SPE Special Publications 16,264, pp. 233246. Strubel, G., 1966. Die hydrothermale Loslichkeit von Colestin in

209

System SrSO4 NaClH 2 O. Neues Jahr. Miner. Monatsh., 99107. Strubel, G., 1967. Zur Kenntnis und genetischen Bedeutung des Systems BaSO4 NaClH 2 O. Neues Jahr. Miner. Monatsh., 223234. Templeton, C.C., 1960. Solubility of barium sulfate in sodium chloride solutions from 25 to 958C. J. Chem. Eng. Data, 514516. Uchameyshvili, N.Y., Malinin, S.D., Khitarov, N.I., 1966. Solubility of barite in concentrated chloride solutions of some metals at elevated temperatures in relation to problems of the genesis of barite deposits. Geochem. Int. 10, 951963. Vetter, O.J.G, Vanderbroek, I., Nayberg, J., 1983. SrSO4 : the basic solubility data. Int. Symp. on Oilfield and Geothermal Chemistry, Denver, CO, June 13, 1983. Am. Soc. Pet. Eng., 271281. Weare, J., 1987. Models of mineral solubility in concentrated brines with application to field observation. In: Carmichael, I.S.E., Eugster, H.P. Eds.., Thermodynamic Modeling of Geological Material: Minerals, Fluids and Melts. Rev. Miner. 17, pp. 143174, Miner. Soc. Am., 499 pp. Whitfield, M., 1975. The extension of chemical models for seawater to include trace components at 258C and 1 atm. pressure. Geochim. Cosmochim. Acta 39, 15451557. Yuan, M.D., Todd, A.C., 1991. Prediction of sulfate scaling tendencies in oilfield operations. SPE Prod. Eng., 6372.

Vous aimerez peut-être aussi