Anda di halaman 1dari 17

Geochimica et Cosmochtmtca Acta Vol 52, pp 821-837 OOI6-7037j88j$3.00 +.

00
Copyright © 1988 Pergamon Press pic Pnnted III USA

The prediction of mineral solubilities in natural waters: A chemical equilibrium model for the
Na-Ca-CI-S04-H20 system, to high temperature and concentration
NANCY M0'LLER
Department of Cheimistry, University of California, SanDiego, La Jolla, CA 92093, U.S.A.

(Received January 19, 1987; accepted in revised farm January 13, 1988)

Abstract-This paper describes a chemical equilibrium model for the Na-Ca-CI-S04-H20 system which calculates
solubilities from 25°Cto 250°C and from zero to high concentration (I - 18. m) within experimental uncertainty.
Theconcentration and temperature dependence of the model were established by fitting available activity (solubility,
osmotic andemf)data. Asingle ioncomplex, CaSO~, which increases instrength with temperature, isincluded explicitly
in the model.
Thevalidation of model accuracy bycomparison to laboratory and field solubility data isincluded. Applications of
the model are also given. Phase diagrams constructed forthe Na-Ca-CI-S04-H20 system and predicted solubilities of
anhydrite and hemihydrate in concentrated seawater at high temperature are in very good agreement with the data.
Calculations of the temperature of gypsum-anhydrite coexistence as a function of water activity are compared to
reported values, and are used to estimate tbe composition-temperature relation for gypsum-anhydrite transition in a
natural brine evaporation. A preliminary model for barite solubility in sodium chloride solutions at high tempera-
ture (100°C to 250°C), based on this parameterization of the CaS04-NaCI-H20 system, gives good agreement with
the data.

I. INTRODUCTION eludes a description of the solution model parameters which


are adjusted to fit the concentration-dependent solution be-
VARIABLE TEMPERATURE computer models of natural brine havior at constant temperature and pressure. Pitzer and co-
geochemistry have important advantages for application to workers have shown that the temperature-dependence of ac-
studies of salt deposition. They can easily accommodate the tivity and osmotic coefficients for various aqueous salt so-
many chemical interactions and great variability inherent in lutions can be well represented by temperature functions of
natural settings,such as marine vents and geothermal waters. these solution model parameters. Their work on the NaCl-
They can also easily simulate these conditions, which are H20 (PITZER et al.. 1984), Na2S04-H20 (ROOERS and PITZER,
difficult to reproduce in the laboratory. However, to be ap- 1981)and CaCh-H 20 (PHUTELA and PITZER, 1983)systems
plicableto many natural brine-saltinteractions, a geochemical has been incorporated within the framework of the present
model must be reliable to high concentration. Recently, model when possible. PABALAN and PITZER (1987) have re-
HARVIE et al. (1984) have shown that geochemical models cently modeled the quinary Na-K-Mg-CI-S04-H20 system
can be developed which reliably predict solubilities in com- to high temperature. In this model, they evaluated the pa-
plex brines, Na-K-H-Ca-Mg-CI-OH-HS04-S04-C03-HC03- rameters for all the binary (e.g., NaCI-H20) solutions and for
H 20-C02 (g), from zero to high concentration at 25°C. several of the common ion ternary (e.g., NaCl-Na2S04-H20)
The 25°C models of Weare and co-workers (recently re- solutions. They compared calculated solubilities with exper-
viewed in WEARE, 1987) have been successfully applied to imental data but they did not do quarternary and quinary
interpreting the geochemistry of diverse natural settings: solubilitycalculations. The fitting equation used to determine
studies of ancient (Zechstein II: HARVIE et aI., 1980) and the temperature-dependence of the solution model param-
modern (Bocana de Virrila: BRANTLEY et al., 1984) marine eters and of the standard chemical potentials for the various
evaporite systems, studies of the solar evaporation of seawater species in the Na-Ca-CI-S04-H20 model is given in sec-
(M0'LLER et al., in preparation), studies of mineral precipi- tion II.
tation in lakes (Great Salt Lake: SPENCER et al., 1985;Searles Parameterization of the model is described in section III.
Lake (with borate species added): FELMY and WEARE, 1986) Data from relatively simple binary and ternary systems can
and in fluid inclusions (B. LAZAR and H. D. HOLLAND, pers. be used to completely parameterize highly complex brine
commun.). Generalizing these 25°C models to include a models based on the PITZER (1973) equations. The work of
variabletemperature capabilitywillsignificantly broaden their the Oak Ridge group (HOLMES and MESMER, 1986;HOLMES
applicability. et al., 1978, 1981; MARSHALL and SLUSHER, 1966; MAR-
This article describes a variable temperature model which SHALL et al., 1964) provided many of the reliable high tem-
accurately calculates solubilities in the Na-Ca-CI-S04-H20 perature data necessary to parameterize the model. The 25°C
system from 25°C to 250°C and from dilute (0. m) to high solubility models of Weare and co-workers showed that it is
concentration (I '" 18. m) (for 1. atm. pressure to 100°C necessary explicitly to include only strongly associated ion
and for pressuresalong the water vapor saturation curve above complex species (e.g., HS04"; see HARVIE et al., 1984) to
100°C). For solubility modeling, the activity and osmotic model concentrated brine behavior sucessfully, even for
coefficient expressions are the principal thermodynamic complex chemical systems such as borate-seawater (FELMY
quantities of interest. The model expressions, given in section and WEARE, 1986). In modeling the quarternary Na-Ca-CI-
II, are based on those of PITZER (1973). Section II also in- S04-H20 system to high temperature.a CaSO~ complex was
821
822 N. Meller

included in order to fit the data from 25°C to 250°C using


a simple parameterization scheme for the calcium sulfate
interactions. Only solubility data were used to evaluate these
calcium sulfate interaction parameters because binary os-
motic and emf measurements are not feasibledue to the very + L L m.m; (4)tc' + L mal/;cc'a)
c-cc' a
low solubility of the calcium sulfate salts.
Comparisons of model calculations with the data used in + L L mama' (4)~a' + L mcl/;aa'c) + L L mnmcAnc
the model parameterization are given in section III. Section a-ca' c nc

III also compares predictions ofEQ3/6 (WOLERY, 1983)with


the experimental data at high temperature. The EQ3/6 results + L L mnmaAna + L L L mnmcmalnca} , (II-I)
nan c a
at high temperature, similar to those at 25°C, are generally
in fair agreement with experiments in dilute solutions. Newer
versionsof the EQ3NR code provide options to use the Pitzer
equations and various model parameterizations by Pitzer and
a-ca' c a
co-workers as well as the HARVIE et al. (1984) model for the
seawater system at 25°C (T. J. WOLERY, pers. commun.,
1986). n a

Comparisons with data more complex than binary and and


ternary systems and with field data establish the reliability of
the predictive capabilities of the model for natural waters. c a
Section IV describes such comparisons. Phase diagrams for
the CaS04-bearing minerals calculated by the model are in where
good agreement with experimental data not used in the mod-
el's parameterization. The model is also used to predict the
mineral sequence which would occur during the equilibrium
evaporation at 100°C of a seawater-type brine. The agreement c a c-cc' a-ce'
between the model and experimental results (LANGELIER et
In Eqns, (II-I and 11-4), I is the ionic strength. Thesubscripts c and
al., 1950) is excellent. To illustrate another possible appli- a refer to cations andanions in general; whereas, M refers to a specific
cation of the model, the gypsum-anhydrite transition in a cation and X to a specific anion. Nand n designate neutral species
natural brine evaporation undergoing solar heating (HERR. similarly. Themolality and charge of an ionare indicated by m and
MANN et al., 1973) has been calculated. z, respectively.
Z = L Iz,lm, and (11-5)
Sufficientexperimental data are available for the calcium
sulfate system to parameterize a temperature-dependent
model by detailed evaluation of the system's concentration (11-6)
dependence at many temperatures from 25°C to 250°C. Since Eqn. (11-2) issymmetric forcations and anions, the activity
However,to model systemswith insufficient data, parameters coefficient expression foran anion, X, isomitted (see HARVIE et al.,
1984). When the single ion expressions such as Eqn. (11-2), which
must be estimated using information from well-characterized
are defined by convention only, are used to calculate mean activity
chemicallysimilar systems. The success of this approach using coefficients or solubilities they are equivalent to the equations de-
the Pitzer formalism is illustrated in section V, where reliable veloped by Pitzer and coworkers (PITZER, 1973; PITZER and MAY.
calculations of the high temperature behavior of BaS04 in ORGA, 1973, 1974; PITZER and KIM, 1974).
NaC! aqueous solutions are made with a model based on the Thebracketed term in F (Eqn. (11-4» represents the contribution
to theactivity coefficient expression from theelectrostatic, long-range
CaS04 (T) model described here. For a complete high tem- interaction between solution species. Theform selected byPitzer for
perature Na-Ba-CI-S04-H20 model, solubilities in NaC!- theelectrostatic term in thevirial expansion (see first termin brackets
BaCh-H20 and Na2S04-BaS04-H20 solutions must also be of Eqn. (11·1» gave the best agreement of the model and osmotic
accurately modeled. data. Atvery low ionic strength, it reduces to the limiting law. In the
electrostatic term A· (one third the Debye-Hiickellimiting slope) is
dependent on temperature and pressure.
II. MODEL EQUATIONS
A. = ! [21rNOP~]1/2[~]3/2
3 1000 DkT' (11-7)
The solubilitymodeling approach of Weare and co-workers
is based on the PITZER (1973) equation which expresses, in where the variables are p~, the density of pure water, and D, the
virial expansion form, the molality-dependence of the excess dielectric constant of purewater. No is Avogadro's number, k is the
Gibbs free energy, GO<, of an aqueous electrolyte solution. Boltzman constant, and e is the absolute electronic charge.
The second virial coefficients, Band 4>, are dependent on ionic
Expressions for activity and osmotic coefficients, described strength; C·(see Eqn, (11·6», 1/;, Aand I areassumed tobeindependent
in detail in HARVIE et al., (1984) and FELMY and WEARE ofionic strength. Theconcentration-dependence, developed byPitzer
(1986), are derived from appropriate derivatives of the excess andco-workers, foreach oftheabove virial coefficients hasbeen dis-
Gibbs free energy equation. cussed in detail in an earlier paper (HARVIE et al., 1984). J!P and 4>.
in Eqn. (II-I) refer to specific forms forBand 4> found by Pitzer to
Theexpressions fortheosmotic coefficient (et»and fortheactivity give thebest agreement with osmotic data. Gibbs-Duhem integrations
coefficient (In'YM) of a cation (M) and of a neutral species (N) are yield the appropriate forms for the activity coefficient expressions.
given below: In Eqn. (11-2),
Prediction of mineral solubilities 823

(I1-8) where n is the total number of data points, x(l)data is the value of the
where ith experimental point and x(l)catculated is the calculated (model) value
(I1-9) of the ith point.
The parameter, a, is set equal to 2.0 when an electrolyte has a uni-
valent ion. For 2-2 or higher valence electrolytes, a = 104. <P in the Ill. EVALUATION OF PARAMETERS FOR THE
activity coefficient expression (Eqn. (I1-2)) has the following form: TEMPERATURE-DEPENDENT GYPSUM/ANHYDRITE
SOLUBILITY MODEL OF THE
(11-10) Na-Ca-CI-S04-H20 SYSTEM
EO,}(I), which is a function only of ionic strength and electrolyte pair
Sections III-2 through III-7 describe the evaluation to high
type, is nonzero only for nonsymmetric electrolytes (e.g., 1-2 types).
(See PITZER, 1975, and HARVIE and WEARE, 1980, for integrals de- concentration and temperature of the solution model pa-
fining these terms.) In Eqn, (I1-4)the forms of the derivatives, B' and rameters (see section II) and of the standard chemical poten-
<p', are obtained by taking the derivative of the appropriate function tials (f.L°/RT) for the Na·Ca-Cl·S04-H 20 model (for I atm.
with respect to ionic strength. Therefore, at constant temperature pressure to lOO°C and for the pressure of the saturated water
and pressure, the solution model parameters are: {f0l, (fl), {f2l, and
C' for each cation-anion pair; 0 for each unlike cation-cation or vapor above 100°C). Note that for this model the parameters
anion-anion pair; 1/; for each triple ion interaction where the ions are r,
for the various neutral species interactions, A and are zero,
all not of the same sign; ,\ for ion-neutral pairs; and \' for ion-ion- and that a values (see description in section II) do not vary
neutral interactions (see FELMY and WEARE, 1986). The third virial with temperature. Parameterization of the temperature-de-
coefficients, C, 1/;, and \'. are important only at higher concentrations pendence of A¢ (see Eqn. (11-7» is described in section III-I.
(;:::2. m).
For solubility modeling, the standard chemical potentials, IlJ, of In the present model, the standard chemical potentials (f.L °/
the various solution and solid phase model components may also be RT) of the ionic species, Na+, Ca'", cr and S04-, are set
used as parameters. In a solution. these are used along with the activity equal to zero throughout the 25°C to 250°C temperature
and osmotic coefficients to calculate the chemical potentials, Il}, for range. In the 25°C solubility models of Weare and co-workers,
the solute species by
the standard chemical potentials of the species in solution
aG
-
° + RTln "(}m),
= Il} = Il} (II-II) represent the free energy of formation of each ion from the
an} elements, Using the zero standard state convention, the stan-
and for the solvent by dard chemical potentials (f.L°/RT) of the solid phases are the
natural logarithms of the solubility products. Regardless of
eo = IlH,O = Il~,o + R.~\ -
dnH,O
W )
1000 ~ m}<p. (I1-12) the convention selected, the solubility model results are iden-
ticaL
In Eqn, (I1-l2), <P is the osmotic coefficient, R is the universal gas To fit all the data for the binary and ternary subsystems
constant and W is the molecular weight of water (gms.). For a charge- in the Na-Ca-CI-S0 4-H20 model from 25°C to 250°C, a sin-
balanced set of mole numbers, n}, which minimizes the free energy
(see HARVIE et al.. 1984, and references therein) equilibrium con-
gle ion complex, CaSO~(aq), was introduced. The evaluation
ditions are defined at constant temperature and pressure. of its standard chemical potential as a function of temperature
Parameterization of the Na-Ca-Cl-S04-HzO variable temperature is discussed in section III-7, The study of BUSEY and MESMER
model consisted of three principal steps. ( I) Using the above equations, (1978) was used to establish the temperature dependence of
the model parameters were evaluated by minimizing (Via nonlinear
the standard chemical potential for H 20. The remaining
least square analyses) a standard deviation of the variable concen-
tration data in the binary and ternary subsystems at given temper- standard chemical potentials evaluated as a function of
atures from 25°C to 250°C. (2) These preliminary isothermal results temperature were those for halite (NaCl), mirabilite
were studied to test consistency of the various data sets and to suggest (Na ZS04' lOH20), thenardite (Na2S04), gypsum (CaS04'
appropriate data weighting schemes. When possible, all data sets for 2H 20), hemihydrate (CaS04' O.5H20), anhydrite (CaS04),
a given temperature were used. However. when a data set was clearly
inconsistent with the other data sets, it was not used. Priority was
glauberite (Na2S04' CaS04) and labile salt (2Na2S04'
given to sets which yielded smooth temperature variations in the CaS04' 2H 20) (sections IIIA and III-7). The aqueous solu-
parameter values throughout the 25°C to 250°C temperature range. bility of the dihydrate calcium chloride salt is too high (I
(3) Temperature-dependent equations for the various model param- > 20.m) to reliably calculate the temperature variation of its
eters were established by fitting the isothermal parameter values as
standard chemical potential with this model.
a function of temperature with selected terms in the fitting equation
given below (Eqn. (I1-l3)) using linear least square analyses. This Note that a large number of significant figures is reported
fitting equation, for the values of the constants in the variable temperature
equations for the various model parameters (see Tables 1-
Parameter (T) = at + azT + aliT + a4 In T + as/(T - 263.)
3), In the temperature-dependent fitting equation (11-13) there
+ a6Tz + a7/(680. - T) + as/(T- 227.), (I1-l3) is a large amount of cancellation between terms; therefore
is based on those used by ROGERS and PITZER (1981) and PITZER high accuracy in the calculation of each term is required in
et al. (1984) to describe the temperature dependence of the solution order to retain experimental accuracy in the model predic-
model parameters for the Na zS04-HzO and NaCI-HzO systems, re-
tions, The determination of the values of the constants in
spectively. It has been shown by these workers to be consistent with
both activity and thermal data. the temperature-dependent equations for each ofthe Na-Ca-
The relative error of the model's fit to the data is described in terms Cl-S04-H20 model parameters is described below.
of a sigma value, which is defined as:
//I-I) Evaluation of A"
PITZER et al. (1984) give a table of A" VS, temperature values from
(I1-14)
O°C to 300°C (in degrees Kelvin and pressures along the saturation
curve). These tabulated values were put into equation form using
824 N. M011er

Eqn. (11-13) and least square analysis. The resulting fitting constant In building the model described here, it was found that activity
values, listed in Table I, agreedwell (0" = 6.5 X 10-6) with those of curvesgenerated by the Na2S04(T) model of ROGERS and PITZER
PITZER et al. (1981) lead to nonconvex free energy at concentrations below the
solubility ofNa 2S04 at hightemperatures(T,,= 200°C).HOLMES and
MESMER (1986)developed temperature-dependent equationsfor the
III-2) Evaluation of Na, CI interaction parameters aqueous Na2S04 system to 225°C from their isopiestic data. Com-
PITZER et al. (1984) developed a complete equation of state for paring values of In 'Y for Na2S04(aq) against molality at different
NaCI(aq) validfrom O°C to 300°C,saturation pressure to 1000. bars temperatures, the model of HOLMES and MESMER (1986)is in good
and 0.-6.m. They give two sets of fitting constants for their T, P- agreement with ROGERS and PITZER'S (1981) results at 110°C but
dependentequations. One set is to be usedbelow65°C and the other differs from these resultsby 15% at 225°C. In their model of the Na-
setabove65°C. At 200°C (and pressures of200. bars),they estimate K-Mg-CI-S04-H20 system, PABALAN and PITZER (1987) incorporated
that the uncertainty is .004 for osmotic coefficients calculated from the variable temperatureequations of HOLMES and MESMER (1986)
0.-6. m NaCI. PITZER et al. (1984)did not considersolubility data. for Na2S04(aq). SinceHOLMES and MESMER (1986)used a equal to
Therefore, to testtheirsolubility predictions, theycalculated the mean 1.4in their calculations, whereas our modelingapproach(seesection
activitycoefficient for NaCI at saturation molality ('Y~ (T, P») from II) uses a equal to 2.0 for Na2S04, their parameterization was not
the properties of the solidand the infinitely dilute solution as wellas adopted for the present model. For this model, the ion-interaction
from solubility data at I atm. below 100°C and at the saturation parametersfor Na2S04(aq) were evaluatedas a function of temper-
pressure of water above 100°C. Their results agree with the data ature with a equal to 2.0 usingsolubility data in the Na-CI-S04-H20
within the uncertainties of the solubility measurements for 0° < T system (see following section). The resulting parameterization de-
~ 300°C. scribes activities in aqueoussodiumsulfate solutions to concentrations
Giventhe accuracy of the data available in the Na-Ca-CI-S04-H20 above 1.5 m from 25°C to 250°C whilemaintaining necessary con-
system, it wasdecided to generateone set of fittingconstantsfor the vexity relationships.
aqueous NaCI(T) model instead of using the two sets of PITZER et
al. Their tabulated valuesof the Na,CIbinary parametersfrom O°C III-4) Evaluation of parameters in the Na-S04"CI-H20 system
to 300°C at saturation pressure werethereforefit usingEqn. (11-13).
The resulting NaCI(T) model described by the constants in Table I Values of I3lSl.so. and i3lJl.so. weregeneratedfrom 25°C to 2OD°C
and Eqn. (11-13), agrees wellwith PITZER et al.'s tabulated values of by the Na2S04(T) model of ROGERS and PITZER (1981). Values of
I3lSta, i3lJta and C~a, a (C~a,a = 2C; see HARVIE et al., 1984) from ~l.so. were set equal to zero from 25°C to 250°C. Using solubility
0° to 300°C (0" = 3.09 X 10-6 , 0" = 5.36 X 10-7 and 0" = 6.16 X 10-7, data in NaCI-Na2S04-H20 solutions, I3lStso. and i3lJtso. above200°C
respectively). Note that ~l.a is set equal to zero. and C~a.SO., 9a.so., "'C).so•.Na as well as the salt standard chemical
potentials(/Lo jRT's) wereevaluatedfrom 25°C to 250°C.The above
temperature-dependent equations for A" (Table I) and for the Na,
III-3) Evaluation ofNa,S04 interaction parameters CI interaction parameters(Table I) were used in these fits. The sol-
ROGERS and PITZER (1981; see also ROGERS, 1981) developed ubilitydata were taken from LINKE (1965), STEPHEN and STEPHEN
temperature-dependent equationsforthe activity coefficient and other (1963)and BUKSHTEIN et al. (1953) for temperatures up to 100°C
thermodynamicpropertiesof Na2S04 aqueous solutionsfrom 30°C and from SCHROEDER et al. (1935)from 150°Cand above.
to 200°C usingthermal and activitydata. ROGERS (1981)calculated BUSEY and MESMER (1978) give a table of log Qw (Qw
that below 120°C this Na2S0iT) model reproduces osmotic coef- '" [H+][OH-]) vs. temperature (0-300°C) valuesat infinite dilution
ficient data within±2%up to 2.5 m and that above 120°Creasonable (I = 0.). Using Eqn. (11-13), these tabulated valueswerefit in a least
agreement with the data is limited to concentrationsbelow 1.5 m. squares calculation with the constants listed in Table 3 (0" = 2.63

Table 1. Values of the fitting constants (Eq. (II-13» for the Debye-Hiickel model parameter, A., and for the binary interaction parameters L as a
function of temperature (OK) b..

Parameter Constants
a, .., a. 84 a. a~ a? 8R
lA' 3.3690l532d-ol -6.3210043Od-04 9.l4252359dOO -1.35l43986d-02 2.26089488d-03 1.92118597d-06 4.52586464d+01 O.

~~a1)
1.43783204d+Ol 5.60767406d-03 -4.22l85236d+02-2.5l226677dOO O.
-4.83060685d-ol 1.40677479d-03 1.19311989d+02 O. O.
-2.6l7l8l35d-06 4.4385450SdOO -1.70502337dOO
O. O. -4.23433299dOO
l-o., ..a
~oi,a -1.005887l4d-ol -1.805294l3d-05 8.61l85543dOO 1.24880954d-02 O. 3.41172l0Sd-086.83040995d-o2 2.939226l1d-01
8.l6920027d+Ol 3.01 104957Od-02-2.32l93726d+03-1.43780207d+0l-6.66496l 1ld-Ol-1.03923656d-05 O. O.
~ ..so.
1)
..SO. 1.004630l8d+03 5.77453682d-ol -2.l8434467d+04-l.89ll0656d+02-2.035505488-ol-3.23949532d-04 1.46772243d+03 O.
~~a.So. -8.078l6886d+01-3.5452l126d-02 2.0243883Od+03 1.46l9773Od+Ol-9.l697474Od-02 1.43946005d-05-2.42272049dOO O.

P~a -9.4l895832d+01-4.04750026d-02 2.34550368d+03 1.709l2300d+Ol-9.2288584ld-ol 1.5l488l22d-05-1.39082000dOO O.

~~
3.47870000dOO -1.54l70000d-02 O. O. O. 3.l7910000d-05 O. O.
-3.0357873ld+Ol-1.3626472Sd-02 7.64582238d+025.S045806ldOO -3.27377782d-ol 5.69405869d-D6-5.36231106d-Ol O.
13&>,50. 0.15dOO O. O. O. O. O. O. O.
p~so. 3.OdOO O. O. O. O. O. O. O.
Pbto. -1.29399287d+02 4.0043 1027d-o1 O. O. O. O. O. O.
~~.so. O. O. O. O. O. O. O. O.

a. The p(2) values for all binaries except p~so. are equal to zero from 25°C to 250"C. The temperature variation of p~so. given in the table
above was used from 25-50"C to improve the fit to the pure Waler gypsum data (see text).
b. The Na-Cl, Na-S04, Ca-Cl, and Ca-s04 parameters were evaluated from 0°C-300°C, 25°C-250"C, 25°C-250°C, and 25°C-2500C, respec-
tively, with the exception of Pbto.(see above).
c. In Tables (1-3) a high number of significant figures is necessary in the constant values to retain experimental accuracy in the model predictions
(see text, section Ill).
Prediction of mineral solubilities 825

Table 2. Values of the fitting constants L (Eq. (Il-B) for the mixing parameters b. of the ternary solutions as a functionof temperature
(OK) c'.

Parameters Constants
al a2 a3 84 as a6
90050. (25°-150"C) O.07dO O. O. O. O. O.
90 050• (1500-2500C) 5.67983244<1+01 -1.6302l206d-Ql -1.8747982Od+04 5.70511185dOO 8.90099309d+02 9.2l443436<1-Q5
'l'O,sO..Na (25°-150°C) -.009d0 O. O. O. O. O.
'l'o,SO.,Na (15Q0-2S00C) -3.298ll409d+02 -4.424l0302d-02 1.6295735ld+04 5.l6258079d+Ol -3.5334175ld+02 O.
'l'o,so..ea(2S0-2500C) -Q.01SdO O. O. O. O. O.
~..ea(25°-2S00C) O.05dO O. O. O. O. O.
'l'Na,Ca,O (25°-250"C) -Q.OO3dO O. O. O. O. O.
'l'Na,Ca,SO. (25°-250°C) -Q.012d0 O. O. O. O. O.

a. The constants., and as in Eq.(ll-13) are equalto zerofor all the aboveparameters.
b. Note that all ",'s and ~'s for neutralspeciesinteractions (see Fehny and Weare(1986» are zero from 25-2500C.
c. The temperature range listed with each parameterindicatesthe temperatures wherethe associated constants (aca~ are to be used. Note
that data for parameterizing the NaCl-QICl 2- H20 systemextendedonly to 2000C(seetext).

x 10-1) . The resulting temperature-dependent equation was used to sigmas of .038and .078,respectively. Note that above32°C, rnirabilite
calculate equilibrium constants for the hydrated minerals such as becomes metastable. At 50°C, the model calculates thenardite sol-
mirabilite (Na2S0•• IOH 20 ). ubility in pure water 7.6%higher than the lowestexperimental value
Isothermal fitsfrom 25°C to 250°C of the solubilitydata indicated (3.150 m). The agreement of the model with solubilitydata at 100°C
that the mixing parameters (Table 2) could be held constant (8CI•SO, and 150°C (Fig. lb) is also excellent (Ubal = .033 and Ulben = .046 at
= .07 and 1J'cl.SO"Na = - .009) and the data fit within I 0% from 25°C 100°C and Ubal = .020 and Ulben = .035 at 150°C).
to 150°C (see Figs. la, b). Using the resulting model described by Above 150°C, it was necessary to vary the values of 80 •so, and
Tables 1-3, the experimental data for halite solubility,mirabilite sol- 1J'C1,SO"Na with temperature (Table 2) in order to fit the temperature
ubility and thenardite solubility in the NaCl-Na2S0.-H20 solutions variation of the solubility data (see SCHROEDER et al., 1935) within
at 30°C (Fig. la) are fit with sigmas of .020, .028, and .056, respec- 10%. Note that about 230°C, the solubilityofNa2S0. in water drops
tively. At 50°C, the halite and thenardite solubilitydata are fit with sharply (see SCHROEDER et aI., 1935). This break in the solubility

a. b.

TEMPERATURE' 30°C TEMPERATURE' 150°C

20 2.0
o
U'J 1.6
o
";., 1.6
o'" o
z
E 12 E 1.2
o
08 08
04 0.4
O~----:--~+-~~+--<------;~--g, O!:---'--~-"--+-~---;~-'-O
o 2 3 4 5 o 2 4 6
mNaCI mNaCI
c. d.

TEMPERATURE' 200°C TEMPERATURE' 250°C

o
U'J
2 0
(J)
2
N

z'"
o 0
z
E E

'"
l> '"
l>
C
S; --<
'" '"
0
2 4 6 8 0 2 4 6 8
mNaCI mNaCI
FIG. I. A comparison of model calculations (solid lines) of mineral solubilitiesin Na2S0.-NaCl-H20 solutions with
experimental data (see text). (la): mirabilite, thenardite and halite solubilities at 30°C; (I b): halite and thenardite
solubilities at 150°C; (Ie): halite and thenardite solubilitiesat 200°C; (ld): halite and thenardite solubilitiesat 250°C.
826 N. Meller

vs. temperature curve occurs at a temperature corresponding to a a.


change in the stable form of Na2S04(S) (BRODALE and GIAUQUE,
1972). A comparison of solubilities calculated by the model with 5
TEMPERATURE' 50·C
experimental data for 200°C (Uhal = .054 and U'hen = .037) and for
250°C (Uhal = .063 and Uthen = .034) are given in Figs. Ic and Id, 4
respectively. The agreement of the model and experimentaldata for
the solubility of NaCI and Na2S04 in pure water from 25° to 250°C
is excellent (see previous figures). PABALAN and PITZER (1987) report 3
good agreement of their model calculations with the experimental zs'"
c
u
data to about 185°e. They suggest that additional data are needed E 2
to more clearly define the temperature dependency of the model
0
parametersat high temperature.

1II-5) Evaluation of Ca.Cl interaction parameters


0
The activity coefficient for CaCh rapidly increases with concen- 0 2 3 4
tration. The consequent difficulty of modelingat highconcentration mNoCI
solubilities which are dependent on Ca,CI interactions is discussed b.
in HARVIE and WEARE (1980). (Seealso the discussion in PHUTELA
and PITZER, 1983.) PHUTELA and PITZER (1983)established a tem- 5
perature-dependentCaCh-H20 model from 25°C to 200°C by com- TEMPERATURE' lOO·C
bining their 25°C Ca,C1 parameter valueswith those determined by 4
HOLMES et al. (1978) at 108.85°, 140.65°, 172.25° and 200.85°C
from isopiestic data. The maximum concentration of the Phutela
and Pitzer model is 4.3 m. The accuracy of the osmotic coefficients uc'" 3
calculatedby HOLMES et al. is ±.005. u
In the presentmodel, 13&0 is set equalto zeroand the temperature- E
dependent equation ofPhutela and Pitzerfor {jgl,o is usedfrom 25°C 2
to 250°e. Using Eqn. (11-13), temperature-dependentequations for
~.Cl and Ct..Cl wereestablished(Table I) by fittingisothermalvalues 1
determined above200°C from the vapor pressuredata OfZAREMBO
et al. (1980) and WOOD et at. (1984) along with values generated 0
from 25°C to 200°C by the PHUTELA and PITZER (1983)equations. 0 23456
In thesecalculations, the data wereweighted such that the parameter
values ofPhutela and Pitzerwerenot signifcantly degraded. At 200°C, mNoCI
the data OfZAREMBO etal. (1980)and WOOD etal. (1984), converted
FIG. 2. A comparison of model calculations (solid line) of halite
to osmotic coefficient data, are fit with the model (Tables 1-3) with
solubility in CaCh-NaCI-H 20 solutions with experimental data (see
sigmas of .085and .132,respectively. When the lowestconcentration
text). (2a): 50°C; (2b): 100°e.
point in each data set are given low weighting, the ZAREMBO et al.
and the WOOD et al. data are fit with sigmas of .055 and .046, re-
spectively. These fits are comparable to those reported by WOOD et from 25°C to 200°e. With 11 = .05 and'" = -.003, the isopiestic
al. in their isothermal parameterization of the Pitzer equations at data of HOLMES et al. (1981) are fit with U = .0029 at lOO.36°C, U
200°e. At 250°C, the model fits the ZAREMBO et al. data with a = .0031 at 139.9°C, U = .0030 at 172°Candu = .0041 at201.02°e.
sigmaof .052 and the WOOD et al. data with a sigmaof .041.
III-7) Ca,S04 interactions and the Na-Ca-CI-S04-H]O system
III-6} Evaluation 01parameters in the Na-Ca-CI-H]O system
The variations between data setsforthe solubility ofvariouscalcium
The availability of solubilitydata for evaluatingthe mixingparam- sulfate salts have been discussed by several authors (e.g., HARDIE,
eters, I1Na,ea and "'Na,ea.a in the NaCI-CaCh-H20 system is seriously 1967). Figs. 3 and 4 illustrate the variability between several examples
limitedfor temperaturesabove 100°e. For temperaturesup to about of pure water solubility data for gypsum,anhydrite and hemihydrate
130°C,data are found in LINKE (1965)and BUKSHTEIN et al. (1953). from 25°C to 250°e. Note that some of these data were not used to
From 100° to 200°C, the isopiestic study in the NaCI-CaCh-H20 parameterize the model(seebelow). Solubilities of the calciumsulfate
system of HOLMES et al. (1981) was used. Above this temperature, salts in water and in aqueous NaCI solutions are low and decrease
the 200°C-values wereretaineddue to lackofdata fortheirevaluation. with temperature. At 250°C, anhydrite solubilityis only about .0002
Constant values of 11 (.05) and e (-.003) were selected which are m. Since the model parameters for the Ca,S04 binary interactions
suitablefor both the solubilityand isopiestic data (see Table 2). tend to be insensitive to solubility data in non-common ion systems,
Using the model described by Tables 1-3, the NaCI-CaCh-H 20 solubility data in common ion ternary systems(e.g., CaS04-Na2S04-
solubilitydata are fit to approximately 5. m CaCh (I = 15.m)from H20) were used when possible to evaluate these parameters. There
25°C to 200°C within 10%. The chemical potentials for CaCl2' 6H20 are several reportsof such data at temperatures from 25°C to 100°e.
(antarcticite) and CaCh· 2H20 were not determined becauseof the Abovethis temperature rangethe amount of appropriate data is lim-
very high solubility of these salts. The solubility data for halite-sat- ited.
urated calciumchloridesolutionsat 25°C are fitwitha sigmaof .074. The many diverse calcium sulfate solubility data sets from 25°C
The model calculation is about 6.% lower than the data at 5.27 m to 250°C could be fitwithin experimentaluncertainty usingconstant
CaCho At 50°C (seeFig.2a), the sigmafor the halite data set is .067, values for the calcuim sulfate interaction parameters: ~.so"
the difference between the model and experiment being about 3.% ~.so" l3l?~so" aa,so" and"'N a,o.so, by introducing a CaSO~(aq) ion
for the 5.18 m datum. Sigmafor the halite data set at 100°C is .076 pair which increases in strength with temperature. To avoid redun-
(see Fig. 2b). Compared to the 5.12 m CaCh data point, the model dancy problems between ~.so, and p.°/RT for CaSO~, the value of
is about 5.3% low. At 125°C, the difference betweenthe model and ~l.so, wasset equal to 3.0 (~l.so, = 3.1973in HARVIE et al. (1984).
data is 5.2% for the 5.19 m CaCh data point, with the data set being The third virial coefficients, C~ a,so, and "'a.so,.ea, which were both
fit with a sigmaof .052. Above about 5.m CaCh, the difference be- insensitive to the solubilitydata, were set equal to the 25°C-values
tween model calculations and experimental data rapidly increases. of HARVIE et al. (1984)(seeTables 1-2). J3l5l.so, was set equal to zero
Note that this maximum concentration range is similar to that used (seediscussion below). Varyingthe standard chemicalpotentials (p.o/
to establish the PHUTELA and PITZER (1983) binaryCa,C1 parameters RT) of the salts and the CaSO~ ion pair as well as ~.so, and
Predictionof mineral solubilities 827

024 icalpotentials(J.l0 /RT) of the variouscalciumsulfatemineralsin the


model. Note that parameterizations of the gypsum, glauberite and
labilesalt standard chemical potentialsdo not extendto 250°C. Ex-
trapolation much beyond the temperature range listed in the table
.022 is not suggested. Extrapolating the glauberite chemical potential pa-
021 rameterization established from25°Cto 100°Cto higher temperatures
0205 results in inappropriately large amounts of glauberite precipitation
.020 above about 120°C. Therefore, the temperature variation of the
glauberite standard chemical potential above 100°C givenin Table
3 is used in the model to raise the chemical potential artificially in
order to increase glauberite solubility. This avoids the interference
018 ofglauberite precipitation in anhydritesolubility studiesat hightem-
perature. Additional measurements of stable and metastable gypsum,
glauberite and labile salt solubilities are necessary to model the be-
016
havior of these minerals in concentrated brines to 250°C. For the
metastable hemihydrate only solubility data in pure water (e.g.,
>- 015 D'ANS, 1933 and PARTRIDGE and WHITE, 1929), smoothed data
I-
:::; Table II were found to 250°C.
<t
-'
0 .014 In the following sections, the evaluationof the model parameters
::;;
relating to the calcium sulfate system isdiscussed from25°Cto 250°C.
013 Model calculations are also compared with the solubility data used
in the model parameterization from 25°C to 250°C.
012

011

.010 004,...---..----..---,---...,.-----,

009

008 o PW
o D'Ans
....>-
007 ::i

006
i 002

005 ~
70 100 125 o
0 25 50 001

TEMPERATURE (Oe) 0018,...---,---,---,---,----,

FiG. 3. A comparison of model calculations (solid lines)of pure


water solubilities of gypsum and anhydrite with experimental data x DBT
as a function of temperature from 25°C to 125°C. The model cal- .. BD
culationsof the pure water solubility of hemihydrateare compared I> BB
DOl
to the experimental data from 70°C to 125°C. The gypsum data I> 0 PW
(opencircles) are those of:BARBA et al. (1984), BLOCK and WATERS o TR
(1968), Bocx (1961), HILL and WILLS (1938), INNORTA et al.(1980), 0012 • Marshall
MADGIN and SWALES (1956), MARSHALL and SLUSHER (1966), ... Slroub (Linke r p6621
MARSHALL et at. (1964), and OSTROFF and METLER (1966). The
anhydrite data (closed circles) are those of: BLOUNT and DICKSON >-
.... 001
(1969), Bocx (1961), INNORTA et at. (1980), MADGIN and SWALES ...::i
(1956), MARSHALL et al. (1964), PARTRIDGE and WHITE (1929)and ...J
o
POSNJAK (1938). The hemihydrate data (crosses) are those of: Bocx "'0008
(1961), D'ANS (1933), MARSHALL et al. (1964) and PARTRIDGE and
WHITE (1929).
0006

0004
'ltNa.ea.SO. to fitthe calcium sulfate solubility data from 25°Cto 250°C,
indicated that ~so. and 'It~.so. could be held constant (.15 and
-.012, respectively) over the entire temperature range. A simplein- 0002 II!
verse linear relationship (solidline, Fig. 5) between temperatureand o
the ion pair standard chemical potential resultedin good (typically
within 10%) fits to the data sets from 25°C to 150°C. However, 250 275
above 150°C, the strength of the ion pair complex increased more
rapidlywithincreasing temperature(seeFig.5).For comparison, the
In K", values for CaSO~(aq) calculated by YEATTS and MARSHALL FiG. 4. Acomparison ofmodelcalculations (solid lines) ofanhydrite
(1969) as a function of temperature using solubility data in caSo4- and hemihydrate solubilities in pure water with experimental data
Na2S04-NaN03-H20 solutionsare alsoshownin Fig. 5 (dashedline). as a function of temperature from 150°Cto 250°C. The data refer-
Note that in the Na-Ca-CI-S04-H20 model described here, J.l°/RT ences are: BB = BOOTH and BIDWELL (1950); BD = BLOUNT and
(CaSO~) equals In K", since the standard chemical potentials of the DICKSON (1969); D'Ans = D'ANS (1933); DBT = DICKSON et al.
ions have been set equal to zero (seeabove). (1963); Marshall = MARSHALL et al. (1964); PW = PARTRIDGE and
Table3 indicates the temperature range of the solubility data (stable WHITE (1929); Straub = STRAUB (1932); TR = TEMPLETON and
and metastable) usedin establishing the values of the standardchem- ROOERS (1967).
828 N. M011er

TEMP ('CI the model generally falls belowthe majority of data values(seesolid
25 50 75 100 125 150 175 200 225 250 line in Fig. 3). However, the percentage disagreement between the
-30 model and experimentis low. At 25°C, the calculated value (.0148
m) of gypsum-Hjf) solubility is only 2.6% lower than the majority
-40 of measuredvalues(.0152 m).
* The optimal value(asterisk, Fig. 5) for the standard chemicalpo-
0- 50 -----.... - ........ tential (Jlo jRT) of the CaSO~(aq) complexis calculated to be -4.19
~
o~_60
(with j3g~,so. equal to zero)at 25°C. Note that u" jRT(CaSO~) equals
<3 In Kd (seediscussion above). This valuefor In ~ is lessnegative than
.... -70
values reported by other authors at 25°e. For example, this pKd
0;
Q::;

valueof 1.82comparesto 2.032for pKd determined by YEATTS and


~-80
MARSHALL (1969) (dashed line, Fig. 5) from solubility data in the
CaS04-Na2S04-NaN03-H20 system. (See YEATTS and MARSHALL,
-90 1969, for a description of other reported p~ values.) Note that the
negative (f2) values included in the model below 50°C account for
-tOO some additional association. Valuesof pKd are always a function of
the model. It has been our experience that accurate valuesof pKd in
0030 0025 this rangeare difficult to obtain from solubility data (HARVIE, 1981).
I/'K
llI-7b) Evaluation at 35°C
FIG. 5. The variation of the standard chemical potential (JlOjRT)
for the CaSO~ complex species with the reciprocal of temperature At 35°C, the sodium-rich, common ion data of HILL and WILLS
(degrees Kelvin) calculated by the present model (solidline) and by (1938) and MADGIN and SWALES (1956)were used to parameterize
YEATTS and MARSHALL (1969)(dashedline).The asterisk indicates the model. The Hill and Wills data with the double salt,
the optimal value (seetext) of this standard chemical potential cal- 5CaS04. Na2S04 •3H20, were not included because of the scarcity
culated at 25°e. and scatterof the available data at varioustemperatures. The model
isin excellent agreement withthe Hill and Willsdata for halite,then-
ardite, gypsum and g1auberite solubilities (0' = .0118, 0' = .0632, 0'
= .0402 and 0' = .0551, respectively). The fit to the Hill and Wills
1I1-7a) Evaluation at 25°C labile salt data is also within 15% (0' = .1324). The Madgin and
Swales gypsum data set was fit with a sigmaof .0795. MADGIN and
Smoothedsolubility data in the Na-Ca-CI-S04-H20 system, gen- SWALES (1956) also report anhydrite and gypsum solubility data at
erated usingthe HARVIE et al. (1984)parameters, were used in par- 25°e. The model agreement with these data, which were not used
ameterizing the modelat 25°C. At 25°C, our previousmodels(HAR- in the model parameterization, is: O'anh = .2123 and O'gyp = .0173.
VIE and WEARE, 1980; HARVIE et al., 1984) demonstrated that a
CaSO~ ion pair was not required for accurate modeling. However, llI-7e) Evaluation at 40°C
note that a negative I3&so. accounted for the observed negative de-
viation from the limiting law in these models. Use of a negative The NaCl-Na2S04-CaS04-H20 data of BLOCK and WATERS (1968),
l3ru.so. (with or without a caSo~ ion pair) gives the best fit to the the NaCI-CaS04-H20 data of MARSHALL et al. (1964;68 hr. points)
verydilute gypsum solubility data below 50°C. Therefore, the final and MARSHALL and SLUSHER (1966) and the Na2S04-CaS04-H20
model equations include the CaSO~ ion complex species as wellas data of BARBA et al. (1984)were used in parameterizing the model
negative j3g~.so. values (seeTable I) below50°e. at 40°e. The model'scalculationof halite,thenardite,g1auberite and
Using the model described in Tables 1-3, halite, thenardite, an- labilesalt solubility in these aqueous solutionscomparesto the data
hydrite, gypsum, g1auberite and labile salt generated solubilities in withsigmas of .0355, .0637, .1506 and .1239, respectively. The sigmas
the Na-Ca-C1-S04-H20 system at 25°C are fit within less than 5%. for the fitsto the gypsumsolubility data of BARBA et al., BLOCK and
The sigmas for these fits are: .0109, .0083, .0284, .0569, .0251 and WATERS, MARSHALL et al., and MARSHALL and SLUSHER are: .0346,
.0333,respectively. The solubility of gypsum in water calculated by .0534,.0650 and .0650, respectively.

Table 3. Valuesof the fitting constants L (Eq. (11-13» for the standardchemicalpotentials (if) as a function of temperature(OK) b..

Parameters Constants
a a? a. a. a~ lit; a7
K,. (2S0-2S0OC) 1.0403113Od03 4.860928S1d-Ol -3.262243S2d04 -1.90877133d02 -S.3S2048SOd-Ol -2.32009393d-04 S.20S49183dOI
Halite (2S0 -2S00C) S,47825646d03 2.60407906dOO -1.31078912dOS -1.00232512d03 3.5S4S2179dOI -1.l6346S18d-03 7.21138559d02
Thenarditc(25°-250"C) -S.12ISS332d03 -~617S19SOdOO 1.1422S946dOS 9.50777893Od02 -~88441000dOO 1.19461496d-03 -1.907S9987d03
Mirabilite(25°-32°C) -S.7S709707d02 8,40961924d-Ol O. O. O. O. O.
Anhydrite(250-250"C) 3.9383771Od02 -4.30644041d-Ol -3.02S20897d04 -3.60047281dOl S.283SS203dOl 3.96964291d-04 -2.l2549985d03
Hemihydrate(25°-250"C) -2.20114425d03 -1.647S3737dOO 2.9691189Sd04 4.37226849d02 2.376106S2dOI 9.3043S4S2d-04 -2.50170197d03
Gypsum (25°-IIOOC) 1.3S486062d03 2.268779SSd-Ol -6.07006342d04 -2.27071423d02 O. O. O.
Glauberite(25°-1OOOC) z,72477417d04 -2.9S491928dOl -1.4723427Sd06 -2.70490647d03 S.02823869d02 4.80848099d-02 -9.l2078602d05
Glauberite" (IOOOC-2S0°C) -4.8250S7d03 -1.184927dOO 1.477336dOS 8.200287d02 O. O. O.
Labile Salt (25°-7S°C) S.18724792d03 -2.44078217dOI -8.2625S863dOS 4.68697S4Od02 S.38646IS3d02 2.36032300d-02 O.
CaS04'(aq)(25°-ISO"C) -1.4747774SdOI O. 3.26409496<103 O. O. O. O.
CaS04'(aq) (ISO"-2S0OC) 7.16726968dOI -1.0398963Od-01 -1.46847796d04 O. O. O. O.

a. The constant lisin Eq.(II-13)is equal to zero for all the abovepotentials.
b. The temperaturerange listed with each parameterindicatesthe temperatures wherethe associatedconstants(at-Ii?) are to be used.
c. Above about 120OC, Eq. (11-13) and constants a1 - ~ are used in the model only to artificially destabilize the glauberite potential in order to prevent
glauberiteprecipitation(see sectionill-7a).
Prediction of mineral solubilities 829

III-7d) Evaluation at 50 0 e
TEMPERATURE; 50·C
Figures 6a, b show the model's agreementwith the solubility data
at 50°e. The data used in the model's parameterizationat this tem- 003
o
perature were those of OSTROFF and METLER (1966) (see triangles
in Fig. 6a) and HILL and WILLS (1938), with the exception of their
5CaS04' Na2S04' 3H20 data (seeFig.6b). The valuesfor the glaub- ~'¢ 002
erite-thenardite and glauberite-gypsum coexistence points of BARRE <3
(1911)werealso used in the fit. The sigmavaluesfor the fitsto these E
thenardite, gypsum, anhydrite, glauberite and labile salt data are:
.0929,.0604,.0372,.0938and .1295,respectively. BOCK (1961)also
reported data for the solubility of gypsum(Fig. 6a, open circles) and o
anhydrite(Fig. 6a, crosses) in aqueous NaCIsolutionsat 50°e. These
data (lTanh = .2076; lTgyp = .1295) were not used to parameterizethe
model.The data of BOCK (1961)forgypsumand anhydritesolubility
in aqueous NaCl solutions below 50nC were generally inconsistent
with the data used in the model parameterization (see above). At
40°C, the model's fit to these data is lTanh = .1869 and lTgyp = .1305. FIa. 6b.Acomparison ofmodelcalculations (solid lines) ofgypsum,
At 25°C, the sigma valuesfor the fitsto their anhydrite and gypsum anhydrite,glauberite, labilesalt and thenardite solubilities in CaS04'
data are .1591 and .0770, respectively. Figure6b showsthe excellent Na2S04-H20 solutionsat 50°C with experimentaldata (seetext).
agreement of the model with the complexsolubilitydata in the Na-
Ca-CI-S04-H20 system at 50°C of Hill and Wills.
fit by the model with sigmas of .0856, .0566, .0443 and .0864, re-
III-7e) Evaluation at 55°e
spectively. The anhydrite data could also be fit wellisothermally (IT
The data of BLOCK and WATERS (1968) for the NaCI-Na2S04- = .0851). However, establishing the smooth temperature variation
CaS04-H20systemat 55°C, whichincludesthenardite, gypsumand for the anhydrite standard chemical potential given in Table 3, in-
glauberitesolid phases,were used in parameterizing the model. The creases the sigmafor the fit to the anhydrite data to .1568.
sigmasfor the finalmodel's agreementwith thesedata are: IT = .0675
for thenardite solubility, IT = .1090 for glauberite solubility and IT III-7h) Evaluation at 80°C
= .0579 for gypsumsolubility. The Russian data reference (BUKSH-
TEIN et al., 1953) gives smoothed data summarizingvariousdata sets At 80°C,MARSHALL and SLUSHER (1966) reportgypsum solubility
for gypsum solubility in aqueous NaCl solutions. The model agrees data in the NaCI-CaS04·H20 system. The model agrees well with
with these data with IT = .0537. thesedata, whichwereusedin parameterizing the model (IT = .0279).

III-7f) Evaluation at 70° e III-7i) Evaluation at 85°e

The common ion data of BLOCK and WATERS (1968) at 70°C At 85°C, the solubility data of BLOCK and WATERS (1968) with
includes data for the doublesalt,5CaS04· Na2S04'3H20, in addition thenardite, gypsum and glauberiteas the solid phases were used to
to gypsum, glauberite and thenardite solid phase data. Their data, parameterize the model. The sigmas for the fit to the data sets for
without the 5CaS04' Na2S04 . 3H20 salt, as well as NaCI-CaS04- thenardite, glauberite and gypsumare low (IT = .0348, IT = .0996and
H20 data with gypsum as the solid phase (OSTROFF and METLER, IT = .0370, respectively). Their data for the anhydrite solid phase
1966; MARSHALL and SLUSHER, 1966) were used to parameterize exhibitconsiderable scatter. The fit to these data, which has a sigma
the model at 70°e. The agreement with the data is excellent. The of .2307, lies within the spread of the experimentalmeasurements.
sigmas for the fit to the thenardite, gypsum and glauberitedata sets
are .0614, .0503 and .0867, respectively. III-7)) Evaluation at 95°e
At 95°C, the data of MARSHALL and SLUSHER (1966)for [NaCl]
III-7g) Evaluation at 75°e ,,; 2.043 m are in good agreement with the model's prediction (IT
The HILL and WILLS (1938)data for the Na2S04-CaS04-H20 sys- = .0328). However, the model's agreement with the Marshall and
tem (with the exceptionof the data for the 5CaS04' Na2S04' 3H20 Slusherdata over the entire concentration range (to [NaCI] = 5.16
solid phase)were used to parameterizethe model at 75°e. The data m) is considerably poorer (sigma equals .4268). MARSHALL and
sets for thenardite, gypsum,glauberiteand labile salt solubilities are SLUSHER (1966)report in a footnotethat the data for concentrations
of NaCIabove 2.043 m are for solutionssaturated with gypsumand
a double salt ofCaS04 and Na2S04. Moreover, they state that at the
highest molalities NaCI(s) is also present. Note that below95°C, the
model is in good agreement with the data of Marshall and Slusher
over the entire concentration range.

III-7k) Evaluation at lOOoe


<5
en The variation in the measured valuesof MARSHALL et al. (1964),
8 BLOCK and WAtERS (1968) and BLOUNT and DICKSON (1969) for
E anhydritesolubility in aqueous NaClsolutionsat lOO·C is significant
002
TEMPERATURE ; 50·C (see Fig. 7a). The pure water solubilityvalue of anhydrite (see Fig.
3) reportedby MSJand by BD(.00553 m and .00535 m, respectively)
OOt
are in good agreement. The .0131 m pure water value reported by
Block and Waters appears in error (Fig. 3). The MARSHALL et al.
00 2 3 4 5 6 (1964) anhydrite solubilities (Fig. 7a, triangles) are about 30.8% higher
mNaCI than the valuesof BLOUNT and DICKSON (1969)(Fig. 7a, crosses) in
I.m NaCI solutions and about 21.5% higher in 4. m NaCl. The an-
FIa. 6a.Acomparison of modelcalculations (solid lines) ofgypsum, hydrite solubilities of BLOCK and WATERS (1968)(Fig. 7a, open cir-
anhydrite and halitesolubilities in CaSO.-NaCl-H 20 solutions at 50°C cles)are betweenthe BLOUNT and DICKSON (1969)and MARSHALL
with experimentaldata (seetext). et al (1964) valuesabove NaCI = I. m.
830 N. Meller

a. and allowing p.°/RT (anhydrite) to vary, the linear dependence of


P.0 /RT for CaSO~(aq) with I/T (degrees Kelvin) used from 25°C to
Ii
005 HEMIHYDRATE
Ii 100°C (see Table 3 and Fig. 5) can be retained and the solubility
data of BLOUNT and DICKSON at 150°Cfit within 10%. The solubility
data of MARSHALL et al. at 150°C cannot be fit within 15% unless
004
",,~---.:r-~~~~---- the CaS04-H20 point of BLOUNT and DICKSON (1969)is substituted
Ii for their point; or, unless both In Kd (CaSO~(aq» and p.°/RT (an-
c5
(/) 003 "
x ,p 0 hydrite)are allowed to vary. The resultingvalue for In ~ is consid-
0
<..> ANHYDRITE erably less negative than the valuepredicted bythe linear I/T variation
E
002 (see Fig. 5). Since the value of p.°/RT for CaSO~(aq) is not wellde-
TEMPERATURE ' 100·C termined by the data in CaS04-NaCI-H20, it was decided to keep
001 the linearvariation between p.°/RTand I/T(OK) in the presentmodel
and to weight the BLOUNT and DICKSON (1969)data more heavily
0 than the MARSHALL et al. (1964)data.
0 2 3 4 5 6 The agreement with the BLOUNT and DICKSON (1969)data set is
mNoCI very good (u = .0893). The choice of P.0 /RT (anhydrite), resulting
from the weighting scheme selected, caused a larger sigma for the
FIG. 7a. A comparison of model calculations (solid lines) of an-
hydrite and hemihydrate solubilities in CaS04-NaCI-H20 solutions
at 100°C with experimental data (seetext). The dashed line is pre- b.
dicted by EQ3/6. Note that the agreementof the EQ3/6 calculations
0010
with the data in NaCI-CaS04 aqueous solutionsis atypically goodat
100°C(seediscussion in section III).
0008

Varying p.°/RT for anhydrite, each of the above data sets can be ;%'0006
fit well separately. However, since the data sets differ significantly
from one another, the resulting values for the anhydrite standard <3
E
chemical potential also differ. An intermediate value was selected
whichyielded a goodfit to the Na2S04-CaS04-H20 data. This results
in a model(Table3) whichagrees verywell(Fig. 7a; a = .0773)with 0002
the Block and Waters anhydrite data (using the Marshall et al. CaSo4-
H20 point). This model fits the MARSHALL et al. (1964)anhydrite 0
0 04 08 12 16 20 24 28
data with a sigma of .1547 and BLOUNT and DICKSON (1969) an-
hydrite data witha sigma of .1913. At 4. m NaO, the modelcalculates mNo2S04
anhydrite solubilities about 9% higher than the BLOUNT and DICKSON c.
(1969) valueand about II %lowerthan the MARSHALL et al. (1964)
value (Fig. 7a). Therefore, these model calculations are well within
0.024 TEMPERATURE ' 1000C
the spread of the data (see Fig. 7a). Using the p.°/RTvalues for he-
mihydratedescribed in Table 3, the modelfitsthe NaCI-CaS04-H20 NoCI-lm
0.020
data at 100°Cwith hemihydrateas the solidphase of MARSHALL et
al. (1964)with an overallsigmaof .1161 (seeFig. 7a, triangles). c5 0.016
The agreementof the model and the BLOCK and WATERS (1968) (/)
0
common ion Na2S04-CaS04-H20 data with and without NaCI is <..> 0012
E
verygood(seeFigs. 7b-d). Calculated glauberite solubilities compare
with the experimental data with a sigmaof .0787.(The sigmafor the
anhydrite data has been givenabove.)
Predictions of anhydrite solubilities in aqueous NaCI solutions 0004
and in aqueous Na2S04 solutionsat 100°Cby EQ3/6 (seesectionI) 0
are shown by the dashed lines in Fig. 7a and 7b, respectively. The 0 04 08 1.2 1.6 2.0
agreementwith the NaCIsolutiondata at 100°Cshouldnot be taken mNo 2S04
as indicative of EQ3/6's agreement with data in this system at all
temperatures. The agreement is best at 100°C because of the tem- d.
peraturedependence of the EQ3/6parameterization forCaS04-NaO-
H20 solutions. It is important to note that below and above this 003 TEMPERATURE - iOOOC
temperature the difference betweenEQ3/6 calculatedvaluesand the Noel '4m
NaCI-CaS04-H20 experimental data increases significantly (e.g.. see
Fig. 8a). The agreement of EQ3/6 calculations and experiment in
the sulfate-rich solutions is poor over the entire temperature range c5
(/)
(e.g.• Fig. 7b). The addition ofNaO to the Na2S04-CaS04-H20 so- 0
<..>
lutions does not improve the agreement. For example, at 100°C in E
the common ion system withoutNaO, EQ3/6calculates an anhydrite
solubility only about 55% of the experimental valueat 0.5 m Na2S04
(seeFig. 7b). With the addition of 1. m NaCI, the EQ3/6 value (not
shown) is about 53% of the experimental value at 0.5 m Na2S04,
whereas the present model calculates a value about 97% of the ex- 02 0.4 0.6 08
perimental value (see Fig. 7c).
mNo 2S04
III-7l) Evaluation at i50°C
FIG. 7b-d. A comparison of model calculations (solid lines) of
At 150°C, the data of MARSHALL et al. (1964) and of BLOUNT anhydrite, glaubertie and thenardite solubilities in CaS04-NaCI-
and DICKSON (1969)in the NaCI-CaS04-H20 systemwere used to Na2S04-H20 solutions at 100°C with experimental data (see text).
parameterize the model since appropriate common ion data were (7b): NaO = O. m; (7c): NaCI = 1. m; (7d):NaCI = 4. m. The dashed
not found. Usingthe parameterization schemedescribed in Table I line is predictedby EQ3/6 (seediscussion in section III).
Prediction of mineral solubilities 831

model'sfitto the MARSHALL et al. data (u = .2121). The MARSHALL


et al. valuesfor anhydrite solubility are higherthan those calculated 6 TEMPERATURE' 2000C
by the model (12.5% at I. m NaCl and 24.4% at 6. m NaCl). Figure
4 illustrates the significant variation in the reportedanhydritesolubility 5
data in water taken at or near the vapor pressure curve from 150°C
to 250°C. Minimum and maximum solubilities at 150°C differ by
23.1 %. The model calculationof anhydrite solubility in water (solid
line) is compared to the data in Fig. 4.
- 4~
....0
0<1'
en
u
0 3
ANHYDRITE
000

E
2
III-7m) Evaluation at 200°C
Common ion data at 200°C also were not found. Since solubility
data in common ion ternary systems are very important for para-
meterizing the Ca,S04binaryinteractions, it wasdecided to interpolate o0L......'-:0-:!07'04:'"""70.J.:00-8.......".00:t-:-12"..-.'--0""'.0!-;-;16..........0~02.0
data at 200°C usingSTRAUB'S (1932)valuesfor anhydrite solubility mNo 2S04
in the common ion system, CaS04-Na2S04-H20, at 182°C, 207°C,
244°C and 316°C. These interpolated data were used along with FIG. 8b. A comparison of model calculations (solid line) of an-
those of MARSHALL et al. (1964)and BLOUNT and DICKSON (1969) hydrite solubility in CaS04-Na2S04-H20 solutions at 200°C with
in the NaCI-CaS04-H20 systemto parameterizethe modelat 200°C. interpolated experimentaldata (seetext).
Using the resultingtemperature-dependent equations (Tables 1-3),
the model fits the anhydrite solubilitydata in aqueous NaCl (MAR.
SHALL et al. (1964)and BLOUNT and DICKSON (1969), Fig. 8a) with
a sigmaof .1328. is similar, although the percentage difference between the two data
The model (Figs. 8a, b) agrees wellwith the CaS04-H20 point of setsis significant. For example, at 0.5 m NaCIthe anhydritesolubility
BLOUNT and DICKSON (1969) (2.6% higher) whereas, the MARSHALL of BLOUNT and DICKSON (1969) is about 30% less than the TEM·
etal. and interpolated StraubCaS04-H20 pointsare fitless well (11.6% PLETON and ROGERS (1967)value.There is a considerable spreadin
and 8% higher, respectively). At this temperature, the BLOUNT and the purewateranhydritesolubility data takenat 250°Cor extrapolated
DICKSON (1969)point is similarto several other reported values(see to this temperature (seeFig.4). The maximum and minimum values
Fig. 4). Note that maximum and minimum pure water anhydrite differby 85.1%.
solubilities differ by 28.% at 200°C (Fig. 4). Above I. m NaCl, the The TEMPLETON and ROGERS Na2S04-CaS04-H20 data could not
model fits the BLOUNT and DICKSON (1969) data (Fig. 8a, crosses) be fit well, (having sigmas> .3), using a variety of data weighting
with sigmalessthan .06; the total data set has a sigmaof .1062. The schemesas wellas adjusting many model parameters. However, the
MARSHALL et al. data (Fig. 8a, open circles) are fit with a sigma of interpolatedNa2S04-CaS04-H20 data of STRAUB (1932)could be fit
.1984, the model agreeing well (u ,,; .10) with their .0I, .03, 2. and along with the NaCl·CaS04-H20 data of BLOUNT and DICKSON
3. m NaCl points. The model calculationsof anhydrite solubility in (1969) and the NaCl-CaS04-H20 ofTEMPLETON and ROGERS (1967)
the common ion system, CaS04-Na2S04-H20, at 200°C compares (~.3 m NaCl). Using the standard chemical potentials describedby
with the interpolated data of Straub (Fig. 8b) with a sigmaof .0524. Table 3, the TEMPLETON and ROGERS (1967) NaCl-CaS04-H20 data,
The dashedline shownin Fig.8a represents the anhydrite solubilities which were more heavily weighted (~0.3 m NaCl), were fit by the
in aqueous NaCl solutions predicted at 200°C by EQ3/6 (see dis- model(Fig. 9) witha sigmaof .0691. Withthe omission of the CaSo 4-
cussionabove). H20 point, the BLOUNT and DICKSON (1969) data are fit with a
sigmaof .195.At0.5 m NaCl,the modelcalculates anhydritesolubility
about midway between the value of BLOUNT and DICKSON (1969)
III-7n) Evaluation at 250°C (modelis about 17% higher)and the valueof TEMPLETON and ROG-
At 250°C, in addition to the measurementsof BLOUNT and DICK- ERS (1967)(modelis about 19% lower). Asshownin Fig.4, the model
SON (1969) (Fig. 9, crosses) in the CaS04-NaCl-H20 system, there calculation(solidline)of anhydrite solubility in pure water (.000262
are the anhydrite solubility data of TEMPLETON and ROGERS (1967) m) falls about in the middle of the experimental values. The final
for the CaCh and Na2S04 common ion systemsand for the NaCI modelagrees withthe interpolated Straubcommon ion data at 250°C
system(Fig. 9, triangles) as wellas the interpolatedpoints of STRAUB (not shown)with a sigmaequal to .0554.
(1932,see above)for the Na2S04-CaS04-H20 system. The variation Solubilities predictedby the modelin the NaCI-CaCh-CaS04-H20
of anhydritesolubility with NaCIconcentration forboth the BLOUNT systemagree with those of TEMPLETON and ROGERS (1967) with a
and DICKSON (1969)and TEMPLETON and ROGERS (1967)data sets sigma of .1505. Note that these data were not used in the parame-
terization of the model. When the replicate points and the single

0020 x
0016 TEMPERATURE'
250·C
0.016
~ 0012 ~

)'
en 0<1' 0.012
o
u en
E 0008 u
0

E 0008

0.004

2 6 8
0 2 4 6 8
mNoCI
FIG. 8a.Acomparison of modelcalculations (solid line)ofanhydrite
solubility in CaS04-NaCl-H20 solutionsat 200°C with experimental FIG. 9. A comparison of modelcalculations (solid line)of anhydrite
data (seetext). The dashedline is predictedby EQ3/6 (seediscussion solubility in CaS04-NaCl-H20 solutionsat 250°C withexperimental
in section III). data (seetext).
832 N. Meller

highlydivergentpoint (CaS04 solubilityin .0333 m CaCh) of TEM- CI2


PLETON and ROGERS (1967)are removed, the model's predictedsol- 1 8 2 0
ubilities agree well with the TEMPLETON and ROGERS (1967) data
(0- = .1099). H ---ll--
x .... 7,8,9
Comparisons with data more complicated than the binary and
ternary data used in the model's parameterization help to evaluate
the reliability of the model's predictions. They also help to assess the
998 r
x 002

applicability to natural waters. Some predictionsof the model in the


NaCI-Na2S04-CaS04-H20 and NaCl-CaClrCaS04-H20 systems have 996 A
004 Co
been discussed in this section.Further model validation is presented N02
by the applicationsof the model described in the following section.
994 T=100 oC 006
IV. TESTING AND APPLICATIONS OF THE MODEL
t(HA)
Phase diagrams for equilibria involving the CaS04-bearing 992 008
minerals are difficult to determine experimentally because of
the relatively low solubility and reactivity of these minerals. 0 2 8
At 25°C (1 atm. pressure), the phase relations in the 6-com-
ponent system, Na-K-Mg-Ca-CI-S04-H20, and its various
FIG. lOb. Model calculations (solid line) are compared with the
5-, 4-, and 3-component subsystems have been calculated data (crosses) ofGROMOVA (1960) for the Ca-richsideofthe reciprocal
(HARVIE et al., 1982; see also EUGSTER et al., 1980). For the diagram of the quarternary system, Na-Ca-Cl-S04-H20, at 110°C.
quaternary system, Na-Ca-CI-S04-H20, at equilibrium (T, In this diagram, solution compositions are projected from the H20
P) there coexists a maximum of three solids and solution. comer,and the arrowspoint in the directionof decreasing OH,O values.
Because of the low gypsum solubility, the gypsum field dom- The model calculationsare carried out to an ionic strength of about
18. m (see text for the concentration limit of the model parameter-
inates the 25°C reciprocal diagram with solution composi- ization in this system). Point #1 is the halite-anhydrite univariant
tions projected from the H 20 comer (see HARVIE et al., 1982). point shownin Fig. lOa. Crosses #7, #8 and #9 are the halite-calcium
The invariant solutions, saturated with various combinations chloride dihydrate, halite-anhydrite-calcium chloride dihydrate and
of antarcticite (CaClz' 6H 20), halite, gypsum, anhydrite, and anhydrite-calcium chloride dihydrate invariant points, respec-
tively, at 110°C calculated from the Gromova data.
glauberite, thenardite and mirabilite (Na2S04' IOH20), are
located close to the ternary sidelines because of the large dif-
ferences between the mineral solubilities.
HARVIE et al. (1987) was used to calculate the equilibrium
GROMOVA (1960) has reported solubility data for the so-
configuration at 110°C of the NaCl-Na2S04-CaS04-H20 and
lutions: NaCI-Na2S04-CaS04-H20 and NaCi-CaClz-CaS04-
H 20 at 110°C. For comparison with these data, which were NaCI-CaClz-CaS04H20 quaternary systems as functions of
concentration. At 110°C, the anhydrite field dominates the
not used in the model parameterization, the algorithm of
reciprocal diagram and mirabilite and antarcticite no longer
appear on the diagram. The relatively low solubility of an-
hydrite causes the invariant points for the NaCI-NazS04-
GI 2 CaS04-H20 system to be located close to the "Na" sideline.
.996 994 .9 0 Figure lOa expands the "Na" side of the reciprocal diagram

'tfl:
I I !
(

l92
HA 1 in order to more clearly distinguish these points. The very
9 HAGI high solubility of calcium chloride dihydrate results in its
.1-
'; HThGI field lying very close to the CaClz comer ofthe diagram. The
2 "Ca" sideline has been expanded in Fig. lOb.
In Figs. lOa, b the Gromova data (crosses) have been com-
pared to the model calculations (solid lines). In these figures,
x .4 Co the closed circles indicate the calculated compositions of the
GI invariant solutions for the stable equilibria (#2-#9), and for
4 6 the univariant point (#1) in the Na-Ca-Cl-S04 reciprocal sys-
tem at 110°C, projected from H 20. Note that the halite-
.2 x .8 thenardite-glauberite (HThGI) invariant point (#3) is a "eu-
tectic". The arrows in these figures indicate the direction of
o GIA increasing concentration by pointing to lower values of
.004 .008 aH,O' The invariant mineral assemblages and solutions cor-
504 responding to the points (closed circles) given in Fig. lOa are
listed in Table 4 along with the values for the activity ofwater
FIG. lOa. Model calculations (solid lines) are compared with the
data ofGROMOVA (1960)(crosses) for the Na-rich side of the recip- (aH,O)' The experimental data are given in parentheses in
rocal diagram for the quarternary system, Na-Ca-Cl-S04-H20, at Table 4. Figure lOa and Table 4 show that the agreement
110°C. (See also Table 4.) In this diagram, solution compositions between the experimental data and the calculated values is
are projected from the H20 corner and the arrows point in the di- very good (comparable to that found in reciprocal diagrams
rection of decreasing OH,o concentration. The invariant points cal-
culated by the model (#2-#6) and the halite-anhydrite univariant calculated with the HARVIE et al., 1984, model). Note that
point of maximum water activity(OH,O) (#1) are shownby the closed this agreement represents predictions of the model since the
circles. Gromova data were not used in the model parameterization.
Prediction of mineral solubilities 833

Table 4. Invariant PointsL for the Quarternary System, NaC- resulting solution has an ionic strength (.67) as well as Ca'"
NazSOCCaS04-HzO at IIO"C. (Experimental valuesb . are given in and S04- molalities similar to those of seawater.
parentheses)
At 25°C, both the present model and the HARVIE and
Solution Composition WEARE (1980) and HARVIE et al. (1984) models predict the
Point Solid Phases a1\o
mN, me. rna mso same evaporation mineral sequence when the same 4-com-
I" 6.734 .0259 6.734 .0259 H,A .7413 ponent "seawater" (Na-Ca-Cl-S04-H20) composition is used.
2 6.930 .00637 6.664 .1391 H,A,Gl .7401 However, the omission of key seawater components alters
(6.934) (.00590) (6.716) (.1150) this 25°C evaporation sequence from that observed at 25°C
3 7.381 .00121 6.509 .4374 H,Th,G1 .7364 for the 6-component seawater system (HARVIE and WEARE,
4 7.382 .0 6.590 .4366 H,Th .7365
(7.614)
1980) and for the full seawater and carbonate system (HARVIE
(.0) (6.586) (.5143)
5 5.976 .00371 .0 2.992 Th,Gl .9040 et al., 1984). For example, with the 4-component "seawater",
(5.838) (.00312) (.0) (2.922) halite appears in the evaporation path after glauberite (aH20
6 3.292 .00794 .0 1.654 A,Gl .9445 '= .749) instead of before glauberite as in the 6-component
(2.586) (.00835) (.0) (1.301) model.
At lOO°C, gypsum is metastable relative to anhydrite and
a. Point no. 1 is a maximum in a1\Oalong the univariant line of halite-
therefore does not appear in the equilibrium evaporation
anhydritecoexistence.
mineral sequence. The "seawater" (Na-Ca-Cl-S04-H20) only
b. Gromova (1960). has to be concentrated by a factor of 1.22 at lOO°C before
anhydrite precipitates. If anhydrite is withheld in the calcu-
lations at lOO°C (simulating kinetic or other barriers to an-
The calculated univariant point (#1, Figs. lOa, b) is also hydrite precipitation), hemihydrate would appear at: e.F.
given in Table 4. This point is on the univariant line for (concentration factor) '= 2.98, aH20 '= .940 and I '= 1.97.
halite and anhydrite coexistence and is a point where there LANGELIER et al. (1950), in their research on scale control
is a maximum in aH20. At this point, mea '= mso. in the in distillation equipment, studied the solubilities of anhydrite
halite-anhydrite saturated solution. The role of such points and hemihydrate in concentrated seawater brines at lOO°e.
in controlling the eventual solution composition and mineral Their results, which indicated a concentration factor of about
assemblage occurring as a result of brine concentration is 1.5 for anhydrite precipitation and of about 3.05 for hemihy-
discussed for the 6-component (Na-K-Ca-Mg-Cl-S04-HzO) drate precipitation are in very good agreement with those of
seawater system at 25°C in HARVIE et al. (1982). the present model at 100°e. If both anhydrite and hemihy-
The concentration path from the halite-anhydrite point drate are withheld in the model calculations at 100°C, gypsum
(HA, # 1) to the CaCh' 2H 20 region is shown in Fig. lOb. would appear at: CF. = 3.16; aH20 = .936 and I
Using the Gromova data, the ionic strength of the solution = 2.09.
at the halite-anhydrite-calcium chloride dihydrate invariant If the unconcentrated "seawater" medium is "heated"
point (see #8 in Fig. lOb) is about 43. m. The present model, (brought to equilibrium at increasing temperatures), anhydrite
based on a third virial coefficient expansion ofthe free energy would first precipitate at lO8°e. Disallowing the precipitation
expression, does not extend to these high ionic strengths. of anhydrite results in the precipitation of hemihydrate at
Moreover, the NaCI-CaCh-H 20 system was parameterized 144°e. (Gypsum was not parameterized to sufficiently high
by data having maximum CaCh concentrations of about 5. temperatures to calculate at what temperature it would pre-
m (l '= 15. m) (see above). Therefore in Fig. lOb, the solid cipitate.)
line representing model predictions extends to I - 18. m. The present model has been used to calculate the temper-
The agreement with the Gromova data to this concentration ature of gypsum-anhydrite coexistence as a function of the
is excellent. In Fig. lOb, the halite-calcium chloride dihydrate activity of water and ionic strength in CaS04-NaCI-H20 so-
(#7), the halite-anhydrite-calcium chloride dihydrate (#8) and lutions (Fig. 11). For pure solid phases, the gypsum-anhydrite
the anhydrite-calcium chloride dihydrate (#9) invariant points equilibrium at atmospheric pressure is dependent on tem-
at llOoC are from the Gromova data. perature and aH20. In these calculations, NaCI was added to
Gypsum and/or anhydrite are common marine evaporite gypsum-saturated solutions to vary aH20 and the ionic
facies. Therefore, their thermodynamic behavior in NaCl- strength. In Fig. 11, the aH20 vs. transition temperature curve
rich brines has considerable geological importance. The calculated with the present model (solid line) consistently
chemistry of brines containing calcium sulfate can also be predicts higher aH20 values at specific transition temperatures
important to mineral recovery programs and in processes than does HARDIE (1967). On the other hand, the model
involving the manipulation of seawater-type brines at high curve predicts lower QH20 values than predictions based on
temperature. the solubility data of BOCK (1961) (see Fig. 6 in HARDIE,
The equilibrium evaporation mineral sequence at 25°C 1967). In the HARVIE et al. (1984) model, the gypsum stan-
and lOO°C for a 4-component "seawater" solution was cal- dard chemical potential is well-established to high ionic
culated with the present Na-Ca-Cl-S04-H20 solution model. strength using available solubility data (see HARVIE, 1981,
The composition of the "seawater" was derived from the and HARVIE and WEARE, 1980). The standard chemical po-
molal concentrations of seawater (salinity '= 35%0) given in tential of anhydrite was calculated in the HARVIE et al. 25°C
HARVIE (1981, Table 8). The HCO) and CO)- (X2) concen- model using this gypsum standard chemical potential and
trations were added to the Cl" molality and the K+ and Mg''" HARDIE'S (1967) determination of the activity of water at
(X2) concentrations were added to the Na" molality. The gypsum and anhydrite coexistence. Since the present model
834 N. M0IIer

IO,--------..,-----.------,--,-----;r-,-----,--------, brines. Assuming that the chloride ion concentration is con-


servative to halite precipitation, the temperatures of these
salt pan brines are plotted as a function of brine evaporation
ratio (E.R.) in Fig. 12 (crosses). Those brines which were
sampled in the early morning and had much lower solution
temperatures were not included. Note that HERRMANN et af.
report a rise in the air temperature of only about one degree
(31.4 "C to 32.4 "C) during the sampling of the points given
in Fig. 12. Except for canal 12 (E.R. = 5.01) where turbulent
flow is reported, the brine temperatures reported by HERR-
MANN et al. increase with concentration, suggesting a solar
heating effect in this natural system. To simulate this effect,
the model was used to determine the equilibrium configu-
ration of the 4-component "seawater" system as water was
TEMP lOCI removed (evaporation) and the temperature was increased
FiG. II. The variation of the temperature of gypsum-anhydrite simultaneously. A path (solid line, Fig. 12) was produced
coexistence with the activity of water (aH,O) in CaS04-NaCI-H20 which roughly approximates the temperature vs. brine con-
solutions calculated by the present model (seetext; solid line). The centration profile found in the salt works. The algorithm of
variation ofthe coexistence temperaturewithionicstrength (increas-
ing NaCl concentration) is also shown(dashed line). GREENBERG (pers. commun.) was used in these calculations.
In their report, HERRMANN et al. do not evaluate the pre-
cipitated solid phases. However, they report that the measured
uses data generated from the HARVIE et al. model in its pa- calcium ion concentration in the salt works system increases
rameterization at 25°C, the model calculations approach continuously to about pan II (E.R. = 3.84) (see circled point
Hardie's curve as the transition temperature decreases. in Fig. 12), suggesting the start of gypsum precipitation. The
In pure water, the gypsum-anhydrite transition is calculated arrow in Fig. 12 indicates the start of gypsum precipitation
by the present model to occur at 49°C (see the solid line in in the simulation (E.R. = 3.28; T = 34.85°C). For "seawater"
Fig. 11; aH,O = 1.0). A gypsum-anhydrite coexistence tem- evaporation simulated at 25°C (arrow at ordinate in Fig. 12),
perature of 49°C is lower than that calculated by several gypsum precipitates at a similar evaporation ratio (E.R.
workers; e.g., 63.5°e (VAN'T HOFF et al., 1903), 58°C ± 2°C = 3.29). On the other hand, the brine composition at which
(HARDIE, 1967) and 55°C (BLOUNT and DICKSON, 1973). It the gypsum-anhydrite transition takes place is strongly af-
is higher than values generally calculated from solubility data. fected by solar heating. In the 25°C isothermal evaporation
For example, POSNJAK (1938), HILL and WILLS (1938), path this transition occurs at an evaporation ratio of 9.03
MARSHALL and SLUSHER (1966) and BOCK (1961) all predict
values of about 42°C. In assessing the discrepancies of the 1O,-----------,----~----~
published gypsum and anhydrite solubility data below wooe,
HARDIE (1967) argued that since most solubility measure- 9 ANH

ments approached equilibrium only from the undersaturation


side, the transition temperature predicted from these data
would be a minimum value only. However, the model cal- 2 7
culation of 49°C for the gypsum-anhydrite coexistence tem- !«
c::
perature in water is in excellent agreement with the value z 6
o
(49.5°e ± 2'soC) calculated by INNORTA et al. (1980) from ;:::
~ 5
their solubility measurements which approached equilibrium oo,
both from undersaturated and supersaturated solutions. As ...~ 4
more NaCl is added to the solution, the temperature of gyp-
GYP
sum and anhydrite coexistence decreases (see dashed line,
Fig. II). The transition temperature would be 25°C for the
gypsum-saturated solution when sufficient NaCl is added to
reach an ionic strength of 5.45 m. Note that variations in the
30 35 40
specific composition of brines at the same ionic strength will TEMP rC)
slightly alter the predicted transition temperature. For con-
centrated, 4-component "seawater," a transition temperature FiG. 12. Usingthe data (X) of HERRMANN et af. (1973) from the
saltworks ofSecovlje (Yugoslavia), the solution temperature of brines
of 25°C occurs at I = 5.72 m (arrow; Fig. II).
increasingly concentrated duringevaporation are plotted as a function
Models with variable temperature and concentration ca- of temperature. A simulation (using the present CaS04(T) model)
pability can be used to study the effects of changing temper- of this natural evaporation process is shown by the solid line. The
ature on brine evaporation. HERRMANN et al. 's (1973) study arrows (GYP) indicate the start of gypsum precipitation predicted
of salt pan brines (Secovlje, Yugoslavia) undergoing solar byan isothermal evaporation at 25°Cand by the variable temperature
simulation. The arrows (ANH)indicatethe transition from gypsum
evaporation has been used to illustrate such an application. to anhydritepredicted at 25°C and by the variable temperaturesim-
In this study, HERRMANN et al. sampled the major ion con- ulation.The circled datum indicateswhere calciumis first shownto
centrations and the temperatures of the different salt pan be removed from solution in the salt works.
Prediction of mineral solubilities 835

(see arrow at ordinate in Fig. 12); whereas in the solar heated


TEMPERATURE' TEMPERATURE' 150OC
path the transition occurs at E.R. = 5.02. This implies that 100'C
in a natural evaporation system in which there is solar heating,
anhydrite should begin to replace gypsum at a much earlier
/+
stage in the process than predicted by the 25°C models /,..,~-----------­
+
,/~
".
(HARVIE and WEARE, 1980; HARVIE et al .. 1984).

V. MODEL OF BaS04 SOLUBILITY IN AQUEOUS NaCl


SOLUTIONS AT HIGH TEMPERATURE
The Weare solubility modeling approach requires careful TEMPERATURE' 200'C TEMPERATURE' 250'C
evaluation and validation of a large number of parameters 12
/

(similar to the number required by ion pairing models (HAR-


VIE and WEARE, 1980). At present this approach seems to +/ / / / ' / / /
be the only one which can be used with confidence to model
solution behavior accurately above moderately dilute con- + ,.-'
4 //
centration. Since only a limited database exists for many sys- + -:,,/"
tems, it is useful to investigate whether a validated parame-
terization of a well-characterized system can be successfully o
..
/
2 3
transferred to other chemically similar systems. ffiNaCI

BLOUNT (1977) measured barite (BaS04) solubility in


FIG. 13. The prediction of barite solubility in aqueous NaCI so-
aqueous NaCl solutions from 100°C to 300°C and for pres- lutionsat 100°C, 150°C, 200°Cand 250°C using parameters forthe
sures of I bar and higher. To model the Na-Ba-CI-S04-H20 BaS04 system takenfrom the Na-Ca-CI-S04-H20(T) modeland:(I)
system at these high temperatures, would require the eval- the chemical potential (/1°/RT) for baritedetermined from a single
uation of binary and mixing parameters similar to those just BaS04-H20 point(dashed line) or (2)the chemical potential forbarite
determined in step 1 and ONa.Ra fit from the 3. m NaCI point (solid
described for the Na-Ca-CI-S0 4-H20 system. Values of line). The data (X) are thoseof BLOUNT (1977).
11W~.S04' I3ill..S04' 13\l~.S04' eta.SO.' ONa,Ba, ""Na,Ba,Cl, ONa.Ba.SO. and
OCl,SO.,5o from 100°C to 250°C were assumed to be the same
as the corresponding calcium interaction parameter values
rameters specific for the BaS04 system (/oL0 /RT for BaS04(s)
generated with the model described by Tables 1-2. A
and IJNa•Ba) with only two data points at each temperature
BaSO~ ion pair which increased in strength with temperature
yields a model for barite solubility in aqueous NaCI solutions
was introduced as in the calcium sulfate system. Its standard
which is in good agreement with the data to high concentra-
chemical potential values from 100°C to 250°C were taken
tion from 100°C to 250°C. Since this model was not devel-
to be the same as those for the CaSO~ ion pair generated by
oped by evaluating data in the common ion solutions (BaS04-
the model (see Table 3). The Ba, Cl binary interaction pa-
Na2S04-H20 and BaS04-BaCIz-H20), its reliability in pre-
rameter values were taken from the Ca, Cl values. A¢, the
dicting data in such systems is in question (see WEARE, 1987).
Na,CI and Na,S04 binary parameter values and the mix-
When these data, which are particularly important for eval-
ing parameter values for the NaCI-Na2S04-H20 system
uating the strength of the ion complex association, are avail-
were calculated at high temperature as described above
able a reliable model ofthe entire Na-Ba-CI-S0 4-H20 system
(Tables 1-2).
can be developed.
Using the above solution parameter values, the standard
chemical potential (/oL°/RT) for barite was evaluated at each
VI. SUMMARY
temperature by fitting the pure water BaS04-H20 solubility
points of BLOUNT (1977). The resulting /L°/RT values at The variable temperature model for the Na-Ca-Cl-S04-
100°C, 150°C, 200°C and 250°C are: -22.0901, -22.4073, H 20 system presented in this article calculates solubilities
-23.5356, and -25.0543, respectively. In these isothermal from zero to high concentration (I - 18. m) within experi-
fits, /Lo/RT for BaH was set equal to zero. The results of this mental uncertainty over the 25°C to 250°C temperature
model for the BaS04-NaCI-H20 system (dashed lines) are range for which it was parameterized. Generally, this uncer-
compared with the Blount data (crosses) from 100°C to tainty is less than 15%. For well-characterized systems with
250°C in Fig. 13. The fit to the data is within approximately few discrepancies in the data, the disagreement of the model
20% from O. m to a little over 1. m NaCl. and experiment is usually less than 5%. For the complicated
To improve the agreement with the data at high concen- calcium sulfate system, this good agreement of model and
tration ONa,5o was adjusted at each temperature to fit the single experiment was achieved by varying only the standard chem-
BaS04 solubility point in 3. m NaCl using the previously ical potentials of the solids and of a CaSO~ (aq) ion pair.
established values of /oL0 /RTfor BaSOis). The resulting values The discussion of the model parameterization points out
of ONa.Ba at 100°C, 150°C, 200°C and 250°C are: -.0523, several areas where there were problems in the availability
- .0343, -.0269, and -.0217, respectively. This model is and/or accuracy of the data. From 25°C to 150°C, solubility
compared with the Blount data at the different temperatures data in the NaCI-Na2S04-H20 system could be fit well when
in Fig. 13 (solid lines; lTlOO' = .049; lTlW = .056; lT200' = .138 the mixing parameters for this system were held constant.
and lT250' = .217). Therefore, using the parameter values for However, to fit the data above 150°C, it was necessary to
the Na-Ca-CI-S0 4-H20 system and fitting two additional pa- vary these parameters. A very sharp decrease in the solubility
836 N. Meller

vs. temperature curve for Na2S04(aq) is reported (see REFERENCES


SCHROEDER et al.. 1935) at about 230°C. Additional data BARBA D., BRANDANI V. and DI GIACOMO G. (1984) Solubility of
are needed better to define the high temperature behavior of calcium sulfate dihydrate in the system Na2S04-MgCh-H20. J.
Na2S04 solubility in water and sodium chloride solutions. Chern. Eng. Data 29, 42-45.
Data for the common ion NaCl-CaCh-H 20 system above BARRE M. (1911) Etude des sels doubles fermes par les sulfates peu
solubles avec les sulfates alcalins. Ann. Chirn. Phys. 24, 145-201.
200°C were not available and therefore above 200°C the
BLOCK J. and WATERS O. B. JR. (1968) The CaS04-Na2S04-NaCl-
mixing parameters for this system were held constant at the H20 system at 25°C to 100°C. J. Chern. Eng. Data 13, 336-344.
200°C values. BRANTLEY (1987) has recently reported vapor BLOUNT C. W. (1977)Baritesolubilityand thermodynamic quantities
pressure measurements for the NaCl-CaCh-H 20 system at up to 300°C and 1400 bars. Arner. Mineral. 62, 942-957.
these high temperatures. Solubility data for the NaCl-CaCh- BLOUNT C. W. and DICKSON F. W. (1969)The solubilityof anhydrite
(CaS04) in NaCl-H20 from 100 to 450°C and I to 1000 bars.
H 20 system are limited above 25°C. Additional studies are Geochirn. Cosrnochirn. Acta 33, 227-245.
needed to model the behavior of high concentration calcium BLOUNT C. W. and DICKSON F. W. (1973) Gypsum-anhydrite equi-
chloride solutions (CaCh > 6. m) to high temperature. libria in systems CaS04-H20 and CaS04-NaCI-H20. Arner. Min-
To fit the available data in calcium sulfate solutions, it was eral. 58, 323-331.
BOCK E. (1961) On the solubility of anhydrous calcium sulfate and
necessary to allow the strength of the CaSO~(aq) ion pair to
of gypsum in concentrated solutions of sodium chloride at 25°C,
increase significantly above 150°C. Since the aqueous solu- 30°C, 40°C and 50°C. Can. J. Chern. 39, 1746-1751.
bility of calcium sulfate is so low, osmotic and emf experi- BOOTH H. S. and BIDWELL R. M. (1950) Solubilitiesof salts in water
ments are not feasible. Therefore solubility data in solutions at high temperatures. 1. Amer. Chern. Soc. 72, 2567-2575.
with ions common to CaS04 were used when possible to BRANTLEY S. L.( 1987)The chemistryand thermodynamicsof natural
brines and the kinetics of dissolution-precipitation reactions of
evaluate the standard chemical potential of the CaSO~(aq) quartz and water. Ph.D. dissertation, Princeton University.
complex. Such data for the common ion system, CaCh- BRANTLEY S. L., CRERAR D. A., M0LLER N. E. and WEARE J. H.
CaS04-H20, were not found for most temperatures above (1984) Geochemistry of a modern evaporite: Bocana de Virrila,
25°C. Above lOO°C, such data in the Na2S04-CaS04-H20 Peru. J. Sediment. Petrol. 54,447-462.
BRODALE G. E. and GIAUQUE W. F. (1972) The relationship of crys-
system are limited. From 150°C to 250°C, the sulfate-rich
talline forms I, II, IV and V of anhydrous sodium sulfate as de-
common ion data of STRAUB (1932), interpolated at the termined by the third law of thermodynamics. J. Phys. Chern. 76,
temperatures of the isothermal fits, were used to parameterize 737-743.
the model. Solubility data of TEMPLETON and ROGERS (1967) BUKSHTEIN V. M., VALYASHKO M. G. and PEL'SH A. D. (eds.) (1953)
in Na2S04 common ion solutions at 250°C could not be fit Spravochnik po rastvorimosti solevykh sistem, vols. I and II, Izd.
Vses. Nauch.-Issled. Inst. Goz., Goskhimizdat., Moscow-Lenin-
well (having sigmas> .3). However, their 250°C measure- grad, 1270p.
ments in the quaternary NaCI-CaCh-CaS04-H20 system at BUSEY R. H. and MESMER R. E. (1978) Thermodynamic quantities
250°C, which were not used in the model parameterization, for the ionization of water in sodium chloride media to 300°C. J
agreed well with model predictions. In interpreting their Chern. Eng. Data 23,175-176.
D'ANS J. (1933) Die Losungsgleichgewichte der Systeme der Salze
250°C data, Templeton and Rogers postulated the existence
ozeanischer Salzablagerungen. Kali-Forschungsanstalt, Berlin.
ofthree Ca-S04 complexes. The present model includes only DICKSON F. W., BLOUNT C. W. and TUNELL G. (1963). Use ofhy-
a single complex. Additional CaS04 common ion solubility drothermal solution equipment to determine the solubility of an-
data are required to more closely evaluate Ca-S04 complexing hydrite in water from 100°C to 275°C and from I bar to 1000
as a function of temperature. bars of pressure. Arner. J. Sci. 261,61-78.
EUGSTER H. P., HARVIE C. E. and WEARE J. H. (1980) Mineral
Stable and metastable solubility data for most ofthe various equilibria in a six-component seawater system, Na-K-Mg-Ca-CI-
calcium sulfate salts in ternary solutions also were not avail- S04-H20, at 25°C. Geochirn. Cosrnochirn. Acta 44, 1335-1347.
able over the entire 25°C to 250°C temperature range. FELMY A. R. and WEARE J. H. (1986) The prediction of borate
Hemihydrate data in pure water were used from 25°C to mineral equilibria in natural waters: Application to Searles Lake,
250°C. The gypsum, glauberite and labile salt parameteriza- California. Geochirn. Cosrnochirn. Acta 50, 2771-2783.
GROMOVA E. T. (1960) The solubility isotherm of the Na, Ca II CI,
tions ranged from 25°C to only llOoC, lOO°C and 75°C,
S04-H20 system at 110°. Rus. J. Inorg. Chern. 5,1244-1247. See
respectively. Extrapolation outside of these temperature also STEPHEN and STEPHEN (1963), p. 914.
ranges is not advisable. The sensitivity of model predictions HARDIE L. A. (1967) The gypsum-anhydrite equilibrium at one at-
to the quality and availability of data underscores the desir- mosphere. Arner. Mineral. 52,171-200.
HARVIE C. E. (1981) Theoretical investigations in geochemistry and
ability of closely coupling experimental and modeling efforts.
atom surfacescattering. Ph.D. dissertation, Dept. Chemistry, Uni-
The applications of the variable temperature Na-Ca-Cl- versity of California, San Diego.
S04-H20 model discussed here suggest the usefulness of a HARVIE C. E. and WEARE J. H. (1980) The prediction of mineral
full seawater variable temperature model in geochemical and solubilities in natural waters: The Na-K-Mg-Ca-Cl-S04-H20 system
industrial studies. We are presently constructing such a model from zero to high concentration at 25°C. Geochirn. Cosrnochirn
Acta 44, 981-997.
and have recently completed a model of the quinary system, HARVIE C. E., WEARE J. H., HARDIE L. W. and EUGSTER H. P.
Na-K-Ca-Cl-S04-H20, from 25°C to 250°C (GREENBERG (1980) Evaporation of seawater: Calculated mineral sequences.
and M0LLER, in prep.). Science 208, 498-500.
HARVIE C. E., EUGSTER H. P. and WEARE J. H. (1982) Mineral
Acknowledgements-A wish to thank Professor John H. Weare and equilibria in the six-component seawater system, Na-K-Mg-Ca-
Dr. Jerry Greenberg for their generous assistance throughout this CI-S04-H20 at 25°C. II: Composition of the saturated solutions.
work. This work has been supported by funds from the Department Geochim. Cosrnochirn. Acta 46, 1603-1618.
of Energy:DOE DE-AC03-85SF-15522, the National Science Foun- HARVIE C. E., M0LLER N. and WEARE J. H. (1984) The prediction
dation: NSF OCE 85-07902 and the American Chemical Society: of mineral solubilities in natural waters: The Na-K-Mg-Ca-H-CI-
ACS-PRF 14550-AC-2, 5-C. S04-0H-HC03-C03-COrH20 system to high ionic strengths at
25°C. Geochirn. Cosrnochirn. Acta 48,723-751.
Editorial handling: G. R. Holdren, Jr. HARVIE C. E., GREENBERG J. P. and WEARE J. H. (1987) A chemical
Prediction of mineral solubilities 837

equilibrium algorithm for highly non-ideal multiphase systems: PITZER K. S. and MAYORGA G. (1973)Thermodynamics of electro-
Free energyminimization. Geochim Cosmochim. Acta 51, 1045- Iytes. II: Activity and osmotic coefficients for strong electrolytes
1057. with one or both ions univalent. J. Phys. Chem. 77, 2300-2308.
HERRMANN A. G., KNAKE D., SCHNEIDER J. and PETERS H. (1973) PITZER K. S. and MAYORGA G. (1974)Thermodynamics of electro-
Geochemistry of modern seawater and brinesfrom salt pans: Main Iytes. III: Activityand osmotic coefficients for 2-2 electrolytes. J.
components and bromine distribution. Contrib Mtneral. Petrol. Soln. Chem. 3, 539-546.
40, 1-24. PITZER K. S., PEIPER J. C. and BUSEY R. H. (1984)Thermodynamics
HILL A. E. and WILLS J. H. (1938)Ternary systemsXXIV. Calcium properties of aqueous sodium chloride solutions. J. Phys. Chem.
sulfate,sodium sulfate and water. J. Amer. Chem. Soc. 60, 1647- Ref Data 13, 1-102.
1655. POSNJAK E. (1938) The system CaS04-H20. Amer. J. Sci. 235A,
HOLMES H. F. and MESMER R. E. (1986)Isopiestic studiesof aqueous 247-272.
solutionsat elevatedtemperatures. VIII. The alkali-metalsulfates. ROGERS P. S. Z. (1981) Thermodynamics of geothermal fluids. Ph.D.
J. Chem. Thermodynamics 18,263-275. dissertation, University of California, Berkeley.
HOLMES H. E, BAES e. F. JR. and MESMER R. E. (1978) Isopiestic ROGERS P. S. Z. and PITZER K. S. (1981) High temperature ther-
studiesof aqueoussolutionsat elevated temperatures. I. KCI, CaC!2 modynamic propertiesof sodium sulfatesolutions.J. Phys. Chem.
and MgCh. J. Chem. Thermodynamics 10, 983-996. 85, 2886-2895.
HOLMES H. E, BAES e. F. JR. and MESMER R. E. (1981) Isopiestic SCHROEDER W. c, GABRIEL A. and PARTRIDGE E. P. (1935) Sol-
studies of aqueous solutions at elevated temperatures. III. {(I ubilityequilibria of sodiumsulfate at temperaturesof 150to 350°C.
- y)NaCI + yCaCh}. J. Chem. Thermodynamics 13, 101-113. I. Effect of sodium hydroxide and sodium chloride. J. Amer. Chem.
INNORTA G., RABBI E. and TOMADIN L. (1980) The gypsum-an- Soc. 57, 1539-1546.
hydrite equilibrium by solubility measurements. Geochim. Cos- SPENCER R. J., EUGSTER H. P. and JONES B. F. (1985)Geochemistry
mochim. Acta 44, 1931-1936. of Great Salt Lake,Utah. II:Pleistocene-Holocene evolution. Geo-
LANGELIER W. F., CALDWELL D. H., LAWRENCE W. B. and chim. Cosmochim. Acta 49, 739-747.
SPAULDING e. H. (1950) Scale control in sea water distillation STEPHEN H. and STEPHEN T. (1963) Solubilities of Inorganic and
equipment. Indus. Eng. Chem. 42, 126-130. Organic Compounds. The MacMillan Co., New York, N.V.
LINKE W. F. (1965) Solubilities of Inorganic and Metal Organic STRAUB E G. (1932) Solubilityof calcium sulfate and calcium car-
Compounds, 4th. edn. vols. I and 2. Amer. Chem. Soc. bonate at temperatures between 182° and 316°. Ind. Eng. Chem.
MADGIN W. M. and SWALES D. A. (1956)Solubilities in the system 24,914-917. See also BUKSHTEIN et al. (1953), p. 284.
CaS04-NaCI-H20 at 25°C and 35°e. J. Appl. Chem. 6, 482-487. TEMPLETON C. C. and ROGERS J. C. (1967)Solubilityof anhydrite
MARSHALL W. L. and SLUSHER R. (1966) Thermodynamics of cal- in several aqueous salt solutions between 250°C and 350°C. J.
cium sulfate dihydrate in aqueous sodium chloride solutions, 0- Chem. Eng. Data 12, 536-547.
no-c. J. Phys Chem. 70,4015-4027. VAN'T HOFF J. H., ARMSTRONG E. F., HINRICHSEN W., WEIGERT
MARSHALL W. L., SLUSHER R. and JONES E. V. (1964) Solubility F. and JuSTG. (1903)Gips and anhydrite. Z. Physik. Chem. 45,
and thermodynamic relationships forCaS04in NaC!-H20 solutions 257-306.
from 40°C to 200°C, 0 to 4 molal NaC!. J Chem. Eng. Data 9, WEARE J. H. (1987) Models of mineral solubility in concentrated
187-191. brines with application to fieldobservations. Amer, Mineral. Soc.
OSTROFF A.G. and METLER A. V. (1966) Solubility of calciumsulfate Short Course on Thermodynamic Modelingof GeologicSystems,
dihydrate in the system NaCI-MgCh-H 20 from 28°C to 70°e. J Oct. 26-29, 1987, Phoenix, Arizona.
Chem. Eng. Data 11, 346-350. WOLERY T. J. (1983)EQ3NR A computer program for geochemical
PABALAN R. T. and PITZER K. S. (1987) Thermodynamics of con- aqueous speciation-solubility calculations: user's guide and doc-
centrated electrolyte mixtures and the prediction of mineral sol- umentation. Lawrence Livermore Lab. Report UCRL-53414.
ubilities to high temperatures for mixtures in the system Na-K- WOOD S. A., CRERAR D. A., BRANTLEY S. L. and BoRCSIK M. (1984)
Mg-CI-S04-OH-H20. Geochim. Cosmochim. Acta 51, 2429-2443. Mean molal stoichiometric activity coefficients of alkali halides
PARTRIDGE E. P. and WHITE A. H. (1929)The solubilityof calcium and related electrolytes in hydrothermal solutions. Amer. J. Sci.
sulfate from 0 to 200°. J. Amer. Chem Soc. 51, 360-370. 284,668-705.
PHUTELA R. e. and PITZER K. S.(1983)Thermodynamics ofaqueous YEATTS L. B. and MARSHALL W. L. (1969) Apparent invariance of
calcium chloride. J. Soln. Chem. 12,201-207. activity coefficients of calcium sulfate at constant ionic strength
PITZER K. S. (1973) Thermodynamics of electrolytes. I: Theoretical and temperature in the systemCaS04-Na2S04-NaN03-H20 to the
basis and general equations. J Phys Chem. 77, 268-277. critical temperatureof water.Association equilibria. J. Phys Chem.
PITZER K. S. (1975) Thermodynamics of electrolytes. V: Effects of 73,81-90.
higherorder electrostaticterms. J. Soln. Chem. 4, 249-265. ZAREMBO V. I., Lvov S. N. and MATUZENKO M. Yu. (1980) Sat-
PITZER K. S. and KIM J. (1974) Thermodynamics of electrolytes. urated vapor pressureof water and activitycoefficients of calcium
IV:Activity and osmoticcoefficients formixedelectrolytes. J. Amer. chloride in the CaCh-H20 system at 423-623K. Geokhimiya 4,
Chem. Soc. 96, 5701-5707. 610-614.

Anda mungkin juga menyukai