Você está na página 1de 12

Geochimica et Cosmochimica Acta, Vol. 69, No. 13, pp.

3309 3320, 2005


Copyright 2005 Elsevier Ltd
Printed in the USA. All rights reserved
0016-7037/05 $30.00 .00
doi:10.1016/j.gca.2005.01.015

CO2-H2O mixtures in the geological sequestration of CO2. II. Partitioning in chloride


brines at 12100C and up to 600 bar
NICOLAS SPYCHER* and KARSTEN PRUESS
Earth Sciences Division, Lawrence Berkeley National Laboratory, MS 90 1116, 1 Cyclotron Road, Berkeley, CA, USA

(Received September 14, 2004; accepted in revised form January 19, 2005)

AbstractCorrelations presented by Spycher et al. (2003) to compute the mutual solubilities of CO2 and H2O
are extended to include the effect of chloride salts in the aqueous phase. This is accomplished by including,
in the original formulation, activity coefficients for aqueous CO2 derived from several literature sources,
primarily for NaCl solutions. Best results are obtained when combining the solubility correlations of Spycher
et al. (2003) with the activity coefficient formulation of Rumpf et al. (1994) and Duan and Sun (2003), which
can be extended to chloride solutions other than NaCl. This approach allows computing mutual solubilities in
a noniterative manner with an accuracy typically within experimental uncertainty for solutions up to 6 molal
NaCl and 4 molal CaCl2. Copyright 2005 Elsevier Ltd

1. INTRODUCTION the compressed CO2 gas phase, which is one of our objectives.
Nevertheless, in their study, Duan and Sun (2003) present
In Part I of this study (Spycher et al., 2003), a noniterative
useful Pitzer expressions (Pitzer, 1973) and parameters to com-
approach was presented to compute the mutual solubilities of
pute activity coefficients for aqueous CO2, which are needed to
pure H2O and CO2 in a temperature and pressure range most
account for salting-out effects. When taking into account as-
relevant to the geologic sequestration of CO2. The method was
sumptions made by these authors, their activity coefficient
intended primarily for efficient numerical simulations of CO2
formulation can be used with our solubility correlations (Spy-
flows. Most practical modeling applications related to CO2
cher et al. 2003) to yield, in a noniterative manner, CO2
sequestration (e.g., Garcia, 2003; Pruess et al., 2004; Xu et al.,
solubilities with an accuracy comparable to their results, as
2004) deal with subsurface waters containing dissolved salts.
discussed later.
For this reason, in Part II of this study, we extend the solubility
Models presented by Li and Nghiem (1986) and Enick and
model to include moderately saline solutions up to 6 m NaCl
Klara (1990) make use of simpler and more efficient cubic
and 4 m CaCl2. As before, the objective is to compute in the
EOS. Duan et al. (2003) show that the model of Li and Nghiem
most efficient manner the solubility of CO2 in the aqueous
phase, as well as the H2O solubility in the compressed-gas (1986), based on a Peng-Robinson EOS, is not as accurate as
phase at equilibrium with the aqueous phase. theirs and, therefore, this model is not further evaluated here.
Several theoretical studies of the CO2-H2O-NaCl system Enick and Klara (1990) fitted a large number of solubility
have been published to date, following various approaches and measurements to a Henrys-law/Peng-Robinson-EOS model,
equations of state (EOS). Two of the more comprehensive using an approach similar to that followed in Spycher et al.
solubility models developed for this system (Bowers and (2003). They then extended their solubility model to include
Helgeson, 1983; Duan et al., 1995) lie outside the pressure- electrolyte solutions by correlating solubilities with total dis-
temperature-composition (P-T-X) range of interest here. These solved solids (on a weight percent basis), using a large set of
studies covered elevated pressures and temperatures mostly experimental and field data. In their correlation, however, these
relevant to the study of hydrothermal systems and fluid inclu- authors did not distinguish between the various types of dis-
sions. More recently, Duan and Sun (2003) presented a model solved salts in the solutions for which CO2 solubilities were
for CO2 solubility in NaCl and other electrolyte solutions regressed (NaCl, CaCl2, and unspecified brines). Because their
applicable to a wide P-T-X range (0 2000 bar, 0 to 260C, correlation for saline solutions inevitably shows significant
0 4.3 m NaCl), overlapping low temperatures and moderate scatter and is approximate at best, it was not considered further.
pressures applicable to the investigation of geologic CO2 se- Rumpf and Maurer (1993) presented a comprehensive solu-
questration. Their model relies on an exhaustive set of exper- bility model for CO2 in electrolyte solutions. Their model relied
imental data from the literature and reproduces published sol- on solution-chemistry correlations similar to those adopted in
ubilities of CO2 in NaCl and CaCl2 solutions, as well as in the present study, but used a standard virial EOS truncated after
seawater, with accuracies close to experimental uncertainty. the second term to compute fugacity coefficients. Rumpf et al.
One drawback of Duan and Suns model, however, is that it (1994) fitted this model to their own measurements of CO2
relies on a fifth-order virial EOS (Duan et al., 1992) that cannot solubility from 40 to 160C and up to 100 bar, in 4 and 6 m
be efficiently implemented in numerical flow simulations. Also, NaCl solutions. These authors used Pitzer expressions similar
this model was not intended to compute the H2O solubility in to those adopted by Duan and Sun (2003) to compute activity
coefficients. As shown later in this study, their activity coeffi-
cient formulation can be extended and combined with our
* Author to whom correspondence should be addressed (nspycher@ solubility model (Spycher et al., 2003) to cover our entire P-T
lbl.gov). range of interest (12 to 100C and up to 600 bar) with an
3309
3310 N. Spycher and K. Pruess

accuracy as good as, if not better than, that of Duan and Suns ventions and standard states. In the P-T range of interest here,
model. dissolved salts are essentially nonvolatile and, therefore, the
Because reasonable aqueous CO2 solubilities can be deter- extended formulation hardly changes. The water mole fraction
mined without considering non-ideal mixing effects in the gas in the CO2-rich phase (yH2O) and the CO2 mole fraction in the
phase, other studies have focused primarily on the aqueous aqueous phase (xCO2) are respectively expressed as:


phase without integrating an accurate solubility model for the
compressed gas phase. Such studies include those of Ellis and KH0 2O aH2O
(P P0)V H2O
Golding (1963), Malinin and Savelyeva (1972), Malinin and y H2O exp (1)
H2OPtot RT
Kurovskaya (1975), Drummond (1981), and Cramer (1982), all
of whom conducted their own CO2 solubility experiments. and


These authors determined Henrys constants for pure H2O and
saline solutions, from which salting-out effects could be quan- CO21 y H2OPtot (P P0)V CO2
tified, as further discussed later and in Appendix A-1. Nesbitt xCO2 
exp (2)
55.508 K
x
0
CO2 g RT
(1984) and Barta and Bradley (1985) developed similar models
using experimental data from others, but further expanded the In these equations, K0 is the thermodynamic equilibrium con-
formulation of activity coefficients for aqueous CO2 as speci- stant for each component at temperature T and reference pres-
fied later. sure P0 1 bar, for respective reactions H2O(l) ` H2O(g) and
In all studies discussed above, the solubility of CO2 in water CO2(aq) ` CO2(gl). P is total pressure, V the average partial
was expressed through some form of gas-liquid partitioning molar volume of each pure condensed phase over the pressure
function relating either directly or indirectly to a thermody- range P0-P, is the fugacity coefficient of each component in
namic equilibrium constant. If an EOS was used, this was done the CO2-rich (compressed gas) phase, and R is the gas constant.
merely to account for gas phase non-ideality. With The effect of dissolved salts is expressed through aH2O, the
this Henrys law or solution chemistry approach, the effect activity of liquid water, and x, an activity coefficient for
of dissolved salts on CO2 solubility is treated by using an aqueous CO2. For consistency with the formulation and values
activity (or salting-out) coefficient expressing the departure of K0 developed in our initial model, x is on a mole fraction
from solubility in pure water. This approach is followed here as scale and is unity when no salts are present (i.e., x 1 as
well. In the last decade or so, however, other solubility models xsalt0). Note that by equating the water activity to water mole
have been developed that make use of one single EOS (e.g., fraction and setting x 1, Equations (1) and (2) revert to the
Peng-Robinson) to compute the properties of both the aqueous formulation for pure water presented in Spycher et al. (2003).
and compressed-gas phases at equilibrium. These models, re- Parameters and equations for computing K0 and , as well as
ported primarily in the chemical engineering literature, have values of V , were reported in Spycher et al. (2003) and are
been applied to pure CO2-H2O systems (e.g., Shyu et al., 1997) unchanged here. Values of aH2O are approximated as discussed
and to systems including electrolyte solutions (e.g., Harvey and below, and various existing models for computing x are re-
Prausnitz, 1989; Zuo and Guo, 1991; Sreide and Whitson, viewed later in this paper.
1992; Srensen et al., 2002; Masoudi et al., 2004). The main Accurate values of water activity for solutions of various
drawback of these models is that an iterative procedure must be electrolytes can be obtained using the Pitzer ion interaction
used to solve the EOS twice, once for the gas-phase composi- model (Pitzer 1973) (e.g., Rumpf and Maurer, 1993). However,
tion and a second time for the liquid-phase composition. Fur- in a salinity range up to ionic strength around 6 molal (below
thermore, while conceptually elegant, these models have not halite saturation), the model can be simplified by assuming that
been shown to be superior to traditional Henrys law models the water activity equals its mole fraction on the basis of a fully
and produce large errors when their parameters are extrapolated
ionized salt. Deviations in computed yH2O values resulting from
beyond the range of measured solubilities, particularly in mul-
this assumption are evaluated later ( 8% at ionic strength
ticomponent systems (e.g., Patel et al., 2001). For these rea-
6 m). Note that the effect of these deviations on computed
sons, and because of their incompatibility with the more tradi-
CO2 solubilities (through Eqn. 2) is one to two orders of
tional correlations adopted in Spycher et al. (2003), these
magnitude smaller, because yH2O in Eqn (2) typically remains
models were not considered further.
quite small (a few percent or less at temperatures below 100C
and pressures above 25 bars or so).
2. SOLUBILITY MODEL Equations (1) and (2) are solved as previously by setting


The formulation of the basic model (Spycher et al., 2003) is
extended with an activity coefficient for aqueous CO2 and a KH0 2O (P P0)V H2O
A exp (3)
correction to the activity of water to account for the effects of H2O Ptot RT


dissolved salts. The extended formulation is presented below,
followed by a review of available models and/or data from CO2 Ptot
(P P0)V CO2
which activity coefficients for aqueous CO2 can be derived. B exp (4)
55.508xKCO
0
2 g
RT

2.1. Extended Solubility Correlations Taking the water mole fraction as a reasonable approximation
of water activity, we rewrite Equation (1) as
The reader is referred to Spycher et al. (2003) for the
derivation of the basic correlations and a description of con- y H2O A1 xCO2 xsalt (5)
CO2 solubility in chloride brines 3311

and the mutual solubilities are then computed as

1 B xsalt
y H2O (6)
(1 A B)

xCO2 B1 y H2O (7)


In Equations (5) and (6), xsalt is the mole fraction of the
dissolved salt on a fully ionized basis and including dissolved
CO2. It is defined as:

msalt
xsalt (8)
55.508 msalt mCO2(aq)
where m stands for molality and is the stoichiometric number
of ions contained in the dissolved salt (i.e., 2 for NaCl, 3 for
CaCl2, etc.). Accordingly, the CO2 molality is expressed from
the mole fraction as

xCO2(msalt 55.508)
mCO2 (9)
(1 xCO2)
or also as

xCO255.508
mCO2 (10)
x H2O
It is more practical to use the salt molality instead of mole
fraction as an input parameter because it is independent from
Fig. 1. Activity coefficients for aqueous CO2 from various sources
the CO2 solubility. This is done (see Appendix B) by substi- (see text and Appendix A.2). Data from Cramer (1982) (and regressed
tuting Eqn (7) into Eqn (9), then substituting the resulting by Battistelli, 1997) and Nesbitt (1984) are on a mole fraction scale;
expression for CO2 molality into Eqn (8), which then yields other data are on a molality scale. As discussed in the text, data from
xsalt in terms of msalt, B, and yH2O. The resulting expression for these sources are not directly comparable, yielding significant scatter.
xsalt is then substituted into Eqn (6) which, after some rear-
rangement, yields: puted activity coefficients on a molal scale from measurements
1 B55.508 of Henrys law ratios (Equation A-1). They used Henrys
y H2O (11) constants (kH) for saline solutions determined from their own
(1 A B) ( msalt 55.508) msaltB experimental data, but a Henrys constant at infinite dilution
By computing fugacity coefficients as presented in Spycher et (kH0 ) determined by others (Plummer and Busenberg, 1982).
al. (2003), Eqns (7) and (11) can be solved without iteration to Their activity coefficients deviate significantly from other val-
provide the mutual solubilities of CO2 and H2O-salt solutions. ues shown in Figure 1, possibly because of inconsistencies
between their measurements and adopted kH0 value. These au-
2.2. Activity Coefficients for Aqueous CO2 thors fitted their activity coefficients to a Pitzer formulation
with temperature-dependent ion-interaction parameters. Their
Studies by Duan and Sun (2003), Rumpf et al. (1994), He regression with temperature appears to have been conducted on
and Morse (1993), Barta and Bradley (1985), Nesbitt (1984), the same number of data points as fit parameters, which could
Cramer (1982), and Drummond (1981) present data and/or explain the somewhat wavy behavior displayed in Figure 1. For
equations from which activity coefficients can be derived for these reasons, their activity coefficient formulation was not
aqueous CO2 in NaCl and other electrolyte solutions. Activity considered further.
coefficient values from these studies, however, are not all Barta and Bradley (1985) regressed the experimental data of
directly comparable and yield a significant scatter (Fig. 1), for Ellis and Golding (1963) and Drummond (1981) for solutions
various reasons discussed below and in Appendix A.1. Note up to 6.5 m NaCl, covering a P-T range around 20 to 400C and
that all these studies show a temperature dependence of salting- 1 400 bar, using a Pitzer ion-interaction model. These authors
out effects, but only that of Duan and Sun (2003) also include fitted these data to expressions yielding Henrys constants, as a
a pressure dependence to reasonably reproduce experimental function of temperature. Activity coefficients on a molality
solubilities. scale (m ), with the convention that m 1 as msalt0, were
He and Morse (1993) determined activity coefficients of derived from their Henrys law formulation (using Equation
aqueous CO2 at 1.013 bar (1 atm) from 0 to 90C for solutions A-1). The values obtained in this way are within the range of
of various compositions (0.13 m HCl, 0.5 6 m NaCl, 0.15 m other data (Fig. 1) but display a convex instead of concave
KCl, 0.15 m CaCl2, 0.15 m MgCl2, 0.013 m Na2SO4, trend with temperature. These authors were not able to regress
0.01 0.9 m K2SO4 and 0.1 4 m MgSO4). These authors com- simultaneously data for pure water and for NaCl solutions, and
3312 N. Spycher and K. Pruess

possibly overfitted their results. Therefore, their model was not compute activity coefficients through Eqn A-2. Cramers salt-
considered either. ing-out coefficients were derived on the basis of Henrys con-
In the remaining cited studies, CO2 solubilities in pure water stants (Eqn A-1) expressed in terms of mole fraction (and later
and in saline solutions were simultaneously considered, pre- correlated to a Setchenow equation expressed in terms of
senting an advantage by reducing the risk of inconsistencies molality). Therefore, the Cramer/Battistelli model provides ac-
when evaluating salting-out effects. Duan and Sun (2003) si- tivity coefficients on a mole fraction scale (x), with the con-
multaneously fitted a Pitzer ion-interaction model and a ther- vention that x1 as xsalt0, which can be entered directly
modynamic equilibrium formulation to a large number of sol- into Eqns (2) and (4) without further correction. Note that
ubility data for pure water and electrolyte solutions, as a Cramer (1982) did not see a need to take pressure into account
function of both pressure and temperature. The experimental in his fit of salting-out coefficients.
data regressed by Duan and Sun (2003) cover a wide P-T-X Using primarily his own experimental data, Drummond
range spanning 0 to 260C, 0 2000 bar, 0 6.5 m NaCl and (1981) provided a useful regression of Henrys constants as a
0 3.9 m CaCl2. Their activity coefficient formulation is given function of both temperature and salt molality for solutions up
in Appendix A.2. This model, however, does not yield x to around 6.5 m NaCl and a P-T range covering 20 to 400C
values that can be used directly in Eqn (2). This is because and 1 400 bar. He expressed Henrys law in terms of molality,
these authors assumed ideal mixing and did not take into and obtained a reasonable fit of Henrys constants without
account salt effects when expressing the partial pressure of introducing a pressure dependency. Activity coefficients on a
H2O in the gas phase (however, these effects were indirectly molal scale (m ) can be derived from his data, with the con-
absorbed in their model by separate fit parameters expressing vention that m 1 as msalt0 (Appendix A.2). Conversion to
reference chemical potential). Therefore, in their regression, the a mole fraction scale with the convention that x1 as xsalt0
CO2 partial pressure remains the same over pure and saline (Appendix A.2) yields activity coefficients within 5% of the
water, such that the pressure terms in Henrys ratios in Equa- values given by the Cramer/Battistelli regression.
tion A-1 cancel out. As such, their activity coefficient formu- Nesbitt (1984) derived a simple relation to compute activity
lation yields the quantity * mo/m, where mo is the CO2 coefficients on a mole fraction scale (x) with the convention
molality in pure water and m the molality in saline water (at the that x1 as xH2O1 (Appendix A.2). This author fitted his
same given P and T). This relationship was verified from formulation to existing experimental data covering a P-T-X
computed values of mo and m tabulated by Duan and Sun range 0 to 500C, 11500 bar, and 0 6 m NaCl. For use with
(2003) and values of * calculated with their Pitzer expression Eqn (4), his model must be corrected for the convention that
(Equation A-4). Their activity coefficient formulation can thus x1 as xsalt0 (Appendix A.2).
be implemented with our solubility model by first applying our Note that the activity coefficient formulation of Nesbitt
correlations for pure water (setting xsalt 0 and x 1) to (1984), as well as the molality- to mole fraction-scale conver-
determine mo, then computing the CO2 solubility in saline sion required with activity coefficients derived from Drum-
solutions as m mo/*. The composition of the gas phase is mond (1981) and Rumpf et al. (1994), introduce a dependency
then determined directly from Eqn (5) after converting the salt of the activity coefficient on CO2 solubility. This requires
and aqueous CO2 molalities to mole fractions (Eqn. 8 and 10). further manipulation of Eqns (7) and (11) to solve these equa-
tions directly, because B in Eqn. 11 is no longer independent of
Rumpf et al. (1994) fitted the correlations of Rumpf and
Maurer (1993) to their CO2 solubility measurements in the xCO2. A more practical alternative, adopted here, is to simply
iterate between Eqns (7) and (11). Because the coupling be-
range 40 to 160C and 596 bar, at 4 and 6 m NaCl. In doing
tween B and xCO2 is weak, values of x always converge within
so, these authors determined Pitzer ion-interaction parameters
3 4 iterations without any provisions to speed convergence.
for activity coefficient expressions similar to those imple-
mented by Duan and Sun (2003) (Appendix A.2). In their
3. RESULTS
limited P-T range, Rumpf et al. (1994) neglected the effect of
pressure on these parameters and included a temperature de- CO2 solubilities calculated1 using Eqns (3) to (11), in com-
pendence only on the binary interaction parameter. In their bination with the various activity coefficient formulations dis-
analysis, Rumpf et al. (1994) used a Henrys constant for pure cussed earlier and in Appendix A.2, were compared to both
solutions previously fitted to published data using the same experimental and computed data reported in the literature (Ta-
model (Rumpf and Maurer, 1993), thus providing consistency. ble 1). Results of these comparisons are discussed below.
Their expression yields activity coefficients on a molality scale Most experimental data on CO2 solubility in aqueous elec-
(m ) with the convention that m 1 as msalt0 (Appendix trolyte solutions have been summarized in Scharlin (1996).
A.2). For use with Equations (2) and (4), these coefficients Very few experimental data have been published on CO2
were converted to a mole fraction scale with the convention solubility in saline solutions at low temperatures and high
that x1 as xsalt0, as shown in Appendix A.2. pressures. Prutton and Savage (1945) reported solubility for
Battistelli et al. (1997) regressed, as a function of tempera- CaCl2 solutions near 76 and 101C up to pressures of 650
ture, salting-out coefficients determined by Cramer (1982) for bar. For NaCl solutions, usable data at temperatures below
NaCl solutions from 0 to 300C. Cramer (1982) used his own 100C are limited to pressures up to 200 bar. These data include
experimental data up to 240C, 62 bar, and 1.95 m NaCl, as CO2 solubilities measured by Bando et al. (2003), Kiepe et al.
well as data from others up to 330C and 200 bar at similar
NaCl concentrations. The regression equation from Battistelli 1
A Fortran computer program to carry out these calculations is avail-
et al. (1997) is given in Appendix A.2 and can be used to able from the authors upon request.
CO2 solubility in chloride brines 3313

Table 1. Deviations, as root mean square error (RMSE), between predicted CO2 solubilities (Equations 311) and available experimental and
computed data below 100C for NaCl and CaCl2 solutions, using various activity coefficient formulations from the literature, as discussed in the
text and in Appendix A.2.

Model deviations for various activity coefficient


formulations

P-T-X range covered below 100C Rumpf Cramer/


Duan and et al. Battistelli Nesbitt
Temperature Pressure Nb. Sun RMSE RMSE RMSE Drummond RMSE
Solubility Data Source (C) (bar) Aqueous Solution Points (%) (%) (%) RMSE (%) (%)

Computed Data
Duan and Sun (2003) 3090 1600 pure water 30 1.6 1.6 1.6 1.6 1.6
3090 1600 1 m NaCl 30 1.6 2.9 3.0 3.2 3.7
3090 1600 2 m NaCl 30 1.6 4.2 4.1 4.6 6.0
3090 1600 4 m NaCl 30 1.7 6.7 8.2 9.4 10.0
Experimental Data
Drummond (1981) 31100 3958 pure water 11 7.8 7.8 7.8 7.8 7.8
27100 4265 1 m NaCl 14 7.3 7.8 8.8 6.2 8.8
28100 4268 2 m NaCl 15 3.6 2.2 2.2 4.3 2.3
22100 4067 3 m NaCl 17 3.1 3.2 2.3 2.4 6.1
20100 3559 4 m NaCl 17 7.9 4.1 10 11 2.1
24100 3554 6 m NaCl 16 5.6 9.2 13 11 15

Malinin and Savelyeva 25100 48 pure water 5 4.3 4.3 4.3 4.3 4.3
(1972) and
Malinin and Kurovskaya 25100 48 0.41.1 m NaCl 6 4.9 4.7 4.5 5.8 4.7
(1975) 25100 48 1.82.2 m NaCl 5 4.5 4.2 4.5 4.7 5.4
25100 48 2.63.2 m NaCl 5 4.7 4.0 4.8 6.1 5.7
25100 48 45.9 m NaCl 7 5.5 4.8 11 9.3 11

Rumpf et al. (1994) 50 1158 pure water 7 0.96 0.96 0.96 0.96 0.96
4080 596 4 m NaCl 22 4.1 2.2 5.3 6.3 5.5
4080 692 6 m NaCl 16 6.2 2.1 18 16 6.0

Nighswander et al. 8081 23102 pure water 8 3.7 3.7 3.7 3.7 3.7
(1989) 8081 4099 0.18 m NaCl 8 3.4 3.4 3.5 3.3 3.5

Bando et al. (2003) 3060 100200 pure water 12 2.8 2.8 2.8 2.8 2.8
3060 100200 0.170.53 m NaCl 36 2.0 2.0 2.4 1.7 2.1

Kiepe et al. (2002) 40100 0.193 pure water 36 9.8 9.8 9.8 9.8 9.8
4080 0.1101 0.52 m NaCl 24 8.6 8.3 8.0 9.1 8.0
4080 0.1101 2.5 m NaCl 23 6.8 5.4 5.5 6.8 5.7
4080 0.1100 44.3 m NaCl 23 10 9.2 5.9 6.2 13

Prutton and Savage 101 58623 pure water 14 5.0 5.0 5.0 5.0 5.0
(1945) 76101 16628 1 m CaCl2 28 6.0 5.7 20 17 16
76101 22657 2.3 m CaCl2 25 5.5 4.3 40 26 39
76101 15638 4 m CaCl2 20 6.8 8.7 61 41 77

Malinin and Savelyeva 25100 48 pure water 5 4.3 4.3 4.3 4.3 4.3
(1972) and
Malinin and Kurovskaya 20100 48 0.21.3 m CaCl2 10 4.3 3.7 19 16 13
(1975) 20100 48 1.92.4 m CaCl2 5 3.6 3.3 40 32 31
20100 48 33.4 m CaCl2 2 7.1 9.5 58 48 51
20100 48 3.96.0 mCaCl2 3 6.5 6.1 72 49 95

(2002), Rumpf et al. (1994), Nighswander et al. (1989), Drum- model to all the experimental data mentioned above, except for
mond (1981), Malinin and Kurovskaya (1975), and Malinin those of Bando et al. (2003) and Kiepe et al. (2002), and to data
and Savelyeva (1972) (Table 1). from other sources covering a much wider P-T range than
Because of the lack of experimental solubility data at low considered here. Their model reproduces experimental data
temperatures and high pressures, we first compare our results to close to or within experimental uncertainty (estimated by Duan
the computed CO2 solubilities tabulated by Duan and Sun and Sun at around 7%).
(2003) at 30, 60, and 90C and 1 600 bar for solutions up to Combining the activity coefficient formulation of Duan and
4 m NaCl (Fig. 2). Duan and Sun (2003) fitted their solubility Sun (2003) (Appendix A.2) and the solubility correlations
3314 N. Spycher and K. Pruess

Fig. 2. Predicted mutual solubilities of CO2 and H2O (Eqn. 311) using the activity coefficient formulations of Duan and
Sun (2003) (solid lines) and Rumpf et al. (1984) (dashed lines) (see text and Appendix A.2). CO2 solubilities computed by
Duan and Sun (2003) using their full solubility model are also shown for comparison (symbols). Note that computed
solubilities of H2O in CO2 at 30C (below critical temperature) display a sharp discontinuity as gaseous CO2 changes to
liquid CO2 with increased pressure (see Spycher et al., 2003).

presented earlier, the CO2 solubilities reported by Duan and 0.4, 5, 15, and 30% for solutions 1, 2, 3, and 4 m CaCl2 (ionic
Sun (2003) (i.e., computed using their full model) are repro- strength 3, 6, 9, and 12), respectively. For these ranges in
duced with a root-mean-square error (RMSE) around 2% at all composition, differences between activity and mole fraction
salt molalities (Table 1 and Fig. 2). This deviation expresses due solely to temperature and pressure remain below 1% for
differences between the pure-water solubilities computed by NaCl solutions and below 10% for CaCl2 solutions (the
each model. The other activity coefficient formulations (Ap- temperature effect being largely dominant). These approximate
pendix A.2) produce somewhat larger deviations but still rea- deviations in yH2O do not include the average fit error around
sonably good results, with an RMSE increasing from 3% at 5% (Spycher et al., 2003) determined from experimental data
1 m NaCl up to 10% at 4 m NaCl (Table 1). Note that Duan for the pure CO2-H2O system.
and Sun (2003) limit their model application to 4 m NaCl, Results of our solubility model combined with the various
although these authors seem to have fitted solubilities at con- activity coefficient formulations (Appendix A.2) were also
centrations up to 6 m NaCl. compared directly to experimental solubilities (Table 1 and
No experimental or computed data were available for com- Figs. 35). For pure water, most of the solubility data listed
paring water concentrations in the compressed gas phase (yH2O) in Table 1 are reproduced with RMSE values less than 5%,
for saline CO2-H2O solutions (Fig. 2). As discussed earlier, the even though these data were not included in the original
accuracy of yH2O directly relates (through Eqn. 1) to the accu- regression (Spycher et al., 2003). The experimental CO2
racy of the liquid water activity, which is approximated in our solubilities of Drummond (1981) for pure water are repro-
model by the liquid water mole fraction (on the basis of a fully duced with RMSE values 8% (Fig. 3). Solubilities reported
ionized salt). The deviation between water mole fraction and by Kiepe et al. (2002) for pure water deviate somewhat from
activity was evaluated using osmotic coefficient values re- the rest of the data and are reproduced with RMSE values
ported by Pitzer et al. (1984) for sodium chloride solutions and around 10%.
by Ananthaswamy and Atkinson (1985), Phutela and Pitzer For NaCl solutions, the various activity coefficient formula-
(1983), and Holmes et al. (1994) for calcium chloride solutions. tions reproduce experimental data with RMSE values generally
In our PT range of interest, assuming that aH2O xH2O leads to less than 10%. The activity coefficient formulations from Duan
computed yH2O values deviating positively from true values and Sun (2003) and Rumpf et al. (1994) provide the best results
by 1, 3, 5, and 8% for solutions 3, 4, 5, and 6 m NaCl, and (Table 1). The latter reproduces best the solubilities measured
CO2 solubility in chloride brines 3315

Fig. 3. Deviations between model results (Eqn. 311) and experimental CO2 solubilities in NaCl solutions (Drummond
1981), using various activity coefficient formulations (symbols) as discussed in the text and Appendix A.2. The P-T range
of the experimental data are 20 to 100C and 35 68 bar (see Table 1). Note the different scale of the vertical axis at 6 m
NaCl.

by Rumpf et al. (1994) at 6 m NaCl (RMSE 2%) (Fig. 4), showing similar RMSE values within a 3%10% range for
which is to be expected, because their activity coefficient solutions up to 6 m CaCl2, compared to deviations up to 95%
formulation was fitted to these data. The more recent data of with the other formulations (Table 1 and Fig. 6). Note that even
Bando et al. (2003) at low salinity (0.5 m NaCl) up to 200 bar though the computed CO2 solubility is reasonably accurate up
are also reproduced with RMSE values 2% (Table 1 and Fig. to 6 m CaCl2, large errors in computed yH2O values (30%) are
5). Data from Kiepe et al. (2002) at higher salinity but lower expected to occur at concentrations 4 m CaCl2 because aH2O
pressures are reproduced with RMSE values mostly below 10% is approximated by xH2O in Eqn. 1.
(Table 1 and Fig. 4).
The various activity coefficient formulations were also tested 3. DISCUSSION AND CONCLUSIONS
against data for CaCl2 solutions. Duan and Sun (2003) ex-
tended their formulation to deal with salts other than NaCl by For NaCl solutions up to 2 molal, all activity coefficient
assuming that interaction parameters for ions with the same formulations provide reasonable CO2 aqueous solubilities
charge have roughly the same values. The same method was within 10% of the experimental and computed reference data
used, here, to extend the formulation of Rumpf et al. (1994) to shown in Table 1. At molalities up to 6 m NaCl and 4 m CaCl2,
CaCl2 solutions (Appendix A.2). The activity coefficient for- the activity coefficient expressions of Duan and Sun (2003) and
mulations from Drummond (1981) and Cramer (1982) were Rumpf et al. (1994) provide more accurate results, reflecting
extended to CaCl2 solutions by using ionic strength in place of the superiority of the Pitzer formulation over simpler models.
salt molality. The formulation of Nesbitt (1984) was extended Overall, the activity coefficient formulation of Rumpf et al.
using CaCl2 mole fractions (on a fully ionized basis) in place of (1994) yields slightly smaller RMSE values relative to exper-
NaCl mole fractions. As expected, the activity coefficient ex- imental data, most visibly (and expectedly) so when comparing
pression of Duan and Sun (2003) and Rumpf et al. (1994) are results against the solubility measurements made by these
clearly superior in reproducing solubilities in CaCl2 solutions, authors at 6 m NaCl (Fig. 4). However, in the absence of more
3316 N. Spycher and K. Pruess

Fig. 4. Comparison between model results (Eqn. 311) (lines) and experimental CO2 solubilities in NaCl solutions
(Rumpf et al., 1994, solid symbols and Kiepe et al., 2002, open symbols), using various activity coefficient formulations
as discussed in the text and Appendix A.2.

Fig. 5. Comparison between model results (Eqn. 311) (lines) and experimental CO2 solubilities in NaCl solutions
(Bando et al., 2003) (symbols), using various activity coefficient formulations as discussed in the text and Appendix A.2.
CO2 solubility in chloride brines 3317

Fig. 6. Comparison between model results (Eqn. 311) (lines) and experimental CO2 solubilities in CaCl2 solutions
(Prutton and Savage, 1945) (symbols), using various activity coefficient formulations as discussed in the text and Appendix
A.2.

experimental data at these salinities and at high pressures, it is tions (Appendix A.2) with our solubility correlations, mutual
difficult to assess which of these two formulations is more CO2-H2O solubilities in the range 12 to 100C, 1 600 bar, and
accurate. Note that the formulation of Rumpf et al. (1994) 0 6 m NaCl (or 4 m CaCl2) can be computed in a direct manner,
yields good results within a P-T range significantly larger than yielding CO2 solubilities with an accuracy (typically 7%) within
the range covered by these authors when regressing their data the spread of experimental data. We tend to favor the formulation
(pressures limited to 100 bar and no temperatures below of Duan and Sun (2003) because it was fitted over a wider P-T
40C). Also, their formulation (which is only a function of range than the formulation of Rumpf et al. (1994) and is somewhat
temperature) yields an accuracy similar to that of Duan and easier to implement. It must be recalled, however, that CO2
Suns formulation (which is pressure and temperature depen- aqueous solubilities computed at pressures above 100 bar and
dent), indicating that within the pressure range considered here temperatures below 100C may have a large uncertainty, be-
(up to 600 bars) a pressure correction to activity coefficients cause they rely on very few experimental data points. As more
is not necessary. experimental data become available in this P-T range, a refit of
Using the activity coefficient formulation of Duan and Sun ion-interaction parameters, as done previously by Duan and Sun
(2003) allows computing the mutual solubilities of CO2 and H2O (2003) and Rumpf et al. (1994), but using our solubility correla-
in a noniterative manner. As discussed earlier, the formulation of tions, may increase the model confidence and accuracy for appli-
Rumpf et al. (1994) is most easily implemented using a few cations to geologic CO2 sequestration.
iterations because of the (weak) dependence of activity coeffi-
cients on the CO2 solubility (introduced by the conversion from a
molality to a mole fraction scale). This is not seen as a disadvan- AcknowledgmentsWe thank two anonymous reviewers and the As-
tage, however, because the type of iterative procedure and equa- sociate Editor, D. Wesolowski, for their useful comments. We also
tions involved would not significantly decrease performance when thank John Apps and Carl Steefel for their internal review of this paper.
implemented into large numerical simulations. In any case, using This work was supported by the US Department of Energy through the
either approach, the EOS required to calculate mutual solubilities Office of Basic Energy Sciences under Contract No. DE-AC03-
76SF00098.
is solved noniteratively (Spycher et al., 2003).
By combining either of these two activity coefficient formula- Associate editor: D. J. Wesolowski
3318 N. Spycher and K. Pruess

REFERENCES trolyte solutions and its applications to gas hydrates. Fluid Phase
Equilib. 215, 163174.
Bando S., Takemura F., Nishio M., Hihara E., and Akai, M. (2003) Nesbitt H. W. (1984) Calculation of the solubility of CO2 in NaCl-rich
Solubility of CO2 in aqueous solutions of NaCl at 30 to 60C and hydrothermal solutions using regular solution equations. Chem.
10 to 20 MPa. J. Chem. Eng. Data 48, 576 579. Geol. 43, 319 330.
Barta L., and Bradley D. J. (1985) Extension of the specific interaction Nighswander J. A., Kalogerakis N., and Mehotra A. K. (1989) Solu-
model to include gas solubilities in high temperatures brines. bilities of carbon dioxide in water and 1 wt% NaCl solution at
Geochim. Cosmochim. Acta 49, 195203. pressures up to 10 MPa and temperatures from 80 to 200C. C.
Battistelli A., Calore C., and Pruess K., (1997) The simulator J. Chem. Eng. Data 34, 355360.
TOUGH2/EWASG for modeling geothermal reservoirs with brines Patel N. C., Abovsky V., and Watanasiri S. (2001) Calculation of
and non-condensible gas. Geotherm. 26, 437 464. vapor-liquid equilibria for a 10-component system: comparison of
Bowers T. S., and Helgeson H. C. (1983) Calculation of the thermo- EOS, EOS-GE and GE-Henrys law models. Fluid Phase Equilib.
dynamic and geochemical consequences of nonideal mixing in the 185, 297 405.
system H2O-CO2-NaCl on phase relations in geologic systems: Phutela R. C., and Pitzer K. S. (1983) Thermodynamics of aqueous
Equation of state for H2O-CO2-NaCl fluids at high pressures and calcium chloride. J. Solut. Chem. 12, 201207.
temperatures. Geochim. Cosmochim. Acta 47, 12471275. Pitzer K. S. (1973) Thermodynamics of electrolytes: I. Theoretical
Cramer S. D. (1982) The solubility of methane, carbon dioxide and basis and general equations J. Phys. Chem. 77, 268 277.
oxygen in brines from 0 to 300C. Report of Investigations 8706, Pitzer K. S., Peiper J. C., and Busey R.H. (1984) Thermodynamic
U.S. Department of the Interior, Bureau of Mines. properties of aqueous sodium chloride solutions. J. Phys. Chem.
Denbigh K. (1983) The Principles of Chem. Equilibrium, 4th. ed. Ref. Data 13, 1 64.
Plummer N. L., and Busengerg E. (1982) The solubility of calcite,
Cambridge Univ. Press, 494 p.
aragonite and vaterite in CO2-water solutions between 0 and 90C
Diamond L. W., and Akinfiev N. N. (2003) Solubility of CO2 in water
and an evaluation of the aqueous models for the system CO2-H2O-
from 1.5 to 100C and from 0.1 to 100 MPa: evaluation of
CaCO3. Geochim. Cosmochim. Acta 46, 10111040.
literature data and thermodynamic modeling. Fluid Phase Equilib.
Pruess K., Garcia J., Kovscek T., Oldenburg C., Rutvist J., Steefel C.
208, 265290.
and Xu T. (2004) Code intercomparison builds confidence in nu-
Drummond S. E. (1981) Boiling and mixing of hydrothermal fluids:
merical simulation models for Geologic disposal of CO2. Energy
chemical effects on mineral precipitation. Ph.D. thesis, Pennsylva-
29, (9 10), 14311444.
nia State University.
Prutton C. F., and Savage R. L. (1945) The solubility of carbon dioxide
Duan Z., and Sun R. (2003) An improved model calculating CO2 in calcium chloride-water solutions at 75, 100, 120C and high
solubility in pure water and aqueous NaCl solutions from 257 to pressures. J. Am. Chem.Soc. 67, 1550 1554.
533 K and from 0 to 2000 bar. Chem. Geol. 193, 257271. Rumpf B., Nicolaisen H., Ocal C., and Maurer G. (1994) Solubility of
Duan Z., Mller N., and Weare H. (1995) Equation of state for the carbon dioxide in aqueous solutions of sodium chloride: experi-
NaCl-H2O-CO2 system: Prediction of phase equilibria and volu- mental results and correlation. J. Sol. Chem. 23, 431 448.
metric properties. Geochim. Cosmochim. Acta 59, 2869 2882. Scharlin P. (1996) Carbon dioxide in water and aqueous electrolyte
Duan Z., Mller N., and Weare J. H. (1992) An equation of state for the solutions. Solubility Data Series. International Union of Pure and
CH4-CO2-H2O system: II. Mixtures from 50 to 1000C and 0 to Applied Chemistry, Oxford University Press, 383 p.
1000 bars Geochim. Cosmochim. Acta 56, 2619 2631. Shyu G.-S., Hanif N. S. M, Hall K. R., and Eubank P. T. (1997) Carbon
Ellis A. J., and Golding R. M. (1963)The solubility of carbon dioxide dioxide-water phase equilibria results from the Wong-Sandler com-
above 100C in water and in sodium chloride solutions. Am. J. Sci. bining rules. Fluid Phase Equilib. 130, 73 85.
261, 47 60. Sreide I., and Whitson C. H. (1992) Peng-Robinson predictions for
Enick R. M., and Klara S. M. (1990) CO2 solubility in water and brine hydrocarbons, CO2, N2 and H2S with pure water and NaCl brine.
under reservoir conditions. Chem. Eng. Comm. 90, 2333. Fluid Phase Equilib. 77, 217240.
Garcia J. E. (2003) Fluid dynamics of carbon dioxide disposal into Srensen H., Pedersen K. S., and Christensen P. L. (2002) Modeling of
saline aquifers. Ph.D. thesis, University of California, Berkeley. gas solubility in brine. Org. Geochem. 33, 635 642.
LBNL Report 54280, Lawrence Berkeley National Laboratory, Spycher N., Pruess K., and Ennis-King J. (2003) CO2-H2O mixtures in
Berkeley, California. the geological sequestration of CO2. I. Assessment and calculation
Harvey A. H., and Prausnitz J. M. (1989) Thermodynamics of high- of mutual solubilities from 12 to 100C and up to 600 bar.
pressure aqueous systems containing gases and salts. AIChE J. 35, Geochim. Cosmochim. Acta 67, 30153031.
635 644. Wolery T. (1992) EQ3/6: software package for geochemical modeling
He S., and Morse W. (1993) The carbonic acid system and calcite of aqueous systems: package overview and installation guide (ver-
solubility in aqueous Na-K-Ca-Mg-Cl-SO4 solutions from 0 to sion 7.0). Lawrence Livermore National Laboratory Report UCRL-
90C. Geochim. Cosmochim. Acta 57, 35333554. MA-110662 PTI. Livermore, California.
Holmes H. S., Busey R. H., Simonson J. M., and Mesmer R. E. (1994) Xu T., Apps J. A., and Pruess K. (2004) Numerical simulation of CO2
CaCl2(aq) at elevated temperatures. Enthalpies of dilution, isopies- disposal by mineral trapping in deep aquifers. Appl. Geochem. 19,
tic molalities and thermodynamic properties J. Chem. Thermodyn. 917936.
26, 271298. Zuo Y., and Guo T. M. (1991) Extension of the Patel-Teja equation of
Kiepe J., Horstmann S., Fisher K., and Gmehling J. (2002) Experimen- state to the prediction of the solubility of natural gas in formation
tal determination and prediction of gas solubility data for CO2 water. Chem. Eng. Sci. 46, 32513258.
H2O mixtures containing NaCl or KCl at temperatures between
313 and 393 K and pressures up to 10 MPa. Ind. Eng. Chem. Res.
41, 4393 4398. APPENDIX AACTIVITY COEFFICIENT
Li Y., and Nghiem L. X. (1986) Phase equilibria of oil, gas and FORMULATIONS FOR AQUEOUS CO2
water/brine mixtures from a cubic equation of state and Henrys
A.1. General Considerations
law. Can. J. Chem. Eng. 64, 486 496.
Malinin S. D., and Kurovskaya N. A. (1975) Investigations of CO2 The usual approach to determining activity coefficients of dissolved
solubility in a solution of chlorides at elevated temperatures and gases in water is to make use of the relationship
pressures of CO2. Geokhym. 4, 547551.
Malinin S. D., and Savelyeva N. I. (1972) The solubility of CO2 in ln() ln(kH kH0 ) (A-1)
NaCl and CaCl2 solutions at 25, 50 and 75C under elevated CO2
pressures. Geokhym. 6, 643 653. where kH and kH0 is the ratio of gas fugacity ( f) to aqueous molality (m)
Masoudi R., Tohidi B., Danesh A., and Todd A. C. (2004) A new or mole fraction (x) (Henrys constant) in saline and pure water,
approach in modeling phase equilibria and gas solubility in elec- respectively (i.e., kH f/m or kH f/x, as discussed below).
CO2 solubility in chloride brines 3319

When using Eqn (A-1), the definition and concentration units used 0.411370585 6.07632013 104T 97.5347708 T
for expressing kH and kH0 are crucial in the definition of . Fundamen-
tally, kH0 should be a true equilibrium constant extrapolated to condi- 0.0237622469P T 0.0170656236P (630 T)
tions of infinite dilution (at 1 bar, or water saturation pressure above 1.41335834 105T ln (P)
100C), in which case Eqn (A-1) yields a true activity coefficient,
depending not only on salt concentration but also deviating from unity 3.36389723 104 1.98298980 105T
at non-zero concentrations of the dissolved gas (i.e., the convention that
1 as xCO2 or mCO20). Practically, however, kH0 is often taken as the 2.12220830 103P T 5.24873303 103P (630 T)
ratio of the gas fugacity to the gas solubility in pure water but not where T is temperature in degrees Kelvin (ranging from 273 to 533 K),
extrapolated to infinite dilution. In this case, the activity coefficients P is pressure in bar (ranging from 0 to 2000 bar), m are molalities (for
obtained with Eqn (A-1) are unity in pure water (i.e., =1 as xsalt or ionic strength ranging from 0 to 4.3 m, but up to 6 m NaCl and 4 m
msalt0), and can be related to the salting-out coefficient, ks, through CaCl2 in our P-T range of interest). The activity coefficient calculated
the Sechenow equation in this way is not a true activity coefficient and is related to the CO2
solubility by the following relationship (see main text):
ln( ) ksmsalt (A-2)
mCO
0
2
mCO2 (A-5)
Note that the salt molality (msalt) in Eqn A-2 is commonly replaced by
0
ionic strength for salts other than 1:1 electrolytes. where m is the aqueous CO2 molality in pure water at P and T and mCO2
CO2
The difference between and = can be evaluated using expressions is the aqueous CO2 molality in a saline solution with a composition defined
presented by Diamond and Akinfiev (2003) to compute the activity by mNa, mK, mCa, mMg, mCl and mSO4 at the same P and T.
coefficient of CO2 in pure water. In our P-T range of interest, the
difference is always less than 7%. Note that this difference is absorbed Rumpf et al. (1994)
into the values of KCO0
in Eqn. (2) and (4), because these values were This model makes use of a Pitzer formulation similar to that adopted
2
fitted to experimental data assuming unit activity coefficient in the by Duan and Sun (2003). Their model yields activity coefficients on the
absence of dissolved salts (Spycher et al., 2003). Therefore, for con- molality scale, with the convention that m 1 as msalt0. Rumpf et al.
sistency, values of activity coefficients x introduced into Eqn. (2) (1994) use solubility correlations (Rumpf and Maurer, 1993) without
and (4) should be corrected for the convention that x1 as xsalt0, simplifications, and their activity coefficients are appropriate for use in
even though the correction is not very large. Eqns (2) and (4) after conversion to a mole fraction scale, as shown
At elevated salt concentrations, a more important difference in below. The activity coefficients are given by
activity coefficient values obtained with Eqn (A-1) results from the
concentration units in the expression of Henrys law. If Henrys con- ln (m ) 2msaltB(0) 3msalt
2
(A-6)
stants are defined in terms of mole fractions (e.g., Cramer, 1982), then
with
the activity coefficients are on a mole fraction scale (x), even though
they might be related to a salting-out coefficient (through Eqn A-2) B(0) 0.254 76.82 T 10656 T2 6312 103 T3
using molality (as was done by Cramer, 1982). If Henrys constants are
defined in terms of molality (e.g., Drummond 1981), then the activity 0.0028
coefficients derived through Eqn (A-2) are on a molality scale (m).
The relationship between both types of coefficients can be derived from where T is temperature in degrees Kelvin (ranging from 313 to 433 K)
fundamental thermodynamic relations (e.g., Denbigh, 1983, p. 278) and and msalt is NaCl molality (up to 6 molal). For extension to other
is given by chloride solutions, we extend their model using the same simplifica-


tions as proposed by Duan et al. (1992) (and applied by Duan et al.
m x 1
mi xx H2O (A-3)
2003), based on the assumption that interaction parameters for ions
55.508 with the same charge have roughly the same values. Doing so yields

ln (m ) 2B(0)(mNa mK 2mCa 2mMg)


where the summation of molalities mi is across all dissolved species in
solution. As mentioned previously, values of activity coefficients x in 3mCl(mNa mK mCa mMg) (A-7)
Eqns (2) and (4) should always be on a mole fraction scale, for consistency
with the derivation of these equations (Spycher et al., 2003). where m stands for the molality of each individual ion in solution. For
Smaller differences in activity coefficient values derived through compatibility with Eqs (2) and (4), the activity coefficients are con-
Eqn (A-1) may also arise if experimental kH0 values are not extrapolated verted (Eqn A-3) to a mole fraction scale and convention that x1 as
to water saturation pressures (or 1 bar at T 100C). Other small xsalt0 using the following relationship:


differences may also result from the method used to calculate the gas
fugacity necessary to determine kH and kH0. If partial pressure (Pi) is mCO2(aq) mi CO2(aq) mCO2(aq)
x m 1 1 (A-8)
used instead of fugacity ( fi i Pi i yi Ptotal), errors in values 55.508 55.508
calculated through Eqn A-1 will be primarily caused by the difference
between the CO2 gas mole fraction (yi) over pure and saline water (the Cramer (1982)/Battistelli et al. (1997)
difference between fugacity coefficient values CO2 over pure and This model is derived from salting-out coefficients (ks) reported by
Cramer (1982), for NaCl solutions, and regressed as a function of
saline water is negligible in our P-T range of interest). temperature by Battistelli et al. (1997). The model takes the form:

A.2. Activity Coefficient Expressions from the Literature log10 (x) ksmsalt (A-9)
with
Duan and Sun (2003)
The activity coefficient formulation presented by Duan and Sun ks 1.19784 101 7.17823 104T 4.93854 106T2
(2003) is a Pitzer formulation fitted to experimental solubility data. It
1.03826 108T3 1.08233 1011T4
takes the following form:
where T is temperature in degrees centigrade (ranging 0 to 350C) and
ln () 2(mNa mK 2mCa 2mMg) msalt is NaCl molality (ranging 0 1.95 m). As discussed earlier, values
mCl(mNa mK mCa mMg) 0.07mSO4 (A-4) of x  in this particular case are on a mole fraction scale, with the
convention that x 1 as xsalt0, and can be used directly in Eqns (2)
with and (4) without further correction.
3320 N. Spycher and K. Pruess

Drummond (1981) (A-12) the result of that equation for xsalt 0. These manipulations
Drummond (1981) reported regression equations to compute yield the following revised expression:
Henrys law constants as a function of temperature and NaCl molality.
From his Henrys law expressions and Eqn A-1, the following activity ln (x) A(2xNaCl 2xNaClxCO2(aq) xNaCl
2
) (A-13)
coefficient formulation can be derived:
where A remains the same function of temperature as shown above.
 3
ln ( ) (1.0312 1.2806 10 T 255.9 T)ms
m
APPENDIX BDERIVATION OF EQN. 11
(0.4445 1.606 103T)ms (ms 1) (A-10)
First, the CO2 molality is expressed by substituting Eqn (7) into
Eqn (9)
where T is temperature in degrees Kelvin (for 20 to 400C) and ms
is NaCl molality (for 0 6.5 m). The activity coefficient values given by B1 y H2O(msalt 55.508)
this equation are on a molality scale with the convention that m 1 as mCO2 (B-1)
ms0. Identical expressions have been derived by others for imple- 1 B1 y H2O
mentation into geochemical modeling codes (e.g., Wolery 1992). For The salt mole fraction is then obtained by substituting Eqn (B-1) into
compatibility with Eqns (2) and (4), these activity coefficients are Eqn (8)
converted to a mole fraction scale, with the convention that x1 as
xsalt0, using Eqn A-8. msalt
xsalt (B-2)
B(1 y H2O)(msalt 55.508)
Nesbitt (1984) msalt 55.508
Nesbitt (1984) developed a simple formulation applicable to NaCl 1 B(1 y H2O)
solutions, yielding true activity coefficients on a mole fraction scale as
Eqn (B-2) is then expanded to yield
follows:
msalt1 B(1 y H2O)
ln (x) A(xH2 2O 1) (A-12) xsalt
msalt1 B(1 y H2O) 55.5081 B(1 y H2O)
with (after corrections of typographical errors on the original paper) B(1 y H2O)(msalt 55.508) (B-3)
which, after further expansion and cancellation of several terms becomes
A exp(1.281 105T2 9.606 103T 9.445) (1.98717T)
msalt msaltB msaltBy H2O
xsalt (B-4)
where T is temperature in degrees Kelvin (ranging from 273 to 440 K; msalt 55.508
the author provides other coefficients for higher temperatures) and xH2O
is the water mole fraction (on the basis of a fully ionized salt, up to 6 m Substituting Eqn (B-4) into Eqn (6) and rearranging yields final
NaCl). Conversion for compatibility with Eqn (2) and (4) with the Eqn (11)
convention that x1 as xsalt0 is carried out by expressing the water
mole fraction in Eqn (A-12) in terms of the mole fractions of other (1 B)55.508
y H2O
components (i.e., xH2O 1xsalt xCO2), then subtracting from Eqn (1 A B)(msalt 55.508) msaltB

Você também pode gostar