Sei sulla pagina 1di 26

Fluid Phase Equilibria, 94 (1994) 89- 114 89

Elsevier Science B.V.

A gE model for single and mixed solvent electrolyte


systems
1. Model and results for strong electrolytes

Jiding Li , fans-Gavin Polka b and Jiirgen Gmebling b,*


tl Department of Chemical Engineering, Tsinghua University, Peking (China)
b Lehrstuhl fiir Technische Chemie (FB 9), Universittit Oldenburg, Postfach 2503,
D-261 I1 Oldenburg (Germany}
(Received March 31, 1993; accepted in final form August 24, 1993)

ABSTRACT

Li, J., Polka, H.-M. and Gmehling, J., 1994. A gE model for singfe and mixed solvent
electrolyte systems. 1. Model and results for strong electrolytes. Fluid Phase Equilibria, 94:
89-l 14.

A gE mode1 for electrolyte systems has been developed which is based on results from
statistical thermodynamics and takes into account the interactions between all species present
in the electrolyte solution. In this model the electrolyte solution is treated as a nonelectrolyte
solution plus charge interactions, and it is assumed that the charge interactions show two
effects, direct and indirect. The indirect effect differs not only from the long-range electro-
static interaction but also from the short-range interaction for nonelectrolyte solutions. With
the help of a large data base the required parameters (164 in number) have been fitted for
10 solvents, 18 cations and 10 anions. The model parameters have been used to calculate the
VLE behavior, osmotic coefficients and mean ion activity coefficients for a large number of
systems with high accuracy.

Keywords: theory, excess functions, vapour-liquid equilibria, electrolytes.

INTRODUCTION

The knowledge of the phase equilibria of single-solvent or mixed-solvent


electrolyte systems is very important for separation processes in the chemi-
cal industry, hydrometallurgy, water pollution control and so on. A general
thermodynamic model for the calculation of activity coefficients of solvents
and electrolytes is required in order to be able to predict phase equilibria for
these systems reliably.

* Corresponding author.

0378-3812/94~$07.00 e) 1994 - Elsevier Science B.V. All rights reserved


SSDZ 0378-3812(93)02423-K
90 J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114

For single-solvent electrolyte systems, in particular aqueous electrolyte


systems, Debye and Hiickel presented their theory of interionic attraction in
1923. This theory first made it possible to calculate the ion activity co-
efficients at infinite dilution. In order to apply the Debye-Hiickel theory to
higher electrolyte concentrations, Stokes and Robinson (1948) proposed the
concept of ionic hydration, and derived a two-parameter equation for single
electrolyte solutions. However, this theory suffers from two serious defects:
the ionic hydration numbers are not additive and it cannot be used for
mixed electrolytes. Glueckauf (1955), Robinson et al. (1971), and Stokes
and Robinson (1973) improved ionic hydration theory, so that it can be
used for the calculation of the activity of water in concentrated aqueous
electrolyte solutions. The expression used to calculate the activity of water
is, however, very complicated.
Since the early 197Os, integral techniques based on the Ornstein-Zernicke
equation have been used to solve the primitive model of electrolyte solutions.
Blum ( 1975) used the method proposed by Baxter ( 1971) for hard-sphere
solutions and square-well potentials, and obtained a solution for the mean
spherical approximation (MSA). Planche and Renon ( 1981) generalized the
method of Baxter by using the formalism of Blum, taking into account
short-range forces between the molecules in order to get a nonprimitive
representation of electrolyte solutions. The resulting analytical expression for
the Helmholtz energy of the mixture can be used to represent osmotic
coefficients in aqueous systems with strong electrolytes. The potential used
in the MSA is, however, still far from that of actual electrolyte solutions.
Pitzer (1973) developed general equations for the thermodynamic proper-
ties of aqueous electrolytes based on the Debye-Hiickel theory. These
general equations can be used to calculate accurately activity coefficients of
both pure and mixed electrolytes in water up to a concentration of several
moles of salt per kg of solvent. The Pitzer equations are therefore widely
used for aqueous electrolyte systems. However, if the Pitzer parameters are
applied to the high concentration range (for example, IO-20 M) very large
deviations are obtained in most cases. In addition, in the Pitzer equations,
as in many other models, the interactions between solvent and solvent, and
solvent and solute have not been taken into account, and the solvent is only
treated as a continuum, which makes it impossible to extend the Pitzer
equations to mixed-solvent electrolyte systems.
Up to now most of the research on electrolyte solutions has been
restricted to aqueous electrolyte systems. Electrolyte systems with more than
one solvent have received only minor attention.
Mock et al. (1986) applied the NRTL model to electrolyte systems. Their
model provides a consistent thermodynamic framework for the representa-
tion of the phase equilibrium of mixed solvent electrolyte systems. Different
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89- 114 91

parameters are, however, required to calculate the VLE at different tempera-


tures (pressures). This means that the model can only be used to correlate VLE
of mixed-solvent systems containing strong electrolytes for a given tem-
perature (pressure), Sander et al. (1986) proposed an extended UNIQUAC
model for mixed-solvent/salt systems. Their model can be used to describe and
predict the behaviour of vapour-liquid equilibria for various mixed solvent/
salt systems. Unfortunately, in most cases only poor results are obtained
for osmotic coefficients and mean ion activity coefficients. In addition, there
seems to be an inconsistency in the Sander model concerning the derivation of
the Debye-Hiickel term for a solvent in a mixed-solvent system.
Macedo et al. (1990) used a modified Debye-Htickel term derived from
the McMillan-Mayer solution theory to replace the Debye-Htickel term in
the Sander model.
The gE model is presented in this work in order to describe VLE, osmotic
coefficients and mean ion activity coefficients for single- and mixed-solvent
electrolyte systems, and to overcome most of the weaknesses inherent in the
current models. The present model is based on the interactions between the
different species (ions and molecules) as well as the results of statistical
thermodynamics, in which solvent/solvent, solvent/solute and solute/solute
interactions are taken into account. The complicated electrolyte solution
containing single or mixed solvents will be treated as a nonelectrolyte solution
with charge interactions. In addition, Debye-Hiickel theory and the UNI-
QUAC model are used in this model.

THERMODYNAMIC RELATIONS OF SINGLE-SOLVENT AND MIXED-SOLVENT


SYSTEMS WITH STRONG ELECTROLYTES

For strong (100% dissociated) electrolytes, the mole fractions of solvent


s(x,) and ion j(x,) in the liquid phase are defined in the following way:

(1)

xl = nj 1 nso1+ 2 %m (2)
sol LO >
where n is the number of moles, and sol and ion represent the solvent and
the ion respectively.
The salt-free mole fraction of solvent s is defined as

4 = n, 1 nsol (3)
1 sol
The pure liquid is chosen as the reference state of the solvent. Different
reference states (pure ion and infinite dilution) are used for the ions in this
92 .I. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114

work. If the concentration scales are the same, the relations for the activity
coefficient and chemical potential at the standard states in different reference
states can be written as
P/ =luj +RTlny& (7)
In yy = In yg - In yjOtBj (8)
where /-I, and & are the chemical potentials of the standard states.
Superscript indicates that the refernece state is the pure liquid. Superscript
# refers to the reference state B. The state B can be not only a traditional
reference state such as infinite dilution, but also a state in the multicompo-
nent system. This reference state is called the working reference state (Li et
al., 1990), and is determined flexibly by practical considerations.
However, if the reference states are the same, for example the reference
state x~=x,-,l,~+O,y, + 1, the relationship of the activity coefficients
between mole fraction and molality scale is represented by

(9)

where the superscript (x) denotes the mole fraction scale and (m) the
molality scale. M, is the mean molecular weight of the mixed solvent (m)
(kg mol-), and M, the molecular weight of the pure solvent s (kg mol-I).
The osmotic coefficient is calculated from

(10)

where M, is the molecular weight of the solvent s (kg mol-) and ysis the
activity coefficient of the solvent s.
Assuming ideal vapour phase behaviour, VLE can be calculated from the
equation

GYX = Y.3 (11)


where y, is the vapour phase mole fraction of solvent s and P: is the vapor
pressure of pure solvent s at system temperature.

THEORETICAL BASIS

The following interactions have to be taken into account in a treatment


of electrolyte systems.
(i) Coulomb (charge-charge) interactions between ions:

u12 = e1 ezlDr,2 (12)


J. Li et al. / Fluid Phase Equilibria 94 (1994) 89- 114 93

Since the potential u12 varies with r-r, its effect is felt over a very long
range.
(ii) Charge-induced dipole interactions between ions:
U12= +z:az + z&)/o& (13)
where z1 and z2 are the charges, and aI and a2 are the ion polarizabilities.
(iii) Charge-dipole interactions between ion and molecule
u12= -zlep2 cos e/Dry2 (14)
where p2 is the electric moment of the dipole.
(iii) Ion/ion, ion/molecule and molecule/molecule dispersion interactions:
U12= - 1Sal c(*ZrZJ(Z1 + Z&z (15)
where Z is the ionization potential.
(iv) Ion/ion, ion/molecule and molecule/molecule core repulsion interac-
tions. This repulsion can usually be approximated by
u12= Ar;; n: 18, 161 A >O (16)
or by a hard-core model.
The ion/ion, ion/molecule and molecule/molecule interactions in elec-
trolyte solutions should be divided into two groups: the charge interactions
given by eqns. (12), (13) and (14), and the noncharge interactions repre-
sented by eqns. (15) and (16). The charge interactions include charge-
charge, charge-induced dipol and charge-dipole interactions. All noncharge
interactions can be considered as short-range interactions of nonelectrolyte
solutions. The electrolyte solution should therefore be treated as a nonelec-
trolyte solution with additional charge interactions. For example, for the
simplest nonelectrolyte solution the potential energy between species i and j
is given by
Uij = 03 (y < 4

= 0 (y > Q) (17)
If Coulomb forces (a charge interaction) act on it, it becomes the electrolyte
solution:
Uij = 00 (y < a)
= ziz,e2/Dr (r > 4 (18)
From statistical thermodynamics (Rasaiah and Friedman, 1968; Ra-
manathan and Friedman, 1971), the osmotic pressure of a solution can be
written as

l-I - ckT = l/6 cc cicj (19)


ij s0
94 J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

Where II is the osmotic pressure, c, and c, are concentrations, c is the total


concentration, g, is the radial distribution function, and k is the Boltzmann
factor. If the potential of eqn. (18) is used, the following equation is
obtained:

n- CkT = (e2/6D) 11 C,CizlZJ Ixgcj(r)4nr dr


1J s 0

+ 2/3W3W 11 c,cJg,J(a) (20)


1J

Pitzer ( 1973) combined eqn. (20) with the traditional Debye-Hiickel the-
ory, and derived the following important formula:
4 - 1 = n/ckT - 1 = - ti3/24nc( 1 + KLZ)
+ c[2na3/3 + (1/48n)(~a/c*( 1 + K)~)] (21)
Where e5 is the osmotic coefficient, and K* is defined as

K? = (4ne2/DkT) 1 z?c, (22)


D is the dielectric constant and e is the electron charge.
In eqn. (21) the first term on the right-hand side describes the long-range
electrostatic interactions, which represent traditional Debye-Hiickel theory.
The second term is a so-called hard-core term, which depends on the ionic
strength of the solution. However, it is well known that the Debye-Hiickel
theory only considers long-range electrostatic interactions. The first term
therefore only represents the effect of charge-charge interactions. The
second term represents the effects that arise as a consequence of introducing
the charges into the simple nonelectrolyte solution described by eqn. (17). It
is an indirect effect because it is not concerned with the effect of the direct
interactions between the charges but with the effect which becomes impor-
tant under the influence of the electric field of the charges. As a result we
may assume that the charge interactions have two effects: direct and indirect
effects, both depending on the ionic strength of the solution. Furthermore,
the quantitative expression for the indirect effect given by the second term
in eqn. (21) will be generalized to describe the indirect effects, not only
between ions but also between solvent and ion, because up until now it has
not been possible to treat the complicated potential functions between
species (ions or molecules) by means of statistical theories.
Apart from the theoretical analysis described above, the following theo-
retical results are used in this work.
(A) It can be readily shown by statistical thermodynamics that the first
virial coefficient represents an ideal gas, the second virial coefficient is
J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114 95

caused by the interactions between pairs of species, and the third virial
coefficient arises from three-body interactions. The second virial coefficient
is of greater interest because it is related to the radial distribution function.
If the interaction only depends on the distance r, the equation for the
second virial coefficient is

B = 27rN, no[ 1 - exp( --~(~)jkT)]r~ dr (23)


s0
where U(Y)is the potential.
(B) If the potential in eqn. ( 18) is used, the radial distribution functions
derived by statistical mechanics can be calculated numerically by the Monte
Carlo method or by other methods. For l-l electrolytes the radial distribu-
tion functions at contact distance a are shown in Fig. 1.
If the mixture is ideal and completely random the radial distribution
functions g, are always equal to unity. In Fig. 1, the g, + (a) = g__(a) values
are always less than unity for ionic strengths between 0 and 4; the interac-
tions between ions of same charge can thus be ignored over a wide
concentration range. However, the g++(a) = g-_(u) values increase slowly
with increasing ionic strength, which also means that the interactions
between ions of the same charge have to be considered at very high

Fig. 1. Radial distribution function as a function of ionic strength for the primitive potential
model of an electrolyte solution from Monte Carlo calculations (Card and Valleau, 1970).
96 J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

concentrations. In contrast, the g+_ values are always greater than unity.
Thus, the interactions between ions of opposite charge must be taken into
account across the whole concentration range.

THE NEW gE MODEL

Various types of interactions have to be taken into account for electrolyte


solutions. In our model the excess Gibbs energy is calculated as the sum of
three contributions:
GE = GER + GsR + GFR (24)
The first term on the right-hand side of the equation represents the
long-range (LR) interaction. contribution caused by the Coulomb electro-
static forces, and mainly describes the direct effects of charge interactions.
The second term represents the indirect effects of the charge interactions.
In order to distinguish the indirect effects from the noncharge short-range
interactions of the nonelectrolyte solution, and because some of the charge
interactions (such as the charge-dipole interactions and charge-induced
dipole interactions) are proportional to ye2 and re4, we shall refer to this
term as the middle-range (MR) interaction contribution.
The third term expresses the contribution of the noncharge interactions,
which is identical to the short-range (SR) interactions in nonelectrolyte
solutions.
The GER term
The direct effects of the charge interactions arise mainly from the electro-
static forces. G& can thus be calculated in terms of the Debye-Hiickel
theory as modified by Fowler and Guggenheim ( 1949):
ion

GER = -(30) - ,F, siz;e2r(~u) (25)


where si is the number of ions i in the system, D is the mixed-solvent
dielectric constant, a is the distance of closest approach between two ions
and r(x) is defined as
r(x) = 3x -3[ln( 1 + X) - x + x2/2] (26)
where x = k-a.
The expressions for the long-range interaction contribution to the activity
coefficients for ion j and solvent s follow from the appropriate derivations
of eqn. (25):
In yk = -[z_fe/(2DkT)][~/( 1 + x)] + [iij/(24CV,,a3)]x3a(x) (27)
In yt = [i7J(24~NOa3)]x30(x) (28)
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-l 14 91

where ijj and V, are the partial molar volumes for ion j and solvent s, and
cr.(x)is calculated by the following equation:
a(x)=3x-3[1+~-(l+~)-1-21n(l+x)] (29)
Robinson and Stokes (1965) have shown that the second term on the
right-hand side of eqn. (27) is not very important for the activity coefficient
of ion j and may be neglected. Thus, the formula for In y) can be
approximated to
In y, = - +412/( 1 + bl2) (30)
If the partial molar volume of the solvent s in the solution is approximated
as the molar volume of the pure solvent s, the equation for In rb can be
derived from eqn. (28):

ln Y, = [2AM,d/b3d,][ 1 + bI2 - (1 + bI*) - - 2 ln( 1 + bI/)] (31)


where M, is the molecular weight of the solvent s (kg mol-) and d, is the
molar density of the pure solvent s evaluated from the DIPPR Tables (1984)
in kmol mW3. d is the mixed-solvent molar density calculated from the
following empirical formula:

(32)

where cpiO,is the salt-free volume fraction of solvent sol in the liquid
phase, and is defined by

(33)

In eqns. (31) and (32) I is the ionic strength, and A and b are the
Debye-Hiickel parameters calculated as

I = 0.5 C m,z:

A = 1.327757 x 10sdO~s/(DT).s (35)


b = 6.359696d0.s/(DT)0.s (36)
where T is the absolute temperature in Kelvin and D is the dielectric
constant for mixed solvents. Equations (35) and (36) are based on a
value of 4 8, for the distance of closest approach. For a binary mixture,
by using Osters mixing rule (Franks, 1973) the formula for D can be
written as
D = D, + [(& - 1)(2& + 1)/20* - (0, - l)]C& (37)
98 J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

For a multicomponent mixture, D can be estimated from the expression

where Dsol is the dielectric constant of the pure solvent sol, and can be
obtained from the Table of Maryott and Smith (1951).

The GbR term

The G& term is the contribution of the indirect effects of the charge
interactions to the excess Gibbs energy. For a solution containing n, moles
ofsolventn(n=1,2,... , sol) and ni moles of ion j (j = 1,2, . . . , ion) GLR
is obtained from the following equation:

GE&T = h&'~ c Bklnkn,


k I

Where nkg denotes n kg solvent, and Bkl is the interaction coefficient between
the species k and I (ion or molecule) and is similar to the second virial
coefficient representing the indirect effects. By using the simplest potential
model for electrolyte solutions (eqn. (18)) and from radial distribution
theories of statistical thermodynamics, one can obtain an expression for the
dependence of the so-called indirect effects on the ionic strength (eqn. (21)).
Furthermore, in eqn. (21) the first term in brackets, 2rcu3/3, is independent
of the ionic strength, and the second term in brackets depends on the ionic
strength in the solution. Combining these facts with eqn. (23), it seems
reasonable to assume that Bkl, which respresents all indirect effects caused
by the charges, can be described by the following simple formula:
B,(I) = b,j + Cijexp(all* + ~121) (40)
where b, and c,, are the so-called middle-range interaction parameters
between species i andj (6, = bjl, Cii= Cji).a, and a2 are constants which were
determined empirically using a number of reliable experimental data for
single solvent and mixed solvent electrolyte systems. The best results were
obtained with the values given in eqns. (41) and (42):
B lon,lon= bion,ion+ ~ion,lon
exp( -P2 + O-131) (41)
Bion,sorvent
= bion.sorvent
+ C,on,solvent
exp( - 1,2P* + 0.131) (42)
Here, the formula for Bion,ion_ is different from the formula for Bion,solventy
which is reasonable because the ion/ion charge interactions (eqns. (12) and
(13)) are different from the ion/solvent interactions (eqn. (14)).
On the basis of the approximate results of the radial distribution func-
tions between ions, it follows that the so-called middle-range interactions
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-l 14 99

between like-charged ions can be ignored at moderate concentrations. In


addition, there appear to be no middle-range interactions between solvents,
so that eqn. (39) can be simplified to

GLIRT = (nkg) - C C Bsor,ion(I)nsolnion


[ solion
+ C C Bc,(Z)n,n,
c a 1
where the c indices cover all cations and the a indices cover all anions.
(43)

By differentiating eqn. (43) with respect to the number of moles of solvent


and ions, one obtains

ln Y? = c B,,,,,(W,,, - (MS/Mm) 1 C [Bs,i.i,,(I)


Kin solion

+ IB~,,,,,,(I)lx~,,m,,, - Ms c c VU0 + K_AWv, (44)


c LI

ln Y? = (Mm) - 2 B,,,,i(OxL + [Z,?/(2Mm)I C C BHol,ion(I)x&1mion


sol sol Km
+ 1 Bj.a(Oma + <zf /2) 1 C BLa(Omcmu- B,,s(I)IMv (45)
a c c1
where x&,, is the salt-free mole fraction of solvent sol, B(1) is equal
to dB(Z)/dl, and M, is the mean molecular weight of mixed solvent
(kg mol-I), and is calculated as

M, = c x~cl,~SOl (46)

The GtR term

G& expresses the contribution of the short-range interactions to the


excess Gibbs energy, and can be calculated using the UNIQUAC model.
The expression for the activity coefficient of solvent s can be written as
(47)

(48)

(49)

(50)

(51)

(52)
100 J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

In these equations i and k cover all solvents and ions, r, and qS are the van
der Waals volumes and surface areas of the solvent S, and aij are the
UNIQUAC interaction parameters between species i and j, whereby aij is
different from aj,.
The relationship for the activity coefficient of ion j normalized to the
infinite dilution reference state can be obtained from eqn. (8):

lnyjSR=InyF-lny~(B)+lny~-Inyy(ZI) (53)
where In yc and In yp can be written as eqns. (48) and (49). In y&,
I. I and
are- calculated from the following equations:

= 1 - rj /r, + ln(rj /VS> - %j [ 1 - rjqs lCrsqj) + ln(rjq3 /Crsqj ))I (54)


= qj ( 1 - In YSj - YJS) (55)

The complete expressions for the activity coeficients of solvents and ions

The complete equation for calculation of the activity coefficient of sol-


vent s is

ln yS = In yFR + In yy + In yf (56)
where the concentration scale of the solvent s is the mole fraction, the
reference state xl = x, + 1, ys+ 1, and the standard state is pure solvent s
at the temperature and pressure of the system. In rk, In y,, In yf are
calculated from eqns. (31), (44) and (47) respectively.
With the help of eqn. (9) the complete formula for the activity co-
efficient of ion j is obtained:

In yj = (In 7, + In YrR + In yf) - In MS /Mm + MS C mien (57)


( ion >
where the concentration scale of ion j is the molality scale, the refer-
ence state xl =xS-+l, Z+O, 1, and the standard state is the hypo-
Yj +

thetical ideal solution of unit molality at system temperature and pressure.


In y:, In y,v, In 7; are calculated from eqns. (30), (45) and (53)
respectively.
The mean activity coefficient of a salt MX is calculated by

ln yMXCt) = ( l/v>bM In YM + VX In 7x1 (58)


where vM and vx represent the stoichiometric factors for the cation M and
the anion X. v = vM+ vx is the stoichiometric number of moles of ions for
one mole of salt MX. The terms In yM and In yx are calculated from eqn.
(57).
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114 101

ESTIMATION OF PARAMETERS

A computerized data base (293 binary (one solvent, one salt) and 154
ternary (two solvents, one salt) data sets) were used to fit the model
parameters for 10 solvents, 18 cations and 10 anions (Fig. 2). The VLE data
types included in the database are: (1) X, y, T, P = constant; (2)
x, y, P, T = constant; (3) x, P, T = constant; (4) x, T, P = constant. Fur-
thermore, osmotic coefficients (5) x, 4, T = constant and mean activity
coefficients (6) x, y f , T = constant have also been included. The maximum
salt concentration was 29 mol kg- for 1: 1 salts, 14 mol kg- for 2: 1 salts.
Unfortunately, since a thermodynamic consitency test on VLE with salts is
not possible, it is difficult to evaluate the quality of the data sets.
The model parameters were determined by minimization of the following
objective function using the Simplex-Nelder-Mead method (Nelder and
Mead ( 1965)) :

F(Qij,ali, bij, c,,>


=C 1 ge(Qexp - Qcad2 = min
njl nt

n fitted

0 not fitted

q larger data base desirable

use zero if needed

Fig. 2. Present status of the model parameter matrix.


102 J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114

where Q represents y, y, T, P, cj and y &, and is a weighting factor for


gQ

Q. np and nt are the number of data points and data types respectively. The
subscript exp refers to experimental, and talc to calculated data.
The fitted interaction parameters are given in Table 1. Four parameters
(two short-range aji and ail; two middle-range b, and cii) are required to

TABLE 1
The interaction parameters of the model

MeOH 289.6 - 181.0


EtOH 162.4 - 14.5
2-PrOH 83.60 110.5
I-PrOH 130.0 106.2
1-BuOH - 23.46 308.8
Acetone -45.90 337.9
Li+ 621.4 - 584.6 - 7.443 1.574
Na+ 219.4 - 299.8 - 7.432 1.576
Kf 537.5 -452.3 - 7.475 1.577
Rb+ -258.8 58.10 1.576
CS+ 30.9 3.427 1.576
NH; 16.19 -342.1 - 10.36 1.576
Ag+ 300.5 378.3 1.576
Mg2+ 663.3 - 679.3 3.155
Ca+ 1137.0 - 759.9 - 19.84 3.149
Sr2+ 798.0 - 527.7 - 19.91 3.152
Ba2 430.0 - 557.5 - 19.79 3.152
Mn2+ -623.5 - 689.7 3.152
Fe?+ 943.5 -643.1 3.152
co2+ 697.5 - 807.0 3.151
Ni2+ 721.7 - 705.5 - 19.48 3.152
cl.l2+ 687.3 - 765.6 - 18.73 3.152
Zn2+ 481.6 -20.44 -20.38 3.134
Cd2+ j 326.0 - 308.3 - 19.76 3.142
F- 11.39 120.1 7.052 - 1.576
Cl- 22.93 159.1 7.387 - 1.576
Br 22.41 282.5 5.952 - 1.576
I- 21.47 341.7 7.483 - 1.576
ClO; -21.46 327.4 - 1.576
ClO, -28.14 316.5 7.675 - 1.576
CH, COO - 16.86 287.4 7.445 - 1.576
NO; - 17.99 347.6 7.462 - 1.576
so:- - 364.0 789.4 14.28 -3.151
OH- - 17.24 193.6 - 1.576
EtOH - 101.7 130.2
2-PrOH 39.7 -26.7
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89- I14 103

TABLE 1 (continued)

i j a, a,I

MeOH 1-PrOH - 13.2 57.1


MeOH 1-BuOH 40.39 121.5
MeOH CH,COOMe - 37.75 311.7
MeOH CHCl, - 137.5 658.0
MeOH THF -77.12 316.9
MeOH LIf 298.6 -634.8 5.760 1.117
MeOH Na+ 495.5 601.3 5.702 1.198
MeOH Kf 555.3 641.1 5.555 1.452
MeOH Rb+ 653.9 -486.8 1.508
MeOH CS+ 145.2 - 530.2 1.547
MeOH NH,+ 782.5 895.7 0.4216 1.118
MeOH Ca*+ 1988.0 919.2 2.446 1.936
MeOH cl_l*+ 167.8 626.1 4.284 1.969
MeOH Zn*+ 387.1 347.4 10.65 2.906
MeOH F- -7.100 - 186.7 - 6.220 -1.132
MeOH Cl- - 378.2 -231.6 - 5.620 - 1.006
MeOH Br- -92.80 -62.15 -8.211 - 1.263
MeOH I- -489.8 - 672.8 - 5.352 - 1.367
MeOH ClO; - 92.08 -116.9 - 1.253
MeOH CH3COO- - 295.0 671.8 - 5.647 - 1.090
MeOH so:- - 227.7 - 105.1 - 12.06 -3.591
EtOH 2-PrOH - 176.5 241.9
EtOH 1-PrOH - 69.6 126.7
EtOH 1-BuOH 162.3 16.39
EtOH Li+ - 313.2 -661.6 3.998 1.071
EtOH Na+ - 172.2 - 304.8 3.811 1.435
EtOH K+ 280.5 - 75.24 3.637 1.158
EtOH NH: 271.2 1023.0 - 3.870 1.071
EtOH Ca*+ 2102.0 1334.0 - 5.429 1.760
EtOH Sr*+ 799.3 701.5 - 5.503 2.358
EtOH Ba*+ 1082.0 - 276.8 - 5.097 1.921
EtOH cu*+ 1140.0 -713.8 -2.615 0.6357
EtOH Zn*+ 319.9 - 692.2 - 6.633 2.761
EtOH Cd* 1075.0 - 296.5 - 5.341 2.618
EtOH F- 177.5 200.4 -4.078 -2.071
EtOH Cl- 1519.0 - 164.9 - 3.620 - 1.206
EtOH Br- 57.68 103.8 - 7.279 - 1.457
EtOH I- 314.3 194.5 - 3.434 - 1.340
EtOH ClO, 1986.0 -381.2 -4.382 -0.8019
EtOH CH,COO- 1098.0 - 20.09 - 3.487 - 1.267
EtOH NO; - 37.62 - 3.430 - 3.537 - 1.092
2-PrOH 1-PrOH - 200.9 289.4
2-PrOH 1-BuOH 845.5 -511.2
2-PrOH Li+ -482.6 515.8 5.250 1.227
2-PrOH Na+ - 327.3 131.2 5.062 2.114
104 J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

TABLE 1 (continued)

i i aY
aJr btj lj m nlax

2-PrOH Ca2+ 376.2 45.45 -6.382 2.718


2-PrOH Cl- 2204.0 - 100.0 -4.962 - 1.506
2-PrOH Br- 867.3 452.3 -9.708 - 1.626
I-PrOH 1-BuOH 449.1 - 338.7 -
1-PrOH Li+ -613.2 - 777.0 0.000 2.522 !
I-PrOH Na+ - 358.4 412.1 5.215 1.963 -
1-PrOH K+ -641.2 778.9 5.173 3.071 -
1-PrOH NH: 874.8 318.6 -4.789 2.078 -
1-PrOH Ca2+ 1071.0 1282.0 -6.184 2.550 -
I-PrOH srZ+ 2261.0 357.1 - 6.359 2.742 -
I-PrOH Ba*+ 1463.0 - 113.4 - 5.956 2.879 _
I-PrOH NiZ+ 464.7 392.0 -4.980 2.551 -
I-PI-OH Cd*+ 180.6 389.8 - 6.845 10.52 !
1-PrOH Cl- 2204.0 - 550.0 - 4.962 - 1SO6 _
1-PrOH NO; 2176.0 -499.5 -4.791 - 1.841 -
I-BuOH Na+ 841.7 944.7 -7.001 1.493 -
1-BuOH Ca2+ -288.2 202.3 - 14.16 2.368 -
I-BuOH Cl- 769.6 845.1 6.936 - 1.406 -
Acetone Li+ 1555.0 1163.0 3.065 1.256 -
Acetone K+ -741.7 152.2 2.799 1.649 -
Acetone Cl- 739.9 - 333.3 - 2.653 - 1.206 -
CH3COOMe Na+ 503.5 - 674.4 0.05276 2.024 _
CH3COOMe Ca2+ 1492.0 -498.4 -21.22 2.013 !
CH3COOMe Zn*+ -493.3 455.6 -2.382 0.8312 _
CH3COOMe Cl- 1213.0 636.3 0.7034 -1.246 -
CH3COOMe CH,COO- 238.1 81.83 0.3891 - 1.246 -
CHCl, Li+ - 986.6 495.7 7.768 1.853 -
CHCl, Cl- 595.5 205.6 - 6.403 - 1.806 _
THF K+ - 165.9 836.5 -4.565 1.144 _
THF CH,COO- 904.5 111.5 4.821 -1.165 -
Li+ Cl- 33.12 - 164.6 0.3416 -0.6285 19.2
Li+ Br- -812.4 -51.08 0.4646 -0.4535 20.0
Li+ I- -711.8 -663.5 0.2691 1.808 10.1
Li+ ClO,- 196.0 -512.9 0.2360 0.9890 4.5
Li+ CH,COO- 104.6 - 237.2 0.07441 0.1587 4.0
Li+ NO; 405.7 442.5 0.1331 -0.2879 20.0
Na+ F- 1.302 - 1.276 0.00967 - 0.02895 1.0
Na+ Cl- 89.17 -483.7 0.1925 0.1165 6.0
Na+ Br- 46.60 - 143.8 0.2174 -0.03856 9.0
Na+ I- - 220.4 563.5 0.2591 - 0.03670 12.0
Na+ ClO; - 33.96 - 535.0 0.09547 0.1751 3.0
Na+ ClO; 128.3 - 294.6 0.1077 0.09876 6.0
Na+ CH,COO- - 85.86 279.1 0.2029 0.2139 3.5
Na+ NO; -292.1 - 1376.0 -0.2313 3.537 10.8
Na+ so;- 269.3 -919.1 0.063 11 0.3924 4.0
Na+ OH- 574.6 -407.1 0.3108 - 0.4205 29.0
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114 105

TABLE 1

i i aY b<J

K+ F- 343.5 - 70.36 0.1766 -0.2147 17.5


K+ Cl- -417.2 244.7 0.09109 -0.1161 5.6
K+ Br- - 835.2 - 149.1 0.1476 0.4943 5.5
Kf I- - 537.8 161.4 0.1098 0.2092 5.6
K+ ClO, 2.784 7.470 -0.2228 0.07244 0.7
K+ CH,COO- -85.10 215.6 0.2147 0.2440 3.5
K+ NO, 504.1 -315.1 -0.08355 -0.1635 3.6
K+ so:- 10.27 196.7 -0.1177 -0.8661 0.9
K+ OH- - 106.5 308.5 0.3213 - 0.3930 20.0
Rb+ cl- 16.65 106.9 0.1044 -0.2138 6.5
Rb+ Br- -263.3 -11.06 0.1035 -0.1555 5.0
Rb+ I- -351.5 134.7 0.1155 -0.1810 4.1
CS+ Cl- -251.9 33.75 0.1344 -0.4132 11.0
CS+ Br- 561.6 573.4 0.08069 -0.3573 5.9
CS+ I- -584.1 - 14.31 0.1293 -0.2586 3.0
NH; Cl- 21.83 332.2 0.03971 -0.1091 8.0
NH: ClO; 199.9 427.4 -0.01045 -0.6339 2.1
NH: NO: 696.0 -783.2 -0.00124 -0.00444 26.0
Ag+ NO; - 945.6 127.2 0.03663 -0.07459 13.0
Mg2f cl- - 199.8 -644.6 0.3565 1.985 5.0
Mg2+ Br- - 1009.0 - 86.33 0.6979 1.168 5.0
Mg2+ I- 311.0 - 656.8 0.5940 2.563 5.0
Mg2+ ClO; -1147.0 - 191.6 0.7995 2.264 4.0
Mg2+ NO; - 108.1 -357.3 0.2825 1.261 2.0
Ca2+ cl- 339.0 0.000 0.4088 -0.2575 12.0
Ca2+ Br- - 1021.o 431.8 0.5323 0.8322 9.0
Ca2+ I- 425.5 461 .O 0.3401 1.113 2.9
Ca2+ NO; -932.5 1629.0 0.07833 0.1674 8.0
Sr2+ cl- 18.29 -204.2 0.3667 1.054 4.0
Sr2+ Br- 288.9 -429.0 0.4056 2.108 3.3
Sr2+ I- 463.4 372.6 0.5565 0.8553 4.1
Sr2+ NO; 857.7 111.0 -0.01566 0.4758 4.0
Ba2 cl- 400.7 -450.2 0.2056 0.6651 1.4
Ba2+ Br- 505.3 - 266.4 0.3163 0.5158 3.4
Ba2+ I- 128.0 -660.3 0.4980 1.744 2.0
Ba2+ NO; 51.26 108.1 - 0.2870 - 0.8653 0.4
Mn2+ Cl- 607.8 -224.1 0.1434 -0.1572 8.0
Fe+ cl- - 57.56 -237.5 0.2870 1.746 2.0
Co2f Cl- 386.4 - 344.7 0.2229 -0.3854 6.0
co2+ NO, -1143.0 - 155.9 0.2462 0.8404 5.0
Ni2+ cl- 231.1 - 181.1 0.3768 0.08065 5.0
cu*+ ClV -118.9 -115.7 -0.05795 0.3295 6.0
Cd+ NO, - 984.9 41.08 0.2968 - 0.02787 8.0
Zn2+ Cl- -72.91 - 126.7 0.2042 -0.00557 22.0
Zn2+ Br- 555.4 1002.0 0.2517 -0.01957 20.0
106 J. Li et al. 1 FIuid Phase Equilibria 94 (1994) 89-114

TABLE 1 (continued)

i j a, 41 bl, c 11 *mix
Zn2+ I- 455.8 873.6 0.3870 -0.2412 12.0
Zn2+ ClO, 37.23 -782.3 0.9374 2.020 4.0
Zn2+ NO; 578.1 940.1 0.5017 0.03591 6.0
Zn+ so;- 46.40 - 1182.0 0.2585 1.718 3.5
Cd+ CI- -541.3 -729.7 0.04152 0.08395 6.0
Cd+ Br- -832.8 - 927.0 0.1498 -0.00026 4.0
Cd+ I- 224.6 - 1275.0 0.3809 - 3.394 2.5
Cd+ NO; 301.1 -329.0 - -0.1397 6.086 2.5

!: The use of the parameters at high concentration of salt can lead to erroneous results.
THF: tetrahydrofuran.

describe the cation-anion interactions. As can be seen from eqn. (44) for
single-solvent systems, only the derivation of Bsor,ionwith respect to the ionic
strength is necessary for the calculation, which in accordance with eqn. (42)
contains only the c parameter. Therefore, owing to a lack of mixed-solvent
electrolyte data, only three parameters (Q, Uji and cl,) for the ion-solvent
interaction were estimated in a few cases. For the solvent-solvent interac-
tion two short-range interaction parameters are always necessary without
any middle-range parameters. The short-range interaction parameters, and
the volume and surface area parameters for the solvents were taken directly
from Gmehling and Onken (1977) and Gmehling et al. (1982). The values
are listed in Table 2. The volume and surface area parameters for the ions
can be estimated from the ionic radii. However, if crystallographic radii are
used, values between 0.1 and 0.5 for the cations are obtained, which reduces
the flexibility of the short-range contribution significantly. It was therefore
decided to fix the volume and surface parameters for all ions at a value of
1.O, which is typical for small molecules like water (r = 0.92; q = 1.4).

TABLE 2
UNIQUAC volume (rs) and surface area (q,v) parameters of solvents

J 5 rs

Hz0 0.920 1.400 1-BuOH 3.454 3.052


MeOH 1.431 1.432 Acetone 2.574 2.336
EtOH 2.106 1.972 CH,COOMe 2.804 2.576
2-PrOH 2.779 2.508 CHCl, 2.870 2.410
1-PrOH 2.780 2.512 THF 2.942 2.720

THF: tetrahydrofuran.
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-l 14 107

In the parameter matrix shown in Fig. 2, the black square means that
the corresponding parameters have been fitted using a sufficiently large
data base and a wide temperature range. The shaded square refers to
lack of good-quality data or data covering only a small temperature
range.
The parameters in Table 1 can be used to predict the behaviour of a
large number of single and mixed solvent systems with salts at different
temperatures and pressures. For practical applications it is very impor-
tant to estimate the maximum electrolyte concentration. It is, however,
difficult to give an accurate range for every parameter. In the last column
of Table 1 only the maximum salt concentration is given from the
water-salt data sets. In addition, the parameters marked ! should be
used carefully because of the limited data base available for parameter
fitting.
If the pressure is less than 1.2 bar and the salt concentration does not
exceed moderate ranges (up to 6 molal solutions), the unmarked parame-
ters in this work can be used for reliable predictions of vapour phase
compositions, temperatures, pressures, activity coefficients of the solvents
and osmotic coefficients. The predictions for the mean activity coefficients
of salts have to be limited to the temperature range 20-30C because of
the limited data base used (25C).

RESULTS AND DISCUSSION

The overall results for single solvent and mixed solvent electrolyte sys-
tems calculated by the model proposed in this paper are summarized in
Table 3. The detailed results and a comprehensive comparison with the
results of other models are presented by Polka et al. (1993).
For mixed solvent/salt VLE data sets (data type 1, 2) the total mean
absolute deviation between experimental and calculated mole fractions in
the vapour phase is 0.018, the temperature 1.04 K and the pressure
0.79 kPa. These results show that the model presented in this paper can be
used reliably to predict VLE for mixed solvent/salt systems. It must be
pointed out here that the above results include all reliable data sets (185
data sets). Only those data sets (19 data sets) were excluded in which the
salt concentration as given by the authors is obviously larger than its
solubility. A few typical prediction results for mixed solvent/salt VLE
data sets are presented in Figs. 3 and 4. It can be seen that the model
provides reliable results for aqueous (A, B, C, D, E) and nonaqueous
(F, G, H) systems, whereby (A, F, H) the salt concentration varied from
zero to saturation.
108 .I. Li et al. 1 Fluid Phase Eq~iIi~ria 94 (1994) 89-114

TABLE 3
Deviations between experimental and predicted osmotic coefficients, mean ion activity
coefficients, binary and ternary VLE

Data type Data sets Ay AT (K) BP (KPa) A4 AYf

(1) x, y, T, P = const 157 0.0180 1.04


(2) x, y, P, T = coonst 28 0.0169 0.79
(3) x, T, P = const 15 0.28
(4) x, P, T = const 223 0.20
(5) x, 4, T = const
m,, < 7.9 59 0.009
m,,, > 7.9 23 0.052
(6) x, y +, T = const
mmax< 7.9
and y +,,, < 30.0 36 0.008
mnlBX 7.9
and y +,,, > 30.0 21 0.149

A: total mean absolute deviation.

Fig. 3. Experimental and predicted y-x phase diagrams for different mixed solvent/salt
systems, (A) methanol( l)-water(2)/NaCl at 100.59 kPa, NaCl molalities 0.2-6.0; (B)
ethanol( I)-water(Z)/CaCl, at 101.325 kPa, CaClz molality 2.0; (C) ethanoh I)-water(2)/
NH,CI at 100.525 kPa, NH4C!1 molalities 0.2- 11.1; (D) 2-propanol( I) -water(Z)/LiBr at
101.325 kPa, LiBr molality 0.6; (E) acetone( I)-water(2)fKCl at 101.325 kPa, KC1 molality
1.0; (F) acetonef I)-methanol~2)~LiCl at 101.325 kPa, LiCI molality 2.8-3.8; (G) methyl
acetate{ 1)-methanol(2)~NaOOCCH~, at 101.325 kPa, NaOOCCH, molality 0.0-2.0; (H)
CHCI,( l)-methanol(2)~~iCl, at 101.325 kPa, LiCl molality 2.5 to 3.8 ( -) model, (A)
experimental values from (A) Johnson and Furter ( 1960), (B) Dobroserdov et al. (1958), (C)
Furter (1958), (D) Sada et al. (1975), (E) Kupriyanova (1973) (F) Tatsievskaya et al.
(1982), (G) Park et al. (1973), (H) Tatsievskaya et al. (1982).
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114 109

I I / I I 5, I / I I I I I I, I / I
5ooo 00 0, 06 08 02 01 06 08 02 01 08 08 02 01 06 08 10

r,,ss,t ,rer,. y, x,(mlt rrec,.Y, X,bll herb. y, I,,.dL free). y,

Fig. 4. Experimental and predicted T-y -_x phase diagrams for different mixed solvent/salt
systems. (A) methanol( 1) -water( 2)/NaCI at 100.59 kPa, NaCl molalities 0.2-6.0; (B)
ethanol( 1) -water( 2) /CaCl, at 101.325 kPa. CaCI, molality 2.0; (C) ethanol( 1) - water( 2)/
NH,Cl at 100.525 kPa, NH,Cl molality 0.2- 11.1; (D) 2-propanol( 1) -water( f)/LiBr at
101.325 kPa, LiBr molality 0.6; (E) acetone(l)-water(2)/KCl at 101.325 kPa, KC1 molality
1.0; (F) acetone( 1) -methanol(2)/LiCl at 101.325 kPa, LiCl molality 2.8-3.8; (G) methyl
acetate( 1) -methanol( 2)/NaOOCCH,, at 101.325 kPa, NaOOCCH, molality 0.0-2.0; (H)
CHCl,( I)-methanol(2)/LiCl, at 101.325 kPa, LiCl molality 2.5-3.8. (-) model, ( A, Cl)
expermental values from (A) Johnson and Furter ( 1960), (B) Dobroserdov et al. (1958), (C)
Furter ( 1958), (D) Sada et al. ( 1975), (E) Kupriyanova (1973), (F) Tatsievskaya et al.
(1982) (G) Park et al. (1973), (H) Tatsievskaya et al. (1982).

For single solvent/salt VLE data sets (data type 3,4) the absolute
deviations of pressure and temperature are within the experimental error in
most cases. If the solvent is water there is almost perfect agreement with the
experimental data. If the solvent is nonaqueous the absolute deviations
exceed the experimental error in a few cases.
The total mean absolute deviation of osmotic coefficients (data type 5) is
0.009, and the relative deviations are within 1% in most cases at moderate
salt concentrations (mm._ < 8). Even at high salt concentrations
(8 <mmax < 29) the total mean absolute deviation is smaller than 0.052, the
relative deviations being approx. 3%. Figure 5 shows examples for the
calculation of osmotic coefficients at moderate and high salt concentrations.
The results for the mean ion activity coefficients (data type 6) are even
better than those for osmotic coefficients across the whole salt concentration
range. In most cases the calculated results for the mean ion activity
co-efficients are in perfect agreement with the experimental ones for moderate
110 J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

1.4

-0.2
00 2.0 40 6.0 5.0 10.0 15.0 20.0 25.0 : ,o
m m

Fig. 5. Experimental and predicted osmotic coefficients for aqueous electrolyte solutions at
25C. (A) salt molalities 0.0-7.9; (B) salt molalities 0.0-29.0. (--) model, ( +, a, 0)
experimental values from (A) MgBr?, Stokes (1948); Co(N03),, Stokes (1948); NH,Cl,
Hamer and Wu (1972); (B) NaOH, Hamer and Wu ( 1972); NH,NO,, Hamer and Wu
(1972); CaCI, at 5OC, Duckett et al. (1986).

and high salt concentrations. Figure 6 shows some results in which the salt
molalities vary from 0 to 20 and the values for the mean ion activity
coefficients from 0.2 to 500.
The model gives quite large deviations for water-2-propanol-LiCI, l-
propanol-2-propanol-CaClz and methanol-ethanol-CaCI, systems (mean
absolute deviation for the mole fraction of vapour phase 0.0252, for the
temperature 1.88 K and for the pressure 1.50 kPa). Better results can be
obtained by using different constants for a, and a2. We suggest that the
constants (a, and a2) may not be the best for all solvent-ion pairs,

-1.2

-J.6

8.0 5.0 10.0 15.0 50 JO.0 15 0 * 0..O

m m m

Fig. 6. Experimental and predicted mean activity coefficients for aqueous electrolyte solu-
tions at 25C. (A) salt molalities 0.0-6.5, y f 0.2- 1.05; (B) salt molalities O.O- 19.5, y k
0.5-60; (C) salt molalities 0.0-19.5, y + 0.2-500. (----) model, (+, A, 0) experimental
values from (A) NaCl, Hamer et al. (1972); KI, Hamer and Wu (1972); Sr(NO,),, Stokes
(1948); (B) CaCl,, Stokes (1948); LiCl, Hamer and Wu (1972); MgBr,, Stokes (1948); (C)
LiBr, Hamer and Wu (1972); CaBr,, Stokes (1948).
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114 111

CONCLUSION

A general gE model describing the behaviour of single and mixed solvent


systems with salts has been proposed. In this model the electrolyte solution
has been treated as a nonelectrolyte solution with additional charge interac-
tions. The charge interactions are assumed to have direct and indirect effects
on the electrolyte solutions. The direct effect has been represented by
Debye-Hiickel theory. The quantitative results of the indirect effect of
Coulomb interactions derived from statistical thermodynamics are general-
ized to describe the indirect effects of all charge interactions.
With the help of a large data base, 164 parameter groups have been fitted;
these include 10 solvents, 18 cations and 10 anions. The model parameters
have been used to predict different types of data sets. For 185 mixed
solvent/salt VLE data sets, the total mean absolute deviation for the vapour
phase mole fraction is 0.018, for the temperature 1.04 K and for the
pressure 0.79 kPa. For 238 single solvent/salt VLE data sets, the absolute
deviations are within the experimental errors in most cases. For 139
water-salt osmotic coefficient and mean ion activity coefficient data sets, the
total mean relative deviations are no more than 1% for moderate salt
concentrations (mmax< 8) and approx. 3% for the high salt concentrations
(8 < m,,, < 29). The results show that the model presented in this paper can
be used to predict VLE for single solvent and mixed solvent/salt systems,
osmotic coefficients and mean activity coefficients for water-salt systems
with high accuracy.

NOMENCLATURE

ai, UNIQUAC interaction parameter between species i and j


A Debye-Hiickel parameter
b Debye-Hiickel parameter
b, middle-range interaction parameter between species i and j
B second virial coefficient
middle-range interaction parameter between species i and j
2 molar density (kmol rn-)
D dielectric constant
e electron charge
F objective function
g radial distribution function
g molar Gibbs energy
gQ weighting factor for thermodynamic quantity Q
G Gibbs energy
I ionic strength, ionization potential
J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

Boltzmann constant
molality
molecular weight
molar number
Avogadro number
pressure
surface area parameter
thermodynamic quantity
volume parameter
distance between two particles
number of ions of type i
absolute temperature
potential function
partial molar volume
liquid phase mole fraction
salt-free liquid phase mole fraction
vapour phase mole fraction
charge number

Greek symbols

Y activity coefficient of solvent


YI!I mean activity coefficient of salt
K inverse Debye length
fl chemical potential
stoichiometric factor
rr osmotic pressure
d, osmotic coefficient
9 volume fraction

Superscripts

C combinatorial
E excess property
LR long range
MR middle range
R residual
S saturation
SR short range
# working reference state

Subscripts

a anion
J. Li et al. 1 Fluid Phase Equilibria 94 (1994) 89-114 113

c cation
ion ion
j ionic species
m solvent mixture
max maximum
s solvent s
SO1 solvent sol

Supplementary material available: the parameters can be obtained on a


diskette, at cost, from the authors.

REFERENCES

Baxter, R.J., 1971. Physical Chemistry: An Advanced Treatise, Vol. VIII A, Chapter 4, In:
H. Eyring, D. Henderson and W. Jost (Eds.), Academic Press, New York.
Blum, L., 1975. Mol. Phys., 30: 1529.
Card, D.N. and Valleau, J.P., 1970. J. Chem. Phys., 52: 6232.
Debye, P. and Hiickel, E., 1923. Theory of electrolytes. 1. Freezing point lowering and
related phenomena. Physik Z., 24: 185.
DIPPR Tables of Physical and The~odynamic Properties of Pure Compounds, AIChE,
New York, 1984.
Dobroserdov, L.L., 1958. Tr. Leningr. Tekhnol. Inst. Pishch. Promsti., 15: 55.
Duckett, L.M., Hollifield, J.M. and Patterson, C.S., 1986. J. Chem. Eng. Data, 31: 213.
Fowler, R.H. and Guggenheim, E.A., 1949. Statistical Thermodynamics, Chapter 9, Cam-
bridge University Press.
Franks, F., 1973. Water, A Comprehensive Treatise, Vol. 2, Chapter 7, Plenum, New York.
Furter, W.F., 1958. Ph.D. Thesis, University of Toronto, Ontario.
Glueckauf, E., 1955. Trans. Faraday Sot., 51: 1235.
Gmehling, J. and Onken, U., 1977. Vapor-Liquid Equilib~um Data Collection,
DECHEMA Chemistry Data Series, Vol. 1, Part 2a, DECHEMA, Frankfurt.
Gmehling, J., Onken, U. and Arlt, W., 1982. Vapor-Liquid Equilibrium Data Collection,
DECHEMA Chemistry Data Series, Vol. 1, Part 2c, DECHEMA, Frankfurt.
Hamer, W.J. and Wu, Y.C., 1972. J. Phys. Chem. Ref. Data, 1: 1047.
Johnson, A.I. and Furter, W.F., 1960. Can. J. Chem. Eng., 38: 78.
Kupriyanova, Z.N., 1973, Zh. Prikl. Khim. (Leningrad), 46: 234.
Li, J., Li, Y., Lu, J. and Teng, T., 1990. J. Chem. Ind. Eng. (in Chinese), 1: 66-73.
Macedo, E.A., Skovborg, P. and Rasmussen, P., 1990. Chem. Eng. Sci., 45: 875-882.
Maryott, A. and Smith, E.R., 195 1. Table of Dielectric Constants of Pure Liquids, National
Bureau of Standards Circular.
Mock, B., Evans, B. and Chen, C.C., 1986. AIChE J., 32: 16.55-1664.
Nelder, J.A. and Mead R.A., 1965. Comput. J., 7: 308.
Park, W.K., Do, KS., Bae, H.K. and Shim, H.S., 1973, Rep. Inst. Ind. Technol., 1: 33.
Pitzer, K.S., 1973. J. Phys. Chem., 77: 268-277.
Planche, H. and Renon, H.J., 1981. Phys. Chem., 85: 3924.
Polka, H.-M., Li, J. and Gmehling, J., 1994. Fluid Phase Equilibria, 94: 115-127.
Ramanathan, P.S. and Friedman, H.L., 1971. J. Chem. Phys., 54: 1086.
Rasaiah, J.C. and Friedman, H.L., 1968. J. Chem. Phys., 48: 7242.
114 J. Li et al. / Fluid Phase Equilibria 94 (1994) 89-114

Robinson, R.A. and Stokes, R.H., 1965. Electrolyte Solutions, Butterworth, London.
Robinson, R.A., Duer, W.C. and Bates, R.G., 1971. Anal. Chem., 43: 1862.
Sada, E., Morisue, T. and Yamaji, H., 1975. Can. Chem. Eng., 53: 350.
Sander, B., Fredenslund, Aa. and Rasmussen, P., 1986. Chem. Eng. Sci., 41: 1171- 1183.
Stokes, R.H., 1948, Trans. Faraday Sot., 44: 295.
Stokes, R.H. and Robinson, R.A., 1973. J. Solution Chem., 2: 173.
Tatsievskaya, G.I., Vitman, T.A., Kushner, T.M. and Serafimov, L.A., 1982. Russ. J. Phys.
Chem., 56: 1668.

Potrebbero piacerti anche