You are on page 1of 15

A comprehensive model for calculating phase equilibria and

thermophysical properties of electrolyte systems

P. Wang, A. Anderko*, R.D. Young and R.D. Springer


OLI Systems Inc.
108 American Road
Morris Plains, NJ 07950, USA

ABSTRACT

A thermodynamic model has been developed for calculating phase equilibria and
other properties of multicomponent electrolyte systems. The model has been designed to
reproduce the properties of both aqueous and mixed-solvent electrolyte systems ranging
from infinite dilution to solid saturation or pure solute limit. The model incorporates
formulations for the excess Gibbs energy and standard-state properties coupled with an
algorithm for detailed speciation calculations. The excess Gibbs energy model consists of
a long-range interaction contribution represented by the Pitzer-Debye-Hckel expression,
a second virial coefficient-type term for specific ionic interactions and a short-range
interaction term expressed by the UNIQUAC equation. The accuracy of the model has
been demonstrated for common acids and bases and for multicomponent systems
containing aluminium species in various environments.

*
Author to whom correspondence should be addressed. E-mail: aanderko@olisystems.com

1 of 15
INTRODUCTION

Electrolyte thermodynamics is of great importance for the simulation and


optimization of hydrometallurgical processes. In many processes, multicomponent
electrolyte solutions are encountered at high concentrations under diverse conditions of
temperature and pressure. Such solutions are challenging for computational models
because of their complex chemical behaviour and strong nonideality.

Various models for electrolyte solutions have been reviewed by Zemaitis et al.
(1), Pitzer (2), Rafal et al. (3), Loehe and Donohue (4) and Anderko et al. (5). A
characteristic feature of electrolyte systems is the fact that phase equilibria and other
thermodynamic properties are often inextricably linked to speciation equilibria, which
may be due to ion pairing, acid-base reactions, complexation and other phenomena. In
many industrial processes, speciation-related properties such as pH, oxidation-reduction
potential, distribution of complexed or hydrolyzed species, etc., are of primary
importance. Thus, self-consistent treatment of speciation and phase equilibria is of
utmost importance for realistic simulation of electrolyte systems.

Recently, a general, speciation-based thermodynamic model has been developed


for mixed-solvent electrolyte solutions (6). This model was shown to reproduce
simultaneously vapor-liquid equilibria, solid-liquid equilibria, speciation, caloric and
volumetric properties of electrolytes in water, organic, or mixed solvents. The model is
valid for salts from infinite dilution to the fused salt limit and for various completely
miscible inorganic systems (such as acid-water mixtures) over a full concentration range.
Also, the model is capable of representing phase equilibria in multicomponent inorganic
systems containing multiple salts, acids, and bases (7). Complex phase behavior such as
formation of multiple hydrated salts, double salts, or the presence of eutectic points can
be accurately represented. The model is referred to as the mixed-solvent electrolyte
(MSE) model because it is equally valid for classical aqueous systems, those with more
than one distinct solvent and mixtures in which a given component may continuously
vary from being a solute to being a solvent (e.g., in acid-water mixtures).

In this study, we examine the applicability of this model to selected systems that
are related to hydrometallurgical processes. Rather than focusing on particular processes,
we select systems that provide a stringent test for the model. First, the model is used to
represent the properties of fundamentally important acid-water and base-water solutions.
Then, it is applied to systems containing aluminum in various aqueous environments.

THERMODYNAMIC MODEL

The thermodynamic framework has been developed (6) by combining an excess


Gibbs energy model for mixed-solvent electrolyte systems with a comprehensive
treatment of chemical equilibria. In this framework, the excess Gibbs energy is expressed
as

2 of 15
ex
G ex G LR G ex G ex
= + MR + SR (1)
RT RT RT RT

ex ex
where G LR represents the contribution of long-range electrostatic interactions, GMR
ex
represents specific, primarily ionic (i.e., ion-ion and ion-molecule) interactions and G SR
ex
is a short-range contribution resulting from intermolecular interactions. For brevity, GMR
is referred to as the middle-range term. Activity coefficients are obtained from eq. (1) by
differentiation following standard thermodynamic definitions. The derivation of eq. (1)
was discussed in detail by Wang et al. (6) and, therefore, only the basic features of the
model are summarized here.

The long-range interaction contribution is calculated from the Pitzer-Debye-


Hckel formula (8, 9) expressed in terms of mole fractions and symmetrically
normalized, i.e.,



ex
G DH 4 Ax I x 1 + I x1 2
= ni ln (2)
RT i
0
x i [ 1 + I x ,i ( ) 12
]
i

where the sum is over all species, Ix is the mole fraction-based ionic strength, Ix,i0 denotes
the ionic strength when the system composition reduces to a pure component i, i.e., Ix,i0
= zi2; is related to the hard-core collision diameter ( = 14.0) and Ax is given by

32
1 e2
Ax = (2N A d s )1 2 (3)
3 4 0 s k B T

where ds and s are the molar density and the dielectric constant of the solvent,
respectively.

For the short-range interaction contribution to the excess Gibbs energy, the
UNIQUAC equation (10) is used.

ex ex ex
GUNIQUAC G combinator G residual
ial
= + (4)
RT RT RT

with the combinatorial and residual terms expressed as:

ex
G combinator ial Z
= ni x i ln i + q i x i ln i (5)
RT i i xi 2 i i

3 of 15
ex
G residuall
= ni q i xi ln j ij (6)
RT i i j

where

qi xi
i = (7)
qjxj
j
ri x i
i = (8)
rj x j
j

a ji
= exp (9)
ji RT

where qi and ri are the surface and size parameters, respectively, for the species i; Z is a
constant with a value of 10; aij is the binary interaction parameter between species i and j
(aij aji). The UNIQUAC model has been extensively verified for systems composed of
neutral molecules.

The middle-range interaction contribution is calculated from an ionic strength-


dependent, symmetrical second virial coefficient-type expression (6):

ex
G MR
= ni xi x j Bij (I x ) (10)
RT i i j

where Bij (I x ) = B ji (I x ) , Bii = B jj = 0 and the ionic strength dependence of Bij is given by

Bij (I x ) = bij + cij exp( I x + a1 ) (11)

where bij and cij are adjustable parameters and a1 is set equal to 0.01. In general, the
parameters bij and cij are calculated as functions of temperature as

bij = b0,ij + b1,ij T + b2,ij / T (12)

cij = c0,ij + c1,ij T + c2,ij / T (13)

In practice, the middle-range parameters are used to represent ion-ion and ion-
neutral molecule interaction. The short-range parameters are used primarily for
interaction between neutral molecules. For electrolyte systems encountered in
hydrometallurgy, the middle-range contribution is by far the most important.

4 of 15
Speciation effects due to the formation of ion pairs and complexes are explicitly
taken into account using chemical equilibria. Chemical equilibrium calculations require
the standard-state chemical potential, i0, for all species together with activity
coefficients. Thus, for a chemical reaction:

aA + bB + ... = cC + dD + ... (14)

the equilibrium conditions can be determined from

G 0 x c x d ... Cc Dd ...
= ln Ca Db (15)
RT x x ... a b ...
A B A B

with
G 0 = i i0 (16)
i

where i0 is the standard-state chemical potential of species i, the sum is over all species
participating in the chemical reaction, and i is the stoichiometric coefficient of species
i in Eq. (14) with positive values for the species on the right-hand side of the equation
and negative values for those on the left-hand side.

The standard-state chemical potentials for aqueous species are calculated as


functions of temperature and pressure using the Helgeson-Kirkham-Flowers equation of
state (11). The parameters of this model are available for a large number of aqueous
species including ions, associated ion pairs, and neutral species (12-15). It should be
noted that standard-state properties calculated from the model of Helgeson et al. are
based on the infinite-dilution reference state and on the molality concentration scale. At
the same time, the activity coefficients obtained from eq. (1) are symmetrically
normalized and based on the mole fraction scale. To make the speciation calculations
consistent, the following conversions are performed: (i) the activity coefficients
calculated from Eq. (1) are converted to those based on the unsymmetrical reference
state, i.e. at infinite dilution in water, as described by Wang et al. (6), and (ii) the
molality-based standard-state chemical potentials are converted to a corresponding mole
fraction-based quantity.

Evaluation of model parameters

The parameters of the model are determined on the basis of various kinds of
experimental data. In particular, data of the following type may be included in the
regression of model parameters:

1. Vapor-liquid equilibria
2. Osmotic coefficients (or activity of water) in aqueous solutions and activity
coefficients, especially in completely dissociated systems;

5 of 15
3. Solubility of salts in water and other solvents
4. Properties of electrolytes at infinite dilution, such as acid/base dissociation and
complexation constants
5. Data that reflect ionic equilibria in solutions, such as pH measurements
6. Densities
7. Heats of mixing and dilution
8. Heat capacities.

Among these types of data, vapor-liquid equilibria and salt solubilities are of
paramount importance. Whereas vapor-liquid equilibrium data directly constrain the
excess Gibbs energy, solid-liquid equilibrium data can be used to constrain both the
excess Gibbs energy and the properties of solid phases that are in equilibrium with the
liquid phase. The use of multiple data types is important to ensure the accuracy of model
parameters. For example, caloric data (heat of mixing or dilution and heat capacities) are
useful to determine the temperature dependence of model parameters. This makes it
possible to make reliable predictions of solubilities well outside of the temperature range
of experimental solubility data.

The experimental data cover the concentration ranges from infinite dilution in
water to the solubility or fused salt limit, whichever applies. The combined model is
typically applicable for temperatures up to 0.9Tc, where Tc is the critical temperature of
the mixture. This is due to the fact that the excess Gibbs energy model is valid only for
liquid-like systems and, therefore, cannot approach the critical locus too closely. Thus,
the properties of relatively dilute aqueous systems can be represented up to ca. 300 C. In
concentrated electrolyte systems, the critical temperature is typically higher and the
model may be valid well beyond 300 C. For example, the model has been successfully
applied up to 500 C for concentrated sulfuric acid solutions as described below.

RESULTS AND DISCUSSION

In this study, we first apply the model to fundamental systems that often
determine the electrolyte environment in practical hydrometallurgical systems. Figure 1
shows vapor-liquid equilibria for the system HCl H2O at various temperatures. At low
and moderate temperatures, this system shows liquid phase splitting, which is indicated
by horizontal lines. The horizontal lines connect the compositions of two coexisting
phases and show the equilibrium pressure of a three-phase liquid-liquid-vapor system.
Below the three-phase pressure, Figure 1 shows the compositions of coexisting vapor and
liquid phases. The correct representation of vapor-liquid and liquid-liquid equilibria
demonstrates that the model accurately predicts the nonideality of hydrochloric acid
solutions.

Figure 2 shows vapor-liquid equilibria in the sulfuric acid water sulfur


trioxide system. Unlike in the case of hydrochloric acid, the volatility of sulfuric acid is
very small and the partial pressure of water in this system is close to the total pressure. In

6 of 15
Figure 2, the mole fraction of sulfur trioxide is used as a concentration variable. This
allows us to show the behavior of oleum mixtures (SO3 H2SO4) in addition to that of
sulfuric acid (H2SO4 H2O) on one plot. In view of the stoichiometric relation SO3 +
H2O = H2SO4, pure H2SO4 corresponds to x(SO3) = 0.5 in Figure 2. It is noteworthy that
the model accurately reproduces the behavior of the system ranging from pure water to
pure sulfur trioxide at temperatures from 0 C to approximately 300 C. At temperatures
between ca. 300 C and 500 C, the model correctly reproduces the phase behavior of
concentrated solutions.

Figure 3 shows solid-liquid equilibria in the system NaOH H2O as a function of


temperature. In this case, multiple hydrated solid phases are observed. The model
accurately predicts the topology of the phase diagram, including eutectic points (e.g., the
ice - NaOH7H2O eutectic) and congruently melting solid hydrate phases (e.g.,
NaOH3.5H2O and NaOHH2O) at temperatures ranging from ca 40 C to 300 C. It
should be noted that the model reproduces other properties such as vapor-liquid
equilibria, enthalpies of dilution and heat capacities in addition to solid-liquid equilibria.

After verifying the performance of the model for fundamental electrolyte


systems, calculations have been made for a large number of binary and multicomponent
systems containing aluminum. Figure 4 shows the solubility behavior of mixed systems
containing aluminum oxides or hydroxides (expressed as weight % of Al2O3), sodium
hydroxide (expressed as weight % of Na2O) and water. The characteristic solubility
maximum in this system is due to a transition between the stability of aluminum
hydroxide or oxyhydroxide solid phases at lower alkalinities and sodium aluminate at
higher alkalinities. The model accurately reproduces this behavior for temperatures
ranging from 0 C to at least 200 C (which is the upper limit of experimental data).

While Figure 4 illustrates the solubility behavior of aluminum at strongly alkaline


conditions, Figure 5 shows a solubility diagram for aluminum species at pH values
ranging from 1 to 11. At the conditions shown in this figure, boehmite (AlOOH) is the
stable phase. The pH effect on the phase behavior is of particular interest here because of
a shift in the pH dependence as the temperature is increased from ca. 100 to 290 C. Both
the temperature and pH dependence of solubility are accurately reproduced by the model.

A particular advantage of the model is its ability to reproduce the properties of


multicomponent systems on the basis of a limited amount of data. Of particular interest is
the prediction of solubilities in systems containing multiple salts. An example of such
calculations is shown in Figures 6 and 7. Figure 6 shows the solubility of solid phases in
the binary system Al2(SO4)3 H2O. The parameters obtained by fitting the
thermodynamic properties of this system and the Na2SO4 H2O binary were further used
to calculate the solubility behavior of the ternary mixture Al2(SO4)3 Na2SO4 H2O.
The results are shown in Figure 7. It is noteworthy that the ternary phase diagram shows
the presence of a solid phase that does not occur in either of the binary subsystems (i.e.,
AlNa(SO4)2). Such double salts are common in many mixed electrolyte systems and can
be accurately predicted by the model.

7 of 15
100

10

-20C -10C 5C 15C 25C 35C 45C


1
Pressure, atm

0.1

0.01

0.001

0.0001
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x (HCl)
Figure 1 - Calculated and experimental vapor-liquid equilibria in the HCl H2O system.
The horizontal lines at high pressures connect the compositions of two
coexisting liquid phases. Experimental data are from references (16-22).
1.0E+03

500C 400C
1.0E+02 300C 200C

1.0E+01 100C

1.0E+00 50C
25C
1.0E-01
0C
P, atm

1.0E-02

1.0E-03

1.0E-04

1.0E-05

1.0E-06

1.0E-07

1.0E-08
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0

x SO3

Figure 2 - Calculated and experimental vapor-liquid equilibria in the sulfuric acid/oleum


system as a function of the mole fraction of SO3. Literature data are taken from
Gmitro and Vermeulen (23), Brand and Rutherford (24), Schrage (25), and
Giazitzoglou and Wuster (26).

8 of 15
100
90
80
NaOH
70
NaOH, weight %

60
NaOH.H2O
50
NaOH.2H2O
40
NaOH.3.5H2 O
30
NaOH.4H2O
20 NaOH.7H2O
10
Ice
0
-40 -20 0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300
Temperature, C

Figure 3 Calculated and experimental solubility of NaOH in aqueous solutions as a


function of temperature. Experimental data are taken from references (27) and
(28).

45

40
200 C
35 Al(OH)3 (s)
or AlOOH(s)
30 Na2 O.Al2 O3.2.5H2 O
Al2O3 wt%

25 110C

20
95C
15
50C
10
30C
5 25C
0
0 10 20 30 40 50 60 70

Na2O wt%

Figure 4 - Representation of solubilities of solids in the ternary system Al2O3 Na2O


H2O. The experimental data are from references (29-36).

9 of 15
1.E+01
1.E+00
150 C
1.E-01 203.3 C

1.E-02 250 C 101.5 C


1.E-03
m (AlOOH)

1.E-04 290.2 C

1.E-05
1.E-06
1.E-07
1.E-08
1.E-09
1.E-10
0 1 2 3 4 5 6 7 8 9 10 11 12
pH

Figure 5 Representation of the solubility of boehmite (AlOOH) in 0.03 m NaCl as a


function of pH. The experimental data are from references (37) and (38).
60

50
wt% Al2(SO4)3

40
Al2 (SO4 )3 .16H2 O
30

20
ice
10

0
-20 -10 0 10 20 30 40 50 60 70 80 90 100 110 120
t/C
Figure 6 Calculated and experimental solubilities of aluminum sulfate as a function of
temperature. The experimental data are from references (27) and (39-43).

10 of 15
40
Al2 (SO4 )3 .16H2 O
35

30
wt% Al2 (SO4 )3

25 AlNa(SO4 )2 .12H2 O

20
Na2 SO4
15
25C 30C
10
0C
5
Na2 SO4 .10H2 O
0
0 5 10 15 20 25 30 35
wt%-Na2SO4

Figure 7 Calculated and experimental solubility of solids in the system Al2(SO4)3


Na2SO4 H2O. The experimental data are from Linke and Seidell (27).

CONCLUSIONS

A recently developed comprehensive mixed-solvent electrolyte model has been


applied to calculate phase equilibria in multicomponent systems containing salts, acids
and bases in wide concentration and temperature ranges. The model is based on a
combination of an excess Gibbs energy formulation with a treatment of standard-state
properties and an algorithm for comprehensive speciation calculations. The model is
valid for aqueous systems ranging from infinite dilution to molten salts or pure acids at
temperatures from the freezing point to 300 C. In favorable cases, the model can be
extended to even higher temperatures (up to 500 C) as long as the system of interest is
subcritical. In particular, the model has been shown to be very useful for reproducing
complex solid-liquid equilibrium diagrams, which may involve multiple hydrates,
eutectic points formed by salt hydrates and ice, double salts, congruently melting solid
phases, etc. The model has been verified for a large number of systems including
common acids, bases and salts of first- and second-group elements, aluminum, tin, and
iron. It is also applicable to organic and organic salt mixtures. Work is in progress on
developing model parameters for a large class of systems that are of interest in
hydrometallurgy.

11 of 15
ACKNOWLEDGEMENT

The work reported here was supported by Alcoa, Chevron, DuPont, Mitsubishi
Chemical, Rohm & Haas and Shell.

REFERENCES

1. J.F. Zemaitis Jr., D.M. Clark, M. Rafal and N.C. Scrivner, Handbook of Aqueous
Electrolyte Thermodynamics, DIPPR, AIChE, New York, 1986.

2. K.S. Pitzer, Ed., Activity Coefficients in Electrolyte Solutions, 2nd ed., CRC
Press, Boca Raton, FL, 1991.

3. M. Rafal, J. W. Berthold, N.C. Scrivner and S. L. Grise, Models for Electrolyte


Solutions, in Models for Thermodynamic and Phase Equilibria Calculations, S.I.
Sandler (ed.), M. Dekker, New York, 1994, p. 601.

4. J. R. Loehe, M. D. Donohue, Recent Advances in Modeling Thermodynamic


Properties of Aqueous Strong Electrolyte Systems, AIChE J., Vol. 43, 1997,
180-195.

5. A. Anderko, P. Wang and M. Rafal, Electrolyte Solutions: From


Thermodynamic and Transport Property Models to the Simulation of Industrial
Processes, Fluid Phase Equilibria, Vol. 194-197, 2002, 123.

6. P. Wang, A. Anderko and R.D. Young, A Speciation-Based Model for Mixed-


Solvent Electrolyte Systems, Fluid Phase Equilibria, Vol. 203, 2002, 141-176.

7. P. Wang, R.D. Springer, A. Anderko and R.D. Young, Modeling Phase


Equilibria and Speciation in Mixed-Solvent Electrolyte Systems, Fluid Phase
Equilibria, Vol. 222-223, 2004, 11-17.

8. K. S. Pitzer, Thermodynamics of Electrolytes. I. Theoretical Basis and General


Equations, J. Phys. Chem., Vol. 77, 1973, 268-277.

9. K. S. Pitzer, Electrolytes. From Dilute Solutions to Fused Salts, J. Am. Chem.


Soc. Vol. 102, 1980, 2902-2906.

10. D. S. Abrams and J. M. Prausnitz, Statistical Thermodynamics of Liquid


Mixtures: A New Expression for the Excess Gibbs Energy of Partly or
Completely Miscible Systems, AIChE J., Vol. 21, 1975, 116-128.

12 of 15
11. H. C. Helgeson, D. H Kirkham and G.C. Flowers, Theoretical Prediction of the
Thermodynamic Behavior of Aqueous Electrolytes at High Pressures and
Temperatures Parts I through IV, Amer. J. Sci. Vol. 274, 1974, 1089-1198;
ibid., Vol. 274, 1974, 1199-1261; ibid., Vol. 276, 1976, 97-240; ibid., Vol. 281,
1981, 1249-1516.

12. E. L. Shock, H. C. Helgeson and D. A. Sverjensky, Calculation of the


thermodynamic and transport properties of aqueous species at high pressures and
temperatures: Standard partial molal properties of inorganic neutral species,
Geochim. Cosmochim. Acta, Vol. 53, 1989, 2157-2183.

13. E. L. Shock and H. C. Helgeson, Calculation of the thermodynamic and


transport properties of aqueous species at high pressures and temperatures:
Correlation algorithms for ionic species and equation of state predictions to 5 kb
and 1000C, Geochim. Cosmochim. Acta, Vol. 52, 1988, 2009-2036.

14. E. L. Shock, D. C. Sassani; M. Willis and D. A. Sverjensky, Inorganic species in


geologic fluids: Correlations among standard molal thermodynamic properties of
aqueous ions and hydroxide complexes, Geochim. Cosmochim. Acta, Vol. 61,
1997, 907-950.

15. D. A. Sverjensky; E. L. Shock, H. C. Helgeson, Prediction of the


thermodynamic properties of aqueous metal complexes to 1000C and 5 kb,
Geochim. Cosmochim. Acta, Vol. 61, 1997, 1359-1412.

16. M. Massucci, S.L. Clegg, and P. Brimblecombe, "Equilibrium Partial Pressures,


Thermodynamic Properties of Aqueous and Solid Phases, and Cl2 Production
from Aqueous HCl and HNO3 and Their Mixtures", J. Phys. Chem. A, Vol. 103,
1999, 4209-4226.

17. J.T.F. Kao, "Vapor-Liquid Equilibrium of Water - Hydrogen Chloride System", J.


Chem. Eng. Data, Vol. 15, 1970, 362-367.

18. H.F. Holmes, R.H. Busey, J.M. Simonson, R.E. Mesmer, D.G. Archer, and R.H.
Wood, "The Enthalpy of Dilution of Aqueous Hydrogen Chloride to 648 K and
40 MPa. Thermodynamic Properties", J. Chem. Thermo., Vol. 19, 1987, 863-890.

19. R.H. Perry and C.H. Chilton, "Chemical Engineers Handbook", 5th Edition,
McGraw-Hill, New York 1973.

20. J.J. Fritz and C.R. Fuget, "Vapor Pressure of Aqueous Hydrogen Chloride
Solutions, 0 to 50 C", Ind. Eng. Chem., Vol. 1, 1956, 10-12.

13 of 15
21. C.F. Pan, "Activity and Osmotic Coefficients in Dilute Aqueous Solutions of
Uni-Univalent Electrolytes at 25C", J. Chem. Eng. Data, Vol. 26, 1981, 183-184.

22. W.J. Hamer, Y.C. Wu, "Osmotic Coefficients and Mean Activity Coefficients of
Uni-Univalent Electrolytes in Water at 25 C", J. Phys. Chem. Ref. Data, Vol. 1,
1972, 1047-1099.

23. J.I. Gmitro and T. Vermeulen, "Vapor - Liquid Equilibria for Aqueous Sulfuric
Acid", Supplementary Data UCRL 10886, 1963.

24. J.C.D. Brand and A. Rutherford, "The Vapour Pressure of the System Sulphuric
Acid - Disulphuric Acid", J. Chem. Soc., 1952, 3916-3922.

25. J. Schrage, "Dampfdrucke von H2SO4 - SO3", Gemischen Fortschritt Berichte,


VDI, Vol. 3, Number 208, 1990.

26. Z. Giazitzoglou and G.Wuster, "Das Binare System H2SO4 - H2O; P, V, T -


Daten und Dampfdruckkurven in Weiten Konzentrationsbereichen bei Erhohtem
Druck", Ber. Bunsenges. Phys. Chem., Vol. 85, 1981, 127-132.

27. W.F. Linke, A.S. Seidell, "Solubilities of Inorganic and Metal Organic
Compounds K-Z", Volumes I and II, 4th Edition, 1965.

28. V.V. Vyazova and A.D. Pelsha, "Handbook of Experimental Solubility Data for
Binary Aqueous and Non-Aqueous Systems Containing Group I Elements",
Volume 3, Number 1-2, 1961.

29. J.W. Sprauer and D.W. Pearce, "Equilibria in the Systems Na2O - SiO2 - H2O
and Na2O - Al2O3 - H2O at 25 C", J. Phys. Chem., Vol. 44, 1940, 909-916.

30. D.J. Wesolowski, "Aluminum Speciation and Equilibria in Aqueous Solution: I.


The Solubility of Gibbsite in the System Na - K - Cl - OH - Al(OH)4 from 0 to
100 C", Geochim. Cosmochim. Acta, Vol. 56, 1992, 1065-1091.

31. G.A. Panasko and P.V. Yashunin, "Calculations for the Na2O - Al2O3 - H2O
System", Russ. J. Appl. Chem., Vol. 37, 1964, 285-289.

32. N.M. Chaplygina, L.S. Itkina, E.V. Petrova, "Solubility and Solid Phases of the
Al(OH)3 - NaOH - H2O and Al(OH)3 - KOH - H2O Systems at 50 C", Russ. J.
Inorg. Chem., Vol. 19, 1974, 762-766.

33. G. Qiu and N. Chen, "Phase Study of the System Na2O - Al2O3 - H2O", Can.
Metall. Quart., Vol. 36, 1997, 111-114.

34. R. Fricke and P. Jucaitis, Zeit. Anorg. Allgem. Chem., Vol. 191, 1930, 129.

14 of 15
35. P. Jucaitis P., Zeit. Anorg. Chem., Vol. 220, 1934, 257.

36. Y.-F. Zhang, Y. Li, and Y. Zhang, "Phase Diagram for the System Na2O - Al2O3
- H2O at High Alkali Concentration", J. Chem. Eng. Data, Vol. 48, 2003, 617-
620.

37. P. Benezeth, D.A. Palmer, and D.J. Wesolowski, "Aqueous High - Temperature
Solubility Studies. II. The Solubility of Boehmite at 0.03 m Ionic Strength as a
Function of Temperature and pH as Determined by In-situ Measurements",
Geochim. Cosmochim. Acta, Vol. 65, 2001, 2097-2111.

38. W.L. Bourcier, K.G. Knauss, and K.J. Jackson, "Aluminum Hydrolysis Constants
to 250 C from Boehmite Solubility Measurements", Geochim. Cosmochim. Acta,
Vol. 57, 1993, 747-762.

39. W.F. Ehret W. F. and F.J. Frere, "Ternary Systems Involving Water and
Aluminum Fluoride with Aluminum Nitrate, Sulfate or Chloride", J. Amer.
Chem. Soc., Vol. 67, 1945, 68-71.

40. J.T. Dobbins and R.M. Byrd, "A Study of the Soda-Alum System", J. Phys.
Chem., Vol. 35, 1931, 3673-3676.

41. E.A. Gee, "The System Aluminum Sulfate - Ethanol - Water at 30 and 80 C", J.
Amer. Chem. Soc., Vol. 67, 1945, 179-182.

42. H.A. Horan and J.A. Skarulis, "The System Li2SO4 - Al2(SO4)3 - H2O at 0 C",
J. Amer. Chem. Soc., Vol. 61, 1939, 2689-2691.

43. A. Apelblat and E. Korin, "The Molar Enthalpies of Solution and Vapor Pressures
of Saturated Aqueous Solutions of AlCl3, Al(NO3)3, and Al2(SO4)3", J. Chem.
Thermodynamics, Vol. 34, 2002, 1919-1927.

15 of 15

You might also like