You are on page 1of 7

Fluid Phase Equilibria 290 (2010) 88–94

Contents lists available at ScienceDirect

Fluid Phase Equilibria


journal homepage: www.elsevier.com/locate/fluid

Thermodynamic modeling and experimental measurement of


calcium sulfate in complex aqueous solutions
G. Azimi, V.G. Papangelakis ∗
Department of Chemical Engineering and Applied Chemistry, University of Toronto, 200 College St., Toronto, ON, M5S 3E5 Canada

a r t i c l e i n f o a b s t r a c t

Article history: A newly developed database for the Mixed Solvent Electrolyte (MSE) model of the OLI Systems software
Received 6 July 2009 was employed to model the solid and aqueous phase equilibria of calcium sulfate hydrates in electrolyte
Received in revised form solutions containing NiSO4 , H2 SO4 , MgSO4 , Fe2 (SO4 )3 , LiCl and HCl from 25 to 90 ◦ C. The MSE model is
24 September 2009
a variant of an excess Gibbs free energy model for Mixed Solvent Electrolyte systems which takes into
Accepted 28 September 2009
account long-range electrostatic interactions from a Pitzer–Debye–Hückel equation, middle-range inter-
Available online 7 October 2009
actions from a second virial coefficient-type equation and short-range interactions from the UNIQUAC
model. The effect of cations with similar anions on the solubility was investigated and it was found that
Keywords:
Calcium sulfate
the solubility depends mainly on the anion type while the cation has a minor effect. The effect of acid
Dihydrate (DH) addition and the acid type was also studied. The addition of both HCl and H2 SO4 increases the solubility;
Hemihydrate (HH) however H2 SO4 has a less pronounced effect due to the common ion effect. Furthermore, the effect of
Anhydrite (AH) phase transitions between different calcium sulfate hydrates was studied. The transformation of CaSO4
Phase transformation dihydrate to anhydrite results in a significant decrease in the solubility, which complicates the chem-
Solubility istry of the system. Since it is not practical to measure solubility data under all conditions of interest,
Chemical modeling the model developed and the experimental results obtained serves the purpose of assessing the calcium
sulfate scaling potential for a wide variety of complex aqueous industrial streams.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction are dealing with the formation of CaSO4 hydrates as undesirable


byproducts, mostly as scale. As the scale layer becomes increasingly
The knowledge of the phase equilibria and solubility of inorganic thicker, it reduces the production capacity and process efficiency
salts in electrolyte solutions is essential in the development, design, because of increased heat transfer resistance, reduction of material
optimization and operation of many chemical processes including flow and operating volumes, corrosion and wearing out of construc-
environmental applications (gas treatment, wastewater treat- tion materials.
ment, or chemical disposal), separation processes (crystallization, The stability regions of the CaSO4 hydrates depend on the solu-
extractive distillation and seawater desalination), electrochemical tion conditions. Each crystalline phase can be stable, metastable or
processes (corrosion or electrolysis), energy production sources unstable at certain temperatures and compositions. The transfor-
(scaling in production wells) and hydrometallurgical processes mation of calcium sulfate dihydrate (DH) to anhydrite (AH) results
[1]. in a significant decrease in the solubility level that makes the
Several inorganic salts exist in more than one crystalline form, prediction and control of calcium sulfate formation complicated.
the stability of which depends on the solution conditions in Therefore, understanding the chemistry of CaSO4 phase equilibria
terms of temperature or composition. Calcium sulfate is one of and being able to estimate its scaling potential in industrial pro-
the most common salts in this category, occurring as three dif- cesses involving electrolytes is of great theoretical significance and
ferent hydrates: dihydrate (DH: CaSO4 ·2H2 O), hemihydrate (HH: practical importance [5,7].
CaSO4 ·0.5H2 O) and anhydrite (AH: CaSO4 ). The purpose of this work is to use a recently developed database
Many industrial processes including wastewater treatment, for the Mixed Solvent Electrolyte (MSE) model of the OLI Systems
desalination, sulfur dioxide removal from the exhaust gas of coal- software [8,9] and map the effect of temperature and acidity as
driven power plants [2,3] and hydrometallurgical processes [4–6] well as sulfate and chloride concentrations on the phase equilibria
of CaSO4 hydrates and provide practical guidelines for solubility
prediction of the thermodynamically stable phase under a partic-
∗ Corresponding author. ular solution composition and temperature. To this end, a number
E-mail address: vladimiros.papangelakis@utoronto.ca (V.G. Papangelakis). of experiments were conducted in electrolyte solutions containing

0378-3812/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.fluid.2009.09.023
G. Azimi, V.G. Papangelakis / Fluid Phase Equilibria 290 (2010) 88–94 89

2.3. Procedure

Solutions of known composition were placed in the reaction ves-


sels with an excess of saturating solid phase. Experiments were
started by heating the charged reactors to temperature under vig-
orous agitation and allowing sufficient time to reach equilibrium.
Samples were withdrawn after equilibration, immediately filtered,
and diluted by 5% HNO3 and kept in test tubes at room temperature.
In the systems studied the solubility of calcium sulfate decreases
with decreasing temperature. Therefore, samples withdrawn at
higher temperatures, e.g., 90 ◦ C, will form dihydrate (DH) precip-
itate upon cooling to room temperature. In order to avoid this
phenomenon, samples were diluted by 5% nitric acid.
In each sampling step, 5 mL of the solution was withdrawn
which does not affect the molarity of different elements in the
system. The concentration of other elements than Ca was also mon-
itored through out all experiments to confirm that the solution
compositions remained unchanged. During solid samples with-
drawal, only ∼1 g of the solid was taken, and since dihydrate was
added in excess (∼50 g) initially, solid sampling would not have a
significant effect on the overall composition of the solid phase.
The Ca concentration was determined by Inductively Coupled
Plasma (ICP-OES) analysis. The density of all liquid samples was
measured at temperature using a portable density meter (DMA
Fig. 1. Scheme of the glass reactor used in this work.
35N ) from Anton Paar. Samples of the equilibrating solid phase were
withdrawn at various temperatures for X-ray diffraction, thermo
gravimetric (TGA) and scanning electron microscope (SEM) analy-
NiSO4 , H2 SO4 , MgSO4 , Fe2 (SO4 )3 , LiCl and HCl. The specific objec-
sis. They were filtered, and washed with alcohol and dried at below
tives are:
40 ◦ C in an oven under vacuum.

• To investigate whether the type of cation has a significant effect


on the solubility of calcium sulfate or all cations in the solution 2.4. Equilibration time
can be substituted with an arbitrary pseudo cation.
• To study the effect of acid addition on the solubility of CaSO4 Equilibration time in solubility measurements varies from sev-
hydrates and to investigate the model performance in predict- eral hours to several days depending on the dissolution rate of the
ing the chemistry of calcium sulfate phase equilibria in acidic solid phase under applied conditions. In this work, several kinetic
solutions. tests were conducted at various temperatures and the results
• To study the effect of the phase transformation on the solubility showed that around 24 h was necessary to achieve saturation.
of calcium sulfate. Reproducibility tests showed that the experimentally measured
data are accurate to within ±5%.

2. Experimental
3. Theory
2.1. Reagents
The following reactions govern the solubility of CaSO4 hydrates:
In this work, all solutions were prepared by dissolving reagent
grade chemicals directly without further purification. Both CaSO4 CaSO4 ·nH2 O(s) = Ca2+ + SO4 2− + nH2 O (1)
solids phases (DH and AH) used in this work were from J.T. Baker. X-
ray powder diffraction was carried out on both solids using a Philips Ca2+ + SO4 2− = CaSO4(aq) o (2)
PW3719 diffractometer. Result showed 100% dihydrate (DH) and
The solubility of calcium sulfate is:
anhydrite (AH), respectively. No traces of hemihydrate (HH) or
anhydrite (AH) were found in the dihydrate (DH) solid powder.
[Ca]total = [Ca2+ ] + [CaSO4(aq) o ] (3)

2.2. Apparatus To obtain the equilibrium constants of reactions (1) and (2) at tem-
perature T and pressure P, the standard state chemical potentials
The experiments were performed inside 1 L double-layer glass of the products and reactants are required. These data are widely
reactors where heating was provided through a circulating oil available in standard thermodynamic compilations. To extrapolate
jacket. Temperature was controlled within ±1 ◦ C of the set-point. the standard state thermodynamic properties to high temperatures
The reactor slurry was kept suspended by a shaft stirrer. Sam- and pressures, the HKF model, developed by Tanger and Helge-
ples were withdrawn through a dip rubber tube using preheated son [10], embedded in the OLI software was employed. To account
syringes, and filtrations were performed using 0.22 ␮m PTFE for the non-ideality (excess properties) of the electrolyte solutions
syringe filters from Fisher. Fig. 1 represents a scheme of the studied, the MSE activity coefficient model [11,12], which is also
experimental set-up used in this work. To avoid solution evapo- embedded in the OLI Systems software, was used. In the end, a self-
ration during the runs, the stirrer bushing was fully sealed using consistent and calibrated model was produced. More details on the
Dow Corning® high vacuum grease, which basically contains poly- modeling methodology and the database developed are available
dimethylsiloxane. elsewhere [8,9,13].
90 G. Azimi, V.G. Papangelakis / Fluid Phase Equilibria 290 (2010) 88–94

Fig. 2. CaSO4 solubility as a function of MSO4 (M = Ni, Mg, Mn) concentration; MgSO4
experimental data are from [14–17]; NiSO4 experimental data are from [17,18];
MnSO4 experimental data are from [17,19]; curves are model predictions. Fig. 3. CaSO4 solubility in CaSO4 –MgSO4 –HCl (0.5 M)–H2 O solution; experimental
data are from this work, and curves are the model predictions.

4. Results and discussion


In industrial applications, particularly in hydrometallurgy, solu-
4.1. Effect of divalent cations on the solubility of CaSO4 hydrates tions sometimes contain several cations for which there are no
experimental data available. By knowing that cation type does not
Fig. 2 presents a comparison between the solubility of have a significant effect on the solubility level, all divalent cations
CaSO4 dihydrate (DH) in ternary solutions of CaSO4 –NiSO4 –H2 O, can be substituted with a certain cation for which experimental
CaSO4 –MgSO4 –H2 O and CaSO4 –MnSO4 –H2 O along with the model data are available during the simulation of the process.
“predictions” at two different temperatures: 25 and 75 ◦ C. It is
clear that the difference between CaSO4 solubilities in these sys- 4.2. Effect of different acids on the solubility of CaSO4 hydrates
tems at both temperatures is always less than ∼15% indicating
that the cation type does not have a dramatic effect on the CaSO4 To compare the effect of H2 SO4 and HCl on the solubility of
solubility. For all three cations, as the metal sulfate concentra- calcium sulfate dihydrate (DH), two different sets of experiment
tion increases from pure water, the solubility initially drops and were performed: first, the solubility was measured as a function of
then increases gradually. After passing a maximum, the solubility MgSO4 concentration in 0.5 M HCl solutions. Second, the solubility
decreases smoothly with further increasing the MSO4 (M = Mg, Mn, was measured as a function of NiSO4 concentration in 0.5 M H2 SO4
Ni) concentration. The initial drop is due to the common ion effect solutions. The measured data are presented in Tables 1 and 2.
which shifts the dissolution reaction (Eq. (1)) to the left; the subse- Figs. 3 and 4 show the measured solubility data in comparison
quent increase is attributable to the association of Ca2+ and SO4 2− with the model prediction results from 25 to 90 ◦ C. In both sys-
ions and formation of calcium sulfate neutral species (Eq. (2)). The tems, the solubility first decreases with increasing MSO4 (M = Ni,
solubility decrease in concentrated solutions is due to the salting- Mg) concentration due to the common ion effect of the added
out effect due to the decreased number of free water molecules in SO4 2− ions. This effect is nullified by further increasing MSO4 con-
the solution to dissolve calcium sulfate. centration because of the association of Ca2+ and SO4 2− ions and

Table 1
Solubility of CaSO4 dihydrate in 0.5 M HCl solutions at various MgSO4 concentrations.

MgSO4 (mol/L) 25 ◦ C 45 ◦ C 70 ◦ C 90 ◦ C

CaSO4 (mol/L) Density (g/mL) CaSO4 (mol/L) Density (g/mL) CaSO4 (mol/L) Density (g/mL) CaSO4 (mol/L) Density (g/mL)

0.0 0.0810 1.007 0.0940 1.005 0.1300 1.004 0.1560 0.997


0.1 0.0456 1.021 0.0594 1.013 0.0785 1.008 0.0987 0.999
0.5 0.0168 1.057 0.0205 1.047 0.0231 1.038 0.0292 1.030
1.0 0.0110 1.112 0.0139 1.100 0.0153 1.105 0.0188 1.080
1.5 0.0087 1.162 0.0105 1.158 0.0113 1.102 0.0135 1.128

Table 2
Solubility of CaSO4 dihydrate in 0.5 M H2 SO4 solutions at various NiSO4 concentrations.

NiSO4 (mol/L) 25 ◦ C 45 ◦ C 70 ◦ C 90 ◦ C

CaSO4 (mol/L) Density (g/mL) CaSO4 (mol/L) Density (g/mL) CaSO4 (mol/L) Density (g/mL) CaSO4 (mol/L) Density (g/mL)

0.00 0.0190 1.030 0.0258 1.022 0.0403 1.010 0.0545 1.002


0.10 0.0180 1.044 0.0234 1.037 0.0351 1.026 0.0454 1.015
0.25 0.0166 1.060 0.0210 1.056 0.0298 1.044 0.0383 1.029
0.50 0.0166 1.106 0.0196 1.095 0.0266 1.085 0.0329 1.071
1.00 0.0148 1.180 0.0172 1.170 0.0234 1.157 0.0314 1.141
1.50 0.0131 1.238 0.0158 1.230 0.0194 1.218 0.0245 1.202
G. Azimi, V.G. Papangelakis / Fluid Phase Equilibria 290 (2010) 88–94 91

Fig. 5. CaSO4 solubility in CaSO4 –NiSO4 (1.4 M)–H2 SO4 (1 M)–Fe2 (SO4 )3
Fig. 4. CaSO4 solubility in CaSO4 –NiSO4 –H2 SO4 (0.5 M)–H2 O solution; experimental (0.2 M)–LiCl (0.3 M)–H2 O solutions at various retention times and different
data are from this work, and curves are the model predictions. temperatures.

formation of calcium sulfate neutral species. Also, the solubility the equilibrating solid phase was estimated semi-quantitatively by
of dihydrate (DH) is significantly higher in the presence of HCl measuring the peak intensities of the different X-ray diffraction
(Fig. 3) compared to that in H2 SO4 (Fig. 4). This effect is most pro- patterns and was confirmed by thermo gravimetric analysis (TGA).
nounced near zero metal sulfate concentrations. The addition of Hemihydrate was not detected in any of the samples despite a care-
both H2 SO4 and HCl increases the solubility of CaSO4 dihydrate ful search for this phase. The experimentally measured solubility
(DH) due to the formation of bisulfate ions; however, in the case
of H2 SO4 , there is a common ion effect due to the produced SO4 2−
from the second dissociation of H2 SO4 hindering the dissolution
reaction.
In both systems, model “predictions” are in good agreement
with the experimental data. It should be emphasized that no
extra fitting were perfumed in these systems which proves the
predictability of the model in multicomponent systems using
the interaction parameters obtained in binary and ternary sys-
tems, i.e., CaSO4 –H2 O, CaSO4 –MgSO4 –H2 O, CaSO4 –NiSO4 –H2 O,
CaSO4 –HCl–H2 O and CaSO4 –H2 SO4 –H2 O which have been studied
previously [8,9].

4.3. Effect of phase transition on the solubility of CaSO4 hydrates

4.3.1. CaSO4 –NiSO4 –H2 SO4 –Fe2 (SO4 )3 –LiCl–H2 O System


Nickel processing solutions typically contain NiSO4 , H2 SO4 ,
Fe2 (SO4 )3 and small amounts of a chloride salt. In this work, the
solubility of calcium sulfate dihydrate (DH) was measured in a
solution of 1.4 M NiSO4 , 1 M H2 SO4 , 0.2 M Fe2 (SO4 )3 and 0.3 M LiCl.
Solubility measurements were carried out based on heating from
Fig. 6. CaSO4 solubility in CaSO4 –NiSO4 (1.4 M)–H2 SO4 (1 M)–Fe2 (SO4 )3
25 to 90 ◦ C followed by subsequent cooling. At a given tempera- (0.2 M)–LiCl (0.3 M)–H2 O solutions; solid curves are the model predictions,
ture, the relative amount of dihydrate (DH) and anhydrite (AH) in and the dashed line shows the phase transition region.

Table 3
Calcium sulfate concentration, saturated solution density and solid phase composition in 1.4 M NiSO4 , 1 M H2 SO4 , 0.2 M Fe2 (SO4 )3 , 0.3 M LiCl solutions at various temperatures
and retention times.

Time (h) T (◦ C) Density (g/mL) CaSO4 (mol/L) Method Solid phase composition

24 25 1.333 0.0079 Heating 100% DH


48 45 1.318 0.0106 Heating 100% DH
72 60 1.315 0.0136 Heating 100% DH
96 70 1.305 0.0155 Heating 100% DH
120 70 1.305 0.0127 Heating 80% DH + 20% AH
144 70 1.305 0.0089 Heating 100% AH
168 90 1.298 0.0085 Heating 100% AH
192 80 1.300 0.0084 Cooling 100% AH
216 60 1.315 0.0082 Cooling 100% AH
240 35 1.326 0.0082 Cooling 100% AH
264 25 1.333 0.0080 Cooling 25% DH + 75% AH
92 G. Azimi, V.G. Papangelakis / Fluid Phase Equilibria 290 (2010) 88–94

Fig. 8. CaSO4 solubility vs. temperature in CaSO4 –MnSO4 –H2 SO4 –H2 O solutions;
the experimental data are from [20]; the solid curves are the predicted values and
the dashed lines show the phase transition region.

Powder X-ray diffraction analysis of the equilibrating solid


phase showed only dihydrate (DH) during heating up to 70 ◦ C;
however, the solid sample withdrawn at 70 ◦ C after 3 days reten-
tion time and the one withdrawn at 90 ◦ C showed 100% anhydrite
(AH) in the composition. The solid sample withdrawn at 70 ◦ C
after 2 days retention time contained a mixture of dihydrate
(DH) and anhydrite (AH). On cooling, anhydrite remained the
dominant equilibrating solid phase above 25 ◦ C, but at 25 ◦ C
the conversion of anhydrite (AH) to dihydrate (DH) occurred.
Thus, the major drop in the solubility of calcium sulfate is due
to the transformation of dihydrate (DH) into anhydrite (AH), a
phase which has a significantly different solubility–temperature
relationship.
Fig. 6 shows CaSO4 solubility data as a function of temperature
along with the model predictions. As can be seen, the model pre-
dicts the solubility of dihydrate and anhydrite precisely. However,
the phase transition area, which is not thermodynamically stable,
cannot be predicted by the model and is marked by a dashed line
in the figure.
The SEM images of the solid samples withdrawn at 70 ◦ C after 24,
48 and 72 h retention times are presented in Fig. 7(a)–(c), respec-
tively. As mentioned above, the first solid sample presented in
Fig. 7(a) shows 100% dihydrate in composition, whereas, the sam-
ple shown in Fig. 7(b) reveals a mixture of 80% DH and 20% AH,
as estimated from XRD and TGA. Fig. 7(c) presents a solid sample
Fig. 7. SEM images of the solid phases at 70 ◦ C showing the different crystal mor- with 100% AH after complete transformation of dihydrate in the
phologies of (a) 100% dihydrate (DH) withdrawn after 24 h retention time (b) system. As is clear, the morphologies of calcium sulfate dihydrate
transitional of 80% dihydrate (DH)–20% anhydrite (AH) mixture withdrawn after
48 h and (c) 100% anhydrite withdrawn after 72 h.
and anhydrite are very different: DH crystals are in a monoclinic
form, whereas AH crystals have an orthorhombic structure and are
needle shaped.
data, saturated solution densities and solid phase compositions at
various temperatures are summarized in Table 3. 4.3.2. CaSO4 –H2 SO4 –MnSO4 –H2 O system
Fig. 5 presents the solubility data as a function of the retention There are some difficulties involved in the electrolytic process
time along with solid sample compositions at different tempera- for the winning of manganese dioxide from H2 SO4 /Mn2+ elec-
tures. As is shown, the solubility increases with temperature up to trolytes where CaSO4 scale develops in heat exchangers, pipes
70 ◦ C. The system was kept at 70 ◦ C for 3 days and slurries were and other equipment. These processes require regular cleaning to
periodically sampled. The results showed a gradual decrease in the maintain efficient process operation. Farrah et al. [20] studied the
solubility in the second and the third samples at 70 ◦ C withdrawn effect of H2 SO4 , MnSO4 and temperature on the solubility of CaSO4
on the day 2 and day 3, respectively. The solubility values measured hydrates over the temperature range of 30–105 ◦ C.
at 90 ◦ C had the same order of magnitude as the third sample with- Fig. 8 shows their experimental solubility data along with the
drawn at 70 ◦ C after 3 days. On cooling, a relatively flat solubility predicted results obtained from the new model, which are in very
curve is noted up to 25 ◦ C. good agreement. It is obvious from the figure that dihydrate (DH)
G. Azimi, V.G. Papangelakis / Fluid Phase Equilibria 290 (2010) 88–94 93

solubility decreases with an increase in MnSO4 concentration at solubility is significantly higher in the presence of HCl compared to
fixed H2 SO4 concentrations due to the common ion effect. Also, that in H2 SO4 because in the case of H2 SO4 , there is a common ion
there is a systematic increase in the solubility with temperature effect due to the produced SO4 2− from the second dissociation of
in the presence of MnSO4 up to 80 ◦ C because of the tendency of H2 SO4 hindering the dissolution reaction.
the sulfate ions to form bisulfate. Above 80 ◦ C, calcium sulfate sol- Furthermore, the effect of the phase transformation of calcium
ubility declines sharply as a result of dihydrate (DH) to anhydrite sulfate dihydrate (DH) into anhydrite (AH) on the solubility was
(AH) transformation. It can be seen from the figure that the sol- studied. The transformation takes place at higher temperatures
ubility of anhydrite is also depressed by subsequent addition of and acid concentrations and has a complex effect on the solubil-
MnSO4 . ity. In this work, complete transformation was achieved at 70 ◦ C
after 3 days retention time in the presence of 1 M H2 SO4 . The solu-
5. Conclusions bility of anhydrite is much lower than that of dihydrate (∼50–80%
lower); therefore process solutions which are saturated with dihy-
Many industries are dealing with calcium sulfate scale for- drate (DH) at ambient temperature and are recycled to autoclaves,
mation and need regular removal of precipitated calcium sulfate operating at temperatures above 100 ◦ C, need to be processed to
hydrates. At elevated temperatures, the transformation between decrease their calcium content and make it less than anhydrite sat-
the calcium sulfate phases has a complex effect on the solubility, uration level inside an autoclave. Otherwise, scale formation will
making the behaviour of calcium sulfate difficult to predict and be unavoidable.
control. Chemical modeling is a practical asset to map calcium sulfate
In this work, a recently developed database for the MSE model chemistry in electrolyte solutions over wide ranges of temperature
of the OLI Systems software was used to study the effect of dif- and concentration. This, in turn, results in gaining comprehensive
ferent electrolytes on the CaSO4 solubility. It was found out that insight for process improvements and optimization in various pro-
it is mainly the anion which controls the solubility of calcium sul- cesses.
fate, while cations have a minor effect. The solubility behaviour
of CaSO4 is the same in all metal sulfate electrolytes with differ- Acknowledgements
ent cations: as the metal sulfate concentration increases from pure
water, the solubility initially drops due to the common ion effect The authors would like to acknowledge the financial support
and then increases gradually due to the formation of calcium sul- provided by Anglo American plc., Barrick Gold Corporation, OLI
fate neutral species. After passing a maximum, solubility decreases Systems Inc., Sherritt International Corporation, Vale Inco Ltd., the
smoothly with a further increase in the concentration due to the Ontario Graduate Scholarship (OGS) and the Natural Sciences and
salting-out effect and decreased number of free water molecules in Engineering Research Council of Canada (NSERC) for this project.
the solution.
The solubility of CaSO4 dihydrate was measured in solutions Appendix A. MSE middle-range interaction parameters
containing NiSO4 , H2 SO4 , MgSO4 and HCl. It was found that the (OLI-version 8.1.3)

Species i Species j BMD0 BMD1 BMD2 CMD0 CMD1 CMD2 Temperature


range (◦ C)

MnSO4 –H2 O Mn2+ SO4 2− −716.157 0.9059 93031.7 255.511 – – 0–180


CaSO4 –H2 O Ca2+ SO4 2− 10887.73 −16.973 −1770400 −15416.42 24.215 2508590 0–400

Ca2+ Mn2+ 683.490 −2.0005 – −868.843 2.5173 –


CaSO4 –MnSO4 –H2 O 25–100
CaSO4(aq) Mn2+ 2134.468 −2.8306 −394938 – – –

Ca2+
Mg 2+
−2602.165 4.8761 333009 3630.73 −6.8234 −462726
CaSO4 –MgSO4 –H2 O 25–175
CaSO4(aq) Mg2+ 678.644 −1.529 −62232.5 −562.19 1.7298 –

CaSO4 –Na2 SO4 –H2 O Ca2+ Na+ 25.171 −0.0262 – −33.189 0.0210 – 25–300

Ca2+
HSO4 − 3715.460 −6.0144 −618224 −4472.033 7.3066 733093
CaSO4 –H2 SO4 –H2 O 25–300
CaSO4(aq) HSO4 − 393.149 −1.6809 11323.61 −671.232 2.5821 –

CaSO4 –CaCl2 –H2 O/CaSO4 – Cl− SO4 2− 465.363 −0.8069 −62105.4 −511.001 0.9236 64179.32
HCl–H2 O/CaSO4 –NaCl– Cl− HSO4 − −148.074 0.3573 – 176.162 −0.4188 – 22–300
H2 O/CaSO4 –MgCl2 –H2 O Cl− CaSO4(aq) 9.4550 0.0367 – −55.0627 0.02547 –

Ca2+ Ni2+ 269.924 −0.7478 97192 −1160.872 1.9865 –


CaSO4 –NiSO4 –H2 O 25–175
CaSO4(aq) Ni2+ −656.530 1.5199 – 863.322 −1.8770 –

CaSO4 –Fe2 (SO4 )3 – Ca2+ Fe3+ −172.067 −9.8085 955584 313.487 13.8038 −1362390
25–90
ZnSO4 –H2 SO4 –H2 O CaSO4(aq) Fe3+ 663.006 −1.1448 −104326 1099.014 −1.1435 −202962
94 G. Azimi, V.G. Papangelakis / Fluid Phase Equilibria 290 (2010) 88–94

References [10] I.V. Tanger, H.C. Helgeson, Am. J. Sci. 288 (1988) 19–98.
[11] P. Wang, A. Anderko, R.D. Young, Fluid Phase Equilib. 203 (1–2) (2002) 141–
[1] A. Anderko, P. Wang, M. Rafal, Fluid Phase Equilib. 194–197 (2002) 123–142. 176.
[2] H. Dathe, A. Jentys, P. Haider, E. Schreier, R. Fricke, J.A. Lercher, Phys. Chem. [12] P. Wang, R.D. Springer, A. Anderko, R.D. Young, Fluid Phase Equilib. 222–223
Chem. Phys. 8 (13) (2006) 1601–1613. (2004) 11–17.
[3] K. Lee, B. Teong, M.J. Subhash, R. Abdul, Energy Sources 28 (13) (2006) [13] H. Liu, V.G. Papangelakis, Ind. Eng. Chem. Res. 45 (2006) 39–47.
1241–1249. [14] K.K. Tanji, Environ. Sci. Technol. 3 (7) (1969) 656–661.
[4] J.E. Dutrizac, Hydrometallurgy 65 (2–3) (2002) 109–135. [15] A. Arslan, G.R. Dutt, Soil Sci. 155 (1) (1993) 37–47.
[5] J.E. Dutrizac, A. Kuiper, Hydrometallurgy 82 (2006) 13–31. [16] Y. Umetsu, B.K. Mutalala, K. Tozawa, J. Mining Metall. Jpn. 6 (1989) 13–22.
[6] J.E. Dutrizac, A. Kuiper, Hydrometallurgy 92 (2008) 54–68. [17] G. Wollmann, W. Voigt, J. Chem. Eng. Data 53 (6) (2008) 1375–1380.
[7] G.H. Nancollas, M.M. Reddy, F. Tsai, J. Cryst. Growth 20 (1973) 125–134. [18] A.N. Campbell, N.S. Yanick, Trans. Faraday Soc. 28 (1932) 657–661.
[8] G. Azimi, V.G. Papangelakis, J.E. Dutrizac, Fluid Phase Equilib. 260 (2) (2007) [19] B.I. Zhelnin, G.I. Gorshtein, L.K. Bezprozvannaya, J. Appl. Chem. USSR 46 (3)
300–315. (1973) 534–537.
[9] G. Azimi, V.G. Papangelakis, J.E. Dutrizac, Fluid Phase Equilib. 266 (1–2) (2008) [20] H.E. Farrah, G.A. Lawrance, E.J. Wanless, Hydrometallurgy 86 (2007) 13–21.
172–186.

You might also like