Comparison among Pitzer-type Models for the Osmotic and Activity

Aug 2, 2011 - Facultad de Ciencias Bбsicas Ingenierнa y Tecnologнa, Universidad Autуnoma de Tlaxcala, Apizaco, Tlax. C.P. 90300, Mйxico. G. A. Ig...
0 downloads 0 Views 916KB Size
CORRELATION pubs.acs.org/IECR

Comparison among Pitzer-type Models for the Osmotic and Activity Coefficients of Strong Electrolyte Solutions at 298.15 K F. Perez-Villase~nor* and S. Carro-Sanchez Facultad de Ciencias Basicas Ingeniería y Tecnología, Universidad Autonoma de Tlaxcala, Apizaco, Tlax. C.P. 90300, Mexico

G. A. Iglesias-Silva Departamento de Ingeniería Química, Instituto Tecnologico de Celaya, Celaya, Gto. C.P. 38010, Mexico

bS Supporting Information ABSTRACT: In this work three Pitzer-type models were compared: the original one, the Archer model (an extended Pitzer model) with ionic strength dependence in the third virial coefficient, and the PIH model that uses as an adjusting parameter the closest approach constant. The performance of these models in correlating the osmotic and activity coefficient of 1:1, 1:2, 1:3, 1:4, 2:1, 2:2, 3:1, 3:2, and 4:1 aqueous electrolyte solutions was tested using experimental data sets for 245 single electrolyte aqueous solutions. The absolute average percentage deviations were 0.78, 1.18, and 3.56% for the Archer, PIH, and Pitzer models, respectively.

’ INTRODUCTION Modeling and optimization of industrial processes with electrolyte solutions17 require rigorous activity models in order to take into account nonidealities in this kind of systems. Several theories and equations have been developed with this objective.815 Among them, the most employed models are those developed by Pitzer911 and Chen et al.12 Perez-Villase~ nor et al.16,17 proposed a modified Pitzer model (PIH model) considering the closest approach parameter in the DebyeH€uckel term as an adjusting parameter. This modification improves the correlative capacity of the equation even when the second parameter of the virial coefficient BMX is removed. Archer18 proposed an extended Pitzer model (Archer model) adding the ionic strength dependence into the third virial coefficient. With this extension, the Archer model correlates accurately the activity and osmotic coefficients of aqueous solutions of NaBr,18 NaCl,19,20,24,25 MgSO4,21 KCl,22 NaNO3,23 and K2SO4.25 Rard and Wijesinghe26 developed a useful method to convert the parameters of the Archer modification to those of the Pitzer model. However, these parameters are not the optimum values for the representation of the nonideality. In this work, the osmotic and activity coefficients for 245 single electrolytes aqueous solutions were correlated to the Archer model. The performance of this equation was compared with that of the Pitzer and PIH models.

Table 1. Parameters for Three Different Variants of Pitzertype Models parameter

G ¼ f ðIÞ þ 2mM mX ½BMX þ zM mM CMX  wS RT

ð1Þ

This equation is also the basis of all modifications of the Pitzer model. From eq 1, the osmotic coefficient can be obtained from r 2011 American Chemical Society

Archer

PIH

bMX

1.2

1.2

fit to experimental

R1

2.0/1.4b

2.0c/1.4

not required

R2 R3

12a not required

1232Aϕ 0.42.5c

not required not required

βMX(0)

fit to experimental

fit to experimental

fit to experimental

data

data

data

βMX(1)

fit to experimental

βMX(2)

fit to experimental

fit to experimental

data

data not required

data fit to experimental

dataa

not required

data

CMX(0)

fit to experimental data

fit to experimental data

fit to experimental data

CMX(1)

not required

fit to experimental

not required

data a Not required for electrolytes with univalent ions. b For electrolytes without univalent ions. c Most common value reported.

its first derivative with respect to the solvent weight: ϕ1 ¼ 

’ EQUATIONS Pitzer expressed the excess Gibbs energy of a single electrolyte in an aqueous solution as E

Pitzer

þ

1

RT

∑i



∂Gex mi ∂wS



¼ jzM zX jf ϕ þ P, T, ni

2ðvM vX Þ3=2 ϕ mMX 2 CMX vMX



 2vM vX ϕ mMX BMX vMX

ð2Þ

Received: December 11, 2010 Accepted: August 2, 2011 Revised: July 28, 2011 Published: August 02, 2011 10894

dx.doi.org/10.1021/ie102466b | Ind. Eng. Chem. Res. 2011, 50, 10894–10901

Industrial & Engineering Chemistry Research

CORRELATION

Table 2. Reported Values for Parameters Used in BMX and CMX Exponential Expansions R1

electrolyte

a

R2

R3

R4a

ref

1.34

53

νM νX 1=νMX ln γ( ¼ jzM zX jf γ MX ¼ ðγM γX Þ



 2νM νX 2ðνM νX Þ3=2 þ mMX 2 CγMX mMX BγMX þ νMX vMX

Ca(NO3)2

2.0

CaCl2 Cs2SO4

1.4 2.0

2.5/1.0 2.5

54 64

Cu(NO3)2

2.0

1.5

58

Fe2(SO4)3

2.0

2.5

71

H2SO4

2.0

2.5

60,71

K2SO4

2.0

2.5

25,63

Mg(NO3)2

2.0

MgCl2

2.0

MgSO4 NaBr

2.0/1.4 2.0

2.5/1.0 2.0/1.7

NaCF3SO3

2.0

2.5

38

NaCl

2.0

2.5

20,25,62,65,69

NaNO3

2.0

2.5

23

NdCl3

0.5

1.5

67

(NH4)2SO4

2.0

2.5

57

Rb2SO4

2.0

2.5

64

SrCl2 ZnCl2

2.0 1.1

2.5 2.5

69 70

ZnSO4

1.4

32Aϕ

2.5

2.0 30.65Aϕ

ð3Þ Functions f(I), fϕ, and fγ arising in eqs 13 correspond to the electrostatic interactions expressed by

68

0.4

12

Also, the mean activity coefficient can be calculated from the definition of the activity coefficient:

0.28

f ðIÞ ¼ 

59 60,21 18,56

1.1551/1.1

fϕ ¼ 

ϕ

ð0Þ

ð1Þ

1=2

!

ð6Þ

In general, the second virial coefficient, BMX, and its derivatives are functions of the ionic strength:10,11,21 ð1Þ

ð0Þ

BMX ðIÞ ¼ βMX þ

þ CMX e  R4 l 

2βMX 1=2 ½1  eR1 I ð1 þ R1 I 1=2 Þ R1 2 I

1=2

ð2Þ

þ Table 3. Archer Model Performance at Different Values for R3 ϕ

2βMX 1=2 ½1  eR2 I ð1 þ R2 I 1=2 Þ R2 2 I

ð0Þ

ð1Þ

BMX ðIÞ ¼ βMX þ βMX eR1 I

average error (%) R3 = 2.0

R3 = 2.5

R3 fitted to experimental data

1:1

0.8156

0.7708

0.6053

[1:2 + 2:1]

0.5132

0.6334

0.2862

[1:3 + 3:1]

1.1109

1.4806

1.1972

[1:4 + 4:1]

3.9876

3.5783

2.8897

2:2

0.0884

0.0836

0.0860

3:2

0.5904

1.5911

1.0908

Global

0.7815

0.8949

0.6294

electrolyte data set

ð5Þ

I 1=2 2 þ lnð1 þ bMX I 1=2 Þ 1=2 bMX 1 þ bMX I

f ¼ Aϕ

61,66,70 ð2Þ

ð4Þ

Aϕ I 1=2 1 þ bMX I 1=2

γ

Parameter required when an expanded form of CϕMX is used: CMX ¼ 2jzM zX j1=2 ½CMX þ CMX e  R3 l

4Aϕ I lnð1 þ bMX I 1=2 Þ bMX

BγMX ðIÞ

1=2

ð2Þ

þ βMX eR2 I

ð7Þ

1=2

ð8Þ

" # ð1Þ 2βMX e R1 I 1=2 2 1=2 1 ¼ þ ð1 þ R1 I  R1 IÞ R1 2 I 2 2 3  R2 I1=2 ð2Þ 2βMX 4 e ð1 þ R2 I 1=2  R2 2 IÞ5 þ 2 1 2 R2 I ð0Þ 2βMX

ð9Þ

Table 4. Parameters of the Archer Model for Electrolytes without Monovalent Ions electrolyte

β(0) MX

σ(β(0) MX)

β(1) MX

σ(β(1) MX)

β(2) MX

σ(β(2) MX)

C(0) MX

σ(C(0) MX)

C(1) MX

σ(C(1) MX)

2:2 electrolytes BeSO4

0.31939 8.28  103 2.84052 1.49  101 58.2017 3.34

0.00871 1.83  103

MnSO4 NiSO4

0.22299 6.85  10 3.11328 1.28  10 59.4446 3.34 0.19103 7.42  103 3.20830 4.24  102 59.1036 9.40  101

4

UO2SO4

0.31412 3.00  103 1.64522 9.08  102 23.8120 2.15

0.16963 1.53  10

2

0.00136 5.40  104 0.00371 2.10  103 1

CuSO4

2

2.63361 2.53  10

3

3

45.1019 9.21  10

1

1

ZnSO4

0.17429 8.19  10

2.84574 1.23  10

54.7203 2.93

Al2(SO4)3

0.61964 9.19  103 10.45783 3.35  101 106.5028 1.10  101

0.00871 1.21  103

0.00306 4.41  10 0.38506 1.69  101 0.00757 6.40  104 0.50026 1.03  101 0.00401 1.42  104

0.22768 1.16  101

4

0.21342 1.88  101

0.00879 5.77  10

3:2 electrolytes Cr2(SO4)3

2

0.58651 1.77  10

[Co(Ethylene-diamine)3]2(SO4)3 0.367358 0.02184 fixed parameters

b = 1.2

1

8.90254 6.36  10

53.5839 2.10  10

4.076823 0.530765

31.81012 5.656867 R1 = 1.4 10895

1

0.00217 7.80  104 0.74607 3.98  101 0.00275 1.36  103

1.88548 7.02  101

0.00672 0.00111

5.643153 0.9625531

R2 = 12

R3 = 2.0

dx.doi.org/10.1021/ie102466b |Ind. Eng. Chem. Res. 2011, 50, 10894–10901

Industrial & Engineering Chemistry Research

CORRELATION

(0) Figure 1. Relationship of β(1) MX and βMX for electrolytes using the Pitzer model.

(0) Figure 2. Relationship of β(1) MX and βMX for electrolytes using the Archer model.

(

Similarly, the third virial coefficient11,21 is calculated as ð0Þ

CMX ðIÞ ¼ CMX þ

ð1Þ

4CMX 1=2 ½6  eR3 I ð6 þ 6R3 I 1=2 R3 4 I 2

þ 3R3 2 I þ R3 3 I 3=2 Þ ϕ

CγMX ðIÞ

ð0Þ

ð1Þ

1=2

ð11Þ

ð0Þ

 4CMX h 1=2 6  eR3 I 6 þ 6R3 I 1=2 4 2 R3 I ð1Þ

CMX þ

þ 3R3 2 I þ R3 3 I 3=2 

ð10Þ

CMX ðIÞ ¼ 2jzM zX j1=2 ½CMX þ CMX eR3 I 

¼ 3jzM zX j

1=2

R3 4 I 2 2

 ð12Þ

Table 1 shows the fitting and fixed parameters for the models used in this work. 10896

dx.doi.org/10.1021/ie102466b |Ind. Eng. Chem. Res. 2011, 50, 10894–10901

Industrial & Engineering Chemistry Research

CORRELATION

(0) Figure 3. Relationship of C(1) MX and CMX for electrolytes using the Archer model.

Figure 4. Osmotic (a) and activity (b) coefficient plots of 1:1 electrolytes comparing the Archer model . b, HNO3 ; 2, NH 4NO 3 ; 9, KOH.

Figure 5. Residuals for the osmotic and activity coefficients of electrolytes with monovalent ions using the Archer, PIH, and Pitzer models. 10897

dx.doi.org/10.1021/ie102466b |Ind. Eng. Chem. Res. 2011, 50, 10894–10901

Industrial & Engineering Chemistry Research

CORRELATION

Table 5. Average Percentage Error for Pitzer Model and its Variants model

ϕ percent error

γ percent error

Archer

0.6951

0.8650

0.7800

Pitzer PIH

3.0564 0.8686

4.1483 1.3113

3.6024 1.0899

average percent error

electrolytes with monovalent ions

electrolytes without monovalent ions Archer

1.0853

0.5993

0.8423

Pitzer

1.5828

0.5891

1.0859

PIH

5.4707

8.0478

6.7592

Archer

0.7042

0.8588

0.7815

Pitzer

3.0318

4.0889

3.5604

PIH

0.9454

1.4237

1.1846

global

the Pitzer model (see Table 1). For the parameter R 3 , three different cases were considered: to fix the parameter to a value of 2.5, to fix it to a value of 2.0 and, finally, to consider it as a fitting parameter. To quantify the model performance, the absolute average percent deviation was used erroravg

Figure 6. Residuals for the osmotic and activity coefficients of 2:2 and 3:2 electrolytes using the Archer, PIH, and Pitzer models.

’ RESULTS Experimental sources and molality ranges for each aqueous electrolyte solution are shown in Table 2s of the Supporting Information. Adjusting parameters of the Archer model have been estimated using a nonlinear least-squares method in GREG27 subroutine by minimizing the objective function: SðyÞ ¼

n

cal 2 ∑ ðyobs i  yi Þ i¼1

ð13Þ

In eq 13, yi is either the osmotic coefficient, the mean ionic activity coefficient, or both. The collection of parameters for Pitzer and PIH models has been reported early16,17,28,29 using the same data sets. When Archer model is used the values of the set of parameters R 1 , R 2 , and R 3 are not the same for each electrolyte solution reported in Table 2, in contrast with Pitzer model, which has well-defined values for R 1 and R 2 (see Table 1). However, the parameter R 1 has the same value than the Pitzer model in 75% of the systems. Although there are few values for the parameter R 2 , it is around 12.0. The parameter R 3 has a value of 2.5 for half of the reported electrolyte solutions. Taking into account these facts, the parameters R 1 and R 2 have been fixed to the values given in

   100 n yobs  ycal i i  ¼    n i ¼ 1  yobs i



ð14Þ

As shown in Table 3, the best results are obtained if R 3 is fitted; however, in order to maintain the number of adjusting parameters for Archer model as small as possible, it is desirable to fix R 3 . In this case, better results were obtained for the systems used in this work when R 3 = 2.0. Therefore, final calculations were performed with R 3 = R 1 = 2.0. Table 5s of the Supporting Information shows the results for electrolyte solutions with monovalent ions, while Table 4 shows the results for 2:2 and 3:2 electrolyte solutions. Figures 1 and 2 show the relationship between β (0) MX and (1) β MX for the Archer and Pitzer models in which the majority of the parameters lie down between 1 and 1. This agrees with the results of Pitzer and Mayorga. 30 As seen in Figure 3, (1) it seems to be a linear relation between C(0) MX and C MX . However, a detailed analysis shows that parameters have values around (0.05, and the distribution is similar for β(0) MX and β(1) MX as shown in Figures 1 and 2. To evaluate the performance of the three models, results previously reported by Perez-Villase~nor 28 and BedollaHernandez 29 for the Pitzer model were considered. The results for the PIH model obtained by Perez-Villase~ nor et al., 16,17 using the same data sets given in Table 2s in the Supporting Information, were used. Table 5 shows that the best results are achieved when the Archer model is used. This model correlates the experimental data within an absolute average percentage deviation of 0.7815 while for the PIH and Pitzer models it is within 1.1846 and 3.5604%, respectively. However, for electrolytes without monovalent ions, results for Pitzer and Archer models are very close and both of them are clearly better than PIH model. 10898

dx.doi.org/10.1021/ie102466b |Ind. Eng. Chem. Res. 2011, 50, 10894–10901

Industrial & Engineering Chemistry Research

CORRELATION

In Figure 4a and b, it is shown that the Archer model follows the correct behavior for the osmotic and activity coefficients of HNO3 , KOH, and NH 4 NO 3 aqueous solutions, even at molalities as high as 25. For electrolyte with monovalent ions, residuals for the osmotic and activity coefficients were compared using the three Pitzer-type models. As shown in Figure 5, the residuals fluctuate around (1% for the Archer model, while for the PIH and Pitzer models, they are around (5 and (10%, respectively. For these kinds of systems, the only difference among the Archer and Pitzer models is the exponential term in the third virial coefficient, and it seems to be the reason of the best performance when the Archer model is used. For electrolyte solutions without monovalent ions, the correlative behavior of the Archer model is slightly better than the Pitzer model. Figure 6 shows the percentage deviations for the three models with respect to the experimental measurements. The residuals for the Archer and Pitzer models are mostly within (1% but at low molalities, they are within (5%. In this case, their performance is clearly better than that of the PIH model.

BMX BMXϕ

’ CONCLUSIONS It was shown that the Archer model correlates the experimental osmotic and activity coefficients of strong electrolyte solutions at 298.15 K better than the Pitzer and PIH models. Although the Archer model contains two more characteristic parameters than the Pitzer model, it has been found that by setting up R3 = R1 = 2.0, the set of fixed parameters can be reduced from four (b, R1, R2, R3) to three (b, R1, R2) without changing the performance of the equation. The PIH model contains only three adjusting parameters and compares well with the Archer model for electrolytes with monovalent ions. The absolute average percentage deviations for the Archer, PIH, and Pitzer models are 0.78, 1.18, and 3.56, respectively. However, for electrolytes without monovalent ions, the average error increases up to 6.75% for PIH model compared with 0.84 and 1.08% for the Archer and Pitzer models, respectively.

Greek letters

’ ASSOCIATED CONTENT

b S

Supporting Information. Experimental data sources for 245 single electrolyte solutions (Table 2s). Parameters for the Archer model for electrolytes with monovalent ions (Table 5s). This material is available free of charge via the Internet at http:// pubs.acs.org.

’ AUTHOR INFORMATION Corresponding Author

*Tel.: (52) 241-417-25-44. Fax: (52) 241-417-58-44. E-mail: [email protected].

’ ACKNOWLEDGMENT The authors are grateful to the Universidad Autonoma de Tlaxcala, Instituto Tecnologico de Celaya, and CONACyT for the financial support in the realization of this work. ’ NOTATION Aϕ bMX

DebyeH€uckel coefficient Closest approach parameter in the Debye H€uckel term

Second virial-type coefficient Second virial-type coefficient form for the osmotic coefficient Third virial-type coefficient Third virial-type coefficient form for the osmotic coefficient Excess Gibbs free energy on a molality basis Ionic strength function form of the Debye H€uckel term Ionic strength function form of the Debye H€uckel term used for the osmotic coefficient Ionic strength function form of the Debye H€uckel term used for the activity coefficient Ionic strength (mol/kg of solvent) Molality of species i (mol/solvent kg) Number of moles in solution of ion i Universal gas constant Absolute temperature (K) Solvent weight (kg) Cation valence Anion valence

CMX CMXϕ Gex f(I) fϕ fγ I mi ni R T wS zM zX R1, R2, R3, R4 ν (0) (1) (2) , βMX , βMX βMX (0) (1) , CMX CMX

ϕ ( γMX

Non linear parameters for second and third virial coefficients in Archer and Pitzer Equation Stoichiometric coefficient Adjustable parameters for Pitzer, Archer or PIH models Adjustable parameters for Pitzer, Archer or PIH models Osmotic Coefficient Mean ionic activity coefficient for MX neutral electrolyte

Subscripts

M X MX

Cation Anion Neutral electrolyte

’ REFERENCES (1) Klauda, J. B.; Sandler, S. I. Global Distribution of Methane Hydrate in Ocean Sediment. Energy Fuels 2005, 19, 459–470. (2) Byrne, R.; Laurie, H. Influence of pressure on chemical equilibria in aqueous systems with particular reference to sea water. Pure Appl. Chem. 1999, 71, 871–980. (3) Jorgensen, T. C.; Weatherley, L. R. Ammonia removal from wastewater by ion exchange in the presence of organic contaminants. Water Res. 2003, 37, 1723–1728. (4) Tan, T.; Gan, S. Vapour-Liquid Equilibrium of Water/Ethanol/ 1-Butanol/Salt: Prediction and Experimental Verification. Chem. Eng. Res. Des. 2005, 83, 1361–1371. (5) Roa, V.; Tapia, M. Estimating water activity in systems containing multiple solutes based on solute properties. J. Food Sci. 1998, 63, 559–564. (6) Sheikholeslami, R.; Ng, M. Calcium Sulfate Precipitation in the Presence of Nondominant Calcium Carbonate: Thermodynamics and Kinetics. Ind. Eng. Chem. Res. 2001, 40, 3570–3578. (7) Nehrke, G.; Reichart, G.; Van Cappellen, P.; Meile, C.; Bijma, J. Dependence of calcite growth rate and Sr partitioning on solution stoichiometry: Non-Kossel crystal growth. Geochim. Cosmochim. Acta 2007, 71, 2240–2249. (8) Meissner, H.; Tester, J. Activity Coefficients of Strong Electrolytes in Aqueous Solutions. Ind. Eng. Chem. Process Des. Dev. 1972, 11, 128–133. 10899

dx.doi.org/10.1021/ie102466b |Ind. Eng. Chem. Res. 2011, 50, 10894–10901

Industrial & Engineering Chemistry Research (9) Pitzer, K. S. Thermodynamics of Electrolytes. I. J. Phys. Chem. 1973, 77, 268–277. (10) Pitzer, K. S.; Mayorga, G. Thermodynamics of electrolytes. III. Activity and osmotic coefficients for 2:2 electrolytes. J. Solution Chem. 1974, 3, 539–546. (11) Pitzer, K. S. Theory: Ion Interaction Approach. In Activity Coefficients in Electrolyte Solutions, 2nd ed.; Pitzer, K. S., Ed.; CRC Press: Boca Raton, FL, 1990. (12) Chen, C.; Britt, H.; Boston, J.; Evans, L. Local Composition Model for the Excess Gibbs Energy of Aqueous Electrolyte Systems. AIChE J. 1982, 28, 588–596. (13) Haghtalab, A.; Vera, J. Nonrandom Factor Model for Electrolyte Solutions. AIChE J. 1991, 37, 147–149. (14) Jaretum, A.; Aly, G. New Local Composition Model for Electrolyte Solutions. Fluid Phase Equilib. 1999, 163, 175–193. (15) Zhao, E.; Yu, M.; Sauve, R.; Khoshkbarchi, M. Extension of the Wilson Model to Electrolyte Solutions. Fluid Phase Equilib. 2000, 173, 161–175. (16) Perez-Villase~nor, F.; Iglesias-Silva, G.; Hall, K. Osmotic and Activity Coefficients Using a Modified Pitzer Equation for Strong Electrolytes 1:1 and 1:2 at 298.15 K. Ind. Eng. Chem. Res. 2002, 41, 1031–1037. (17) Perez-Villase~nor, F.; Bedolla-Hernandez, M. L.; Iglesias-Silva, G. A. Osmotic and Activity Coefficients of 1:1, 1:3, 1:4, 2:1, 2:2, 3:1, 3:2 and 4:1 Strong Electrolytes at 298.15 K Using a Modified Pitzer Equation. Ind. Eng. Chem. Res. 2007, 46, 6366–6374. (18) Archer, D. Thermodynamic Properties of the NaBr+H2O System. J. Phys. Chem. Ref. Data 1991, 20, 509–555. (19) Archer, D. Thermodynamic Properties of the NaCl+H2O System. l. Thermodynamic Properties of NaCl(cr). J. Phys. Chem. Ref. Data 1992, 21, 1–21. (20) Archer, D. Thermodynamic Properties of the NaCl+H2O System. II. Thermodynamic Properties of NaCl(aq), NaCl 3 2H2O(cr), and Phase Equilibria. J. Phys. Chem. Ref. Data 1992, 21, 793–829. (21) Archer, D.; Rard, J. Isopiestic Investigation of the Osmotic and Activity Coefficients of Aqueous MgSO4 and the Solubility of MgSO4 3 7H2O(cr) at 298.15 K:Thermodynamic Properties of the MgSO4 + H2O System to 440 K. J. Chem. Eng. Data 1998, 43, 791–806. (22) Archer, D. Thermodynamic Properties of the KCl+H2O System. J. Phys. Chem. Ref. Data 1999, 28, 1–16. (23) Archer, D. Thermodynamic Properties of the NaNO3+H2O System. J. Phys. Chem. Ref. Data 2000, 29, 1141–1156. (24) Archer, D.; Carter, R. Thermodynamic Properties of the NaCl + H2O System. 4. Heat Capacities of H2O and NaCl(aq) in Cold-Stable and Supercooled States. J. Phys. Chem. B. 2000, 104, 8563–8584. (25) Archer, D.; Kirklin, D. Enthalpies of Solution of Sodium Chloride and Potassium Sulfate in Water. Thermodynamic Properties of the Potassium Sulfate + Water System. J. Chem. Eng. Data 2002, 47, 33–46. (26) Rard, J.; Wijesinghe, A. Conversion of parameters between different variants of Pitzer’s ion-interaction model, both with and without ionic strength dependent higher-order terms. J. Chem. Thermodyn. 2003, 35, 439–473. (27) Stewart, W.; Caracotsios, M.; Sorensen, J. GREG; Department of Chemical Engineering, University of Wisconsin—Madison, USA, 1990. (28) Perez-Villase~nor, F. Prediction of Excess Properties of Multicomponent Strong Electrolyte Solutions. Ph.D. Dissertation, Instituto Tecnologico de Celaya, Celaya, Gto. Mexico, 2002 (in Spanish). (29) Bedolla-Hernandez, M. L. Osmotic and Activity Coefficients of Strong Electrolytes 1:1, 1:3, 1:4, 2:1, 2:2, 3:1, 3:2 and 4:1 at 298.15 K using a Modified Pitzer Equation. Master Thesis, Universidad Aut onoma de Tlaxcala, Mexico, 2007 (in Spanish). (30) Pitzer, K.; Mayorga, G. Thermodynamics of electrolytes. II. Activity and osmotic coefficients for strong electrolytes with one or both ions univalent. J. Phys. Chem. 1973, 77, 2300–2308. (31) Bonner, D. Osmotic and Activity Coefficients of MethylSubstituted Ammonium Perchlorates at 298.15 K. J. Chem. Eng. Data 1982, 27, 62–64.

CORRELATION

(32) Lindenbaum, S.; Boyd, G. Osmotic and Activity Coefficients for the Symmetrical Tetraalkyl Ammonium Halides in Aqueous Solutions at 25 °C. J. Phys. Chem. 1964, 68, 911–917. (33) Wen, W.; Saito, S.; Lee, C. Activity and Osmotic Coefficients of four Symmetrical Tetraalkylammonium Fluorides in Aqueous Solutions at 25 °C. J. Phys. Chem. 1966, 70, 1244–1248. (34) Bonner, D. Osmotic and Activity Coefficients of MethylSubstituted Ammonium Nitrates at 298.15 K. J. Chem. Eng. Data 1981, 26, 148–149. (35) Bonner, D. Osmotic and Activity Coefficients of Some Tetramethylguanidinium Salts at 298.15 K. J. Chem. Eng. Data 1979, 24, 211–212. (36) Hamer, W.; Wu, Y. Osmotic Coefficients and Mean Activity Coefficients of Uni-univalet Electrolytes. J. Phys. Chem. Ref. Data 1972, 1, 1047. (37) Bonner, D. Osmotic and Activity Coefficients of Lithium Chlorate and Lithium Bromate. J. Chem. Eng. Data 1979, 24, 210–211. (38) Rard, J.; Palmer, D.; Albright, J. Isopiestic Determination of the Osmotic and Activity Coefficients of Aqueous Sodium Trifluoromethanesulfonate at 298.15 and 323.15 K, and Representation with an Extended Ion-Interaction (Pitzer) Model. J. Chem. Eng. Data 2003, 48, 158–166. (39) Goldberg, R. Evaluated Activity and Osmotic Coefficients for Aqueous Solutions: Thirty-six Uni-bivalent Electrolytes. J. Phys. Chem. Ref. Data 1981, 10, 671. (40) Robinson, R. Stokes, R. Electrolyte Solutions, 2nd ed.; Butterworths: London, 1965. (41) Wynveen, R.; Dye, J.; Brubaker, C. Activity Coefficient and Conductivity Measurements of High-charge (31, 13, 32) Electrolytes. I. J. Am. Chem. Soc. 1960, 82, 4441–4445. (42) Goves, K.; Dye, J.; Brubaker, C. Activity Coefficients and Conductances of High-charge (41, 14, 12) Electrolytes. II. J. Am. Chem. Soc. 1960, 82, 4445–4448. (43) Bonner, D.; Kim, S. The Osmotic and Activity Coefficients of Some Dibolaform Electrolytes. J. Phys. Chem. 1969, 73, 1367–1370. (44) Goldberg, R. N. Evaluated Activity and Osmotic Coefficient for Aqueous Solutions: Bi-Univalent Compounds of Zinc, Cadmium, and Ethylene Bis(Trimethylammonium) Chloride and Iodide. J. Phys. Chem. Ref. Data 1981, 10, 1–52. (45) Libus, Z.; Sadowska, T. Coordination and Association Equilibria in Aqueous Electrolyte Solutions. I. Osmotic and Activity Coefficients of Divalent Metal Perchlorates. J. Phys. Chem. 1969, 73, 3229–3236. (46) Bonner, D.; Carey, R.; Torres, A. The Osmotic and Activity Coefficients of Some Bolaform Electrolytes. J. Phys. Chem. 1968, 72, 4290–4295. (47) Rard, J.; Weber, H.; Spedding, F. Isopiestic Determination of the Activity Coefficients of Some Aqueous Rare Earth Electrolyte Solutions at 25 °C. 2. The Rare Earth Perchlorates. J. Chem. Eng. Data 1977, 22, 187–201. (48) Rard, J.; Spedding, F. Isopiestic Determination of the Activity Coefficients of Some Aqueous Rare-Earth Electrolyte Solutions at 25 °C. 5. Dy(NO3)3, Ho(NO3)3, and Lu(NO3)3. J. Chem. Eng. Data 1981, 26, 391–395. (49) Spedding, F.; Weber, H.; Saeger, V.; Petheram, H.; Rard, J.; Habenschuss, A. Isopiestic Determination of the Activity Coefficients of Some Aqueous Rare Earth Electrolyte Solutions at 25 °C. 1. The Rare Earth Chlorides. J. Chem. Eng. Data 1976, 21, 341–360. (50) Rard, J.; Shiers, L.; Heiser, D.; Spedding, F. Isopiestic Determination of the Activity Coefficients of Some Aqueous Rare Earth Electrolyte Solutions at 25 °C. 3. The Rare Earth Nitrates. J. Chem. Eng. Data 1977, 22, 337–347. (51) Rard, J.; Spedding, F. Isopiestic Determination of the Activity Coefficients of Some Aqueous Rare-Earth Electrolyte Solutions at 25 °C. 6. Eu(NO3)3, Y(NO3)3, and YCl3. J. Chem. Eng. Data 1982, 27, 454–461. (52) Rard, J.; Miller, D. Isopiestic Determination of the Activity Coefficients of Some Aqueous Rare Earth Electrolyte Solutions at 25 °C. 10900

dx.doi.org/10.1021/ie102466b |Ind. Eng. Chem. Res. 2011, 50, 10894–10901

Industrial & Engineering Chemistry Research 4. La(NO3)3, Pr(NO3)3, and Nd((NO3)3. J. Chem. Eng. Data 1979, 24, 348–354. (53) Macaskill, J.; Robinson, R.; Bates, R. Osmotic Coefficients and Activity Coefficients of Guanidinium Chloride in Concentrated Aqueous Solutions at 25°C. J. Chem. Eng. Data 1977, 22, 411–412. (54) Oakes, C.; Felmy, A.; Sterner, S. Thermodynamic properties of aqueous calcium nitrate {Ca(NO3)2} to the temperature 373 K including new enthalpy of dilution data. J. Chem. Thermodyn. 2000, 32, 29–54. (55) Rard, J.; Clegg, S. Critical Evaluation of the Thermodynamic Properties of Aqueous Calcium Chloride. 1. Osmotic and Activity Coefficients of 010.77 mol.kg1 Aqueous Calcium Chloride Solutions at 298.15 K and Correlation with Extended Pitzer Ion-Interaction Models. J. Chem. Eng. Data 1997, 42, 819–849. (56) Moonis, R. Ally. Solute and Solvent Activities of CaCl2(aq) Solutions from the Adsorption Isotherm Treatment. J. Chem. Eng. Data 1999, 44, 792–797. (57) Rard, J.; Archer, D. Isopiestic Investigation of the Osmotic and Activity Coefficients of Aqueous NaBr and the Solubility of NaBr 3 2H2O(cr) at 298.15 K: Thermodynamic Properties of the NaBr + H20 System over Wide Ranges of Temperature and Pressure. J. Chem. Eng. Data 1995, 40, 170–185. (58) Clegg, S.; Milioto, S.; Palmer, D. Osmotic and Activity Coefficients of Aqueous (NH4)2SO4 as a Function of Temperature, and Aqueous (NH4)2SO4 and H2SO4 Mixtures at 298.15 and 323.15 K. J. Chem. Eng. Data 1996, 41, 455–46. (59) Albright, J.; Rizzo, P.; Rard, J. Isopiestic determination of the osmotic and activity coefficients of Cu(NO3)2(aq) at the temperature 298.15 K. J. Chem. Thermodyn. 1998, 30, 327–352. (60) Wang, P.; Pitzer, K.; Simonson, J. Thermodynamic Properties of Aqueous Magnesium Chloride Solutions from 250 to 600 K and to 100 MPa. J. Phys. Chem. Ref. Data 1998, 27, 971–991. (61) Rard, J.; Clegg, S. Isopiestic determination of the osmotic and activity coefficients of [zH2SO4 + (1-z)MgSO4](aq) at T = 298.15 K. II. Results for z = (0.43040, 0.28758, and 0.14399) and analysis with Pitzer’s model. J. Chem. Thermodyn. 1999, 31, 399–429. (62) Albright, J.; Rard, J.; Serna, S.; Summers, E.; Yang, M. Isopiestic determination of the osmotic and activity coefficients of ZnSO4(aq) at T = 298.15 K, and the standard potential of the electrochemical cell ZnHgx(two phase)|ZnSO4(aq)|PbSO4(s) |PbHgx(two phase). J. Chem. Thermodyn. 2000, 32, 1447–1487. (63) Archer, D.; Carter, R. Thermodynamic Properties of the NaCl + H2O System. 4. Heat Capacities of H2O and NaCl(aq) in Cold-Stable and Supercooled States. J. Phys. Chem. B. 2000, 104, 8563–8584. (64) Palmer, D.; Archer, D.; Rard, J. Isopiestic Determination of the Osmotic and Activity Coefficients of K2SO4(aq) at the Temperatures 298.15 and 323.15 K, and Revision of the Thermodynamic Properties of the K2SO4 + H2O System. J. Chem. Eng. Data 2002, 47, 1425–1431. (65) Palmer, D.; Rard, J.; Clegg, S. Isopiestic determination of the osmotic and activity coefficients of Rb2SO4(aq) and Cs2SO4(aq) at T = (298.15 and 323.15) K, and representation with an extended ioninteraction (Pitzer) model. J. Chem. Thermodyn. 2002, 34, 63–102. (66) Rard, J.; Clegg, S.; Platford, R. Thermodynamics of [zNaCl+ (1-z)Na2SO4](aq) from T=278.15 K to T=318.15 K, and representation with an extended ion-interaction (Pitzer) model. J. Chem. Thermodyn. 2003, 35, 967–1008. (67) Miladinovic, J.; Ninkovic, R.; Todorovic, M.; Jovanovic, V. Correlation of Osmotic Coefficient Data for ZnSO4(aq) at 25°C by Various Thermodynamic Models. J. Solution Chem. 2003, 32, 371–383. (68) Oakes, C.; Rard, J.; Archer, D. Enthalpies of Dilution of NdCl3(aq) at Temperatures from 297.89 to 372.08 K and an Extended Pitzer Ion-Interaction Model for the NdCl3 + H2O System. J. Chem. Eng. Data 2004, 49, 313–32. (69) Rard, J.; Wijesinghe, A.; Wolery, T. Review of the Thermodynamic Properties of Mg(NO3)2(aq) and Their Representation with the Standard and Extended Ion-Interaction (Pitzer) Models at 298.15 K. J. Chem. Eng. Data 2004, 49, 1127–1140. (70) Clegg, S.; Rard, J.; Miller, D. Isopiestic Determination of the Osmotic and Activity Coefficients of NaCl + SrCl2 + H2O at 298.15 K

CORRELATION

and Representation with an Extended Ion-Interaction Model. J. Chem. Eng. Data 2005, 50, 1162–1170. (71) Ninkovic, R.; Miladinovic, J.; Todorovic, M.; Grujic, S.; Rard, J. Osmotic and Activity Coefficients of the [x ZnCl2+(1- x)ZnSO4](aq) System at 298.15 K. J. Solution Chem. 2007, 36, 405–435. (72) Tosca, N.; Smirnov, A.; McLennan, S. Application of the Pitzer ion interaction model to isopiestic data for the Fe2(SO4)3H2SO4H2O system at 298.15 and 323.15 K. Geochim. Cosmochim. Acta 2007, 71, 2680–2698.

10901

dx.doi.org/10.1021/ie102466b |Ind. Eng. Chem. Res. 2011, 50, 10894–10901