Skip to main content
Log in

A Two-Parameter Theoretical Model for Predicting the Activity and Osmotic Coefficients of Aqueous Electrolyte Solutions

  • Published:
Journal of Solution Chemistry Aims and scope Submit manuscript

Abstract

A thermodynamic model of electrolyte solutions is proposed. The model consists of a long-range term expressed by the Pitzer–Debye–Hückel equation (PDH) and a short-range term expressed by the molecular interaction volume model (MIVM). The new model was fitted with 39 different types of single electrolyte systems and was compared with the Pitzer equation, and the mean standard deviation (SD) and the mean average relative deviation (ARD%) are 0.0264, 0.0040 and 2.09%, 0.40%, respectively. Meanwhile, the physical meaning of the two electrolyte-specific interaction parameters (\(B_{ca,s}\) and \(B_{s,ca}\)) of the new model is also discussed. By further comparison with the Pitzer equation and a state-of-the-art model, eUNIQUAC-NRF, the new model exhibits relatively robust extrapolation capability, and also shows the potential ability to predict the activity coefficients of individual ions. In addition, only using binary parameters to predict 29 different types of ternary systems, the overall prediction results of the new model are slightly better than those of the Pitzer equation, and the mean SD and ARD% are 0.0288, 0.0396 and 2.88%, 3.81%, respectively. For some cases involving Rb and Cs, the Pitzer equation needs two ternary adjustable parameters (\(\theta\) and \(\psi\)) to achieve the prediction accuracy of the new model. Furthermore, we also compared the predictions of the new model with the eUNIQUAC-NRF model for several ternary systems; the new model also shows better performance, and its overall prediction accuracy was about twice that of the eUNIQUAC-NRF model, with the average SD and ARD% values being 0.0261, 0.0546 and 2.63%, 5.80%, respectively.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

References

  1. Loehe, J.R., Donohue, M.D.: Recent advances in modeling thermodynamic properties of aqueous strong electrolyte systems. AIChE J. 43, 180–195 (1997)

    Article  CAS  Google Scholar 

  2. Robinson, R.A., Stokes, R.H.: Electrolyte Solutions, 2nd edn. Dover Publications Inc, New York (2002)

    Google Scholar 

  3. Pitzer, K.S.: Thermodynamics of electrolytes. I. Theoretical basis and general equations. J. Phys. Chem. 77, 268–277 (1973)

    Article  CAS  Google Scholar 

  4. Pitzer, K.S., Kim, J.J.: Thermodynamics of electrolytes. IV. Activity and osmotic coefficients for mixed electrolytes. J. Am. Chem. Soc. 96, 5701–5707 (1974)

    Article  CAS  Google Scholar 

  5. Voigt, W.: Chemistry of salts in aqueous solutions: applications, experiments, and theory. Pure Appl. Chem. 83, 1015–1030 (2011)

    Article  CAS  Google Scholar 

  6. May, P.M., Rowland, D.: Thermodynamic modeling of aqueous electrolyte systems: current status. J. Chem. Eng. Data 62, 2481–2495 (2017)

    Article  CAS  Google Scholar 

  7. Renon, H., Prausnitz, J.M.: Local compositions in thermodynamic excess functions for liquid mixtures. AIChE J. 14, 135–144 (1968)

    Article  CAS  Google Scholar 

  8. Abrams, D.S., Prausnitz, J.M.: Statistical thermodynamics of liquid mixtures: a new expression for the excess Gibbs energy of partly or completely miscible systems. AIChE J. 21, 116–128 (1975)

    Article  CAS  Google Scholar 

  9. Chen, C.C., Britt, H.I., Boston, J.F., Evans, L.B.: Local composition model for excess Gibbs energy of electrolyte systems. Part I: Single solvent, single completely dissociated electrolyte systems. AIChE J. 28, 588–596 (1982)

    Article  CAS  Google Scholar 

  10. Chen, C.C., Evans, L.B.: A local composition model for the excess Gibbs energy of aqueous electrolyte systems. AIChE J. 32, 444–454 (1986)

    Article  CAS  Google Scholar 

  11. Walas, S.M.: Phase Equilibria in Chemical Engineering. Butterworth–Heinemann, Oxford (1985)

    Google Scholar 

  12. Thomsen, K., Rasmussen, P., Gani, R.: Correlation and prediction of thermal properties and phase behaviour for a class of aqueous electrolyte systems. Chem. Eng. Sci. 51, 3675–3683 (1996)

    Article  CAS  Google Scholar 

  13. Thomsen, K., Rasmussen, P.: Modeling of vapor−liquid−solid equilibrium in gas−aqueous electrolyte systems. Chem. Eng. Sci. 54, 1787–1802 (1999)

    Article  CAS  Google Scholar 

  14. Thomsen, K.: Modeling electrolyte solutions with the extended universal quasichemical (UNIQUAC) model. Pure Appl. Chem. 77, 531–542 (2005)

    Article  CAS  Google Scholar 

  15. Haghtalab, A., Peyvandi, K.: Electrolyte-UNIQUAC-NRF model for the correlation of the mean activity coefficient of electrolyte solutions. Fluid Phase Equilib. 281, 163–171 (2009)

    Article  CAS  Google Scholar 

  16. Haghtalab, A., Peyvandi, K.: Generalized Electrolyte-UNIQUAC-NRF model for calculation of solubility and vapor pressure of multicomponent electrolytes solutions. J. Mol. Liq. 165, 101–112 (2012)

    Article  CAS  Google Scholar 

  17. Klamt, A.: Conductor-like screening model for real solvents: a new approach to the quantitative calculation of solvation phenomena. J. Phys. Chem. 99, 2224–2235 (1995)

    Article  CAS  Google Scholar 

  18. Klamt, A., Eckert, F., Arlt, W.: COSMO-RS: an alternative to simulation for calculating thermodynamic properties of liquid mixtures. Annu. Rev. Chem. Biomol. Eng. 1, 101–122 (2010)

    Article  CAS  PubMed  Google Scholar 

  19. Eckert, F., Klamt, A.: Fast solvent screening via quantum chemistry: COSMO-RS approach. AIChE J. 48, 369–385 (2002)

    Article  CAS  Google Scholar 

  20. Diedenhofen, M., Klamt, A.: COSMO-RS as a tool for property prediction of IL mixtures: a review. Fluid Phase Equilib. 294, 31–38 (2010)

    Article  CAS  Google Scholar 

  21. Klamt, A.: COSMO-RS for aqueous solvation and interfaces. Fluid Phase Equilib. 407, 152–158 (2016)

    Article  CAS  Google Scholar 

  22. Toure, O., Audonnet, F., Lebert, A., Dussap, C.-G.: COSMO-RS-PDHS: a new predictive model for aqueous electrolytes solutions. Chem. Eng. Res. Des. 92, 2873–2883 (2014)

    Article  CAS  Google Scholar 

  23. Toure, O., Audonnet, F., Lebert, A., Dussap, C.-G.: Development of a thermodynamic model of aqueous solution suited for foods and biological media. Part A: prediction of activity coefficients in aqueous mixtures containing electrolytes. Can. J. Chem. Eng. 93, 443–450 (2015)

    Article  CAS  Google Scholar 

  24. Toure, O., Lebert, A., Dussap, C.-G.: Extension of the COSMO-RS-PDHS model to the prediction of activity coefficients in concentrated water−electrolyte and water−polyol solutions. Fluid Phase Equilib. 424, 90–104 (2016)

    Article  CAS  Google Scholar 

  25. Ingram, T., Gerlach, T., Mehling, T., Smirnova, I.: Extension of COSMO-RS for monoatomic electrolytes: modeling of liquid–liquid equilibria in presence of salts. Fluid Phase Equilib. 2012(314), 29–37 (2012)

    Article  CAS  Google Scholar 

  26. Gerlach, T., Müller, S., Smirnova, I.: Development of a COSMO-RS based model for the calculation of phase equilibria in electrolyte systems. AIChE J. 64, 272–285 (2018)

    Article  CAS  Google Scholar 

  27. Tao, D.P.: A new model of thermodynamics of liquid mixtures and its application to liquid alloys. Thermochim. Acta 363, 105–113 (2000)

    Article  CAS  Google Scholar 

  28. Wilson, G.M.: Vapor−liquid equilibrium. XI. A new expression for the excess free energy of mixing. J. Am. Chem. Soc. 86, 127–130 (1964)

    Article  CAS  Google Scholar 

  29. Prausnitz, J.M., Lichtenthaler, R.N., de Azevedo, E.G.: Molecular Thermodynamics of Fluid-phase Equilibria, 3rd edn. Pearson Education, New York (1998)

    Google Scholar 

  30. Dai, H., Tao, D.P.: A statistical thermodynamic model with strong adaptability for liquid mixtures. Fluid Phase Equilib. 473, 154–165 (2018)

    Article  CAS  Google Scholar 

  31. Tao, D.P.: Prediction of all component activities in iron-based liquid ternary alloys containing phosphorus, titanium, and vanadium. Metall. Mater. Trans. B. 45, 2232–2246 (2014)

    Article  CAS  Google Scholar 

  32. Tao, D.P.: A pseudo-multicomponent approach to important ternary silicate melts. Metall. Mater. Trans. B. 43, 1247–1261 (2012)

    Article  CAS  Google Scholar 

  33. Dai, H., Tao, D.P.: Application of the molecular interaction volume model (MIVM) and its modified form to organic vapor−liquid equilibria. Fluid Phase Equilib. 484, 74–81 (2019)

    Article  CAS  Google Scholar 

  34. Pitzer, K.S.: Electrolytes from dilute solutions to fused salts. J. Am. Chem. Soc. 102, 2902–2906 (1980)

    Article  CAS  Google Scholar 

  35. Pitzer, K.S., Simonson, J.M.: Thermodynamics of multicomponent, miscible, ionic systems: theory and equations. J. Phys. Chem. 90, 3005–3009 (1986)

    Article  CAS  Google Scholar 

  36. Marcus, Y.: Ionic volumes in solution. Biophys. Chem. 124, 200–207 (2006)

    Article  CAS  PubMed  Google Scholar 

  37. Pitzer, K.S.: Activity Coefficients in Electrolyte Solutions, 2nd edn. CRC Press, London (1991)

    Google Scholar 

  38. Kim, H.T., Frederick, W.J.: Evaluation of Pitzer ion interaction parameters of aqueous electrolytes at 25 °C. 1. Single salt parameters. J. Chem. Eng. Data 33, 177–184 (1988)

    Article  CAS  Google Scholar 

  39. Hamer, W.J., Wu, Y.C.: Osmotic coefficients and mean activity coefficients of uni-univalent electrolytes in water at 25 °C. J. Phys. Chem. Ref. Data 1, 1047–1099 (1972)

    Article  CAS  Google Scholar 

  40. Macaskill, J.B., Bates, R.G.: Osmotic coefficients and activity coefficients of aqueous hydrobromic acid solutions at 25 °C. J. Solution Chem. 12, 607–619 (1983)

    Article  CAS  Google Scholar 

  41. Olynyk, P., Gordon, A.R.: The vapor pressure of aqueous solutions of sodium chloride at 20, 25 and 30° for concentrations from 2 molal to saturation. J. Am. Chem. Soc. 65, 224–226 (1943)

    Article  CAS  Google Scholar 

  42. Robinson, R.A.: The vapor pressures of solutions of potassium chloride and sodium chloride. Trans. R. Soc. N. Z. 75, 203–217 (1945)

    CAS  Google Scholar 

  43. Rard, J.A., Archer, D.G.: Isopiestic Investigation of the osmotic and activity coefficients of aqueous NaBr and the solubility of NaBr·2H2O(cr) at 29815 K: thermodynamic properties of the NaBr + H2O system over wide ranges of temperature and pressure. J. Chem. Eng. Data 40, 170–185 (1995)

    Article  CAS  Google Scholar 

  44. Hornibrook, W.J., Janz, G.J., Gordon, A.R.: The thermodynamics of aqueous solutions of potassium chloride at temperatures from 15–45° from Emf measurements on cells with transference. J. Am. Chem. Soc. 64, 513–516 (1942)

    Article  CAS  Google Scholar 

  45. Tien, H.T.: The activity coefficients of rubidium and cesium fluorides in aqueous solution from vapor pressure measurements. J. Phys. Chem. 67, 532–533 (1963)

    Article  CAS  Google Scholar 

  46. Rard, J.A.: Isopiestic determination of the osmotic and activity coefficients of aqueous MnCl2, MnSO4, and RbCl at 25 °C. J. Chem. Eng. Data 29, 443–450 (1984)

    Article  CAS  Google Scholar 

  47. Rard, J.A., Miller, D.G.: Isopiestic determination of the osmotic and activity coefficients of aqueous CsCl, SrCl2, and mixtures of NaCl and CsCl at 25 °C. J. Chem. Eng. Data 27, 169–173 (1982)

    Article  CAS  Google Scholar 

  48. Partanen, J.I.: Mean activity coefficients and osmotic coefficients in aqueous solutions of salts of ammonium ions with univalent anions at 25 °C. J. Chem. Eng. Data 57, 2654–2666 (2012)

    Article  CAS  Google Scholar 

  49. Covington, A.K., Irish, D.E.: Osmotic and activity coefficients of aqueous ammonium bromide solutions at 25 °C. J. Chem. Eng. Data 17, 175–176 (1972)

    Article  CAS  Google Scholar 

  50. Childs, C.W., Downes, C.J., Platford, R.F.: Thermodynamics of aqueous sodium and potassium dihydrogen orthophosphate solutions at 25 °C. Aust. J. Chem. 26, 863–866 (1973)

    Article  CAS  Google Scholar 

  51. Popović, D.Ž., Miladinović, J., Todorović, M.D., Zrilić, M.M., Rard, J.A.: Isopiestic determination of the osmotic and activity coefficients of K2HPO4(aq), including saturated and supersaturated solutions, at T = 298.15 K. J. Solution Chem. 40, 907–920 (2011)

    Article  CAS  Google Scholar 

  52. Rard, J.A., Clegg, S.L., Palmer, D.A.: Isopiestic determination of the osmotic and activity coefficients of Li2SO4(aq) at T = 298.15 and 323.15 K, and representation with an extended ion-interaction (Pitzer) model. J. Solution Chem. 36, 1347–1371 (2007)

    Article  CAS  Google Scholar 

  53. Rard, J.A., Clegg, S.L., Palmer, D.A.: Isopiestic determination of the osmotic coefficients of Na2SO4(aq) at 25 and 50 °C, and representation with ion-interaction (Pitzer) and mole fraction thermodynamic models. J. Solution Chem. 29, 1–49 (2000)

    Article  CAS  Google Scholar 

  54. Palmer, D.A., Archer, D.G., Rard, J.A.: 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 47, 1425–1431 (2002)

    Article  CAS  Google Scholar 

  55. Palmer, D.A., Rard, J.A., Clegg, S.L.: 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 ion-interaction (Pitzer) model. J. Chem. Thermodyn. 34, 63–102 (2002)

    Article  CAS  Google Scholar 

  56. Clegg, S.L., Milioto, S., Palmer, D.A.: Osmotic and activity coefficients of aqueous (NH4)2SO4 as a function of temperature, and aqueous (NH4)2SO4−H2SO4 mixtures at 298.15 K and 323.15 K. J. Chem. Eng. Data 41, 455–467 (1996)

    Article  CAS  Google Scholar 

  57. Rard, J.A., Miller, D.G.: Isopiestic determination of the osmotic and activity coefficients of aqueous magnesium chloride solutions at 25 °C. J. Chem. Eng. Data 26, 38–43 (1981)

    Article  CAS  Google Scholar 

  58. Rard, J.A., Clegg, S.L.: Critical evaluation of the thermodynamic properties of aqueous calcium chloride. 1 Osmotic and activity coefficients of 0–10.77 mol·kg−1 aqueous calcium chloride solutions at 298.15 K and correlation with extended Pitzer ion-interaction models. J. Chem. Eng. Data 42, 819–849 (1997)

    Article  CAS  Google Scholar 

  59. Goldberg, R.N., Nuttall, R.L.: Evaluated activity and osmotic coefficients for aqueous solutions: the alkaline earth metal halides. J. Phys. Chem. Ref. Data 7, 263–310 (1978)

    Article  CAS  Google Scholar 

  60. Rard, J.A., Miller, D.G.: Mutual diffusion coefficients of aqueous MnCl2 and CdCl2, and osmotic coefficients of aqueous CdCl2 at 25 °C. J. Solution Chem. 14, 271–299 (1985)

    Article  CAS  Google Scholar 

  61. Rard, J.A.: Isopiestic investigation of water activities of aqueous NiCl2 and CuCl2 solutions and the thermodynamic solubility product of NiCl2·6H2O at 298.15 K. J. Chem. Eng. Data 37, 433–442 (1992)

    Article  CAS  Google Scholar 

  62. Goldberg, R.N.: Evaluated activity and osmotic coefficients for aqueous solutions: bi-univalent compounds of lead, copper, manganese, and uranium. J. Phys. Chem. Ref. Data 8, 1005–1050 (1979)

    Article  CAS  Google Scholar 

  63. Miladinović, J., Ninković, R., Todorović, M., Jovanović, V.: Osmotic coefficient of the ZnCl2(aq) at T = 298.15 K. J. Chem. Thermodyn. 35, 1073–1082 (2003)

    Article  CAS  Google Scholar 

  64. Rard, J.A., Miller, D.G.: Isopiestic determination of the osmotic and activity coefficients of ZnCl2(aq) at 298.15 K. J. Chem. Thermodyn. 21, 463–482 (1989)

    Article  CAS  Google Scholar 

  65. Archer, D.G., Rard, J.A.: Isopiestic Investigation of the osmotic and activity coefficients of aqueous MgSO4 and the solubility of MgSO4·7H2O(cr) at 298.15 K: thermodynamic properties of the MgSO4+H2O system to 440 K. J. Chem. Eng. Data 43, 791–806 (1998)

    Article  CAS  Google Scholar 

  66. Miller, D.G., Rard, J.A., Eppstein, L.B., Robinson, R.A.: Mutual diffusion coefficients, electrical conductances, osmotic coefficients, and ionic transport coefficientslij for aqueous CuSO4 at 25 °C. J. Solution Chem. 9, 467–496 (1980)

    CAS  Google Scholar 

  67. Albright, J.G., Rard, J.A., Serna, S., Summers, E.E., Yang, M.C.: 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. 32, 1447–1487 (2000)

    Article  CAS  Google Scholar 

  68. Chen, C.C., Mathias, P.M., Orbey, H.: Use of hydration and dissociation chemistries with the electrolyte–NRTL model. AIChE J. 45, 1576–1586 (1999)

    Article  CAS  Google Scholar 

  69. Lu, X.H., Maurer, G.: Model for describing activity coefficients in mixed electrolyte aqueous solutions. AIChE J. 39, 1527–1538 (1993)

    Article  CAS  Google Scholar 

  70. Goldberg, R.N.: Evaluated activity and osmotic coefficients for aqueous solutions: thirty-six uni-bivalent electrolytes. J. Phys. Chem. Ref. Data 10, 671–764 (1981)

    Article  CAS  Google Scholar 

  71. Wilczek-Vera, G., Rodil, E., Vera, J.H.: Towards accurate values of individual ion activities: additional data for NaCl, NaBr and KCl, and new data for NH4Cl. Fluid Phase Equilib. 241, 59–69 (2006)

    Article  CAS  Google Scholar 

  72. Wilczek-Vera, G., Rodil, E., Vera, J.H.: On the activity of ions and the junction potential: revised values for all data. AIChE J. 50, 445–462 (2004)

    Article  CAS  Google Scholar 

  73. Dinane, A., El Guendouzi, M., Mounir, A.: Hygrometric determination of water activities, osmotic and activity coefficients of (NaCl + KCl)(aq) at T = 298.15 K. J. Chem. Thermodyn. 34, 423–441 (2002)

    Article  CAS  Google Scholar 

  74. Robinson, R.A.: The osmotic properties of aqueous sodium chloride−cesium chloride mixtures at 25 °. J. Am. Chem. Soc. 74, 6035–6036 (1952)

    Article  CAS  Google Scholar 

  75. Yan, W.D., Zhang, R., Fu, R.R., Han, S.J.: Measurement and correlation of activity coefficients for the ternary system (sodium bromide + sodium acetate + water) at T = 298.15 K. J. Chem. Thermodyn. 34, 1289–1296 (2002)

    Article  CAS  Google Scholar 

  76. Salamat-Ahangari, R., Saemi Pestebaglu, L., Karimzadeh, Z.: Determination of activity coefficients, osmotic coefficients, and excess Gibbs free energies of KCl and KBr aqueous mixtures at 298.15 K. J. Chem. Eng. Data 63, 290–297 (2018)

    Article  CAS  Google Scholar 

  77. Ghalami-Choobar, B., Moghimi, M., Mahmoodi, N., Mohammadian, M.: Potentiometric determination of the thermodynamic properties for the ternary system (KCl + KNO3 + H2O) at T = 298.15 K. J. Chem. Thermodyn. 42, 454–461 (2010)

    Article  CAS  Google Scholar 

  78. Galleguillos, H.R., Graber, T.A., Taboada, M.E., Hernández-Luis, F.F.: Activity of water in the KI + KNO3 + H2O ternary system at 298.15 K. J. Chem. Eng. Data 48, 851–855 (2003)

    Article  CAS  Google Scholar 

  79. Robinson, R.A.: The osmotic properties of aqueous caesium chloride+ potassium chloride and caesium chloride + lithium chloride mixtures at 25 °C. Trans. Faraday Soc. 49, 1147–1149 (1953)

    Article  CAS  Google Scholar 

  80. Bahia, A.M., Lilley, T.H., Tasker, I.R.: The osmotic coefficients of aqueous CsCl and CsCl + KCl mixtures at 298.15 K. J. Chem. Thermodyn. 10, 683–685 (1978)

    Article  CAS  Google Scholar 

  81. Guo, L.J., Han, H.J., Dong, O.Y., Yao, Y.: Thermodynamics and phase equilibrium of the high concentration solid solution−aqueous solution system KCl–RbCl–H2O from T = 298.15K to T = 323.15 K. J. Chem. Thermodyn. 106, 285–294 (2017)

    Article  CAS  Google Scholar 

  82. Zeng, D.W., Wu, Z.D., Yao, Y., Han, H.J.: Isopiestic determination of water activity on the system LiNO3 + KNO3 + H2O at 273.1 and 298.1 K. J. Solution Chem. 39, 1360–1376 (2010)

    Article  CAS  Google Scholar 

  83. Dinane, A., El Guendouzi, M., Mounir, A.: Hygrometric determination of the water activities and the osmotic and activity coefficients of (ammonium chloride + sodium chloride + water) at T = 298.15 K. J. Chem. Thermodyn. 34, 783–793 (2002)

    Article  CAS  Google Scholar 

  84. Deyhimi, F., Ghalami-Choobar, B.: Potentiometric determination of activity coefficients for NH4Cl in the ternary NH4Cl/LiCl/H2O mixed electrolyte system. J. Electroanal. Chem. 584, 141–146 (2005)

    Article  CAS  Google Scholar 

  85. Deyhimi, F., Salamat-Ahangari, R.: Thermodynamic properties of the mixed electrolyte system: the ternary NH4Br + NaBr + H2O system. J. Chem. Eng. Data 54, 227–231 (2008)

    Article  CAS  Google Scholar 

  86. Huang, X.T., Li, S.N., Zhai, Q.G., Jiang, Y.C., Hu, M.C.: Thermodynamic studies of (RbF + RbCl + H2O) and (CsF + CsCl + H2O) ternary systems from potentiometric measurements at T = 298.2 K. J. Chem. Thermodyn. 103, 157–164 (2016)

    Article  CAS  Google Scholar 

  87. Dou, Z.D., Li, S.N., Zhai, Q.G., Jiang, Y.C., Hu, M.C.: Potentiometric Investigation of the thermodynamic properties of mixed electrolyte systems at 298.2 K: CsF + CsBr + H2O and CsF + CsNO3 + H2O. J. Chem. Eng. Data 63, 3801–3808 (2018)

    Article  CAS  Google Scholar 

  88. Zhang, J., Gao, S.Y., Xia, S.P.: Determination of thermodynamic properties of aqueous mixtures of RbCl and Rb2SO4 by the EMF method at T = 298.15 K. J. Chem. Thermodyn. 35, 1383–1392 (2003)

    Article  CAS  Google Scholar 

  89. Popović, D.Ž., Miladinović, J., Rard, J.A., Miladinović, Z.P., Grujić, S.R.: Isopiestic determination of the osmotic and activity coefficients of the {yNa2HPO4 + (1−y)K2HPO4}(aq) system at T = 298.15 K. J. Solution Chem. 45, 1261–1287 (2016)

    Article  CAS  Google Scholar 

  90. Popović, D.Ž., Miladinović, J., Todorović, M.D., Zrilić, M.M., Rard, J.A.: Isopiestic determination of the osmotic and activity coefficients of the {yKCl + (1– y)K2HPO4}(aq) system at T = 298.15 K. J. Chem. Thermodyn. 43, 1877–1885 (2011)

    Article  CAS  Google Scholar 

  91. Mounir, A., El Guendouzi, M., Dinane, A.: Thermodynamic properties of (NH4)2SO4(aq) + Li2SO4(aq) and (NH4)2SO4(aq) + Na2SO4(aq) at a temperature of 298.15 K. J. Chem. Thermodyn. 34, 1329–1339 (2002)

    Article  CAS  Google Scholar 

  92. Wang, D., Yang, Y.Y., Zhang, X.P., Sang, S.H.: Mean activity coefficients of NaCl in NaCl–CdCl2–H2O ternary system at 298.15 K by potential difference method. J. Chem. Eng. Data 61, 3027–3033 (2016)

    Article  CAS  Google Scholar 

  93. Ma, X.C., Li, X.P., He, X.F., Sang, S.H., Lei, N.F., Nie, Z.: Thermodynamic study of the NaCl–CuCl2–H2O ternary system at 298.15 K by the electromotive force method. J. Chem. Eng. Data 64, 90–97 (2018)

    Article  CAS  Google Scholar 

  94. Ghalami-Choobar, B.: Thermodynamic study of the ternary mixed electrolyte (NaCl+NiCl2+H2O) system: application of Pitzer model with higher-order electrostatic effects. J. Chem. Thermodyn. 43, 901–907 (2011)

    Article  CAS  Google Scholar 

  95. Downes, C.J.: Osmotic and activity coefficients for system NaCl−MnCl2−H2O at 25 °C. J. Chem. Eng. Data 18, 412–416 (1973)

    Article  CAS  Google Scholar 

  96. Reddy, D.C., Ananthaswamy, J.: Thermodynamics of electrolyte solutions: activity and osmotic coefficients of the ternary system KCl−BaCl2−H2O at 25, 35, and 45 °C. J. Chem. Eng. Data 35, 144–147 (1990)

    Article  CAS  Google Scholar 

  97. Lindenbaum, S., Rush, R.M., Robinson, R.A.: Osmotic and activity coefficients for mixtures of lithium chloride with barium chloride and cesium chloride with barium chloride in water at 298.15 K. J. Chem. Thermodyn. 4, 381–389 (1972)

    Article  CAS  Google Scholar 

  98. Downes, C.J., Pitzer, K.S.: Thermodynamics of electrolytes. Binary mixtures formed from aqueous NaCl, Na2SO4, CuCl2, and CuSO4, at 25 °C. J. Solution Chem. 5, 389–398 (1976)

    Article  CAS  Google Scholar 

  99. Ninković, R., Miladinović, J., Todorović, M., Grujić, S., Rard, J.A.: Osmotic and activity coefficients of the {xZnCl2 + (1–x)ZnSO4}(aq) system at 298.15 K. J. Solution Chem. 36, 405–435 (2007)

    Article  CAS  Google Scholar 

  100. Christov, C., Dickson, A.G., Moller, N.: Thermodynamic modeling of aqueous aluminum chemistry and solid−liquid equilibria to high solution concentration and temperature. I. The acidic H−Al−Na−K−Cl−H2O system from 0 to 100 °C. J. Solution Chem. 36, 1495–1523 (2007)

    Article  CAS  Google Scholar 

  101. Downes, C.J.: Osmotic and activity coefficients for mixtures of potassium chloride and strontium chloride in water at 298.15 K. J. Chem. Thermodyn. 6, 317–323 (1974)

    Article  CAS  Google Scholar 

  102. Popović, D.Ž., Miladinović, J., Rard, J.A., Miladinović, Z.P., Grujić, S.R.: Isopiestic determination of the osmotic and activity coefficients of the {yK2SO4 + (1−y)K2HPO4}(aq) system at T = 298.15 K. J. Chem. Thermodyn. 79, 84–93 (2014)

    Article  CAS  Google Scholar 

  103. Mounir, A., El Guendouzi, M., Dinane, A.: Hygrometric determination of water activities, osmotic and activity coefficients, and excess Gibbs energy of the system MgSO4–K2SO4–H2O. J. Solution Chem. 31, 793–799 (2002)

    Article  CAS  Google Scholar 

  104. El Guendouzi, M., Mounir, A., Dinane, A.: Thermodynamic properties of the system MgSO4–MnSO4–H2O at 298.15 K. Fluid Phase Equilib. 202, 221–231 (2002)

    Article  Google Scholar 

  105. Tao, D.P.: The universal characteristics of a thermodynamic model to conform to the Gibbs−Duhem equation. Sci. Rep. 6, 35792 (2016)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  106. Marcus, Y.: Thermodynamics of solvation of ions Part 6: The standard partial molar volumes of aqueous ions at 298.15 K. J. Chem. Soc. Faraday Trans. 89, 713–718 (1993)

    Article  CAS  Google Scholar 

Download references

Acknowledgement

This work was financially supported by the National Natural Science Foundation of China under Grant No. 51464022.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dongping Tao.

Ethics declarations

Conflict of interest

The authors declare that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Electronic supplementary material 1 (DOCX 139 kb)

Appendices

Appendix A: Brief Introduction of MIVM

According to statistical thermodynamics, the canonical partition function of pure liquid i is:

$$Q_{i} = Q_{Ti} Q_{pi} /N_{i} !$$
(19)

where \(Q_{Ti} = \left[ {\left( {2\pi mkT} \right)^{1/2} /h} \right]^{{3N_{i} }}\) and \(Q_{pi}\) are the translation partition function and the configurational partition function of i molecules, respectively. \(Q_{pi}\) can be expressed as:

$$Q_{pi} = \int\limits_{{V_{i} }} \ldots \int {\exp \left( { - E_{pi} /kT} \right)} dx_{1} dy_{1} dz_{1} \ldots dx_{{N_{i} }} dy_{{N_{i} }} dz_{{N_{i} }}$$
(20)

where \(V_{i}\) is the volume, \(E_{pi}\) is the potential energy, and \(N_{i}\) is the molecular number of i, \(k\) is the Boltzmann constant. Based on the lattice theory of solutions [29], \(E_{pi}\) may be chosen as:

$$E_{pi} = \frac{1}{2}Z_{i} N_{i} \varepsilon_{ii}$$
(21)

where \(Z_{i}\) is the nearest molecule or first coordination number. \(\varepsilon_{ii}\) is the ii pair potential energy.

The MIVM originates from the physical picture described in the Introduction. Thus, substituting Eq. 21 into Eq. 20 and integrating it, Eq. 20 can be simplified as:

$$\begin{aligned} Q_{pi} &= \left[ {\iiint_{ + \infty } {\exp \left( { - \frac{{Z_{i} \varepsilon_{ii} }}{2kT}} \right)dxdydz}} \right]^{{N_{i} }} \\ &= \left[ {\int\limits_{0}^{\pi } {\sin \theta d\theta \int\limits_{0}^{2\pi } {d\varphi \int\limits_{0}^{{r_{i} }} {r^{2} \exp \left( { - \frac{{Z_{i} \varepsilon_{ii} }}{2kT}} \right)dr} } } } \right]^{{N_{i} }} \\ &= \left( {\frac{{V_{i} }}{{N_{i} }}} \right)^{{N_{i} }} \exp \left( { - \frac{{Z_{i} N_{i} \varepsilon_{ii} }}{2kT}} \right) \\ \end{aligned}$$
(22)

where \(r_{i}\) is the average radius of molecular cell of i.

For a C-component system, the partition functions of real and ideal solutions are expressed as follows:

$$Q = \prod\limits_{i = 1}^{C} {Q_{Ti} Q_{p} } /N_{i} !$$
(23)
$$Q^{id} = \prod\limits_{i = 1}^{C} {Q_{i} }$$
(24)

where C is the number of components, and \(Q_{p}\) is the configurational partition function, expressed as:

$$\begin{aligned} Q_{p} &= \int\limits_{V} \ldots \int {\exp \left( { - \frac{{\varepsilon_{p} }}{kT}} \right)} \prod\limits_{i = 1}^{C} {\left( {dx_{1} dy_{1} dz_{1} ...dx_{{N_{i} }} dy_{{N_{i} }} dz_{{N_{i} }} } \right)} \\ &= \prod\limits_{i = 1}^{C} {\left[ {\iiint_{ \pm \infty } {\exp \left( { - \frac{{\varepsilon_{p} }}{kT}} \right)}dx_{i} dy_{i} dz_{i} } \right]^{{N_{i} }} } \\ &= \left[ {\iiint_{ \pm \infty } {\exp \left( { - \frac{{\varepsilon_{p} }}{kT}} \right)}dxdydz} \right]^{N} \\ &= \left( \frac{V}{N} \right)^{N} \exp \left( { - \frac{{N\varepsilon_{p} }}{kT}} \right) \\ \end{aligned}$$
(25)

where \(V\) and \(N\) are the volume and molecular number of the system, respectively, and \(\varepsilon_{p}\) is the potential energy function of mixing of its C kinds of molecules. According to the relation between Gibbs energy and partition function:

$$G = kT\left[ {V\left( {\frac{\partial \ln Q}{{\partial V}}} \right)_{T} - \ln Q} \right]$$
(26)

One can get the real and ideal Gibbs energies of the system, respectively:

$$G = kT\left[ {N - N\ln \left( \frac{V}{N} \right) + \frac{{N\varepsilon_{p} }}{kT} - \ln \left( {\prod\limits_{i = 1}^{C} {N_{i} !Q_{Ti} } } \right)} \right]$$
(27)
$$G^{{{\text{id}}}} = kT\left[ {\sum\limits_{i = 1}^{C} {N_{i} } - \sum\limits_{i = 1}^{C} {N_{i} \ln \left( {\frac{{V_{i} }}{{N_{i} }}} \right)} + \sum\limits_{i = 1}^{C} {\frac{{Z_{i} N_{i} \varepsilon_{ii} }}{kT}} - \ln \left( {\prod\limits_{i = 1}^{C} {N_{i} !Q_{Ti} } } \right)} \right]$$
(28)

Therefore, the excess Gibbs energy of the system can be expressed as:

$$\begin{aligned} G^{{{\text{ex}}}} &= G - G^{{{\text{id}}}} \\ &= kT\left[ { - N\ln \left( \frac{V}{N} \right) + \sum\limits_{i = 1}^{C} {N_{i} \ln \left( {\frac{{V_{i} }}{{N_{i} }}} \right)} + \frac{N}{2kT}\left( {2\varepsilon_{p} - \sum\limits_{i = 1}^{C} {\frac{{Z_{i} N_{i} \varepsilon_{ii} }}{kT}} } \right)} \right] \\ &= nRT\left[ {\sum\limits_{i = 1}^{C} {x_{i} \ln \left( {\frac{{V_{\text{m}i} }}{{V_{\text{m}} }}} \right)} + \frac{1}{2kT}\left( {2\varepsilon_{p} - \sum\limits_{i = 1}^{C} {\frac{{Z_{i} N_{i} \varepsilon_{ii} }}{kT}} } \right)} \right] \\ &= nRT\left[ {\sum\limits_{i = 1}^{C} {x_{i} \ln \left( {\frac{{\varphi_{i} }}{{x_{i} }}} \right)} + \frac{{\Delta \varepsilon_{p} }}{2kT}} \right] \\ \end{aligned}$$
(29)

where \(n\) is molar number of the system, \(V_{\text{m}i}\) and \(V_{\text{m}}\) are the molar volumes of i and the system, respectively.\(\varphi_{i} = x_{i} V_{\text{m}i} /V_{\text{m}}\) is the molar volume fraction of component i in the system. \(\Delta \varepsilon_{p}\) is the excess potential energy function of the system:

$$\Delta \varepsilon_{p} = 2\varepsilon_{p} - \sum\limits_{i = 1}^{C} {Z_{i} x_{i} \varepsilon_{ii} }$$
(30)

According to multi-liquid theory [29], the potential energy function \(\varepsilon_{p}\) of mixing of molecules can be chosen as:

$$\varepsilon_{p} = \frac{1}{2}\sum\limits_{i = 1}^{C} {Z_{i} } x_{i} \left( {\sum\limits_{j = 1}^{C} {x_{ji} \varepsilon_{ji} } } \right)$$
(31)

If \(\Delta \varepsilon_{p} = 0\), then Eq. 29 may be reduced to:

$$G^{{{\text{ex}}}} = nRT\sum\limits_{i = 1}^{C} {x_{i} \ln \left( {\frac{{\varphi_{i} }}{{x_{i} }}} \right)}$$
(32)

which is the well-known Flory–Huggins equation.

Equation 29 can be considered as the theoretical form of the MIVM; it only contains properties and parameters of the pure components. Please see reference [27] for more detailed information about the MIVM.

Appendix B: The Derivation of the Electrolyte MIVM (eMIVM)

Before deducing eMIVM, let's start with two special features of the MIVM:

  1. (1)

    The local compositions are defined by the local coordination numbers [31].

Suppose that in the C system there are C types of molecular cells whose central molecules are the corresponding molecules of C components. Then, the local coordination numbers of the cell i are \(Z_{ji}\) and \(Z_{ii}\) that are defined as the molecular numbers of components j and i surrounding the central molecule i. They are also proportional to their corresponding Boltzmann’s factor:

$$Z_{ji} = Z_{i} x_{j} \exp \left( { - \frac{{\varepsilon_{ji} }}{kT}} \right),\;Z_{ii} = Z_{i} x_{i} \exp \left( { - \frac{{\varepsilon_{ii} }}{kT}} \right)\;\;\;i = 1, \ldots ,C\;\;j = 1, \ldots ,C$$
(33)

Based on the concept of the local composition of Wilson [28], the local mole fractions of components j and i around central molecule i can be expressed as follows:

$$x_{ji} = \frac{{Z_{ji} }}{{\sum\limits_{j = 1}^{C} {Z_{ji} } }} = \frac{{x_{j} B_{ji} }}{{\sum\limits_{j = 1}^{C} {x_{j} B_{ji} } }},\;x_{ii} = \frac{{Z_{ii} }}{{\sum\limits_{j = 1}^{C} {Z_{ji} } }} = \frac{{x_{i} }}{{\sum\limits_{j = 1}^{C} {x_{j} B_{ji} } }}\;\;\;i = 1, \ldots ,C$$
(34)

where the pair potential energy parameter \(B_{ji}\) is defined as:

$$B_{ji} = \exp \left( { - \frac{{\varepsilon_{ji} - \varepsilon_{ii} }}{kT}} \right)$$
(35)

Further, the relationship between \(x_{ji}\) and \(x_{ii}\) can be obtained:

$$\frac{{x_{ji} }}{{x_{ii} }} = \frac{{Z_{ji} }}{{Z_{ii} }} = \frac{{Z_{i} x_{j} \exp \left( { - \frac{{\varepsilon_{ji} }}{kT}} \right)}}{{Z_{i} x_{i} \exp \left( { - \frac{{\varepsilon_{ii} }}{kT}} \right)}} = \frac{{x_{j} }}{{x_{i} }}B_{ji}$$
(36)

For convenience in representing other local mole fraction ratios, the following expressions are introduced:

$$\frac{{x_{ji} }}{{x_{li} }} = \frac{{x_{j} B_{ji} }}{{x_{l} B_{li} }} = \frac{{x_{j} }}{{x_{l} }}B_{ji,li} ,\;\;l = 1, \ldots ,C$$
(37)
$$B_{ji,li} = \frac{{B_{ji} }}{{B_{li} }} = \exp \left( { - \frac{{\varepsilon_{ji} - \varepsilon_{li} }}{kT}} \right)$$
(38)

Since MIVM is a local composition model, we want to replace the molar volume fraction of component i (\(\varphi_{i}\)) in Eq. 29 with the local molar volume fraction of component i (\(\xi_{i}\)). The local molar volume fraction of component i (\(\xi_{i}\)) can be expressed as:

$$\xi_{i} = \frac{{x_{ii} V_{\text{m}i} }}{{\sum\limits_{j = 1}^{C} {x_{ji} V_{\text{m}j} } }} = \frac{{x_{i} V_{\text{m}i} }}{{\sum\limits_{j = 1}^{C} {x_{j} V_{\text{m}j} B_{ji} } }},\;\;i = 1, \ldots ,C$$
(39)

Similar expressions can also be obtained for cells with a central component j.

  1. (2)

    MIVM defines a clear coordination number.

This feature of the MIVM makes the eMIVM also have a clear coordination number. In the eMIVM, we consider three cases, where the coordination number is 6, 8 and 10, respectively, and find that the fitting quality of the model is not sensitive to the coordination number used. A value of 10 for the coordination number yields the minimum value of the deviation for majority of electrolytes. Therefore, in this paper, the coordination numbers of species are all taken as 10.

The extension of MIVM to electrolyte solutions is mainly according to the idea described in Sect. 2.2. In more detail: since this work assumes that the electrolyte dissociates completely in aqueous solution, three kinds of local cells will appear in the solution, namely, cells with a central ion (either an anion or cation) surrounded by other counterions and molecules, and cells with a central molecule surrounded by anions, cations and molecules. Two key hypotheses proposed by Chen [9, 10] are adopted, namely, the like-ion repulsion and the local electroneutrality. The central ion cells satisfy the like-ion repulsion assumption, i.e., no anions are allowed in cells with a central anion, no cations are allowed in cells with a central cation. The central molecule cells satisfy the local electroneutrality assumption, i.e., the net local charge around a central molecule is zero. For the three kinds of local cells, the following normalizing equations hold:

$$\sum\limits_{c^{\prime}} {x_{c^{\prime}s} } + \sum\limits_{a^{\prime}} {x_{a^{\prime}s} } + \sum\limits_{s^{\prime}} {x_{s^{\prime}s} } = 1\;\;\;\;\left( {\text{central}}\, {\text{molecule}} \, {\text{cells}} \right)$$
(40)
$$\sum\limits_{a^{\prime}} {x_{a^{\prime}c} } + \sum\limits_{s^{\prime}} {x_{s^{\prime}c} } = 1\;\;\;\;\left( {\text{central}} \,{\text{cation}}\, {\text{cells}} \right)$$
(41)
$$\sum\limits_{c^{\prime}} {x_{c^{\prime}a} } + \sum\limits_{s^{\prime}} {x_{s^{\prime}a} } = 1\;\;\;\;\;\left( {\text{central}} \, {\text{anion}}\, {\text{cells}} \right)$$
(42)

where \(s\) and \(s^{\prime}\) denote molecular species, \(a\) and \(a^{\prime}\) denote anionic species, and \(c\) and \(c^{\prime}\) denote cationic species.

By combining Eqs. 36 and 37 and Eqs. 4042, the following expressions for the local mole fractions in term of overall mole fractions may be derived:

$$x_{is} = \frac{{x_{i} B_{is} }}{{\sum\limits_{c^{\prime}} {x_{c^{\prime}} B_{c^{\prime}s} } + \sum\limits_{a^{\prime}} {x_{a^{\prime}} B_{a^{\prime}s} } + \sum\limits_{s^{\prime}} {x_{s^{\prime}} B_{s^{\prime}s} } }}(i = a,c,s)$$
(43)
$$x_{ic} = \frac{{x_{i} }}{{\sum\limits_{a^{\prime}} {x_{a^{\prime}} } B_{a^{\prime}c,ic} + \sum\limits_{s^{\prime}} {x_{s^{\prime}} B_{s^{\prime}c,ic} } }}(i = a,s)$$
(44)
$$x_{ia} = \frac{{x_{i} }}{{\sum\limits_{c^{\prime}} {x_{c^{\prime}} } B_{c^{\prime}a,ia} + \sum\limits_{s^{\prime}} {x_{s^{\prime}} B_{s^{\prime}a,ia} } }}(i = c,s)$$
(45)

To obtain an expression for the excess Gibbs energy of the new model, we need to modify the energy term and the volume term on the right side of Eq. 29. Let’s modify the energy term first. According to the three kinds of local cells in electrolyte solution, Eq. 31 may be changed into:

$$\begin{aligned} \varepsilon_{p} &= \frac{1}{2}\sum\limits_{i = 1}^{C} {Z_{i} x_{i} \left( {\sum\limits_{j = 1}^{C} {x_{ji} } \varepsilon_{ji} } \right)} \\ &= \frac{1}{2}\left[ \begin{gathered} \sum\limits_{s} {Z_{s} x_{s} \left( {\sum\limits_{s^{\prime}} {x_{s^{\prime}s} \varepsilon_{s^{\prime}s} + } \sum\limits_{c^{\prime}} {x_{c^{\prime}s} \varepsilon_{c^{\prime}s} + } \sum\limits_{a^{\prime}} {x_{a^{\prime}s} \varepsilon_{a^{\prime}s} } } \right)} \\ + \sum\limits_{c} {Z_{c} x_{c} \left( {\sum\limits_{s^{\prime}} {x_{s^{\prime}c} \varepsilon_{s^{\prime}c} + } \sum\limits_{a^{\prime}} {x_{a^{\prime}c} \varepsilon_{a^{\prime}c} } } \right)} \\ + \sum\limits_{a} {Z_{a} x_{a} \left( {\sum\limits_{s^{\prime}} {x_{s^{\prime}a} \varepsilon_{s^{\prime}a} + } \sum\limits_{c^{\prime}} {x_{c^{\prime}a} \varepsilon_{c^{\prime}a} } } \right)} \\ \end{gathered} \right] \\ \end{aligned}$$
(46)

The pure component state, taken as the reference state for molecules, and the hypothetical homogeneously mixed, completely dissociated liquid electrolyte mixture is adopted as the reference state for electrolytes. Therefore, the reference potential energy term \(\sum\limits_{i = 1}^{C} {Z_{i} x_{i} \varepsilon_{ii} }\) in Eq. 30 may be changed as:

$$\sum\limits_{i = 1}^{C} {Z_{i} x_{i} \varepsilon_{ii} } = \sum\limits_{s} {Z_{s} x_{s} \varepsilon_{ss} } + \sum\limits_{c} {Z_{c} x_{c} \left( {\frac{{\sum\limits_{a^{\prime}} {x_{a^{\prime}} \varepsilon_{a^{\prime}c} } }}{{\sum\limits_{a^{\prime\prime}} {x_{a^{\prime\prime}} } }}} \right)} + \sum\limits_{a} {Z_{a} x_{a} \left( {\frac{{\sum\limits_{c^{\prime}} {x_{c^{\prime}} \varepsilon_{c^{\prime}a} } }}{{\sum\limits_{c^{\prime\prime}} {x_{c^{\prime\prime}} } }}} \right)}$$
(47)

Substituting Eqs. 46 and 47 into Eq. 30, then the excess potential energy function \(\Delta \varepsilon_{p}\) based on the three kinds of local cells can be expressed as:

$$\begin{gathered} \Delta \varepsilon_{p} = 2\varepsilon_{p} - \sum\limits_{i = 1}^{C} {Z_{i} x_{i} \varepsilon_{ii} } \\ = \sum\limits_{s} {Z_{s} x_{s} \left[ {\left( {\sum\limits_{s^{\prime}} {x_{s^{\prime}s} \varepsilon_{s^{\prime}s} + } \sum\limits_{c^{\prime}} {x_{c^{\prime}s} \varepsilon_{c^{\prime}s} + } \sum\limits_{a^{\prime}} {x_{a^{\prime}s} \varepsilon_{a^{\prime}s} } } \right) - \varepsilon_{ss} } \right]} \\ + \sum\limits_{c} {Z_{c} x_{c} \left[ {\left( {\sum\limits_{s^{\prime}} {x_{s^{\prime}c} \varepsilon_{s^{\prime}c} + } \sum\limits_{a^{\prime}} {x_{a^{\prime}c} \varepsilon_{a^{\prime}c} } } \right) - \left( {\frac{{\sum\limits_{a^{\prime}} {x_{a^{\prime}} \varepsilon_{a^{\prime}c} } }}{{\sum\limits_{a^{\prime\prime}} {x_{a^{\prime\prime}} } }}} \right)} \right]} \\ + \sum\limits_{a} {Z_{a} x_{a} \left[ {\left( {\sum\limits_{s^{\prime}} {x_{s^{\prime}a} \varepsilon_{s^{\prime}a} + } \sum\limits_{c^{\prime}} {x_{c^{\prime}a} \varepsilon_{c^{\prime}a} } } \right) - \left( {\frac{{\sum\limits_{c^{\prime}} {x_{c^{\prime}} \varepsilon_{c^{\prime}a} } }}{{\sum\limits_{c^{\prime\prime}} {x_{c^{\prime\prime}} } }}} \right)} \right]} \\ \end{gathered}$$
(48)

The next step is to modify the volume term. As stated previously, the molar volume fraction (\(\varphi_{i}\)) in Eq. 29 is replaced by the local molar volume fraction (\(\xi_{i}\)), and based on the assumption of the like-ion repulsion, there are no ions of like charge around the central ion (i.e.,\(x_{aa} = x_{cc} = 0\)); therefore, the local molar volume fraction of the central ion cells (modified Eq. 39) are zero:

$$\zeta_{c} = \frac{{x_{cc} V_{\text{m}c} }}{{\sum\limits_{c^{\prime}} {x_{c^{\prime}c} V_{\text{m}c^{\prime}} + \sum\limits_{a^{\prime}} {x_{a^{\prime}c} V_{\text{m}a^{\prime}} + } \sum\limits_{s^{\prime}} {x_{s^{\prime}c} V_{\text{m}s^{\prime}} } } }} = 0$$
(49)
$$\zeta_{a} = \frac{{x_{aa} V_{\text{m}a} }}{{\sum\limits_{c^{\prime}} {x_{c^{\prime}a} V_{\text{m}c^{\prime}} + \sum\limits_{a^{\prime}} {x_{a^{\prime}a} V_{\text{m}a^{\prime}} + } \sum\limits_{s^{\prime}} {x_{s^{\prime}a} V_{\text{m}s^{\prime}} } } }} = 0$$
(50)

while the local molar volume fraction of the central molecule cell (\(\xi_{s}\)) is remains in the volume term, and by combining with Eq. 43, \(\xi_{s}\) can be obtained as follows:

$$\begin{aligned} \zeta_{s} &= \frac{{x_{ss} V_{\text{m}s} }}{{\sum\limits_{s^{\prime}} {x_{s^{\prime}s} V_{\text{m}s^{\prime}} + \sum\limits_{c^{\prime}} {x_{c^{\prime}s} V_{\text{m}c^{\prime}} + } \sum\limits_{a^{\prime}} {x_{a^{\prime}s} V_{\text{m}a^{\prime}} } } }} \\ & = \frac{{x_{s} V_{\text{m}s} }}{{\sum\limits_{s^{\prime}} {x_{s^{\prime}} B_{s^{\prime}s} V_{\text{m}s^{\prime}} + \sum\limits_{c^{\prime}} {x_{c^{\prime}} B_{c^{\prime}s} V_{\text{m}c^{\prime}} + } \sum\limits_{a^{\prime}} {x_{a^{\prime}} B_{a^{\prime}s} V_{\text{m}a^{\prime}} } } }} \\ & = \frac{{x_{s} V_{\text{m}s} }}{{\sum\limits_{k} {x_{k} B_{ks} V_{\text{m}k} } }} \\ \end{aligned}$$
(51)

Now replace the molar volume fraction (\(\varphi_{i}\)) in Eq. 29 with the local molar volume fraction (\(\xi_{i}\)) of Eqs. 4951, and then substitute Eq. 48 into Eq. 29 and combine it with Eqs. 4345. Finally, we can obtain the expression of the molar excess Gibbs energy of the short range term of the new model as shown in Eq. 5.

The relationship between the partial molar and molar quantity (RPMQ) at constant temperature and pressure:

$$\overline{G}_{i}^{{{\text{ex}}}} = RT\ln \gamma_{i} = \left[ {\frac{{\partial \left( {nG_{\text{m}}^{{{\text{ex}}}} } \right)}}{{\partial n_{i} }}} \right]_{{T,p,n_{j \ne i} }} = G_{\text{m}}^{{{\text{ex}}}} + \left( {\frac{{\partial G_{\text{m}}^{{{\text{ex}}}} }}{{\partial x_{i} }}} \right)_{{T,P,x_{k \ne i} }} - \sum\limits_{j = 1}^{C - 1} {x_{j} } \left( {\frac{{\partial G_{\text{m}}^{{{\text{ex}}}} }}{{\partial x_{j} }}} \right)_{{T,P,x_{k \ne j} }}$$
(52)

The activity coefficients of any species can be obtained by combining Eqs. 5 and 52. The activity coefficient equations for any species are given in Eqs. 68.

Here it is worth noting that based on the conclusion that RPMQ and the Gibbs–Duhem equation is equivalent [105], the thermodynamic consistency of these activity coefficient expressions can be verified and confirmed by using only the sum formula of the molar excess Gibbs energy:

$$G_{\text{m}}^{{\text{ex,MIVM}}} = RT\left( {\sum\limits_{c} {x_{c} \ln \gamma_{c}^{{{\text{MIVM}}}} } + \sum\limits_{a} {x_{a} \ln \gamma_{a}^{{{\text{MIVM}}}} } + \sum\limits_{s} {x_{s} \ln \gamma_{s}^{{{\text{MIVM}}}} } } \right)$$
(53)

Appendix C: Model Parameters

3.1 Binary Parameters

By applying the assumption of local electroneutrality to the central molecule cell, one can get:

$$z_{a} x_{as} = z_{c} x_{cs}$$
(54)

Substituting Eq. 36 into Eq. 54, and because the electrolyte solution is electrically neutral (i.e.,\(z_{a} x_{a} = z_{c} x_{c}\)), hence the following relationship can be obtained:

$$\varepsilon_{as} = \varepsilon_{cs}$$
(55)

Since the pair potential energies are symmetric (i.e.\(\varepsilon_{ji} = \varepsilon_{ij}\)), it can be further deduced that:

$$B_{as} = B_{cs} = B_{ca,s}$$
(56)
$$B_{sa,ca} = B_{sc,ac} = B_{s,ca}$$
(57)

For a binary electrolyte system, \(B_{ca,s}\) and \(B_{s,ca}\) are the only two adjustable electrolyte-specific parameters in the new model.

3.2 Multicomponent Parameters

In the same way, the assumption of local electroneutrality applied to the central molecule cell, and the following relation can be obtained:

$$\sum\limits_{c} {x_{cs} } z_{c} = \sum\limits_{a} {x_{as} } z_{a}$$
(58)

Again, substituting Eq. 36 into Eq. 58, and because solution electroneutrality requires \(\sum\limits_{c} {x_{c} } z_{c} = \sum\limits_{a} {x_{a} } z_{a}\), it follows that:

$$\sum\limits_{c} {z_{c} x_{c} B_{cs} } = \sum\limits_{a} {z_{a} x_{a} B_{as} }$$
(59)

because the local composition concept considers only two-body interactions and the like-ion repulsion is assumed. Therefore, Eq. 59 can be generalized to:

$$B_{cs} = \frac{{\sum\limits_{a} {z_{a} x_{a} B_{ca,s} } }}{{\sum\limits_{a^{\prime}} {z_{a^{\prime}} x_{a^{\prime}} } }}$$
(60)
$$B_{as} = \frac{{\sum\limits_{c} {z_{c} x_{c} B_{ca,s} } }}{{\sum\limits_{c^{\prime}} {z_{c^{\prime}} x_{c^{\prime}} } }}$$
(61)

\(B_{ca,s}\) in Eqs. 60 and 61 is the binary parameter.

The variables \(B_{sa,ca}\) and \(B_{sc,ac}\) can be computed from the \(B_{is}\)(Eqs. 60 and 61):

$$\begin{aligned} B_{sa,ca} & = \exp \left( { - \frac{{\varepsilon_{sa} - \varepsilon_{ca} }}{kT}} \right) = \exp \left( { - \frac{{\varepsilon_{sa} - \varepsilon_{ss} }}{kT} - \frac{{\varepsilon_{ss} - \varepsilon_{ca} }}{kT}} \right) \\ & = \frac{{\exp \left( { - \frac{{\varepsilon_{sa} - \varepsilon_{ss} }}{kT}} \right)\exp \left( { - \frac{{\varepsilon_{sa} - \varepsilon_{ca} }}{kT}} \right)}}{{\exp \left( { - \frac{{\varepsilon_{sa} - \varepsilon_{ss} }}{kT}} \right)}} \\ &= \frac{{B_{as} B_{s,ca} }}{{B_{ca,s} }} \\ \end{aligned}$$
(62)

and

$$B_{sc,ac} = \frac{{B_{cs} B_{s,ca} }}{{B_{ca,s} }}$$
(63)

Furthermore, the salt–salt binary interaction parameters are also required in the new model, and these parameters are handled as follows [10]:

$$B_{ca,c^{\prime}a} = \exp \left( { - \frac{{\varepsilon_{ca} - \varepsilon_{c^{\prime}a} }}{kT}} \right) = \frac{1}{{\exp \left( { - \frac{{\varepsilon_{c^{\prime}a} - \varepsilon_{ca} }}{kT}} \right)}} = \frac{1}{{B_{c^{\prime}a,ca} }}$$
(64)

and

$$B_{ca,ca^{\prime}} = \frac{1}{{B_{ca^{\prime},ca} }}$$
(65)

The salt–salt binary interaction parameter is determined by fitting solubility data or activity and osmotic coefficient data in the ternary system. It can be seen that \(B_{ca,s}\),\(B_{s,ca}\),\(B_{ss^{\prime}}\),\(B_{s^{\prime}s}\),\(B_{ca,c^{\prime}a}\),\(B_{c^{\prime}a,ca}\),\(B_{ca,ca^{\prime}}\) and \(B_{ca^{\prime},ca}\) are the adjustable binary parameters for the multicomponent system. Specifically, for a ternary system with only one solvent component and a common cation, the \(B_{ss^{\prime}}\),\(B_{s^{\prime}s}\),\(B_{ca,c^{\prime}a}\) and \(B_{c^{\prime}a,ca}\) are equal to unity.

Appendix D: The Molar Volume of Species

The molar volumes of species are required in the new model. For the sake of simplicity, the ions are treated as spherical, so the molar volumes of ions can be calculated by the following formula [36]:

$$V_{\text{m}i} = \left( {\frac{{4\pi N_{{\text{A}}} }}{3}} \right)r_{i}^{3} = 2522.5r_{i}^{3}$$
(66)

where \(N_{{\text{A}}}\) is the Avogadro constant,\(r_{i}\)(nm) is the ion radius in aqueous solution, and its values are tabulated in Table 7. The molar volume of pure water (\(V_{ms}\)) is 18.07 cm3·mol−1 at 298.15 K.

Table 7 Ionic radii for cations and anionsa

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, C., Xing, Y. & Tao, D. A Two-Parameter Theoretical Model for Predicting the Activity and Osmotic Coefficients of Aqueous Electrolyte Solutions. J Solution Chem 49, 659–694 (2020). https://doi.org/10.1007/s10953-020-00987-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10953-020-00987-z

Keywords

Navigation