Elsevier

Geochimica et Cosmochimica Acta

Volume 88, 1 July 2012, Pages 130-145
Geochimica et Cosmochimica Acta

Prediction of vapor–liquid equilibrium and PVTx properties of geological fluid system with SAFT-LJ EOS including multi-polar contribution. Part II: Application to H2O–NaCl and CO2–H2O–NaCl System

https://doi.org/10.1016/j.gca.2012.04.025Get rights and content

Abstract

The SAFT-LJ equation of state improved by Sun and Dubessy (2010) can represent the vapor–liquid equilibrium and PVTx properties of the CO2–H2O system over a wide PT range because it accounts for the energetic contribution of the main types of molecular interactions in terms of reliable molecular based models. Assuming that NaCl fully dissociates into individual ions (spherical Na+ and Cl) in water and adopting the restricted primitive model of mean spherical approximation to account for the energetic contribution due to long-range electrostatic forces between ions, this study extends the improved SAFT-LJ EOS to the H2O–NaCl and the CO2–H2O–NaCl systems at temperatures below 573 K. The EOS parameters for the interactions between ion and ion and between ion and water were determined from the mean ionic activity coefficient data and the density data of the H2O–NaCl system. The parameters for the interactions between ion and CO2 were evaluated from CO2 solubility data of the CO2–H2O–NaCl system. Comparison with the experimental data shows that this model can predict the mean ionic activity coefficient, osmotic coefficient, saturation pressure, and density of aqueous NaCl solution and can predict the vapor–liquid equilibrium and PVTx properties of the CO2–H2O–NaCl system over the range from 273 to 573 K, from 0 to 1000 bar, and from 0 to 6 mol/kg NaCl with high accuracy.

Introduction

The compositions of fluids in various geological environments, such as metamorphic rocks, pegmatites, carbonatites, geothermal systems, and a wide variety of hydrothermal deposits (e.g. Roedder, 1984, Dubessy et al., 1989, Labotka, 1991, Schmidt and Bodnar, 2000, Anovitz et al., 2004, Heinrich, 2007), can be reasonably approximated by the H2O–NaCl system or the CO2–H2O–NaCl system. Understanding the thermodynamic properties (e.g. PVTx properties, vapor–liquid equilibrium) of the H2O–NaCl and the CO2–H2O–NaCl system over a wide PT range is important for the use of fluid inclusions, for modeling heat and mass transfer in various geological environments, and for many process designs in chemical engineering and environmental engineering, such as enhanced oil recovery using H2O or CO2, CO2 storage in saline aquifers (e.g. Spycher et al., 2003, Hu et al., 2007).

There are two kinds of methods for modeling the thermodynamic properties of aqueous electrolyte solutions, the activity coefficient model and equation of state approach. Most of activity coefficient models for electrolyte solutions are based on the Debye–Hückel theory. The Pitzer model (Pitzer, 1973, Pitzer and Mayorga, 1973), a famous one of these models, can calculate the activity coefficients of many single and mixed aqueous electrolyte solutions and has gained wide application in the field of geochemistry and chemical engineering. However, the activity coefficient models are in terms of the excess Gibbs energy so that they are only useful in predicting activity coefficients. The vapor–liquid equilibrium and PVTx properties of electrolyte solutions cannot be derived from the activity coefficient model itself. Pitzer et al. (1984) combined their model with an empirical EOS for pure water to calculate the density of aqueous NaCl solution. However, a large number of experimental data are required to evaluate the parameters of the Pitzer model and it is difficult to apply the Pitzer model to vapor–liquid equilibria of gas–water–salt systems.

Applying an activity coefficient model or Henry’s law to describe the non-ideality of the aqueous solution, many solubility models, such as the model of Li and Ngheim, 1986, Duan and Sun, 2003, Papaiconomou et al., 2003, Dubessy et al., 2005, Spycher and Pruess, 2005, Akinfiev and Diamond, 2010, have been proposed to calculate CO2 solubilities in aqueous NaCl solution. Although they can give accurate representation of CO2 solubility, solubility models does not allow for the calculation of the composition of vapor phase and PVTx properties of the CO2–H2O–NaCl system.

Equations of state (EOS) are the second type of methods to model the PVTx properties and fluid phase equilibria, which also provide enthalpies and other thermodynamic properties of fluids. There are numerous types of EOS. In the family of EOS, the classical cubic EOS and the virial-type EOS can model the thermodynamic properties of the fluid systems composed of uncharged non-polar or weakly-polar molecules (e.g. N2, CO2, and hydrocarbons) between which the dominant intermolecular forces are the repulsion force and the dispersion force. However, cubic EOS and virial-type EOS are not suitable for modeling highly non-ideal aqueous electrolyte solutions because the long-range electrostatic forces (Coulombic forces) between ions and the hydrogen-bonding forces between H2O molecules are different from the repulsion forces and dispersion forces. In contrast, the molecular-based EOS, which take into account the energetic contribution of various types of intermolecular forces in terms of statistical mechanical approaches, provide the possibility to model the thermodynamic behavior of aqueous electrolyte solutions.

The analytical equations accounting for the energetic contribution of different types of intermolecular forces (e.g. the Coulombic force, the multi-polar force, the repulsion force between hard spheres) have been derived from statistical mechanical approaches (such as perturbation theory, integral equation theory and molecular simulation). These equations form the theoretical basis of molecular-based EOS. Considerable advances have been made in molecular-based EOS for non-electrolyte systems. The SAFT (Statistical Association Fluid Theory) EOS, one popular example of the molecular-based EOS, is suitable for description of the thermodynamic properties of long-chain molecules (e.g. hydrocarbon, alkanol) and molecules with hydrogen bonding. A review on the state of the art of SAFT EOS has been given in our previous paper (Sun and Dubessy, 2010) and is omitted here. Many studies (Jin and Donohue, 1988, Wu and Prausnitz, 1998, Galindo et al., 1999, Liu et al., 1999, Tan et al., 2005) tried to model electrolyte solutions with SAFT EOS or other molecular-based EOS in the last two decades. However, there still remains important work to improve thermodynamic models.

Before modeling the H2O–NaCl system with molecular-based EOS, we should note that the speciation of NaCl (and other strong electrolytes) in aqueous solution changes with temperature (and, to a lesser extent, with pressure). At ambient conditions, NaCl is fully dissociated so that only single ions exist in aqueous solutions. The dominant interaction between ions is the long-range electrostatic force. As temperature increases, ions in aqueous environment tend to associate due to the decrease of the dielectric constant of the solvent resulting from the decrease of hydrogen bonding. At temperatures above 573 K, ion pairs (NaCl° and polyatomic clusters) rather than single ions are the dominant species of NaCl in aqueous solution (Pitzer and Simonson, 1984, Oelkers and Helgeson, 1993). Using an empirical method to treat with salt effect, Bowers and Helgeson (1983) applied a modified cubic EOS to calculate the PVTx properties of the CO2–H2O–NaCl system at temperatures higher than 623 K. Taking account of the repulsion force, dispersion force and dipole–dipole force between H2O and NaCl° based on perturbation theory, Anderko and Pitzer (1993) developed an EOS to calculate the phase equilibrium and PVTx properties of H2O–NaCl system at temperatures above 573 K. Duan et al., 1995, Duan et al., 2003 extended the Anderko–Pitzer EOS to the H2O–NaCl–CO2 and H2O–NaCl–CH4 systems.

Since the long-range electrostatic force is the major difference between strong electrolyte solutions and non-electrolyte systems, one way to develop an EOS for electrolyte solutions is superimposing an equation for the energetic contribution of long-range electrostatic forces on the formula of an existing EOS for non-electrolyte systems. Some models (such as Aasberg-Petersen et al., 1991, Zuo and Guo, 1991, Kiepe et al., 2004) combined the traditional Debye–Hückel term or an electrostatic term derived from a semi-empirical activity coefficient model with a cubic EOS. Although the formulas of these EOS are relatively simple, a satisfactory quantitative description of the thermodynamic properties of aqueous electrolyte solutions is difficult to achieve. Poor agreement with experimental data is due to the fact that the models used to represent the intermolecular interactions in systems they investigated are not exact from the molecular point of view. For instance, the Debye–Hückel model considers the ions as point charges without size, which is only valid at the infinite dilution limit.

In contrast to the Debye–Hückel theory, the mean spherical approximation (MSA) of integral equation approach and the perturbation theory can take account of the ion sizes, and not only the thermodynamic properties but also the detailed structural information can be obtained (Thiery and Dubessy, 1998, Thiery et al., 1998, Galindo et al., 1999). Jin and Donohue (1988) combined the long-range term from Henderson’s perturbation theory (Henderson, 1983) with perturbed anisotropic chain theory (PACT) EOS proposed by Vimalchand and Donohue (1985). Fürst and Renon (1993) tried to combine the long-range term from MSA with a modified Redlich–Kwong–Soave EOS. Considering the poor representation of the cubic EOS for the liquid density, Myers et al. (2002) added the long-range term from MSA to a volume-translated Peng–Robinson EOS. The model of Myers et al. (2002) can represent the activity coefficient, osmotic coefficient and density of aqueous NaCl solution from 273 to 573 K and from 1 to 86 bar. Wu and Prausnitz, 1998, Lin et al., 2007, Inchekel et al., 2008 added the long-range term from MSA to a CPA EOS (cubic EOS plus an association term). In recent years, many studies (Galindo et al., 1999, Liu et al., 1999, Behzadi et al., 2005, Liu et al., 2005, Liu et al., 2008, Tan et al., 2005, Tan et al., 2006, Zhao et al., 2007, Seyfi et al., 2009, Herzog et al., 2010) superimposed the long-range term from MSA on different SAFT-type EOS to model the thermodynamic properties of the H2O–NaCl, H2O–CaCl2 and other water–salt binary systems. The SAFT1E and SAFT2E EOS (Tan et al., 2005, Tan et al., 2006) were extended to some multi-salt–water systems by Ji et al., 2005a, Ji et al., 2006.

Although there are many EOS adopting the long-range term from MSA or perturbation theory as mentioned above, most of them are limited to room temperature and 1 bar or a very narrow PT range except the SAFT2E EOS and Myers’s EOS. The model of Myers et al. (2002) is suitable for the H2O–NaCl system from 273 to 573 K and from 0 to 86 bar, and the SAFT2E EOS improved by Ji and Adidharma (2007) can model the H2O–NaCl system from 273 to 473 K and from 0 to 1000 bar.

Some EOS for water–salts systems below 573 K have been extended to some gas–water–salt ternary systems (Jin and Donohue, 1988, Wu and Prausnitz, 1998, Patel et al., 2003, Ji et al., 2005b). Among them, only the model of Ji et al. (2005b) was designed for the CO2–H2O–NaCl system. Its calculation is accurate but the PT range is small (from 298 to 373 K, up to 200 bar). Jin and Donohue (1988) just calculated the vapor–liquid equilibrium of the CH4–H2O–NaCl system at 375 K. The model of Patel et al. (2003) can represent CH4 solubilities in water (xCH4) from 398 to 603 K but overestimate the content of H2O in vapor (yH2O) by more than 30%. The absolute average deviations of the model of Wu and Prausnitz (1998) for both xCH4 and yH2O are more than 20%. The inaccuracy of these models for ternary systems mainly arises from the inaccuracy for binary systems (water–gas binary system and H2O–salt binary system). In general, there is no model allowing the calculation of the thermodynamic properties of the H2O–NaCl and CO2–H2O–NaCl systems over 273–573 K and 0–1000 bar.

It’s important to develop an EOS suitable for the CO2–H2O–NaCl system over a wide PT range for the studies of fluid inclusion, metamorphic rocks, and CO2 sequestration so on. In the first companion paper (Sun and Dubessy, 2010), we improved the SAFT-LJ EOS proposed by Kraska and Gubbins (1996) and modeled the vapor–liquid equilibrium and PVTx properties of the H2O–CO2 system over a wide PT range. The purpose of this paper is to extend this EOS to the H2O–NaCl and CO2–H2O–NaCl systems below 573 K. Considering that the dissociation constant of NaCl in aqueous solution is greater than 1 at temperatures below 573 K (according to Mesmer et al., 1988) and the lower temperature limit of the EOS of Duan et al. (1995) for the CO2–H2O–NaCl system is 573 K, we chose 573 K as the upper temperature limit of the new EOS. Assuming that NaCl is fully dissociated in aqueous solution at temperatures below 573 K, the restricted primitive model of mean spherical approximation was included in SAFT-LJ EOS to account for the energetic contribution due to long-range electrostatic forces between ions. The theory of our model will be explained in the next section in detail. Section 3 describes the evaluation procedure of the parameters of the model. Section 4 compares the prediction of this model for the phase equilibrium and volumetric properties of the H2O–NaCl and CO2–H2O–NaCl systems with previous studies.

Section snippets

Models and theory

The SAFT-LJ EOS improved by Sun and Dubessy (2010) takes account of the energetic contribution of the main types of molecular interactions in the CO2–H2O system, such as the repulsion force, the dispersion force, the hydrogen-bonding force and the multi-polar forces. It can be extended to the CO2–H2O–NaCl system by adding a new term to account for the energetic contribution due to the long-range electrostatic force between ions considering that the Columbic force is the characteristic of

Evaluation of parameters

There are four species in the CO2–H2O–NaCl system: CO2, H2O, Na+ and Cl because NaCl is considered fully dissociated in aqueous solution below 573 K. The SAFT-LJ parameters for CO2 have been determined in the previous study (Sun and Dubessy, 2010). Because this study made a simplification for the calculation of parameter I in association term, the parameters for H2O and the binary parameters for CO2–H2O were re-evaluated, which are listed in Table 1, Table 2. Besides the energetic contribution

H2O–NaCl system

Fig. 1(a) compares the representation of the mean ionic activity coefficient (λ±m) of aqueous NaCl solution by our model with the experimental data reported by Liu and Lindsay, 1972, Gibbard et al., 1974 at low pressure (1 bar at temperature below 373 K or saturation pressure of pure water at temperature above 373 K). This model can represent the experimental data over a wide range of temperature from 273 to 573 K and a wide range of salinity from 0 to 6 mol/kg with high accuracy. The average

Conclusion

The SAFT-LJ EOS improved by Sun and Dubessy (2010) recently is extended to the H2O–NaCl and CO2–H2O–NaCl systems below 573 K. Assuming that NaCl is fully dissociated in aqueous solution, this study adopted the restricted primitive model of mean spherical approximation to account for the energetic contribution due to long-range electrostatic forces between ions. The association between Na+ and Cl and the association between ion and H2O are neglected. The SAFT-LJ parameters describing the

Acknowledgments

The authors acknowledge TOTAL Company for their financial support. The work is further financially supported by National Natural Science Foundation of China (Project 41073049), by MOST Special Fund from the State Key Laboratory of Continental Dynamics, Northwest University, and by the Science Foundation of Northwest University (No.09NW01). The authors acknowledge Matthew Steel-MacInnis, Denis Zezin and one anonymous reviewer for their detailed, helpful and pertinent comments which improved

References (104)

  • C.A. Haynes et al.

    On converting from the McMillan–Mayer framework I. Single-solvent system

    Fluid Phase Equilib.

    (1998)
  • S. Herzog et al.

    Equation of state for aqueous electrolyte systems based on the semirestricted non-primitive mean spherical approximation

    Fluid Phase Equilib.

    (2010)
  • L. Hnedkovsky et al.

    Volumes of aqueous solutions of CH4, CO2, H2S and NH3 at temperatures from 298.15 K to 705 K and pressures to 35 MPa

    J. Chem. Thermodyn.

    (1996)
  • J. Hu et al.

    PVTx properties of the CO2–H2O and CO2–H2O–NaCl systems below 647 K: assessment of experimental data and thermodynamic models

    Chem. Geol.

    (2007)
  • R. Inchekel et al.

    The simultaneous representation of dielectric constant, volume and activity coefficients using an electrolyte equation of state

    Fluid Phase Equilib.

    (2008)
  • M.B. King et al.

    The mutual solubilities of water with supercritical and liquid carbon dioxide

    J. Supercrit. Fluids

    (1992)
  • W.B. Liu et al.

    A new equation of state for real aqueous ionic fluids based on electrolyte perturbation theory, mean spherical approximation and statistical associating fluid theory

    Fluid Phase Equilib.

    (1999)
  • Z. Liu et al.

    An equation of state for electrolyte solutions by a combination of low-density expansion of non-primitive mean spherical approximation and statistical associating fluid theory

    Fluid Phase Equilib.

    (2005)
  • V. Majer et al.

    Volumetric properties of aqueous sodium chloride solutions from 0.0025 to 5.0 mol kg−1, 323 to 600 K, and 0.1 to 40 MPa

    J. Chem. Thermodyn.

    (1988)
  • C. Schmidt et al.

    Synthetic fluid inclusions: XVI. PVTX properties in the system H2O–NaCl–CO2 at elevated temperatures, pressures, and salinities

    Geochim. Cosmochim. Acta

    (2000)
  • S. Seyfi et al.

    On the prediction of equilibrium phase behavior of amino acids in aqueous and aqueous-electrolyte solutions using SAFT equation of state

    Fluid Phase Equilib.

    (2009)
  • N. Spycher et al.

    CO2–H2O mixtures in the geological sequestration of CO2. II. Partitioning in chloride brines at 12–100 °C and up to 600 bar

    Geochim. Cosmochim. Acta

    (2005)
  • N. Spycher et al.

    CO2–H2O mixtures in the geological sequestration of CO2. I. Assessment and calculation of mutual solubilities from 12 to 100 °C and up to 600 bar

    Geochim. Cosmochim. Acta

    (2003)
  • R. Sun et al.

    Prediction of vapor–liquid equilibrium and PVTx properties of geological fluid system with SAFT-LJ EOS including multi-polar contribution. Part I. application to H2O–CO2 system

    Geochim. Cosmochim. Acta

    (2010)
  • H. Teng et al.

    Solubility of liquid CO2 in water at temperatures from 278 K to 293 K and pressures from 6.44 MPa to 29.49 MPa, and densities of the corresponding aqueous solutions

    J. Chem. Thermodyn.

    (1997)
  • K. Aasberg-Petersen et al.

    Prediction of high-pressure gas solubilities in aqueous mixtures of electrolytes

    Ind. Eng. Chem. Res.

    (1991)
  • D.G. Archer

    Thermodynamic properties of the NaCl + H2O system II. Thermodynamic properties of NaCl(aq), NaCl·2H2O(cr), and phase equilibria

    J. Phys. Chem. Ref. Data

    (1992)
  • F.X. Ball et al.

    Representation of deviation from ideality in concentrated aqueous solutions of electrolytes using a mean spherical approximation molecular model

    AIChE J.

    (1985)
  • S. Bando et al.

    Solubility of CO2 in aqueous solutions of NaCl at (30 to 60) °C and (10 to 20) MPa

    J. Chem. Eng. Data

    (2003)
  • J.L. Bischoff et al.

    Liquid–vapor relations for the system NaCl–H2O: summary of the P–Tx surface from 300 to 500 °C

    Am. J. Sci.

    (1989)
  • L. Blum et al.

    Mean spherical model for asymmetric electrolytes. 2. Thermodynamic properties and the pair correlation function

    J. Phys. Chem.

    (1977)
  • L. Blum et al.

    Analytical solution of the mean spherical approximation for an arbitrary mixture of ions in a dipolar solvent

    J. Chem. Phys.

    (1987)
  • L. Blum et al.

    On the mean spherical approximation for hard ions and dipoles

    J. Chem. Phys.

    (1992)
  • C.-T. Chen et al.

    The apparent molal volumes of aqueous solutions of sodium chloride, potassium chloride, magnesium chloride, sodium sulfate, and magnesium sulfate from 0 to 1000 bars at 0, 25, and 50 °C

    J. Chem. Eng. Data

    (1977)
  • E.C.W. Clarke et al.

    Evaluation of the thermodynamic functions for aqueous sodium chloride from equilibrium and calorimetric measurements below 154 °C

    J. Phys. Chem. Ref. Data

    (1985)
  • J. Dubessy et al.

    Advances in C–O–H–N–S fluid geochemistry based on Raman analysis of fluid inclusions

    Eur. J. Mineral.

    (1989)
  • J. Dubessy et al.

    Modelling of liquid–vapour equilibria in the H2O–CO2–NaCl and H2O–H2S–NaCl systems up to 270 °C

    Oil Gas Sci. Technol.

    (2005)
  • W. Fürst et al.

    Representation of excess properties of electrolyte-solutions using a new equation of state

    AIChE J.

    (1993)
  • A. Galindo et al.

    SAFT-VRE: phase behavior of electrolyte solutions with the statistical associating fluid theory for potentials of variable range

    J. Phys. Chem. B

    (1999)
  • J.A. Gates et al.

    Densities of aqueous solutions of sodium chloride, magnesium chloride, potassium chloride, sodium bromide, lithium chloride, and calcium chloride from 0.05 to 5.0 mol kg−1 and 0.1013 to 40 MPa at 298.15 K

    J. Chem. Eng. Data

    (1985)
  • Gehrig M. (1980) Phasengleichgewichte und pVT-daten ternarer mischungen aus wasser, kohlendioxid und natriumchlorid bis...
  • H.F. Gibbard et al.

    Liquid–vapor equilibrium of aqueous sodium chloride, from 298 to 373 K and from 1 to 6 mol kg−1, and related properties

    J. Chem. Eng. Data

    (1974)
  • R.E. Gibson et al.

    Pressure–volume–temperature relations in solutions. VIII. The behavior of some solutions of electrolytes in water

    Ann. NY Acad. Sci.

    (1949)
  • D.F. Grant-Taylor

    Partial molar volumes of sodium chloride solutions at 200 bar, and temperatures from 175 to 350 °C

    J. Solution Chem.

    (1981)
  • Haas J. L. (1976) Physical properties of the coexisting phases and thermochemical properties of the H2O component in...
  • H.S. Harned

    The electromotive forces of uni-univalent halides in concentrated aqueous solutions

    J. Am. Chem. Soc.

    (1929)
  • H.S. Harned et al.

    The thermodynamic properties of aqueous sodium chloride solutions from 0 to 40°

    J. Am. Chem. Soc.

    (1932)
  • A.H. Harvey et al.

    Thermodynamics of high-pressure aqueous systems containing gases and salts

    AIChE J.

    (1989)
  • A. Hebach et al.

    Density of water + carbon dioxide at elevated pressures: measurements and correlation

    J. Chem. Eng. Data

    (2004)
  • W. Heinrich

    Fluid immiscibility in metamorphic rocks

    Rev. Mineral. Geochem.

    (2007)
  • Cited by (0)

    View full text