Monte Carlo simulations of complete phase diagrams for binary Lennard–Jones mixtures

https://doi.org/10.1016/S0378-3812(01)00378-8Get rights and content

Abstract

Vapor–liquid, vapor–solid, liquid–liquid, and liquid–solid coexistence lines are calculated for binary mixtures of Lennard–Jones spheres with diameter ratio σ11/σ22=0.85, well-depth ratio ϵ11/ϵ22=0.45, and binary interaction parameters δ12=1.0, 0.9, and 0.75, using Monte Carlo simulation and the Gibbs–Duhem integration technique. These calculations allow us to construct complete phase diagrams, i.e. showing equilibrium between vapor, liquid, and solid phases. For the mixture with δ12=1, we find a completely miscible vapor–liquid coexistence region with a eutectic solid–liquid coexistence region. These two regions are separated by a completely miscible liquid phase. For the mixtures with δ12<1, we find that the vapor–liquid and solid–liquid coexistence regions interfere. This interference results in a vapor–solid coexistence region bounded above and below by solid–liquid–vapor coexistence lines. We also find that the mixtures with δ12<1 have a region of liquid–liquid immiscibility that is metastable with respect to the solid–fluid phase equilibria.

Introduction

Phase equilibria for binary mixtures of spherically symmetric molecules has been the subject of intensive investigation for decades. Van Konynenburg and Scott [1], [2] classified the phase diagrams predicted by the van der Waals equation of state for binary mixtures. Remarkably, the simple van der Waals equation of state was found to exhibit five of the six types of fluid phase behavior observed experimentally. This landmark study has been followed by similar analyses for other equations of state, such as the Redlich–Kwong [3], the Carnahan–Starling–Redlich–Kwong [4], the Guggenheim [5], and the Ree [6] equations of state.

Most of the research directed at understanding how intermolecular interactions affect phase behavior has focused exclusively on fluid phase equilibria. However, in real systems solid phases form and often interrupt the complex fluid phase behavior [7], [8]. Phenomenological descriptions of complete phase diagrams (i.e. showing equilibrium between vapor, liquid, and solid phases) have been given by Luks [9], Valyashko [10], [11], and Peters et al. [12]. Valyashko proposed a classification scheme for complete diagrams and introduced new types of complete phase behaviors that have not yet been observed in real systems. Garcia and Luks [13] recently examined the solid–liquid–vapor locus for a series of binary mixtures using the van der Waals equation of state and a simple solid state fugacity model. In their model calculations, they found examples of solid–fluid phase behavior in keeping with what has been observed in real systems, as well as solid–fluid phase behavior that has yet to be verified by experiment. The new possibilities for complete phase behavior proposed by Valyashko and by Garcia and Luks are intriguing and invite further investigation.

Molecular simulation has become a popular way to study the phase behavior of model systems [14]. Numerous simulations of binary mixture phase behavior for the Lennard–Jones intermolecular potential, the quintessential model of a spherically symmetric molecule, have been conducted for both fluid–fluid [15], [16], [17], [18], [19], [20], [21] and solid–liquid [22] phase equilibria. Vlot et al. [23] calculated complete phase diagrams for symmetric (equal diameters, σ11=σ22; equal attractions, ϵ11=ϵ22) Lennard–Jones mixtures by using Monte Carlo simulation for selected state points to determine the excess free energy as a function of composition. This free energy versus composition data was fitted with a two-parameter Redlich–Kister polynomial, and the convex envelope construction method was used to determine the phase diagram.

The Lennard–Jones intermolecular potential is given byuij(r)=4ϵijσijr12σijr6,where uij is the potential energy of interaction between particles i and j, r the distance between particles i and j, ϵij the Lennard–Jones attractive well-depth, and σij the Lennard–Jones diameter. The cross-species interaction parameters (σ12,ϵ12) are determined by the Lorentz–Berthelot combining rules [24]σ12=σ11222,ϵ1212ϵ11ϵ22,where δ12, the binary interaction parameter, accounts for deviations of the unlike-pair attraction from the geometric mean of the like-pair attractions. According to the global phase diagrams associated with the various equations of state [2], [3], [4], [5], [6] for binary mixtures of equal-size, spherically symmetric molecules, liquid–liquid immiscibility occurs whenever the unlike-pair attractions are less than the arithmetic mean of the like-pair attractions. This means that liquid–liquid immiscibility is predicted to occur whenδ12<ϵ11222ϵ11ϵ22.For binary mixtures of unequal-size, spherically symmetric molecules, liquid–liquid immiscibility is predicted to occur whenδ12<ϵ11221122)32ϵ11ϵ221122+1)3.Liquid–liquid immiscibility might not be observed however, because of intervening solid phases, which cannot be predicted by the equations of state. In fact, Rowlinson and Swinton [24] point out that unless the unlike-pair attractions are significantly weaker than the like-pair attractions, the liquid–liquid upper critical solution temperature resides at temperatures below the quadruple point (s1s2lv).

Systematic studies of the effect of δ12 on fluid phase behavior have been conducted for two Lennard–Jones binary mixtures [15], [17] using Gibbs ensemble simulation. In both cases, δ12 was varied over the range 1.0–0.7 to observe the effect of the unlike-pair attractions on the constant temperature Px phase diagram. In the first mixture (σ11/σ22=0.82, ϵ11/ϵ22=0.52), Panagiotopoulos et al. [15] found that liquid–liquid immiscibility occurred for δ12≤0.8 at kT/ϵ11=1.348. In the second mixture (σ11/σ22=0.94, ϵ11/ϵ22=0.73), van Leeuwen et al. [17] found that liquid–liquid immiscibility occurred for δ12≤0.75 at kT/ϵ11=1.218. A possible explanation for why liquid–liquid immiscibility was not observed for δ12=1.0 in these two studies is that the simulations were conducted at fixed temperatures that were higher than the upper critical solution temperature.

In this paper, we calculate complete phase diagrams for binary Lennard–Jones mixtures using Gibbs–Duhem integration combined with semigrand canonical Monte Carlo simulation. Our objective is to explore the effect of δ12 on the phase behavior of a mixture when solid phase formation is included in the calculation. We present complete phase diagrams for binary Lennard–Jones mixtures with diameter ratio σ11/σ22=0.85, well-depth ratio ϵ11/ϵ22=0.45, and binary interaction parameters δ12=1.0, 0.9, and 0.75 at reduced pressure, P=Pσ11311=0.05.

The remainder of this paper is organized as follows. We first outline the Gibbs–Duhem integration method and describe how we applied the procedure to the calculation of complete phase behavior. We then present the resulting complete phase diagrams.

Section snippets

Integration

The coexistence lines were calculated using Gibbs–Duhem integration [25], [26], [27]. In this method, phase coexistence is determined by numerically integrating the Clapeyron differential equation appropriate to the system of interest. Clapeyron equations describe how field variables (variables that must be equal among coexisting phases) change along the phase equilibrium line. The Clapeyron equation for equilibrium between two phases (α and γ) of a binary mixture containing components 1 and 2

Results

In this section, we present the results of our Gibbs–Duhem integration calculations of complete phase behavior for binary Lennard–Jones mixtures.

We calculated three complete phase diagrams at P=Pσ11311=0.05 for diameter ratio σ11/σ22=0.85, well-depth ratio ϵ11/ϵ22=0.45, and binary interaction parameters δ12=1.0, 0.9, and 0.75. According to Eq. (3) (based on equation of state predictions), liquid–liquid immiscibility is predicted to occur for this mixture whenever δ12<1.002. Fig. 2 shows a

Summary

The Gibbs–Duhem integration technique was combined with semigrand canonical Monte Carlo simulations to calculate complete Tx phase diagrams for binary Lennard–Jones mixtures. We studied the effect of unlike-pair attractions on the mixture phase diagram by varying the binary interaction parameter over the range, δ12=1–0.75, while holding the diameter ratio and well-depth ratio constant.

For the mixture with δ12=1, we found a completely miscible vapor–liquid coexistence region and a eutectic

Acknowledgments

We gratefully acknowledge helpful discussions with Professor John M. Prausnitz. This work was supported by the GAANN Computational Sciences Fellowship of the U.S. Department of Education and the Office of Energy Research, Basic Sciences, Chemical Science Division of the U.S. Department of Energy under contract no. DE-FG05-91ER14181. Acknowledgement is made to the Donors of the Petroleum Research Fund administered by the American Chemical Society for partial support of this work.

References (33)

  • K.D. Luks

    Fluid Phase Equilib.

    (1986)
  • C.J. Peters et al.

    Fluid Phase Equilib.

    (1986)
  • D.C. Garcia et al.

    Fluid Phase Equilib.

    (1999)
  • V.I. Harismiadis et al.

    Fluid Phase Equilib.

    (1991)
  • M.E. van Leeuwen et al.

    Fluid Phase Equilib.

    (1991)
  • M. Guo et al.

    Fluid Phase Equilib.

    (1994)
  • R.J. Sadus

    Fluid Phase Equilib.

    (1999)
  • M. Mehta et al.

    Chem. Eng. Sci.

    (1994)
  • R.L. Scott et al.

    Discuss. Faraday Soc.

    (1970)
  • P.H. van Konynenburg et al.

    Phil. Trans. R. Soc. London A

    (1980)
  • U.K. Deiters et al.

    J. Chem. Phys.

    (1989)
  • T. Kraska et al.

    J. Chem. Phys.

    (1992)
  • J. Wang et al.

    Mol. Phys.

    (2000)
  • V.A. Mazur et al.

    Phys. Lett.

    (1984)
  • G.M. Schneider, in: M.L. McGlashan (Ed.), Chemical Thermodynamics, Specialist Periodical Report, Vol. 2, Chemical...
  • R.L. Scott

    Acc. Chem. Res.

    (1987)
  • Cited by (38)

    • A molecular simulation approach to the computation of mutual solubility of water and organic liquids: Application to fatty acids

      2018, Fluid Phase Equilibria
      Citation Excerpt :

      Gibbs ensemble Monte Carlo (GEMC), which was proposed and originally illustrated for vapor-liquid equilibrium and simple osmotic systems by Panagiotopoulos [13,14], remains one of the most widely used techniques for LLE predictions. It has been extended to simple Lennard-Jones systems and complex molecular mixtures [15–27]. One of the key steps in the Gibbs ensemble method is the transfer of molecules between the two phases to achieve chemical equilibrium.

    • A review of molecular simulation applied in vapor-liquid equilibria (VLE) estimation of thermodynamic cycles

      2018, Journal of Molecular Liquids
      Citation Excerpt :

      Furthermore, Budinský et al. [91] studied the vapor-liquid coexistence of mixtures. Lamm and Hall [139] calculated a complete T-x diagram for binary L-J mixtures, and phases including solid, liquid and vapor were presented. Paschek et al. [140] determined activity coefficients of binary L-J mixtures using GDI.

    • A Monte Carlo simulation study of the liquid-liquid equilibria for binary dodecane/ethanol and ternary dodecane/ethanol/water mixtures

      2015, Fluid Phase Equilibria
      Citation Excerpt :

      Additional sampling challenges arise when some of the molecules can aggregate. Thus, most of the molecular simulation studies for LLE have focused on relatively simple model systems [4–15]. More recently, simulation studies have appeared for the LLE of more complex mixtures, but many of these studies report sampling difficulties that could only be overcome with specialized techniques [16–25].

    • Application of the SLV-EoS for representing phase equilibria of binary Lennard-Jones mixtures including solid phases

      2013, Fluid Phase Equilibria
      Citation Excerpt :

      At the same time, Lennard–Jones (LJ) fluids can be considered as very good approximations of the real “simple” molecules objective of this study. Phase diagrams including solid phases for binary mixtures of LJ components have been studied using Monte-Carlo simulation by Lamm, Hall, and Hitchcock [7–10]. Reduced pressure for solid–liquid and solid–gas equilibrium data of the pure LJ fluid were taken from Ref. [12].

    View all citing articles on Scopus
    View full text