A replica Ornstein-Zernike self-consistent theory for mixtures in random pores G. Pellicane, C. Caccamo Dipartimento di Fisica, Università di Messina and Istituto Nazionale per la Fisica della Materia (INFM) D. S. Wilson and L. L. Lee School of Chemical Engineering and Materials Science, University of Oklahoma, Norman, Oklahoma 73019 We present a self-consistent integral equation theory for a binary liquid in equilibrium with a porous medium, based on the formalism of the replica Ornstein-Zernike (ROZ) equations. Specifically, we derive direct formulas for the chemical potentials and the zero-separation theorems (which provide a connection between the chemical potentials and the fluid cavity distribution functions). Next we solve a modified-Verlet closure to the replica OZ equations, which has built-in parameters that can be adjusted to satisfy the zero-separation theorems. The degree of thermodynamic consistency of the theory is kept under control by monitoring all members of the Gibbs-Duhem relation. We model the binary fluid in random pores as a symmetrical binary mixture of non-additive hard spheres in a disordered hard-sphere matrix and consider two different values for the nonadditivity parameter and for the quenched matrix packing fraction, at different mixture concentrations. We compare the theoretical structural properties with Percus-Yevick (PY) and Martinov-Sarkisov (MS) integral equations and assess both structural and thermodynamic properties by performing NVT, standard and biased VT Monte Carlo simulations. Our theory appears superior to the other integral equation schemes here 1 examined and gives reliable estimates of the chemical potentials. This feature should be useful in studying liquid-liquid demixing behavior in general. 2 I. Introduction Phase changes of liquids inside porous media are of both scientific and practical interests. Porous materials, such as activated carbon, silica gels, zeolites, and pillared clay and more recently, aerogels, aminosilicates, aluminophosphate, carbon nanotubes, Vycor glass, microporous BN, and starburst dendrimers, have been extensively used in industry for adsorption, dehumidification, catalysis, gas separation [1] and gas storage[2] [3]. Many of these materials have an amorphous structure, that is, they consist of micro-sized pores that are irregularly distributed throughout the material. For example, aerogels have a cobweb-like structure that is made up of cross interconnecting inorganic or organic colloidal-like particles or polymeric chains with high porosity (75-99%)[4]. When fluids invade the interior, the confinement in narrow dimensions affects the fluid phase behavior, as has been well documented. Few studies have looked into binary fluid mixtures in pores. The simplest model for such a mixture, exhibiting phase changes, is the nonadditive hard sphere mixtures (NAHSM) in random pores. In this work we employ both integral equations (IE) and molecular simulations to characterize the liquidliquid phase behavior of NAHSM under confinement. Such studies should shed light on the size and nonadditivity effects on phase properties in general. The integral equations of the Replica type will be solved for binary nonadditive hard sphere mixtures in pores. For the systems chosen, these equations are afferent. We have applied the replica Ornstein-Zernike (ROZ) [5] equations earlier for pure spheres in pores [6] and obtained accurate results as compared to simulation data. A number of 3 other publications have also investigated the mixtures in pores [7][8]. In our approach, we improve the performance of the OZ integral equations by requiring thermodynamic consistency and conformity to the zero-separation theorems. This combined methodology has been applied and tested in a number of previous studies: on pure hard spheres in pores [9] and in bulk [10], additive [11] and nonadditive hard sphere mixtures in bulk [12], Lennard-Jones molecules [13], and diatomic hard dumbbells [14] as well. In all cases, close agreement was obtained between IE and simulation. The essence of such an approach is the use of the zero separation theorems [15], in addition to the usual thermodynamic consistencies [16], that the values of correlation functions at coincidence (when two particles coincide and the separation distance r=0) obey certain exact conditions that tie them to the thermodynamic properties of the system under study, such as the chemical potentials and isothermal compressibility. This is, in a way, similar to the contact value theorem (e.g., at g(d+), the contact value of the pair correlation g(r)) for hard-core systems [17]. Since these relations are exact, there is all the reason that they should be satisfied for consistency. The importance of ZST is in the “local” nature: that specific values of the correlation functions are required at specific distances r (=0, or =d), in contrast to the “global” (or integral or almost everywhere) nature of the thermodynamic consistencies. Both global and local consistencies reinforce the “accuracy” in the closure formulation. In this work, we develop new zero-separation theorems for our particular system (NAHSM gases in hard sphere pores). There will be five (5) monomer chemical potentials, and seven (7) “dimer” chemical potentials. Consequently, there are eleven 4 (11) coincidence values for the cavity functions (twice for the y10(0) , y20(0) , and y12(0) functions). These zero-separation values are used in constructing the closure relations in the ROZ. In order to test the accuracy of the theoretical calculations, we have carried out three types of Monte Carlo (MC) simulations: the grand canonical ensemble MC, the NVT canonical ensemble MC, and the cavity-biased grand canonical ensemble MC. These simulations are used to generate the correlation functions, chemical potentials, and phase envelopes to check the theoretically derived behavior for the NAHSM in pores. Fifteen conditions are investigated, representing different nonadditivities (Δ) of hard sphere species, matrix porosities (ρ0), gas densities (ρ*), and compositions of mixtures (x). The two hard sphere species, 1 and 2, are of equal size (diameters σ11 = σ22 = σ), except for their nonadditive cross interaction (σ12 = ½(σ11 +σ22 )(1+ Δ)). Two values of Δ = 0.2, and 0.4 are used (i.e., positive nonadditivity). The gas densities ρ* = ρσ3, range from 0.18 (low porosity) to 0.3 (high porosity). The matrix number density ρ0 , reduced by the fluid sphere diameter σ , varies from 0.1 to 0.3. The compositions (mole fractions) are around 0.1, 0.3, and 0.5. This range of variables is needed to characterize the liquidliquid phase diagram. (Note that our conditions of simulation are similar to those of Sierra and Duda [8]). However, we used different MC algorithms— other than the semigrand canonical ensembles, and we used the ZSEP closures for the ROZ equations in addition to the PY or MS closures. In Section II, we present the theory, the Replica OZ as applied to our system, and the ZSEP closure. We give detailed formulas for the zero-separation theorems on the cavity functions, yij(0). The details of our simulation algorithms are also described. In Section 5 III, we show the results of both integral equations and MC simulation data. We also give concluding remarks in the section. II. The Model and the Methods A. Model We consider a binary mixture of non-additive hard-sphere particles: Vij (r ) where r ij 0 r ij 12 (1 ) (1) , ( 11 22 ) 11 , 22 2 and is the non- additive parameter. The fluid mixture is confined in a matrix of random, non-overlapping hard-sphere particles. Within this model for the fluid mixture, a positive non-additivity in the cross interaction between the two species favours homocoordination in comparison to hetero-coordination. As a consequence of the extra-repulsion from unlike interactions, non-additive hard sphere mixtures are expected to separate into two phases at sufficiently high density. We considered this model because it is the simplest mixture possible to exhibit a thermodynamically stable liquid-liquid transition. In all the calculations we assume that the size of the two species is equal to the diameter of the matrix sphere, i.e. 11 = 22=0, briefly stated we considered symmetrical mixtures. Then, we study the effect of the 6 increasing tendency to homocoordynation at 20 % and 40 % compositions ( on the structure and the thermodynamics of the system. B. Theory The formalism of the ROZ equations was derived for the first time for pure fluids by Madden and Gland and Given and Stell [5], and recently extended to binary liquids by Paschinger and Kahl. [7] These equations describe the structure of a binary liquid inside a porous matrix and are formally derived after a partial quenching of one of the components and in the limiting case for s=0 of a fully equilibrated (2s + 1)-component mixture, where s is the number of identical copies of the liquid mixture (the replicas); thus, within this formalism, the mixture confined into the rigid matrix is depicted as a semi-quenched system (partly quenched and partly annealed). The label 0 indicates the matrix species; 1 and 2 the liquid mixture species; 3 and 4 any replica of 1 and 2 (odd numbers for replica of 1, and even numbers for replicas of 2), respectively. In the following, we adopt the matrix notation of [8] for the correlation functions hij(r) and cij(r), and we suppress the argument r in order to improve readability; the Replica OZ equations are so written as (“T” denotes the transpose of a vector and is a convolution ): h00 c00 0 c00 h00 h 01 c 01 h 01 c00 1h 11 c 01 1h 12c 01 h 11 c11 h 01 c T0 1 1h 11 c11 1h 12c12 h 12 c12 h 01 c T0 1 1h 12 c11 1h 12c12 21h 12 c12 where 7 (2) h c ρ 0 , h 01 01 , c 01 01 (3) 1 1 0 ρ2 h 02 c 02 h h 01 11 h 12 h h 12 13 h 14 h 12 c , c11 11 h 22 c12 h 14 c , c12 13 h 24 c14 c12 , c 22 (4) c14 . c 24 The total number of the coupled ROZ equations is 8, while the first one is the bulk OZ equation for the matrix particles. The ROZ equations are solved under different closures, for example, the modified-Verlet closure [18]: Bij (r ) ij 2 (5) , *ij 1 ij * 2 1 ij ij ij where Bij(r) and ij*(r) are the bridge and the renormalized indirect correlation functions, respectively; the renormalization of the indirect correlation function was achieved by adding a soft-Weeks Chandler Andersen (WCA) potential [19] to the ij function 1 * (r ) (r ) f (r ) , where f(r) is the Mayer factor of the repulsive WCA 6:3 potential. 2 The introduction of this term circumvents the discontinuities in the bridge functions when combined with the closure; we slightly modified the indirect correlation function by setting the pseudo-temperature in the Mayer factor to 9999. The 24 adjustable parameters ij, ij and ij (3 parameters for each of the 8 ROZ equations) appear to be excessive and we drastically reduce the number via a consistent procedure by letting them assume “reasonable” values according to the following criteria: while we always set ij=1, we use 8 a two different sets of parameters for the equimolar and asymmetric concentrations, respectively. For the equimolar concentrations we set 11SYM=22 SYM, 12 SYM =14 SYM =12 SYM =14 SYM, 10 SYM =20 SYM ,33 SYM =44 SYM, 11 SYM =22 SYM , 10 SYM =20 SYM ,33 SYM =44 SYM; for the asymmetric concentrations we fix the values of the ij to the symmetrical concentration ones, ij=ijSYM and we allow all the ij to vary under the constraint 12=14. In both cases, we end up with 7 parameters to be determined by enforcing self-consistency theorems. We used zero separation theorems, which relate the cavity functions values at zero distance to the proper expression for the chemical potentials. In a previous paper, one of us reported the formulas of the chemical potentials for common mixtures [12][20]; here we derive these formulas for the system representing the matrix and the mutually noninteracting s copies (replicas) of the fluid molecules. We introduce the following short-hand notation: 1 k ij k dr log y ij hij hij ij hij Bij S ij (6) , 2 where ij(r) and Sij(r) are respectively the indirect correlation functions and the Star functions introduced earlier [20], and k is the number density of specie k. Thus, the monomer chemical potentials are easily written as (all chemical potentials are configurational quantities, in excess of the ideal part): 0 0 00 s 1 01 s 2 02 1 0 10 1 11 (s 1 - 1) 3 13 2 12 (s 2 - 1) 4 14 2 0 20 2 22 (s 2 - 1) 4 24 1 21 (s 1 - 1) 3 23 44 43 (s (7 ) 43 3 0 30 3 33 (s 1 - 1) 3' 33' 4 34 (s 2 - 1) 4' 34 ' 4 0 40 4 44 (s 2 - 1) 4' ' 3 9 1 - 1) 3' ' where 3’ and 4’ are replicas of 1 and 2 other than 3 and 4, respectively, and =1/kBT is the inverse temperature in Boltzmann constant kB units. Now, the reversible works of insertion of dimers made up from pairs of species 1+1, 2+2 , 1+2, 1+0,2+0,1+3,2+4 merged up to zero bond length (at infinite dilution in the mixture) are the chemical potentials for dimers of the corresponding pairs at close contact: 11 0 1 1 0 1 1 11 (s 1 - 1) 3 1 1 3 2 1 1 2 (s 2 - 1) 4 1 1 4 2 2 0 2 2 0 2 2 2 2 (s 2 - 1) 4 2 2 4 1 2 2 1 (s 1 - 1) 3 2 2 3 1 0 0 1 0 0 1 1 0 1 (s 1 - 1) 3 1 0 3 2 1 0 2 (s 2 - 1) 4 1 0 4 2 0 0 2 0 0 2 2 0 2 (s 2 - 1) 4 2 0 4 1 2 0 1 (s 1 - 1) 3 2 0 3 1 2 0 1 2 0 1 1 2 1 (s 1 - 1) 3 1 2 3 2 1 2 2 (s 2 - 1) 4 1 2 4 (8) 13 0 1 3 0 1 1 3 1 3 1 3 3 2 1 3 2 4 1 3 4 (s 1 - 2) 3' 1 3 3' (s 2 - 2) 4' 1 4 4 ' 2 4 0 2 4 0 2 2 4 2 4 2 4 4 1 2 4 1 3 2 4 3 (s 2 - 2) 4' 2 4 4 ' (s 1 - 2) 3' 2 4 3 ' These formulas are general and are applicable to any type of pair potential of interaction. Now we shall specialize the above relations to the case of non-additive hard-sphere mixture in the hard-sphere matrix and we shall also formulate the zero-separation theorems. For the 11 cavity functions: log y11 0 21 11 1 (9) . Thus, we have (upon taking the limit s→0): log y11 0 0 10 1 11 - 3 13 2 12 4 14 (10) . 10 In a similar way, we can build up the zero-separation theorems for the other cavity functions: log y 22 0 0 20 2 22 - 4 24 1 21 - 3 23 log y13 0 0 10 (11) . log y 24 0 0 20 We note that the coincidence value of the mixture-replica cavity functions depends only on the matrix-fluid correlations, similar to the pure case as reported in literature. The case of log y10(0), log y20(0) and log y12(0) is a bit more complicated and we have to distinguish between different cases. For the unlike cavity functions we have: log y12 0 0 20 1 11 2 22 2 24 1 23 log y12 0 0 20 1 11 2 22 1 13 2 14 1 2 1 2 (12) , while for the particle-matrix cavity functions: log y10 0 0 00 1 13 2 14 1 01 2 02 log y10 0 0 01 1 11 1 13 2 12 2 14 log y 20 0 0 00 2 24 1 23 2 02 1 01 log y 20 0 0 02 2 22 2 24 1 12 1 23 1 0 1 0 2 0 (13) . 2 0 Thus, the closure equation () to the ROZ equations is solved numerically under the constraint determined by the 7 eq.s (11)-(14). We also checked the ability of this closure to guarantee internal thermodynamic consistency. Unfortunately, the virial equation of state for the quenched-annealed system is not straightforwardly related to structural functions as for bulk mixtures [21]; in fact for hard-core interactions: 11 P P 0 0 V T 1 2 (14) dg 00 00 ; s 2 2 3 i 0 00 lim i2 3ii g ii ( ii ) 2 0 i 30i g 0i ( 0i ) 21 2 123 g 12 (12 ) s 0 3 ds i i 1, 2 i 1, 2 , where the first term is the infinitesimal change of the virial pressure with respect to the matrix density at constant values of volume, temperature and chemical potentials of the dg 00 00 ; s is the infinitesimal change of the matrix-matrix radial s 0 ds two species, and lim distribution function in the replicated system with respect to the number of replicas in the limit of the quenched-annealed system. Both the two quantities should be evaluated numerically, and in particular the first one implies the calculation of the product of two numerical derivatives as a standard mathematical manipulation of thermodynamic derivatives shows: P P 0 V T 1 2 0 V T N1 N 2 P i 1, 2 i V i 0 V T N1 N 2 (15) T N1 N 2 Thus, we resorted to a different strategy in order to monitor the thermodynamic consistency of the theory, by using the Gibbs-Duhem relation [13]: P V T N 0 i j i i 1, 2 j j 1, 2 V ~c k BT T 1 1 i j c ij (0) (16) , i , j 1, 2 T N k j 0 where the chemical potentials of the two species are differentiated numerically in order to ~c obtain the inverse of the isothermal compressibility and where the c ij (0) are the Fourier Transforms at zero wave vector of the connected part of the direct correlation functions, within the ROZ formalism. 12 The ROZ equations were also solved under the Percus-Yevick (PY) [22] closure: cij (r ) exp Vij ( r ) 1 1 ij (r ) (17) , and the Martinov-Sarkisov (MS) [23] closure: cij (r ) exp Vij ( r ) 1 10.5 ij ( r ) (18) . We solved numerically the different closures to the ROZ equations with a Picard method and adopted a standard mixing procedure for the direct correlation function in order to ensure convergence; normally, 1024 grid points with grid interval 0.010 were used but we also checked the stability of the solution with higher grids of 2048, 4096 points for some thermodynamic state points. Zero separation theorems and the thermodynamic constraint (16) were enforced in the modified-Verlet closure (ZSEP) with a tolerance of 1 %. C. Simulation We have performed Monte Carlo simulations with three algorithms [24]: the grand canonical (T, V, 1, 2) ensemble (GCMC) [25], the standard NVT canonical ensemble, and the cavity-biases grand canonical (CB-GCMC) ensemble [26]. The GCMC is used in order to sample different matrix realizations and fluid mixture particle configurations. Three types of particles moves (displacement, creation or destruction of a particle) were performed randomly with equal probability inside a cubic box with cubic periodic boundaries; typically, averages were taken over 30 matrix realizations and for each realization not less than 100000 trial moves per particle were generated. 13 We also increased the efficiency of the importance sampling by performing simulations in the cavity biased grand canonical (T, V, 1, 2) ensemble (CB-GCMC), where the acceptance probability of a molecular configuration is biased so to attempt insertions of new particles into existing cavities in the system instead of at randomly selected points. In particular, insertion and deletion attempts were accepted with the probabilities: exp V (r N ) V (r N 1 ) Pi min 1,VPcN (r N ) 3 ( N 1) exp V (r N ) V (r N 1 ) Pd min 1, N3 V PcN 1 (r N -1 ) (19) , where PcN (r N ) is the configuration-dependent probability of finding a cavity of diameter c or larger, provided the system consists of N particles and is the DeBroglie thermal wavelength. We estimated PcN (r N ) by implementing a cavity search using a finite grid. We always used a box of side length equal to 13.5, and a uniform grid of 1003 points; the grid size was chosen so as to be computationally manageable while generating a negligible error in the limiting distribution of the Markov chain, since the deletion can occur at any point in the box but the insertions are restricted to the grid points. Finally, we performed standard constant number of particles, constant volume and constant temperature (NVT) simulations from previously equilibrated GCMC simulations in order to evaluate the radial distribution functions. 14 III. Results and discussion We started to assess the structure of integral equation closures against standard NVT simulations. In figures 1-2, from left to right panels, we report radial distribution functions for the particle one-particle one, particle one-particle two and particle onematrix short-range correlations at a equimolar composition. It appears that for a moderate non-additivity and two different matrix porosities, ZSEP gives rise to the best performances, quantitatively describing the simulation profiles for all the correlations. These very good performances by ZSEP maintain unaltered also at a as much asymmetric concentration as 10% (see fig. 3) , with a slight shift up at longer distances for the more dilute specie like-correlations and close to contact for the crossed correlation; however, ZSEP seems to capture the correct phase and oscillations, in agreement with the other integral equation schemes. Moreover, as shown in fig. 4, the particle 2-particle 2 and particle 2-matrix correlations are nicely described by ZSEP. We also monitored the effect of increasing non-additivity in fig. 5, that means increasing the tendency of particles to homocoordination. While ZSEP shows a slight tendency to overestimate and underestimate the like-like and the crossed correlations at longer distances, respectively, it remains definitely the most reliable integral equation closure between the ones here considered. These performances by ZSEP have been checked also at a lower degree of matrix porosity, as reported in fig. 6. After considering structure, we turned our attention to thermodynamics and, in particular, to the ability of the theory to give reliable estimates of the chemical potential, that is a primary ingredient for the determination of liquid-liquid phase separation envelopes. 15 As it appears in fig. 7, where the calculations were performed at fixed equimolar concentration, ZSEP is able to reproduce accurately simulation results up to the mixture densities close to the region of immiscibility in the phase diagram, with a discrepancy not exceeding 2-3 %. This result was confirmed by adopting a different strategy for comparison in figures 8-9: we used as a starting guess for simulations, the theoretical chemical potentials and tried to look at the simulation mean total mixture densities and the mean particle concentrations in comparison with the theoretical values. Again, at the worst case reported in the upper panel of fig. 8, i.e. for the higher matrix density and for the lower mixture non-additivity considered in this work, the percentage discrepancy with simulation was inferior to 3%. In an attempt to reduce this error from theory in comparison with simulation results, we also monitored the effect of imposing the thermodynamic constraint of eq. (16) just for the case =0.2, 0=0.3. To this aim, we used as an additional variational parameter the quantity and it emerged that decreasing the degree of thermodynamic inconsistency by ZSEP from 10% to less than 1% brought a slight change in the calculated chemical potentials (see upper panel of fig. 8). For this reason, we have not tried to enforce systematically the internal thermodynamic consistency of eq. (16) for all the cases considered in this paper. However, as documented in tables I-II, ZSEP was not found to be thermodynamically inconsistent in a significant way, with a percentage discrepancy (%TU) which on the average was of the order of 3-4% and in the worst cases was never exceeding 10-15%. Finally, we calculated the liquid-liquid coexistence lines for the worst case reported in the upper panel of fig. 8 and assessed ZSEP predictions with Cavity-Biased Monte Carlo 16 simulations. Theoretical coexistence lines reported in fig. 10 were calculated by equating the chemical potentials of the two species (second and third equation of (7), respectively), at different concentrations. After performing such a match, we were able to establish the lower consolute critical point, with an error of ~ 3%, as compared to the simulation results. This is in conformity with the earlier observation on the accuracy of the calculated chemical potentials. 17 0 * TU 0.3 1.8 0.6868 1.0111 0.77995 0.8325 0.9553 0.9853 1.1792 0.3 0.5 0.6868 1.0111 0.77995 0.8325 0.9553 0.9848 1.1722 0.2 0.1 0.35 0.5 2.5 0.6726 1.1113 0.8565 1.74695 0.9407 1.0591 2.3981 0.4 0.3 0.18 0.5 2.8 0.5637 1.0687 0.7406 0.7361 0.4 0.1 0.18 0.5 0.7 0.6032 1.6598 1.1328 -0.1468 0.9394 1.5889 0.02915 0.9198 1.0276 1.2120 Table I. Consistency parameters and percentage thermodynamic unconsistency ( for different mixtures at the equimolar concentration. In the top of the table V)V]=N1/( 1 18 0 * 0.3 0.3078 TU 8.6 0.9566 1.0077 0.9530 0.9237 1.0092 13 0.9560 0.3 0.3 0.3090 0.9931 0.9763 1.0905 1.2428 0.9530 0.9941 0.9758 1.08446 1.2443 0.9988 0.9500 1.010 0.9679 0.9922 1.2789 0.7045 1.005 0.9495 1.0087 0.9684 1.0051 1.2776 0.3 0.309- 0.10.31 0.11 0.3 0.3095 0.105 (0.1961) 0.1 0.35 0.3 6.1 0.9427 1.10045 0.9315 1.1004 1.0352 1.9334 2.7364 0.2 0.1 0.35 0.1 4.7 0.9291 1.07895 0.8929 1.1664 0.9702 1.4872 2.7578 0.4 0.3 0.179 0.335 2.9 0.9223 1.0592 0.9098 1.0584 0.9969 1.1145 1.3092 0.4 0.3 0.178 0.1105 4.8 0.9322 1.0360 0.8957 1.1124 0.9616 0.9833 1.3565 0.4 0.1 0.1795 0.310 2.9 0.95505 1.5906 0.92335 1.7837 1.42965 3.9580 -3.1866 0.4 0.1 0.179 1.5 0.9710 0.8856 -3.7338 0.104 1.4072 2.0898 1.2515 8.5566 Table II. Consistency parameters and percentage thermodynamic unconsistency (%TU) for different mixtures. See Table I for the meaning of the symbols. 19 Figure 1. Radial distribution functions for the mixture at =0.2, 0=0.3, =0.5. Full line: ZSEP; long-dashed line: MS; dotted line: PY; open circles: NVT Monte Carlo simulation. 20 Figure 2. Radial distribution functions for the mixture at =0.2, 0=0.1, =0.5. See Fig. 1 for the meaning of the symbols. 21 Figure 3. Radial distribution functions for the mixture at =0.2, 0=0.3, =0.1. See Fig. 1 for the meaning of the symbols. 22 Figure 4. Radial distribution functions for the mixture at =0.2, 0=0.3, =0.1. See Fig. 1 for the meaning of the symbols. 23 Figure 5. Radial distribution functions for the mixture at =0.4, 0=0.3, =0.5. See Fig. 1 for the meaning of the symbols. 24 Figure 6. Radial distribution functions for the mixture at =0.4, 0=0.1, =0.5. See Fig. 1 for the meaning of the symbols. 25 Figure 7. Chemical potential versus total mixture density at =0.2, 0=0.2, =0.5. Triangles: Grand canonical Monte Carlo simulation; circles: Cavity-biased Grand canonical Monte Carlo simulation; full line: ZSEP. 26 Figure 8. Total mixture density versus particle concentration at fixed chemical potentials. =0.2. Triangles: Grand Canonical Monte Carlo simulation; open circles: Cavity-biased Grand Canonical Monte Carlo simulation; diamonds: ZSEP; crosses: ZSEP with thermodynamic consistency. 27 Figure 9. Total mixture density versus particle concentration at fixed chemical potentials. =0.4. See Fig. 7 for the meaning of the symbols. 28 Figure 10. Liquid-liquid phase coexistence boundary in the total mixture densityparticle concentration plane. Error bars: Grand Canonical Monte Carlo simulation; open circles: ZSEP; full line: curve fitting ZSEP points. 29 References [1] B.J.Frisken, F.Ferri, D.S.Cannell, Physical Review letters, 66 (1991) 2754. [2] D.Nicholson, K.E.Gubbins, Journal of Chemical Physics 104 (1996)8126. [3] L.D. Gelb, K.E.Gubbins, R.Radhakrishnan, M. Sliwinska-Batkowiak, Reports Progress Physics 62, (1999) 1573-1659. [4] O. Barbieri, F. Eheburger-Dolle, T.P. Rieker, G.M. Pajonk, N. Pinto, A.V. Rao, J. Non-crystal. Solids, 285, 109 (1001). [5] W. Madden and E. D. Glandt, J. Stat. Phys. 51, 537 (1988) and J. Given and G. Stell, J. Chem. Phys. 97, 4573 (1992). [6] M. J. Fernaud, E. Lomba and L. L. Lee J. Chem. Phys. 111, 10275 (1999). [7] E. Paschinger and G. Kahl, Phys. Rev. E 61, 5330 (2000). [8] O. Sierra and Y. Duda, Phys. Lett. A 280, 146 (2001). A. Meroni, D. Levesque and J. J. Weis J. Chem. Phys. 105, 1101(1996). M. Alvarez , D. Levesque and J. J. Weis , Phys. Rev. E 60, 5495 (1999). [9] M. J. Fernaud, E. Lomba and L. L. Lee J. Chem. Phys. 111, 10275 (1999). [10] L. L. Lee, J. Chem. Phys. 103, 9338 (1995). [11] L. L. Lee and A. Malijevsky J. Chem. Phys. 114, 1 (2001). [12] E. Lomba, M. Alvarez, L.L. Lee, N.G. Almarza, J. Chem. Phys. 104, 4180 (1996). [13] E. Lomba and L.L. Lee, Int’l J. Thermophys. 17, 663 (1996); L.L.Lee, D.M. Ghonasgi, and E. Lomba, J. Chem. Phys. 104, 8058 (1996); L.L. Lee, J. Chem. Phys. 107, 7360 (1997). 30 [14] Y. Duda, L.L. Lee, Y. Kalyuzhnyi, W.G. Chapman, and P. D. Ting, J. Chem. Phys. 114, 8484 (2001); Y. Duda, L.L. Lee, Y. Kalyuzhnyi, W.G. Chapman, and P. D. Ting, Chem. Phys. Lett. 339,89 (2001). [15] See,e.g., E. Meeron and A.J.F. Siegert, J. Chem. Phys. 48, 3139 (1968); .L. Lee and Shing, J. Chem. Phys. 91, 477 (1989). [16] C. Caccamo Phys. Rep. 274, 1 (1996) . C. Caccamo (add references on thermodynamic consistencies by Prof. Caccamo!) [17] See, e.g., L.L. Lee, “Molecular Thermodynamics for Nonideal Fluids” (Butterworth, Boston, 1988) pp.216-218. [18] L. Verlet, Mol. Phys. 41, 183 (1980). [19] J.D. weeks, D. Chandler, and H.C. Andersen, J. Chem. Phys. 54, 5237 (1971). [20] L. L. Lee J. Chem. Phys. 97, 8606 (1992). [21] E. Kieklik, M. Rosinberg, G. Tarjus, P.A. Monson, J. Chem. Phys. 110, 689 (1999). [22] 13 J. K. Percus and G. J. Yevick, Phys. Rev. 110, 1 (1958). [23] G. A. Martynov and G. N. Sarkisov, Mol. Phys. 49, 1495 (1983). [24] See,e.g., M.P. Allen and D.J. Tildesley, “Computer Simulation of Liquids” (Clarendon, Oxford, 1987). [25] D. J. Adams, Mol. Phys. 29, 311 (1975). [26] M. Mezei, Mol. Phys. 40, 901 (1980). M. Mezei, Mol. Phys. 61, 565 (1987). 31