Skip to main content
The Journal of Chemical Physics logoLink to The Journal of Chemical Physics
. 2008 Jun 24;128(24):244511. doi: 10.1063/1.2943318

Kirkwood–Buff theory of four and higher component mixtures

Myungshim Kang 1, Paul E Smith 1,a)
PMCID: PMC2671182  PMID: 18601352

Abstract

Explicit expressions are developed for the chemical potential derivatives, partial molar volumes, and isothermal compressibility of solution mixtures involving four components at finite concentrations using the Kirkwood–Buff theory of solutions. In addition, a general recursion relationship is provided which can be used to generate the chemical potential derivatives for higher component solutions.

INTRODUCTION

Kirkwood–Buff (KB) theory is an exact theory of solutions that relates properties of a solution mixture to radial distribution functions (RDFs) between the different components of the solution.1, 2 KB theory has been widely used to understand the basic properties of solutions,2, 3, 4 the effects of additives on the solubility of solutes (from small hydrocarbons to proteins)5, 6, 7, 8, 9, 10, 11 and biomolecular equilibria,12, 13, 14, 15, 16 to investigate the local composition of solutions in the context of preferential solvation,17, 18 to study the effects of additives on the surface tension of liquids,19, 20 to interpret computer simulation data,13, 21, 22 and to develop models for many of the above effects.23

The central focus of KB theory are the KB integrals (Gij) between the different species i and j in the solution mixture,1

Gij=Gji=4π0[gijμVT(r)1]r2dr, (1)

where gij is the corresponding RDF and r is the intermolecular separation. The above RDFs are defined in a grand canonical (μVT) ensemble open to all species. Chemical potential derivatives for closed or semiopen systems in terms of the KB integrals and number densities (ρi=niV) are then obtained after suitable thermodynamic transformations.2, 24 The KB integrals, together with the corresponding excess coordination numbers, have provided a simple physical picture of changes in the local solution composition around each species.4

Unfortunately, as the number of solution components (n) increases and∕or as one moves from open to semiopen to closed ensembles, the resulting expressions become more cumbersome and involve significant algebraic manipulation.4 Expressions for two component solutions were provided in the original Kirkwood and Buff1 paper. Subsequently, O’Connell25 presented a general matrix formulation of KB theory, and Ben-Naim12 developed a method to simplify the matrices involved for a general n component mixture. O’Connell25 also developed expressions based on the direct correlation function, as defined by the Ornstein–Zernike equation, instead of the total correlation function. This also makes KB theory highly compatible with integral equation theories. However, the physical interpretation of integrals involving the direct correlation function is more complicated than that of the standard KB integral at normal solution densities. Furthermore, the direct correlation function can only be obtained from computer simulations after Fourier transforming the original total correlation function.26

Of course, one could always use the general matrix formulation of KB theory and solve numerically using values for the RDFs or KB integrals provided by some other approach (theory or simulation). However, this tends to obscure the contribution from the different KB integrals and hinder our understanding of specific effects. Therefore, it is often desirable to use explicit expressions that involve combinations of KB integrals and number densities. Explicit expressions for three component solutions have been provided by Ruckenstein and Shulgin (using an algebraic software),27 Ben-Naim,4 and Smith.24 Ben-Naim12 also developed expressions for some properties of four component systems, where several of the components appear at infinite dilution. To our knowledge, explicit expressions for chemical potential derivatives in four or higher component systems have not appeared for the case where all components are present at finite concentrations. Here, we use some of the relationships provided previously by Hall in an alternative derivation of the KB theory28 to generate expressions for four component solutions. A general recursion relationship is then developed for higher component mixtures.

THEORY

General approach

Hall rederived the KB theory using a different approach to Kirkwood and Buff.28 In doing so Hall produced two primary equations from which many of the expressions required here can be generated. However, his approach was still somewhat involved. Here we present a simpler derivation of the Hall equations. The first focuses on changes in the molar concentration of any component. If we consider the species number densities (or molarities) in the grand canonical ensemble to be functions of T and all the chemical potentials (μ), then we can write

dlnρi=j=1n(lnρiμj)T,μkjdμjconstantT, (2)

for any component i at constant T. Here, the summation is over all n components of the solution. We note that all the chemical potentials are independent thermodynamic variables in this open ensemble. The above derivatives can be expressed in terms of KB integrals using the fact that1, 2

(lnρiμj)T,μkj=β(δij+Nij), (3)

which is essentially the starting equation for the KB theory. Here, δij is the Kroenecker delta function, NijjGijNji, β=1∕RT, and R is the Gas constant. From the these two equations one finds

dlnρi=βj=1n(δij+Nij)dμjconstantT. (4)

The above expression is valid for changes in the concentration of any component in any multicomponent system and any (thermodynamically reasonable) ensemble with T constant. This is the equation derived by Hall, however, using a much longer route.28 If one is interested in expressing solution compositions in terms of molalities (mii∕ρ1 to within a conversion factor), then using the fact that d ln mi=d ln ρid ln ρ1 one can write

dlnmi=βj=1n(δij+Nijδ1jN1j)dμjconstantT, (5)

which is also valid for any constant T ensemble. Clearly, in doing so, we have labeled component 1 as the primary solvent and, therefore, it is unique—as it is also experimentally. The consequences of doing this will be discussed later.

In the traditional derivation of KB theory, the set of equations presented in Eq. 4 are converted into matrix form after taking derivatives with respect to ln ρj with all ρkj held constant.4 They can then be solved to obtain a series of expressions involving the quantities

β(μilnNj)T,V,Nkj=β(μilnρj)T,Nkj. (6)

These constant volume derivatives then have to be transformed using a series of thermodynamic relationships into the required and experimentally relevant derivatives at constant P as defined by

μij=β(μilnNj)T,P,Nkj=β(μilnmj)T,P,mkj. (7)

This is clearly the most general approach. However, it has long been recognized that the expressions obtained for higher multicomponent systems (n≥3) involve considerable algebraic manipulation.4, 27 In addition, a significant degree of cancellation of terms in the expressions is often found but not easily recognized in the matrix formulation.

Here, we will adopt a different route which we believe is much simpler for mixtures with a large number of components. Eliminating dμ1 from Eq. 4 using the corresponding Gibbs–Duhem (GD) relationship at constant T and P, j=1nρjdμj=0, provides

dlnρi=βj=2n[δij+Nijmj(δi1+Ni1)]dμjconstantT,P. (8)

This can be used to obtain an expression for changes in the molalities,

dlnmi=βj=2n(δij+Nij+)dμjconstantT,P, (9)

where Nij+=Nij+mj(1+N11Ni1Nj1) and i=2,n. The additional constraint of constant P arises from our use of the corresponding GD expression. This is the equation provided by Hall28 for changes in molal concentrations at constant T and P. It also appears in the original KB paper without derivation.1 The above set of equations can be written in a general (n−1)×(n−1) matrix form for a mixture of n components so that

[1+N22+N23+N24+N2n+N32+1+N33+N42+1+N44+Nn2+1+Nnn+][βdμ2βdμ3βdμ4βdμn]=[dlnm2dlnm3dlnm4dlnmn]. (10)

To continue we will choose our required ensemble and then take derivatives with respect to the molality of one component, in this case ln mj, keeping T, P, and all other mkj constant. This makes the resulting expressions less general than previous approaches, but one can easily recover derivatives with respect to other species by a simple index change. One finds that

[1+N22+N23+N24+N2n+N32+1+N33+N42+1+N44+Nn2+1+Nnn+][μ2jμ3jμ4jμnj]=[δ2jδ3jδ4jδnj]orDnμ=d, (11)

where the matrix elements are given by Dαβ=δαβ+N(α+1)(β+1)+, the vector elements of μα are given by μ(α+1)j, and the vector elements of dα are given by δ(α+1)j with α, β=1, n−1. Hence, we have a set of simultaneous equations which can be solved to give the chemical potential derivatives. Therefore,

μ=Dn1d. (12)

One can express the inverse in terms of cofactors of the original Dn matrix. The chemical potential derivatives are then given by μij=Dnj1,i1Dn, for i,j≠1 and where Dnj1,i1 is a cofactor of Dn. If the chemical potential derivative of species 1 is required, it can be obtained from the GD equations

μ1j=k=2nmkμkj=mjk=2nμjk, (13)

using the solutions to Eq. 11. In the above expression j=1, n, and m1=1.

There are several advantages of this approach. First, it can be applied directly to any number of solution components in any constant T and P ensemble. Therefore, we do not have to transform the subsequent expressions from a constant T and V to a constant T and P ensemble. Second, we have eliminated the chemical potential of species 1 and therefore the resulting set of equations and the corresponding matrix is reduced. Third, the simplicity of the column vector on the right-hand side of Eq. 12 indicates that each chemical potential derivative expression involves only one element of the inverse matrix in the numerator, together with the determinant of Dn in the denominator. This set of combined factors greatly simplifies the resulting expressions.

In the previous sections, we have focused on derivatives of the chemical potentials taken with respect to molality. Derivatives with respect to mole fraction or molarity can be obtained by noting that

dlnxi=dlnρij=1nxjdlnρjconstantT (14)

and

dlnmi=dlnρidlnρ1constantT (15)

for any number of components at constant T. One could develop Eq. 14 in terms of the KB integrals. However, it is much easier, especially for closed systems, to convert the molality based derivatives to mole fraction derivatives after the former has been obtained. For closed systems these equations become

(lnxilnmj)T,P,mkj=δijxj, (16)

with

(lnρilnmj)T,P,mkj=δijρjV¯j, (17)

and therefore

(lnρilnxj)T,P,mkj=δijρjV¯j1xj, (18)

for any number of components at constant T and P.

To determine the corresponding expressions for the partial molar volumes (PMVs) in multicomponent systems, it is sufficient to express the PMVs in terms of the chemical potential derivatives. We will continue to treat species 1 as a unique component. Starting with Eq. 4 for the differential of the number density of species 1 and eliminating dμ1 by using the GD relationship at constant T and P, one finds

dlnρ1=βj=2nmj(1+N11Nj1)dμjconstantT,P. (19)

Obviously, there is a series of similar expressions depending on the initial choice of i in Eq. 4. Taking the derivative with respect to ln mk while keeping T, P, and all other mjk constant provides for k≠1,

ρkV¯k=j=2nmj(1+N11Nj1)μjk=mkj=2n(1+N11Nj1)μkj, (20)

where the appropriate chemical potential derivatives are provided by Eq. 12. If required, the PMV of species 1 can be obtained from the fact that

j=1nρjV¯j=1, (21)

for all mixtures.

Finally, if one starts from Eq. 4 and then takes derivatives with respect to pressure with all mj and T constant, one can show that for any component i,

RTκT=j=1n(δij+Nij)V¯j=V¯i+j=1nNijV¯j, (22)

where κT is the isothermal compressibility. This can be used to derive an expression for the compressibility. If we chose i=1 and then eliminated the PMV of species 1 by using Eq. 21, then

ρ1RTκT=(1+N11)j=2n(1+N11Nj1)ρjV¯j, (23)

which can be written as

ρ1RTκT=(1+N11)j=2n(1+N11Nj1)k=2nmk(1+N11Nk1)μkj, (24)

using the corresponding chemical potential derivatives.

Before leaving this section we note that the PMV of a species can be considered to involve two contributions.2 The first relates to the change in volume of the solution due to the volume occupied by the additional molecule located at a fixed position in the system. The second involves the ideal contribution to the PMV which arises due to the fact that the additional molecule will possess a momentum, corresponding to the particular temperature, which contributes to the pressure of the system when the molecule is released. Under constant P conditions this gives rise to a change in volume according to the compressibility of the solution. Therefore, one can isolate the former change by writing

Vi=V¯iRTκT. (25)

From Eq. 22 we obtain the relationship

ρiVi=N1ij=2n(NjiN1i)k=2nmk(1+N11Nk1)μkj, (26)

which is now a better measure of the actual volume occupied by each species in solution. Another interesting property of solutions is the pseudochemical potential (μ*). The pseudochemical potential plays an important role in solution theory and is defined by the equation2

μi=μi+RTlnΛi3ρi, (27)

where Λ is the thermal de Broglie wavelength. From this equation it is quite easy to show that

β(μilnmj)T,P,mkj=μij(δijρjV¯j), (28)

and also

β(μilnρj)T,P,mkj=β(μilnρj)T,P,mkj1, (29)

which completes our preliminary analysis.

Hence, we have a general set of Eqs. 12, 20, 24 which can be used to derive the KB expressions for the chemical potential derivatives, PMVs, and compressibility of any multicomponent mixture.

Four component mixtures

As an example of the current approach, we will generate the expressions for a four component system where all components appear at finite concentrations. To our knowledge the explicit KB expressions for a four component system have not been presented in the literature. Using the above approach we find

[μ22μ32μ42]=D41[100], (30)

and where the inverse of D4 is given by

D41=1D4[(1+N33+)(1+N44+)N43+N34+N24+N43+N23+(1+N44+)N23+N34+N24+(1+N33+)N34+N42+N32+(1+N44+)(1+N22+)(1+N44+)N42+N24+N24+N23+N34+(1+N22+)N32+N43+N42+(1+N33+)N23+N42+N43+(1+N22+)(1+N22+)(1+N33+)N23+N32+]. (31)

Therefore, the expressions for the chemical potential derivatives in a four component mixture are given by

β(μ2lnm2)T,P,m3,m4=(1+N33+)(1+N44+)N43+N34+D4, (32)
β(μ3lnm2)T,P,m3,m4=(1+N44+)N32+N34+N42+D4,
β(μ4lnm2)T,P,m3,m4=(1+N33+)N42+N43+N32+D4,

with

D4=(1+N22+)(1+N33+)(1+N44+)(1+N22+)N34+N43+(1+N33+)N24+N42+(1+N44+)N23+N32++2N23+N34+N42+, (33)

and where we have also used the fact that ρiNij+=ρjNji+ to simplify the above determinant. The final derivatives (μ12 and μ11) can be obtained after application of the GD relationships

β(μ1lnm2)T,P,m3,m4=m2μ22m3μ32m4μ42 (34)

and

β(μ1lnN1)T,P,N2,N3,N4=μ12μ13μ14. (35)

Derivatives with respect to other species molalities (m3 and m4) are obtained quite easily by either inspection, by noting that mjμjk=mkμkj, or from the fact that

[μ23μ33μ43]=D41[010]and[μ24μ34μ44]=D41[001], (36)

which are also relatively simple to solve.

Application of Eq. 20 and the expressions found above provides the following expressions for the PMVs in four component mixtures,

ρ2V¯2=m2(1+N11N21)μ22+m3(1+N11N31)μ32+m4(1+N11N41)μ42, (37)
ρ3V¯3=m2(1+N11N21)μ23+m3(1+N11N31)μ33+m4(1+N11N41)μ43,
ρ4V¯4=m2(1+N11N21)μ24+m3(1+N11N31)μ34+m4(1+N11N41)μ44.

The above expressions obey Eq. 21 as required. If necessary, the PMV of 1 can then be obtained using Eq. 21. Finally, for the isothermal compressibility we obtain

ρ1RTκT=1+N11(1+N11N21)[(1+N11N21)m2μ22+(1+N11N31)m3μ32+(1+N11N41)m4μ42](1+N11N31)[(1+N11N21)m2μ23+(1+N11N31)m3μ33+(1+N11N41)m4μ43](1+N11N41)[(1+N11N21)m2μ24+(1+N11N31)m3μ34+(1+N11N41)m4μ44], (38)

which can be simplified to provide

ρ1RTκT=1+N11(1+N11N21)2m2μ22(1+N11N31)2m3μ33(1+N11N41)2m4μ442(1+N11N21)(1+N11N31)m3μ322(1+N11N21)(1+N11N41)m4μ422(1+N11N41)(1+N11N31)m3μ34. (39)

Three component mixtures

While three component systems had been studied before,4, 27 it is interesting and informative to compare the limiting expressions provided here with those currently in the literature, especially due to the different notations involved. In addition, this will aid in the development of a general recursive relationship for the derivatives. The limiting forms are quite easy to obtain as we have Nij+0 as ρj→0, and miμij=mjμji→0 as ρi→0 or ρj→0. Therefore, as m4 tends to zero, one obtains the derivatives for a ternary system of 1, 2, and 3. The chemical potential derivatives are then given by

β(μ2lnm2)T,P,m3=(1+N33+)(1+N22+)(1+N33+)N23+N32+, (40)
β(μ3lnm2)T,P,m3=N32+(1+N22+)(1+N33+)N23+N32+.

The corresponding PMV expressions reduce to

ρ2V¯2=m2(1+N11N21)(1+N33+)(1+N11N31)N23+(1+N22+)(1+N33+)N23+N32+, (41)
ρ3V¯3=m3(1+N11N31)(1+N22+)(1+N11N21)N32+(1+N22+)(1+N33+)N23+N32+,

with a compressibility given by

ρ1RTκT=1+N11m2(1+N11N21)2(1+N33+)(1+N22+)(1+N33+)N23+N32++2m2(1+N11N21)(1+N11N31)N23+(1+N22+)(1+N33+)N23+N32+m3(1+N11N31)2(1+N22+)(1+N22+)(1+N33+)N23+N32+. (42)

Two component mixtures

Two component mixtures have clearly been studied before, but not using the present notation. After taking an additional ρ3→0 limit, one finds that for the two component case,

β(μ2lnm2)T,P,N1=11+N22+ (43)

and

ρ2V¯2=m21+N11N211+N22+, (44)

with

ρ1RTκT=1+N11m2(1+N11N21)21+N22+, (45)

which further reduces to the required compressibility equation when ρ2→0.

Open and semiopen systems

Traditionally, KB expressions for open and semiopen systems have been derived starting from the fully closed system results.29, 30 This can be quite tedious. Recently, we suggested starting from expressions for the fully open system and transforming to the required ensemble in a stepwise manner.24, 31 This made the manipulations easier although several steps were still required. However, it is clear from Eqs. 4, 5 that results for open and semiopen systems become almost trivial. As an example we will derive an expression for the preferential binding parameter (∂m3∕∂m2) for ternary mixtures in the T, μ1, and μ3 ensemble, where 1 is the primary solvent, 2 is the biomolecule of interest, and 3 is an additive. Starting from Eq. 5 one immediately finds

(lnm2μ2)T,μ1,μ3=β(1+N22N12) (46)

and

(lnm3μ2)T,μ1,μ3=β(N32N12), (47)

which can be solved to yield

(m3m2)T,μ1,μ3=N23m3N211+N22N12, (48)

and is in agreement with previous results.24 It is clear from Eq. 5 that the same expression is obtained if we have any number of additional components at a constant chemical potential.

Alternatively, one can start from Eq. 9 to obtain an expression for the equivalent property in the T, P, and μ3 ensemble. Hence,

(1+N22+)β(μ2lnm2)T,P,μ3=1 (49)

and

N32+β(μ2lnm2)T,P,μ3=(lnm3lnm2)T,P,μ3, (50)

which can be solved quite easily to give

(m3m2)T,P,μ3=N23+1+N22+, (51)

and is also in agreement with previous results.24 Again, the same expression is valid in the presence of any number of additional species as long as their chemical potentials remain constant.

A general recursion relationship for the chemical potential derivatives

Analysis of the chemical potential derivatives for two, three, and four component systems enables a general recursion relationship to be established. It is clear that the denominator will always contain the determinant ∣Dn∣ for a general n component system. If we focus on the expressions in the numerator, one immediately observes that the numerator of μii is just the determinant of the D matrix for the corresponding n−1 system in which component i has been eliminated. Hence, we have

μnnn=Dn1Dn, (52)

where the superscript indicates a derivative defined in an n component system. A simple change in indices provides equivalent expressions for any μiin where i≠1. The numerators of the other derivatives (μnj,j≠1, or n) also follow a simple pattern. The chemical potential derivatives for the nth component obey the recursive relationship

μnjn=Dn1Dnk=2n1μkjn1Nnk+=μnnnk=2n1μkjn1Nnk+, (53)

which is just a factorization of the Dn matrix that one observes due to the relative simplicity of Eq. 12. Expressions for the μij derivatives where ij≠1 can then be found by inspection.

Five component systems

Using the recursion relationship developed above, one can generate expressions for the chemical potential derivatives in five component solutions. For simplicity, we will only consider the chemical potential derivatives. The following expression is obtained from Eqs. 52, 32 followed by a simple index change (5↔2) in the numerator,

μ22D5=(1+N33+)(1+N44+)(1+N55+)(1+N33+)N45+N54+(1+N44+)N35+N53+(1+N55+)N34+N43++2N34+N45+N53+. (54)

Using Eq. 53 and the set of derivatives for a four component solution provided in Eq. 32, one finds

μ52D5=(1+N33+)(1+N44+)N52++(1+N33+)N42+N54++(1+N44+)N32+N53+N32+N43+N54+N34+N42+N53++N34+N43+N52+. (55)

Consequently, using a simple index change of 5↔3 and 5↔4, one finds

μ32D5=(1+N44+)(1+N55+)N32++(1+N44+)N35+N52++(1+N55+)N42+N34+N42+N54+N35+N45+N52+N34++N45+N54+N32+ (56)

and

μ42D5=(1+N33+)(1+N55+)N42++(1+N33+)N45+N52++(1+N55+)N32+N43+N32+N53+N45+N35+N52+N43++N35+N53+N42+, (57)

respectively. Derivatives with respect to other species can be obtained by inspection. The above expressions are in agreement with those obtained directly via Eq. 12. If required, the corresponding PMVs and compressibility can be obtained from Eqs. 20, 24.

DISCUSSION

We have provided general relationships which can be used to develop explicit expressions for chemical potential derivatives, PMVs, and the isothermal compressibility of any mixture of n components in terms of KB integrals. Our choice of the molality concentration scale makes species 1 unique. Consequently, some of the “symmetry” in the expressions that might be observed for molarity or mole fraction based derivatives is lost using the current notation. We consider this to be an acceptable sacrifice in many cases. It is therefore informative to compare and relate the expressions generated here with those developed by other approaches, especially when considering symmetric ideal solutions. To do this we will refer to the expressions of Smith for ternary solutions which represent the most condensed form for the chemical potential derivatives.24 They are easily expanded to provide the expressions of Ruckenstein and Shulgin27 and Ben-Naim.4 Smith provided the following expressions:

β(μ1lnN2)T,P,N1,N3=ρ2A3ρ1A2A3+ρ2A1A3+ρ3A1A2, (58)
β(μ2lnN2)T,P,N1,N3=ρ1A3+ρ3A1ρ1A2A3+ρ2A1A3+ρ3A1A2,
β(μ3lnN2)T,P,N1,N3=ρ2A1ρ1A2A3+ρ2A1A3+ρ3A1A2,

where the A’s are given by

A1=1+ρ1(G11+G23G12G13), (59)
A2=1+ρ2(G22G23G12+G13),
A3=1+ρ3(G33G23+G12G13).

Comparison with the expressions provided in Eq. 40 indicates that

A1=ρ1N32+ρ2=ρ1N23+ρ3, (60)
A2=1+N22+N32+=1+N22+ρ2A1ρ1,
A3=1+N33+N23+=1+N33+ρ3A1ρ1.

Specific combinations of KB integrals often appear repeatedly in other formulations. For instance, one can define for ij,

ηij=ρi+ρj+ρiρj(Gii+Gjj2Gij)=ρiAj+ρjAi. (61)

In the current notation it is found that

ηi1=ρ1(1+Nii+). (62)

We attempted to find a similar factorization and relationships as found in Eqs. 58, 59 for four component systems, but were unsuccessful.

The application of the KB theory to symmetrical ideal solutions is also of interest. Ben-Naim4 has shown that for a general n component mixture to display symmetric ideal behavior, one must have ΔGij=Gii+Gjj−2Gij=0 for all ij pairs. Therefore, in the current notation, one must have ρ1Nii+=ρi for symmetric ideal solutions. In addition, one finds A1=A2=A3=1 for symmetric ideal ternary solutions.

CONCLUSIONS

Using a new approach we have developed explicit relationships for KB integrals in four and five component solution mixtures. In our opinion, the use of molalities as concentration variables provides the simplest path to expressions in multicomponent solutions. We are currently using this type of approach to study biologically relevant systems containing five or more components.

For a general n component system there are n(n+1)∕2 unique Gij integrals. Determining the integrals from experimental data using the KB inversion approach requires one isothermal compressibility value, n−1 independent PMVs, and n(n−1)∕2 independent μij values as a function of composition. This has been achieved for ternary systems.32 As one moves beyond ternary systems the experimental data becomes increasingly more difficult to obtain. Consequently, we envision that the major use for the expressions provided here will involve either theoretical estimates of the KB integrals or simulated values of the integrals. In either case, the exact KB expressions provide a solid foundation for investigating these complicated solution mixtures.

ACKNOWLEDGMENTS

The project described was supported by Grant No. R01GM079277 from the National Institute of General Medical Sciences. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of General Medical Sciences or the National Institutes of Health.

References

  1. Kirkwood J. G. and Buff F. P., J. Chem. Phys. 10.1063/1.1748352 19, 774 (1951). [DOI] [Google Scholar]
  2. Ben-Naim A., Statistical Thermodynamics for Chemists and Biochemists (Plenum, New York, 1992). [Google Scholar]
  3. Matteoli E. and Mansoori G. A., Fluctuation Theory of Mixtures (Taylor & Francis, New York, 1990). [Google Scholar]
  4. Ben-Naim A., Molecular Theory of Solutions (Oxford, New York, 2006). [Google Scholar]
  5. O’Connell J. P., AIChE J. 17, 658 (1971). [Google Scholar]
  6. Chialvo A. A., J. Phys. Chem. 10.1021/j100113a041 97, 2740 (1993). [DOI] [Google Scholar]
  7. Smith P. E., J. Phys. Chem. B 10.1021/jp983303c 103, 525 (1999). [DOI] [Google Scholar]
  8. Chitra R. and Smith P. E., J. Phys. Chem. B 10.1021/jp012354y 105, 11513 (2001). [DOI] [Google Scholar]
  9. Ruckenstein E. and Shulgin I., Ind. Eng. Chem. Res. 10.1021/ie020348y 41, 4674 (2002). [DOI] [Google Scholar]
  10. Shulgin I. L. and Ruckenstein E., Biophys. Chem. 118, 128 (2005). [DOI] [PubMed] [Google Scholar]
  11. Mazo R. M., J. Phys. Chem. B 110, 24077 (2006). [DOI] [PubMed] [Google Scholar]
  12. Ben-Naim A., J. Chem. Phys. 10.1063/1.431544 63, 2064 (1975). [DOI] [Google Scholar]
  13. Smith P. E., J. Phys. Chem. B 10.1021/jp0474879 108, 18716 (2004). [DOI] [Google Scholar]
  14. Shimizu S. and Boon C. L., J. Chem. Phys. 10.1063/1.1806402 121, 9147 (2004). [DOI] [PubMed] [Google Scholar]
  15. Shimizu S., McLaren W. M., and Matubayasi N., J. Chem. Phys. 10.1063/1.2206174 124, 234905 (2006). [DOI] [PubMed] [Google Scholar]
  16. Pierce V., Kang M., Aburi M., Weerasinghe S., and Smith P. E., Cell Biochem. Biophys. 10.1007/s12013-007-9005-0 50, 1 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Ben-Naim A., Cell Biophys. 12, 255 (1988). [DOI] [PubMed] [Google Scholar]
  18. Marcus Y., Monatsch. Chem. 10.1007/s007060170023 132, 1387 (2001). [DOI] [Google Scholar]
  19. Tronel-Peyroz E., Douillard J. M., Bennes R., and Privat M., Langmuir 10.1021/la00085a011 5, 54 (1989). [DOI] [Google Scholar]
  20. Chen F. and Smith P. E., “Theory and computer simulation of solute effects on the surface tension of liquids,” J. Phys. Chem. B (in press). [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Weerasinghe S. and Smith P. E., J. Phys. Chem. B 118, 10663 (2003). [Google Scholar]
  22. Aburi M. and Smith P. E., J. Phys. Chem. B 108, 7382 (2004). [Google Scholar]
  23. Smith P. E., J. Phys. Chem. B 10.1021/jp046973t 108, 16271 (2004). [DOI] [Google Scholar]
  24. Smith P. E., Biophys. J. 91, 849 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. O’Connell J. P., Mol. Phys. 10.1080/00268977100100031 20, 27 (1971). [DOI] [Google Scholar]
  26. Ramirez R., Mareschal M., and Borgis D., Chem. Phys. 10.1016/j.chemphys.2005.07.038 319, 261 (2005). [DOI] [Google Scholar]
  27. Ruckenstein E. and Shulgin I., Fluid Phase Equilib. 10.1016/S0378-3812(01)00372-7 180, 345 (2001). [DOI] [Google Scholar]
  28. Hall D. G., Trans. Faraday Soc. 10.1039/tf9716702516 67, 2516 (1971). [DOI] [Google Scholar]
  29. Schurr J. M., Rangel D. P., and Aragon S. R., Biophys. J. 10.1529/biophysj.104.057331 89, 2258 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Shulgin I. L. and Ruckenstein E., J. Chem. Phys. 10.1063/1.2011388 123, 054909 (2005). [DOI] [PubMed] [Google Scholar]
  31. Smith P. E., J. Phys. Chem. B 110, 2862 (2006). [DOI] [PubMed] [Google Scholar]
  32. Matteoli E. and Lepori L., J. Chem. Soc., Faraday Trans. 10.1039/ft9959100431 91, 431 (1995). [DOI] [Google Scholar]

Articles from The Journal of Chemical Physics are provided here courtesy of American Institute of Physics

RESOURCES