<<

This is the author’s final, peer-reviewed manuscript as accepted for publication. The publisher-formatted version may be available through the publisher’s web site or your institution’s library.

Kirkwood–Buff integrals for ideal

Elizabeth A. Ploetz, Nikolaos Bentenitis, Paul E. Smith.

How to cite this manuscript

If you make reference to this version of the manuscript, use the following information:

Ploetz, E.A., Bentenitis, N., & Smith, P.E. (2010). Kirkwood–Buff integrals for ideal solutions. Retrieved from http://krex.ksu.edu

Published Version Information

Citation: Ploetz, E.A., Bentenitis, N., & Smith, P.E. (2010). Kirkwood–Buff integrals for ideal solutions. The journal of chemical physics, 132(16), 9.

Copyright: © 2010 American Institute of Physics.

Digital Object Identifier (DOI): doi:10.1063/1.3398466

Publisher’s Link: http://jcp.aip.org/resource/1/jcpsa6/v132/i16/p164501_s1

This item was retrieved from the K-State Research Exchange (K-REx), the institutional repository of Kansas State University. K-REx is available at http://krex.ksu.edu THE JOURNAL OF CHEMICAL PHYSICS 132, 164501 ͑2010͒

Kirkwood–Buff integrals for ideal solutions ͒ Elizabeth A. Ploetz,1 Nikolaos Bentenitis,2 and Paul E. Smith1,a 1Department of Chemistry, Kansas State University, Manhattan, Kansas 66506, USA 2Department of Chemistry and Biochemistry, Southwestern University, Georgetown, Texas 78626, USA ͑Received 20 May 2009; accepted 29 March 2010; published online 22 April 2010͒

The Kirkwood–Buff ͑KB͒ theory of solutions is a rigorous theory of which relates the molecular distributions between the solution components to the thermodynamic properties of the . Ideal solutions represent a useful reference for understanding the properties of real solutions. Here, we derive expressions for the KB integrals, the central components of KB theory, in ideal solutions of any number of components corresponding to the three main scales. The results are illustrated by use of molecular dynamics simulations for two binary solutions mixtures, benzene with toluene, and methanethiol with dimethylsulfide, which closely approach ideal behavior, and a binary mixture of benzene and methanol which is nonideal. Simulations of a quaternary mixture containing benzene, toluene, methanethiol, and dimethylsulfide suggest this system displays ideal behavior and that ideal behavior is not limited to mixtures containing a small number of components. © 2010 American Institute of Physics. ͓doi:10.1063/1.3398466͔

I. INTRODUCTION solutions. In SI solutions the KBIs are neither zero nor inde- pendent of composition. The KBIs for these solutions can be The Kirkwood–Buff ͑KB͒ theory of solutions has pro- ͑␬ ͒ expressed in terms of the isothermal compressibility T and vided a wealth of data and insight into the properties of so- ͑ ͒ 1–13 molar Vi of the pure components at the same T lution mixtures. KB theory can be applied to any stable and P. Expressions for binary, ternary, and quaternary solu- mixture containing any number of components. In particular, tions are available.2,20,21 Ben-Naim has also shown that SI the theory provides exact relationships from the molecular ⌬ solutions including up to four components satisfy Gij=Gii distributions between the various species in solution to the 22 +Gjj−2Gij=0 for all i, j pairs. Our recent analysis of the corresponding thermodynamics of the mixture. The central results obtained from the KB theory of solutions for n=1 to quantities of interest in KB theory are the KB integrals ͑ ͒ 1 4 components proposed the following general expression for KBIs , the KBIs:21 ϱ ␲͵ ͓ ␮VT͑ ͒ ͔ 2 ͑ ͒ n Gij = Gji =4 gij r −1 r dr, 1 SI ␬ ␳ 2 ͑ ͒ 0 Gij = RT T − Vi − Vj + Sn, Sn = ͚ kVk , 2 k=1 where g is the corresponding radial distribution function ij R ␳ N /V ͑rdf͒ between species i and j, and r is their intermolecular where is the constant and i = i are the number separation. In applying KB theory to understand solutions, it of each species. The above expression was postu- i j n is often useful to compare the properties of a real solution lated to be valid for any and combination in any com- with those of a corresponding .2,6,14,15 In par- ponent SI solution. This can be shown to be true for n 21 ticular, deviations from ideal solution behavior provide indi- =1–4 components. Here we provide rigorous proof that cations of specific associations or affinities between the so- this expression is valid for any number of components at 16,17 T P lution components which affect the thermodynamics. constant and . Furthermore, we also determine the corre- Furthermore, assuming ideal behavior for some components sponding expressions for the KBIs which result in ideal be- may be the only possible approach to describe solutions con- havior on the and molarity concentration scales. The taining a large number of components, where experimental results are then illustrated using data obtained from computer data are usually very rare.18,19 Consequently, it is important simulations. to understand the nature of ideal solutions as formulated by KB theory. Ideality in solution mixtures at constant pressure ͑P͒ and II. THEORY ͑ ͒ temperature T can be expressed using a variety of concen- The chemical potential ͑␮͒ of a solute in solution can be tration scales. The most common is the fraction con- expressed in a variety of ways using a series of different centration scale where the corresponding activity coefficients concentration scales. There are three major expressions re- are unity for all species at all solution compositions. These lated to whether the species are represented in ͑ ͒ mixtures are also known as symmetric ideal SI or perfect terms of mole fractions ͑x͒, ͑m͒, or molarities ͑c͒. It is then typical to write a͒ Author to whom correspondence should be addressed. Tel.: 785-532-5109. ␮ ␮o,x ␮o,x ␮ex FAX: 785-532-6666. Electronic mail: [email protected]. i = i + RT ln fixi = i + i + RT ln xi,

0021-9606/2010/132͑16͒/164501/9/$30.00132, 164501-1 © 2010 American Institute of Physics

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-2 Ploetz, Bentenitis, and Smith J. Chem. Phys. 132, 164501 ͑2010͒

␮ץ , ␮ = ␮o,m + RT ln ␥ m i i i i ␤ͩ i ͪ = ␮ , ij ץ ln mj T,P,m ␮ ␮o,c kj i = i + RT ln yici, ␮ ␮ץ i ij 3␳ ͑ ͒ ␤ͩ ͪ = , ͑7͒⌳ ء␮ ␮ ln ␳ ␦ − ␾ ץ i = i + RT ln i i, 3 j T,P,Nkj ij j ͑ ␥ ͒ where a set of activity coefficients f , ,y are used to quan- which are general for any number of components. Alterna- tify deviations from ideal behavior on the , mo- tively, if one starts from Eq. ͑4͒ and then takes derivatives lality, and molarity concentration scales, respectively. The with respect to pressure with all Nj and T constant one finds final equation is the statistical mechanical expression for the that chemical potential in terms of the pseudochemical potential n thermal de Broglie wavelength ͑⌳͒, and number ,͒ء␮͑ ␬ ͑␦ ͒ ͑ ͒ .2 The standard chemical potentials ͑␮o͒ can refer to RT T = ͚ ij + Nij Vj 8 j=1 an infinitely dilute solute or the pure solute at the same T and P. They are constants and independent of the concentration for any i component. of i, whereas the pseudochemical potential is composition The above equations relate the KBIs to properties of the dependent. The following discussion is focused on changes solution. It is easier, albeit less general, if we select a par- in the chemical potentials with composition, and therefore ticular species to continue. We choose i=1 and k1, and ͑ ͒ ͑ ͒ ͑ ␳ ͒ the choice of standard state is largely irrelevant. We also note then write Eqs. 5 and 8 after multiplication by 1 in an RT d ln y when T is constant. It is generally nϫn matrix form so that= ءthat d␮ known that the activity coefficients decrease when the solute ␳ V ␳ V ␳ V ␳ V 1+N ␳ RT␬ displays significant solute self-association, whereas the activ- 1 1 1 2 1 3 ¯ 1 n 11 1 T ␮ ␮ ␮ ␮ ␾ ity coefficients increase when the solute is significantly sol- 12 22 32 n2 N12 − 2 ␮ ␮ ␮ ␮ ␾ vated by the other species in solution. 13 23 33 n3 N13 = − 3 . The general approach used here to determine the re- ΄ ΅΄ ΅ ΄ ΅ ]]] ] quired KBIs involves generating a series of expressions in- ␮ ␮ ␮ ␮ N − ␾ volving derivatives of the chemical potentials which are 1n 2n 3n ¯ nn 1n n valid for real solutions, followed by application of the appro- ͑9͒ priate conditions corresponding to ideality for the various Hence, we have a set of simultaneous equations which can concentration scales. Let us consider the species number be solved quite easily for the required KBIs to give densities in the grand canonical ensemble to be functions of ␳ ␬ temperature and all the chemical potentials. One can then 1+N11 1RT T 23 ␾ write N12 − 2 n −1 ␾ ͑ ͒ N13 = Mn − 3 , 10 ␳ ␤ ͑␦ ͒ ␮ ͑ ͒ d ln i = ͚ ij + Nij d j, 4 ΄ ΅ ΄ ΅ ] ] j=1 ␾ N1n − n which is valid for changes in the number density of any where M is the matrix from Eq. ͑9͒. This is the approach we component in any multicomponent system and any ͑thermo- n used recently for outlining a general KB inversion dynamically reasonable͒ ensemble with T constant. Here, procedure.21 After the appropriate chemical potential deriva- N =␳ G , ␤=1/RT, and ␦ is the Kroenecker delta function. ij j ij ij tives are used in the M matrix, and its inverse has been Two further sets of equations can be derived from the above n determined, the expressions for the KBIs representing our equation. If we take derivatives with respect to ln N , while k series of ideal solutions can be obtained from Eq. ͑10͒. keeping T, P, and all other Njk constant, we find an expres- sion for the partial molar volumes ͑¯V͒, written here in terms ͑␾ ␳ ͒ III. METHODS of fractions i = iVi , n All mixtures were simulated via classical molecular dy- ␦ − ␾ = ͚ ͑␦ + N ͒␮ , ͑5͒ namics techniques using the GROMACS program ͑version ik k ij ij jk 24 j=1 3.3͒. The simulations were performed in the isothermal- isobaric NpT ensemble at the temperature and pressure of for any i species and where we defined 25 interest using Berendsen thermostats and barostats. All sol- ␮ ute and bonds were constrained using the LINCSץ ␤ͩ i ͪ = ␮ . ͑6͒ 26 ij algorithm. A 2 fs time step was used for integration of the ץ ln Nj T,P,Nkj equations of motion. Electrostatic interactions were evalu- 27 The relationships to derivatives on other concentration scales ated using the particle-mesh-Ewald technique, with cutoff are given by distances of 1.2 nm and of 1.5 nm for the real space electro- static and van der Waals interactions, respectively. Details of ␮ ␮ 28–31ץ ␤ͩ i ͪ = ij , the force fields are provided elsewhere. All mixtures ␦ ץ ln xj T,P ij − xj were simulated with enough to occupy a cubic

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-3 KB integrals for ideal solutions J. Chem. Phys. 132, 164501 ͑2010͒

box of length of Ϸ6 nm, starting from random initial con- ␮SI ͑ ͒ d i = RTd ln xi, 11 figurations, and were simulated for between 10 and 20 ns. The simulated KBIs have to be truncated in closed sys- 28,32,33 ␮SI ␦ ͑ ͒ tems. We discussed this issue in detail. Typically, the ij = ij − xj, 12 integral is truncated after several shells ͑1–1.5 nm͒ depending on the size of the solute and solvent. To eliminate and illustrate that the chemical potential of each component statistical variations we averaged the integral values over a is affected by the addition of j through a change in their mole small range of integration distances corresponding to one fractions. For SI solutions ͓defined by Eq. ͑11͔͒ the required solvation shell. The accuracy of this truncation procedure matrix reduces to can be checked by determining the simulated partial molar volumes and compressibility using other approaches.28 Error ␳ ␳ ␳ ␳ 1V1 1V2 1V3 ¯ 1Vn estimates were obtained from multiple 5 ns block averages. − x2 1−x2 − x2 − x2 IV. RESULTS SI ͑ ͒ Mn = − x3 − x3 1−x3 − x3 . 13 A. Ideal solutions on the mole fraction concentration ΄ ΅ ]] scale − xn − xn − xn ¯ 1−xn The most common and useful reference for solution mix- tures is that of SI solutions. The following conditions must The determinant of this matrix equals x1, and the inverse hold: matrix is given by

␳ ͑ ͒ ␳ ͑ ͒ ␳ ͑ ͒ 1 −1+ 1 V1 − V2 −1+ 1 V1 − V3 ¯ −1+ 1 V1 − Vn ␳ ͑ ͒ ␳ ͑ ͒ ␳ ͑ ͒ m2 1+ 2 V1 − V2 2 V1 − V3 2 V1 − Vn ͑ SI͒−1 m ␳ ͑V − V ͒ 1+␳ ͑V − V ͒ ␳ ͑V − V ͒ ͑ ͒ Mn = ΄ 3 3 1 2 3 1 3 3 1 n ΅, 14 ]] ␳ ͑ ͒ ␳ ͑ ͒ ␳ ͑ ͒ mn n V1 − V2 n V1 − V3 ¯ 1+ n V1 − Vn

after evaluation of the required cofactors, or by the binomial It is clear that applying the SI condition to a solution ␳ /␳ inverse theorem. Here, mi = i 1, are dimensionless molali- mixture will not lead to ideality for the alternative concen- ties. Using this result in Eq. ͑10͒ the SI expressions for the tration scales. Consequently, it is interesting to note the con- KBIs can be obtained for any n component solution such that sequences of SI behavior and their effect on the other activity

n coefficients. By taking appropriate derivatives of the expres- ͑ ͒ ␮ ͓ ͑ ͔͒ SI ␳ ␬ ͓ ␳ ͑ ͔͒␾ ͑ ͒ sions in Eq. 3 , applying the SI condition for d i Eq. 11 , 1+N11 = 1RT T + ͚ 1+ 1 Vk − V1 k, 15 k=2 and then equating the results it can be shown that

SI ␥ ץ SI ץ n ͩ ln fi ͪ ͩ ln i ͪ SI =0, =−x , 1 ץ ץ . N = ␳ RT␬ − ͓1+␳ ͑V − V ͔͒␾ + ͚ ␳ ͑V − V ͒␾ 12 2 T 2 1 2 2 2 k 1 k xj T,P mj T,P,m k=3 kj ͑16͒ ln y SI V − V ץ The above equations can be simplified further to provide ͩ i ͪ = j m . ͑19͒ ␳ ␾ץ j T,P,m  1− j n k j SI ␬ ␳ 2 ͑ ͒ G11 = RT T −2V1 + ͚ kVk , 17 k=1 After integration one finds

ؠ n ͚  m V SI ␥SI k j k SI m ͑ ͒ SI ␬ ␳ 2 ͑ ͒ fi =1, i = ؠ , yi = ؠ 20 G = RT T − V1 − V2 + ͚ kVk , 18 ͚ 12 mj + kjmk Vm k=1

which are the expressions provided by the proposed Eq. ͑2͒. for any i on addition of any j component. The zero super- KBIs for the other components can be obtained by a simple script indicates a molality or molar volume of the solution ͑ ͒ ␥ index change. The expressions for the KBIs that result in Vm before addition of the j component. Therefore, i al- Ͼ ideal behavior on the molal and scales ways decreases, while yi will increase if Vj Vm and vice can be found in Appendices A and B. versa.

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-4 Ploetz, Bentenitis, and Smith J. Chem. Phys. 132, 164501 ͑2010͒

B. Deviations from ideal solution behavior S − V − V ␬ 2 1 2 ͑ ͒ G12 = RT T + , 25 KB theory provides a wealth of data on real solutions in 1+f22 terms of the molecular distributions. Thermodynamically, real solutions are characterized by activity coefficients. Gen- S − V − V f G = RT␬ + 2 1 1 − 22 , ͑26͒ eral expressions for changes in the activity coefficients can 11 T ␳ ͑ ͒ 1+f22 1 1+f22 be derived and provide further information concerning the ͒ ץ/ ץ͑ ␮ conditions for ideality and how real solutions deviate from where f22= 22−1= ln f2 ln x2 P,T and the expression for these conditions. Using Eqs. ͑3͒ and ͑4͒, together with a G22 can be obtained from a suitable index change. The sec- ͑ ͒ ͑ ͒ similar manipulation that led to Eq. ͑C5͒, one can show that ond term on the right hand side rhs of Eq. 25 is always negative and typically dominates the first term. Both the sec- n n ond and third terms on the rhs of Eq. ͑26͒ can be negative or ͩ ͪ ␮ ͑ ͒ RTd ln fi =−͚ Nij − ͚ xkNkj d j 21 positive. Alternatively, one could express the above KBIs j=1 k=1 and the deviation from their ideal values using the KB ex- 22 for any real or ideal solution at constant T and P. Alterna- pression for f22. Hence, tively, Eq. ͑C2͒ leads directly to ␬ ͑ ͒͑ ␳ ⌬ ͒ Gij = RT T + S2 − Vi − Vj 1+x1 2 G12 n + ␦ ͑1−x ͒⌬G . ͑27͒ ␥ + ␮ ͑ ͒ ij i 12 RTd ln i =−͚ Nijd j 22 j=2 The above equations illustrate that the expressions for the KBIs in real solutions are in fact strongly related to the SI iϾ T P N+ N for 1 and solutions at constant and . Here, ij= ij expressions under most reasonable conditions. We also find m ͑ N N N ͒ ͑ ͒ + j 1+ 11− i1 − j1 . Finally, Eq. 4 provides that n n SI G − ͑RT␬ + S −2V ͒ G − G ء RTd ln y = d␮ =−͚ N d␮ =−͚ ͑N − m N ͒d␮ ⌬G = 22 T 2 2 Ϸ 22 22 , ͑28͒ i i ij j ij j i1 j 12 ␾2 ␾2 j=1 j=2 1 1 ͑23͒ where the approximation should be valid when the partial molar volumes do not vary significantly with composition. at constant T and P. Clearly, the simplest expression for real The above relationships could be useful for modeling devia- solutions uses the molarity, or pseudochemical potential, ap- tions from ideal solution behavior—as obtained using the proach. Thermodynamically, stable solution mixtures require ␮ Ն ␮ Ͻ 34 Debye–Hueckel theory, for example. ii 0 and ij 0 for all components. Therefore, on in- creasing the concentration of i the molar activity coefficient Ͼ Ͻ C. Molecular dynamics simulations of symmetric of i decreases when Nii 0 and Nij 0, i.e., when self- ideal solutions association of i dominates, and vice versa. Equations ͑21͒–͑23͒ can also be used to obtain the expressions provided The KBIs represent integrals over the rdfs between the in Eq. ͑19͒. various species in solution. The rdfs are closely related to the ͑ ͒ Deviations from ideal behavior for an infinitely dilute potential of mean force Wij=−RT ln gij between species solute ͑2͒ can be expressed in terms of the KBIs and chemi- pairs obtained after averaging over all other molecules in the cal potential derivatives, system, including other i and j molecules. The only way to ϱ gain reliable insights into these rdfs is from integral equation d ln y ϱ ϱ − ͩ 2 ͪ = ͑G − G ͒ + ␳−1 studies or molecular simulations performed with accurate po- d␳ 22 21 2 tentials. Here, we present some of our recent molecular dy- 2 T,P,mk2 namics results using force fields specifically designed to re- n ϫ ␳ ͑ ϱ ϱ ͒␮ ͑ ͒ produce the experimental KBIs. Two binary systems ͚ j G2j − G21 j2, 24 jϾ2 displaying close to SI behavior were studied. The first is the classic example of benzene ͑BEN͒ and toluene ͑TOL͒. The where a general recursive relationship exists for the deriva- second involves mixtures of methanethiol ͑MSH͒ and dim- tives in Eq. ͑24͒.23 Here, the molar activity coefficient will ethylsulfide ͑MSM͒. The resulting KBIs are displayed in tend to decrease if the distribution of solutes around a central Fig. 1. ͑ ͒ solute G22 exceeds the distribution of the primary solvent Figure 1 clearly demonstrates that the simulations repro- ͑ ͒ ͑ ͒ 1 around the solute G21 , but will increase if the distribu- duce, within the statistical errors, the correct KBIs for both ͑ ͒ tion of the other cosolvents G2j exceeds the distribution of mixtures. In addition, the values of the integrals are close to the primary solvent around the solute. those predicted for SI solutions using the pure molar vol- It is also possible to rewrite the usual KBI expressions umes of 91, 109, 54, and 74 cm3 /mol for benzene, toluene, for binary solutions to emphasize the similarities to the SI methanethiol, and dimethylsulfide, respectively. We note that expressions. This is useful as the real partial molar volumes not all force fields accurately reproduce the experimental 32,33,35,36 ͑␤ ͒ are often reasonably independent of composition. Changes in KBIs. The simulated enthalpy of mixing Hm for solute activity then dictate the variation of the real KBIs the BEN/TOL and MSH/MSM mixtures with x2 =0.5 was from the SI expressions. Here, one finds that for real solu- 0.02͑1͒ and 0.01͑1͒, respectively. The experimental value is tions, 0.03 for the BEN/TOL mixture.37 The corresponding simu-

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-5 KB integrals for ideal solutions J. Chem. Phys. 132, 164501 ͑2010͒

-50 3

G 11 MSH-MSH /mol)

3 MSM-MSM G12 -100 2 MSH-MSM (cm

G )

22 r ij ( G ij g -150 1 G11

G12 /mol) 3 -75 0 G22 (cm ij G MSH-MSH MSM-MSM -125 2 G11 pure ) r ( ij /mol) 2000 g 3 1 (cm ij G22 G 0 G12 0 0.2 0.4 0.6 0.8 1 0 0 0.5 1 1.5 x2 r (nm) FIG. 1. KBIs for mixtures of benzene ͑1͒ and toluene ͑2͒ at 313 K ͑top͒, methanethiol ͑1͒ and dimethylsulfide ͑2͒ at 288 K ͑middle͒, and methanol FIG. 3. Center of based rdfs obtained from a simulation of meth- ͑ ͒ ͑ ͒ ͑1͒ and benzene ͑2͒ at 308 K ͑bottom͒. The experimental KBIs are displayed anethiol 1 and dimethylsulfide 2 at 288 K. The rdfs for a mole fraction of ͑ ͒ ͑ ͒ ͑ ͒ x2 =0.5 are displayed in the top panel. The rdfs for a mole fraction of as lines for G11 circles , G22 squares , and G12 triangles , while dashed lines represent the SI solution results ͓Eq. ͑2͔͒, and symbols corre- x2 =0.5 are compared with the distributions obtained for the pure at spond to the simulation data. The experimental KBIs were determined from the same temperature in the bottom panel. literature data ͑Refs. 37, 39,and48͒ as described previously ͑Ref. 29–31͒. the same distances but with small differences in their mag- ⌬ Ϫ ͑ ͒ ͑ ͒ 3 / nitude. In contrast, the MSH/MSM mixture displays similar lated values of Gij were 25 3 and +10 7 cm mol for the two mixtures, compared to Ϫ1 and −10 cm3 /mol ob- magnitudes for the various solvation shells, but these are served experimentally.38,39 All these values indicate close to shifted to slightly different r values. In both cases the perfect SI behavior and provide some confidence in the changes from the pure solution values are small. The differ- simulated rdfs. ent behavior observed for the rdfs in the two solutions can be explained by the fact that the BEN/TOL system will be The corresponding rdfs for mixtures with x2 =0.5 are dis- played in Figs. 2 and 3. The expected similarity between the dominated by ring-ring interactions and packing effects respective rdfs is apparent, although the two solution mix- which will be essentially the same for both components. tures accomplish this in two different ways. The BEN/TOL Hence, the variation of the rdfs with distance is also similar. mixture indicates solvation shells which essentially appear at However, in the case of MSH/MSM the different sizes of both molecules has a larger effect and the rdfs become shifted relative to each other. They display similar peak 3 heights due to a presumed similarity of their interactions. In BEN-BEN both systems g12 is intermediate between g11 and g22 and TOL-TOL further ensures that ⌬G Ϸ0. 2 BEN-TOL ij )

r For comparison, we also included the simulated and ex- ( ij

g perimental results for a nonideal solution involving a mixture 1 of one of the above components ͑BEN͒ with methanol ͑MOH͒. The KBIs and corresponding rdfs are displayed in Figs. 1 and 4. Clearly, the experimental KBIs indicate sig- 0 nificant nonideality, especially at high BEN mole fractions. BEN-BEN This is reproduced by the simulations. The rdfs indicate only TOL-TOL 2 pure minor changes for the BEN-BEN distribution, but significant ) r

( changes in the MOH-MOH distribution. Presumably, these ij g differences arise from the need for MOH molecules to satisfy 1 their hydrogen bonding requirements, which leads to a high degree of self-association at low MOH mole fractions. Simi- 0 lar trends are observed for mixtures of TOL/MOH, MSH/ 0 0.5 1 1.5 29,30 r (nm) MOH, and MSM/MOH. Mixtures of BEN/TOL and, to some extent, MSH/MSM FIG. 2. Center of mass based rdfs obtained from a simulation of benzene ͑1͒ would be expected to display close to ideal behavior. In con- ͑ ͒ and toluene 2 at 313 K. The rdfs for a mole fraction of x2 =0.5 are dis- trast, the behavior of BEN/MSH mixtures would be more played in the top panel. The rdfs for a mole fraction of x2 =0.5 are compared with the distributions obtained for the pure solvents at the same temperature difficult to predict using basic chemical principles. We per- in the bottom panel. formed additional simulations and found that a quaternary

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-6 Ploetz, Bentenitis, and Smith J. Chem. Phys. 132, 164501 ͑2010͒

⌬ 6 TABLE I. Simulated and experimental values of Gij for equimolar binary and quaternary solutions. ͑Estimated errors are displayed in parentheses.͒ BEN-BEN 4 ) MOH-MOH (r ⌬ ij MOH-BEN Gij g 2 ͑cm3 /mol͒

0 System Simulated Experimental BEN-BEN n=2 4 ) MOH-MOH Ϫ ͑ ͒ Ϫ (r pure BEN-TOL 25 3 1 ij

g MSH-MSM 10͑7͒ Ϫ10 2 BEN-MOH 1584͑159͒ 1570 0 n=4 BEN-TOL Ϫ90͑6͒ xMOH = 1.000 10 x = 0.750 BEN-MSH Ϫ5͑8͒ ) MOH (r BEN-MSM Ϫ52͑5͒ ij xMOH = 0.500 g 5 TOL-MSH Ϫ14͑13͒ xMOH = 0.250 TOL-MSM Ϫ68͑3͒ 0 MSH-MSM Ϫ43͑10͒ 0 0.5 1 1.5 r (nm)

FIG. 4. Center of mass based rdfs obtained from a simulation of benzene ͑1͒ for the different component pairs can differ from the SI value and methanol ͑2͒ at 308 K. The rdfs for a mole fraction of x =0.5 are 2 of zero, these differences are negligible in comparison with displayed in the top panel. The rdfs for a mole fraction of x2 =0.5 are com- pared with the distributions obtained for the pure solvents at the same tem- the changes observed for nonideal solutions—assuming one perature in the middle panel. The MOH-MOH rdf is displayed as a function has a well parametrized force field.5,30,36 It is possible that of methanol mole fraction in the bottom panel. ⌬ the KBIs or Gij for solution mixtures may be used to help define a variety of solution behavior. However, this requires

mixture of BEN/TOL/MSH/MSM at xi =0.25 also displays a detailed analysis of a large number of solution mixtures near ideal behavior. The results from this simulation are dis- and is beyond the scope of the current study. played in Fig. 5 and Table I. Clearly, the rdfs are somewhat different and yet they all possess first and second solvation V. DISCUSSION shells at essentially the same distances, and occur with simi- In the previous sections we established expressions for lar probabilities. The enthalpy of mixing is negligible the KBIs, in terms of the isothermal compressibility and mo- ͑␤H =0.00͒. Unfortunately, there is no experimental data m lar volumes of the pure components, which correspond to for this mixture, but the simulations suggest that this would ideal solutions using three different concentration scales. The be an interesting solution mixture for study and that ideal different ideal solutions all result in the same form for the behavior is not limited to trivial solutions with a small num- KBIs, ber of components. ideal ␬ ͑ ͒ Finally, in Fig. 6 and Table I we present a comparison of Gij = RT T − Vi − Vj + Sn, 29 the simulated and experimental ⌬G for all the mixtures con- ij except that when using the molality scale, one takes V =0 sidered here. The simulated, and presumably experimental, 1 for all G and includes a −1/␳ term for the G integral, KBIs displayed in Fig. 6 clearly distinguish between ideal ij 1 11 ⌬ while for the molarity scale the same molar volume is used and nonideal behavior. While the simulated values of Gij

1000 2.5 BEN-BEN TOL-BEN 800 2 BEN-TOL TOL-TOL MOH-MOH BEN-MSH TOL-MSH 1.5 BEN-MSM TOL-MSM 600 (r) ij g /mol) 1 3 400 (cm

0.5 ij 200 0 BEN-BEN MSH-BEN MSM-BEN 0 2

MSH-TOL MSM-TOL simulated G MSH-MSH MSM-MSH 1.5 MSH-MSM MSM-MSM -200 (r) ij g 1 -400 BEN-MOH 0.5 -150 -100 -50 0 3 0 calculated SI G (cm /mol) 0 0.5 1 1.5 0 0.5 1 1.5 2 ij r (nm) r (nm) FIG. 6. A comparison between the simulated KBIs and the SI KBIs ͓Eq. ͑2͔͒ FIG. 5. Center of mass based rdfs obtained from a simulation of a quater- for the two binary ideal solutions ͑open circles͒, one nonideal binary solu- ͑ ͒ ͑ ͒ ͑ ͒ nary mixture xi =0.25 of benzene, toluene, methanethiol, and dimethylsul- tion crosses and labels , and one quaternary solution filled circles used in fide at 300 K. this study. All data refer to equimolar mixtures.

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-7 KB integrals for ideal solutions J. Chem. Phys. 132, 164501 ͑2010͒

͑ ͒ Vi =V0 and Sn =V0 for all species. All the expressions re- the components exhibit nonideal behavior, this affects the ␸ → duce to the compressibility equation as 1 1. The use of KBIs between all of the ideal components as the solutions to /␳ ͑ ͒ ͑ ͒ V1 =0 for all KBIs, and the addition of the −1 1 term for Eq. 10 involve the determinant of the matrix in Eq. 9 . G11, in Im solutions leads to a description of the solvent as a Hence, this type of assumption should be used with care. continuum characterized by a number density, but no struc- The properties of SI solutions and their relationships to tural or volume information. the underlying rdfs and originating potential functions have 2 Only the Sn term in the KBI expressions for SI and ideal been discussed in detail by Ben-Naim. It should be noted molal scale ͑Im͒ solutions depends strongly on composition that all the KBI expressions corresponding to the three ideal ͑␬ T is also composition dependent but only slightly so when solution types represent thermodynamic models of solution away from critical points͒. This dependence is the same for behavior. This is clear from Eqs. ͑11͒, ͑A1͒, and ͑B1͒, which all Gij pairs. The Sn term plays no role in the expressions for are thermodynamic statements of ideality, and are not clearly either the partial molar volumes or chemical potential deriva- related to any physical models of the solution. It is tempting tives as these involve differences in pairs of KBIs.2,15,23 It is to present simple pictures for the underlying potential or rdfs only necessary to obtain the correct expression for the com- which might lead to ideal behavior, but this may not be pos- ͑ ͒ pressibility, as illustrated by Eq. 8 . We note that one could sible, especially for Im solutions where V1 is zero. This is in ͗ 2͘/͗ ͘ also write Sn = V V , where the angular brackets denote contrast, for example, to the physical model of solution mix- an average ͑using mole fractions͒ over the different compo- tures provided by scaled particle theory ͑SPT͒,40,41 or simple nents in the solution mixture, and are not to be considered an hard sphere potentials. On the other hand, a more physical ͗ ͘ ensemble average. Alternatively, Sn = V , if one uses volume model is unlikely to obey all the thermodynamic relation- fractions in the averaging. ships required by Eqs. ͑4͒, ͑5͒, and ͑8͒ for real solutions. There are several implications that arise from the KBI In Figs. 1–3 we provided data for two systems which expressions derived above. The Ben-Naim result,22 approach SI behavior. SI behavior over the whole composi- tion range is rare. In particular, it is unlikely that any salt will ⌬ ideal ͑ ͒ G = Gij + Gjj −2Gij =0, 30 ij display ideal behavior since G22 approaches infinity at low concentrations due to local electroneutrality constraints be- can be shown to be valid for any number of components and 42 any ideal solution as long as i, j1 in the Im case. The tween the ions. Examples of systems exhibiting Im and Ic difference between any two KBIs in ideal solutions is given behavior appear unlikely, but ideality in urea and water mix- by tures on the molarity scale is observed up to 5M–6M urea ͑ Ϸ Ϸ ͒ 16,43 m2 6.5m or x2 0.10 . Here, the activity derivative ideal ideal  Gij − Gkl = Vk + Vl − Vi − Vj = constant i, j 1 for Im presented in Eq. ͑C6͒ deviates from unity by less than 4% ͑31͒ over this range. However, the molar volumes of urea and water are 46 and 18 cm3 /mol, respectively, certainly not ͑ ͒ and is zero for ideal molar scale Ic solutions. Finally, the identical. Ideal behavior is observed because G22−G12 varies difference between the KBIs for SI and Im solutions is sim- from 50 cm3 /mol at very low concentrations to 3 / ply related to properties of the primary solvent, −1 cm mol at 5M. Consequently, when the value of G22 −G is non-negligible the value of ␳ is small, whereas GSI − GIm = ␳ V2 i, j  1, ͑32͒ 12 2 ij ij 1 1 when the urea concentration is much larger the above differ- for any number of components. Another situation of interest ence is small. Furthermore, computer simulations of urea and arises when all the components except for the primary sol- water mixtures indicate that the g22 and g12 rdfs are quite vent ͑1͒ are present at low concentrations, i.e., when they different and suggest that neither the size of the molecules, would be expected to display some ideal behavior. Here, one nor the effective interactions between them, could be consid- ␸ → 32 finds that as 1 1, ered similar. Hence, in our opinion, this is an interesting system but not a clear example of an Ic mixture. ideal ␬ ͑ ͒ Gij = RT T − Vi − Vj + V1 33 Finally, we note that there has been some controversy ͑ ͒ concerning the interpretation of the KBIs in real and ideal for all i and j except for i= j=1 in the Im case . This may be 44,45 particularly relevant for biological systems where the con- solutions. This has primarily revolved around the exact centrations of the peptides and/or proteins of interest are usu- meaning of Nij and the comparison of real and SI KBIs. The ally small. However, it should be noted that ideal behavior results presented here have no bearing on the above issue, for a component appearing at very low concentrations does they merely provide the appropriate expressions which can not imply that the KBIs adopt the corresponding ideal val- be used in either case. ues. For example, salts display large positive values for G22 even at very low salt concentrations. VI. CONCLUSIONS Experimental and theoretical data on solutions contain- ing a large number of components are rare. This is especially Expressions for the KBIs that lead to ideal solution be- true for biological systems. In these situations it is tempting havior using a variety of concentration scales have been pre- to assume that many of the components behave ideally and sented. The relationships are valid for any number of com- that one can then use the expressions provided by Eq. ͑2͒ for ponents covering the full range of composition. The the ideal components. It is immediately apparent, however, expressions for the KBIs are very similar for the different on examining Eqs. ͑9͒ and ͑10͒ that even if only a couple of concentration scales. These expressions correspond to ther-

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-8 Ploetz, Bentenitis, and Smith J. Chem. Phys. 132, 164501 ͑2010͒

modynamic models for solutions which can be used as a APPENDIX B: IDEAL SOLUTIONS ON THE MOLAR reference for understanding real solution behavior. In addi- CONCENTRATION SCALE tion, they provide a basis for studying complicated multi- Finally, ideality on the molar concentration scale ͑Ic͒ component systems for which experimental data might not requires the following relationships to hold: be available. Computer simulations of BEN/TOL and MSH/ MSM mixtures, both displaying close to SI behavior, indi- ␮Ic ␳ ͑ ͒ d i = RTd ln i, B1 cate that different patterns for the rdfs are obtained in these systems and suggest that even these solutions exhibit some ␮Ic = ␦ − ␾ . ͑B2͒ variability in their molecular distributions. However, these ij ij j differences are small in comparison with the behavior of Here, the addition of j affects all the chemical potentials as a nonideal solutions. result of a change in solution volume. These can be used in the matrix shown in Eq. ͑9͒ to obtain expressions for the KBIs which are required for ideality on the molarity scale. ACKNOWLEDGMENTS The initial results are Ic ͑ ␬ ͒/ ͑ ͒ The project described was supported by Grant No. Gij = Vj RT T − Vi Sn, B3 R01GM079277 from the National Institute of General Medi- which are valid for any number of components. These ex- cal Sciences. The content is solely the responsibility of the pressions obey Eqs. ͑5͒ and ͑8͒, but give the appearance that authors and does not necessarily represent the official views  Gij Gji, an impossible outcome according to the well of the National Institute of General Medical Sciences or the known fluctuation formula for the KBIs,1 National Institutes of Health. ͗N N ͘ − ͗N ͗͘N ͘ ␦ G = Vͫ i j i j − ij ͬ. ͑B4͒ ij ͗ ͗͘ ͘ ͗ ͘ Ni Nj Ni APPENDIX A: IDEAL SOLUTIONS ON THE MOLALITY Consequently, the expression provided in Eq. ͑B3͒ can only CONCENTRATION SCALE be true if Vi =Vj, and therefore it is more appropriate to write ͑ ͒ The conditions for ideality depend on the concentration Eq. B3 as scale adopted. If one desires ideal behavior on the molality GIc = V ͑RT␬ − V ͒/S = RT␬ − V , ͑B5͒ scale ͑Im͒ then the following relationships must hold: ij 0 T 0 n T 0 where V is the same for all components. Hence, all G have d␮Im = RTd ln m , ͑A1͒ 0 ij i i to be identical and are independent of composition for ideal behavior on the molarity scale, a much more restrictive con- ␮Im = ␦ , i, j  1, ␮Im =−m , ͑A2͒ ij ij 1j j dition than for SI or Im solutions. This extra restriction arises where the latter is obtained from the Gibbs–Duhem equation as the addition of any component j changes the concentration at constant T and P.23 Clearly, the addition of j has no affect of all the species through the solution volume. The only way on the molality of i unless i= j. The above expressions can be this change can be the same for all j is if they have the same used in the matrix shown in Eq. ͑9͒ to obtain the expressions molar volumes. for the KBIs which are required for ideality on the molality scale. The only additional problem lies in the fact that solv- ͑ ͒ ing Eq. 10 provides only the KBIs to the primary solvent 1, APPENDIX C: PROOF OF THE KBI EXPRESSIONS which occupies a unique position for the molality scale. FOR IDEAL SOLUTIONS However, choosing i=2 and k1 in Eqs. ͑5͒ and ͑8͒ pro- vides a new set of equations which can be solved in the same To demonstrate that Eqs. ͑2͒, ͑A4͒, and ͑B5͒ are indeed  ␮ way as before to generate expressions for i, j 1. We will not correct, we will derive the appropriate expressions for d i present all the matrices in the interest of brevity. The final using the different ideal KBIs in solutions containing any results are number of components. This ensures that the symbolic ma- n trix inversion was performed correctly. The easiest to prove Im ␬ Ј /␳ Ј ␳ 2 is the Ic case. Combining Eqs. ͑4͒ and ͑B5͒, one finds G11 = RT T + Sn −1 1, Sn = ͚ kVk , k=2 n RTd ln ␳ = d␮Ic + ͚ ␳ ͑RT␬ − V ͒d␮Ic. ͑C1͒ Im Im ␬ Ј  i i j T 0 j G1j = Gj1 = RT T − Vj + Sn, j 1, j=1 ⌺ ␳ GIm = RT␬ − V − V + SЈ, i, j  1. ͑A3͒ Using the Gibbs–Duhem equation at constant T and P, j j ij T i j n ␮ d j =0, the summation term disappears and the required re- The above expressions can be generalized by setting V1 =0 to lationship ͓Eq. ͑B1͔͒ is obtained. give To prove ideality on the molality scale one can start with ͑ ͒ ␮ GIm RT␬ V V S ␦ ␦ ␳−1 ͑ ͒ Eq. 4 , eliminate d 1 using the Gibbs–Duhem relationship, ij = T − i − j + n − i1 j1 1 , A4 ␳ and then convert to molalities using d ln mi =d ln i ␳ 1,23,46 which is valid for any number of components. −d ln 1 to provide the following expression:

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions 164501-9 KB integrals for ideal solutions J. Chem. Phys. 132, 164501 ͑2010͒

n 1 J. G. Kirkwood and F. P. Buff, J. Chem. Phys. 19, 774 ͑1951͒. 2 RTd ln m = ͚ ͑␦ + N+͒d␮ , ͑C2͒ A. Ben-Naim, Statistical Thermodynamics for Chemists and Biochemists i ij ij j ͑Plenum, New York, 1992͒. j=2 3 E. Matteoli and G. A. Mansoori, Fluctuation Theory of Mixtures ͑Taylor i n ͓ ͑ ͔͒ & Francis, New York, 1990͒. for =2, . Using the KBIs for Im solutions Eq. A4 one 4 ͑ ͒ +  ͑ ͒ E. Matteoli and L. Lepori, J. Chem. Phys. 80, 2856 1984 . finds that Nij=0 for all i, j 1. Therefore, Eq. A1 is 5 V. Pierce, M. Kang, M. Aburi, S. Weerasinghe, and P. E. Smith, Cell obeyed. Biochem. Biophys. 50,1͑2008͒. Finally, for SI solutions, we start with Eq. ͑4͒ and con- 6 I. L. Shulgin and E. Ruckenstein, J. Chem. Phys. 123, 054909 ͑2005͒. 7 vert to molalities to provide the expression23 S. Shimizu, Proc. Natl. Acad. Sci. U.S.A. 101, 1195 ͑2004͒. 8 J. Rosgen, B. M. Pettitt, and D. W. Bolen, Protein Sci. 16,733͑2007͒. n 9 J. P. O’Connell, Mol. Phys. 20,27͑1971͒. 10 RTd ln m = ͚ ͑␦ + N − ␦ − N ͒d␮ . ͑C3͒ A. A. Chialvo, J. Phys. Chem. 97, 2740 ͑1993͒. i ij ij 1j 1j j 11 ͑ ͒ j=1 Y. Marcus, Monatsch. Chem. 132, 1387 2001 . 12 P. E. Smith and R. M. Mazo, J. Phys. Chem. B 112, 7875 ͑2008͒. 13 Using the SI KBI expressions presented in Eq. ͑2͒ and the F. Chen and P. E. Smith, J. Phys. Chem. B 112, 8975 ͑2008͒. 14 E. Ruckenstein and I. Shulgin, Fluid Phase Equilib. 180,281͑2001͒. Gibbs–Duhem relation, this reduces to 15 P. E. Smith, Biophys. J. 91, 849 ͑2006͒. 16 ␮SI ␮SI ͑ ͒ H. Kokubo, J. Rosgen, D. W. Bolen, and B. M. Pettitt, Biophys. J. 93, RTd ln mi = d i − d 1 . C4 3392 ͑2007͒. 17 A. Ben-Naim, A. M. Navarro, and J. M. Leal, Phys. Chem. Chem. Phys. Noting that changes in the mole fractions can be written as 10, 2451 ͑2008͒. ⌺ 47 ͑ ͒ 18 d ln xi =d ln mi − jxjd ln mj, and using Eq. C4 leads A. Ben-Naim, J. Chem. Phys. 63,2064͑1975͒. gives 19 M. B. Gee and P. E. Smith, J. Chem. Phys. 131, 165101 ͑2009͒. 20 E. Ruckenstein and I. Shulgin, Fluid Phase Equilib. 180,345͑2001͒. n 21 P. E. Smith, J. Chem. Phys. 129, 124509 ͑2008͒. ␮SI ␮SI ͑ ␮SI ␮SI͒ ͑ ͒ 22 Molecular Theory of Solutions ͑ RTd ln xi = d i − d 1 − ͚ xj d j − d 1 . C5 A. Ben-Naim, Oxford University Press, j=1 New York, 2006͒. 23 M. Kang and P. E. Smith, J. Chem. Phys. 128, 244511 ͑2008͒. After a further application of the Gibbs–Duhem expression 24 E. Lindahl, B. Hess, and D. van der Spoel, J. Mol. Model. 7,306͑2001͒. 25 the above equation reduces to Eq. ͑11͒ as required. H. J. C. Berendsen, J. P. M. Postma, W. F. van Gunsteren, A. Dinola, and J. R. Haak, J. Chem. Phys. 81, 3684 ͑1984͒. One can also demonstrate that the KBI expressions for 26 B. Hess, H. Bekker, H. J. C. Berendsen, and J. G. E. M. Fraaije, J. SI, Im, and Ic solutions are correct by applying them to the Comput. Chem. 18, 1463 ͑1997͒. established relationships for binary solutions. The solute 27 T. Darden, D. York, and L. Pedersen, J. Chem. Phys. 98, 10089 ͑1993͒. 28 S. Weerasinghe and P. E. Smith, J. Phys. Chem. B 109, 15080 ͑2005͒. chemical potential derivatives for a binary system of solvent 29 ͑ ͒ ͑ ͒ 2 N. Bentenitis, N. R. Cox, and P. E. Smith, J. Phys. Chem. B 113, 12306 1 and solute 2 are given by ͑2009͒. 30 E. A. Ploetz, N. Bentenitis, and P. E. Smith, Fluid Phase Equilib. 290,43 ͒ ͑ ␮ ␳ + ␳ 1ץ ␤ͩ 2 ͪ = 1 2 = , 2010 . .␳ ␳ ␳ ␳ ⌬ ␳ ⌬ 31 E. A. Ploetz and P. E. Smith ͑unpublished͒ ץ ln x2 T,P 1 + 2 + 1 2 G12 1+x2 1 G12 32 S. Weerasinghe and P. E. Smith, J. Phys. Chem. B 107,3891͑2003͒. 33 S. Weerasinghe and P. E. Smith, J. Phys. Chem. B 118, 10663 ͑2003͒. ͑ ␮ 1 1 34ץ ␤ͩ 2 ͪ I. Prigogine and R. Defay, Chemical Thermodynamics Longmans, Lon- = = , ͒ . ln m 1+N+ 1+m ͑1+␳ ⌬G ͒ don, 1954 ץ 2 T,P 22 2 1 12 35 R. Chitra and P. E. Smith, J. Chem. Phys. 115,5521͑2001͒. 36 A. Perera and F. Sokolic, J. Chem. Phys. 121, 11272 ͑2004͒. ␮ 1 37 S. Murakami, V. T. Lam, and G. C. Benson, J. Chem. Thermodyn. 1,397ץ ␤ͩ 2 ͪ ͑ ͒ = . C6 ͑1969͒. ln ␳ 1+␳ ͑G − G ͒ 38 ץ 2 T,P,N1 2 22 12 K. D. Kassmann and H. Knapp, Ber. Bunsenges. Phys. Chem. 90,452 ͑1986͒. The above derivatives all reduce to unity after inserting the 39 A. W. Jackowkski, Pol. J. Chem. 54, 1765 ͑1980͒. appropriate expressions for the KBIs. This is also true for the 40 H. Reiss, H. L. Frisch, and J. L. Lebowitz, J. Chem. Phys. 31,369 ͑1959͒. corresponding expressions describing ternary and quaternary 41 23 R. A. Pierotti, Chem. Rev. ͑Washington, D.C.͒ 76, 717 ͑1976͒. solutions. Furthermore, the KBI expressions for ideal solu- 42 P. G. Kusalik and G. N. Patey, J. Chem. Phys. 86,5110͑1987͒. tions provided above all lead to the correct expressions for 43 R. H. Stokes, Aust. J. Chem. 20, 2087 ͑1967͒. the partial molar volumes and compressibility for solutions 44 A. Ben-Naim, J. Phys. Chem. B 112, 5874 ͑2008͒. 45 I. L. Shulgin and E. Ruckenstein, J. Phys. Chem. B 112,5876͑2008͒. with any number of components. This can be shown by in- 46 ͑ ͒ ͑ ͒ D. G. Hall, Trans. Faraday Soc. 67, 2516 1971 . serting the ideal KBI expressions into Eq. 5 for the partial 47 P. E. Smith, J. Phys. Chem. B 110, 2862 ͑2006͒. molar volumes and into Eq. ͑8͒ for the compressibility. 48 G. M. Wilson, J. Am. Chem. Soc. 86, 127 ͑1964͒.

Downloaded 04 Aug 2011 to 129.130.37.167. Redistribution subject to AIP license or copyright; see http://jcp.aip.org/about/rights_and_permissions