13
J. Phys. Chem. 1992, 96, 5561-5519 5567 Thermodynamic Description of Micellization, Phase Behavior, and Phase Separatlon of Aqueous Solutions of Surfactant Mixtures Sudhakar Puvvadat and Daniel Blankschtein* Department of Chemical Engineering and Center for Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 021 39 (Received: December 9, 1991; In Final Form: March 24, 1992) We present a thermodynamic framework to describe micellization, phase behavior, and phase separation of aqueous solutions of surfactant mixtures. The theoretical framework consists of evaluating the Gibbs free energy of the solution which is modeled as the sum of three contributions: the free energy of formation of the various solution species, the free energy of mixing micelles, monomers, and water molecules, and the free energy of interactions between the various solution species. By utilizing the conditions of multiple chemical equilibria and thermodynamic stability, all relevant micellar and thermodynamic properties of the solution can be predicted. The predicted properties include (i) the critical micellar concentration (cmc), (ii) the micellar size and composition distribution including its moments, (iii) thermodynamic properties such as the osmotic pressure and compressibility, (iv) the critical line signaling the onset of phase separation, and (v) the coexisting surface bounding the two-phase region of the phase diagram. We find that the predicted properties i-v can be evaluated from knowledge of two molecular contributions: the free energy of mixed micellization, gd, reflecting intramicellar interactions, and a mean-field intermicellar interaction parameter, C , . In this paper, we formulate a simplified phenomenological model for g,, and provide a physical rationalization of the typical values that gdc and Cen can attain for various types of binary surfactant mixtures. Using these typical values of gdc and Ccn, we are able to predict qualitative trends of micellar and phase behavior properties as a function of surfactant type and composition. Specifically, we have derived an expression for the mixture cmc which is identical to the well-known expression derived by Rubingh in the context of the pseudophase separation model using the regular solution theory with an empirical interaction parameter c. Therefore, in our formulation, the molecular basis of the parameter c is clarified. We have also derived expressions for the variations of the optimum micellar composition with solution monomer composition and the weight-average mixed micelle aggregation number with total surfactant composition. In addition, utilizing the conditions of thermodynamic stability, the effects of adding small amounts of a second surfactant on the phase separation characteristics of a single-surfactant solution have been investigated. Specifically, we have derived expressions for the change in the critical temperature with total surfactant composition, as well as for the distribution of the added surfactant between the two coexisting micellar phases. We fmd that all our qualitatiue theoreticalpredictions reproduce very well the experimentally observed trends in aqueous solutions containing nonionic-nonionic, nonionic-ionic, zwitterionic-ionic, and anionic-cationic surfactant mixtures. In addition, we predict some new interesting trends which have not yet been observed experimentally. I. Introduction Surface-active compounds used in commercial applications typically consist of a mixture of surfactants because they can be produced at a relatively lower cost than that of isomerically pure surfactants. In addition, in many surfactant applications, mixtures of dissimilar surfactants often exhibit properties superior to those of the constituent single surfactants due to synergistic interactions between the surfactant molecules.' Indeed, in solutions containing mixtures of surfactants, the tendency to form aggregated structures (mixed micelles) can be substantially different than in solutions containing only the constituent single surfactants. For example, the critical micellar concentration (cmc) of a mixture of anionic and cationic surfactants in aqueous solution is considerably lower than the cmc's of each individual surfactant.* On the other hand, antagonistic interactions, in mixtures of hydrocarbon-based and fluorocarbon-based surfactants in aqueous solution, result in mixture cmc's that can be considerably higher than the cmc's of the constituent single surfactants.j In general, specific interactions (synergistic or antagonistic) between surfactants result in solutions of surfactant mixtures having micellar and phase behavior properties which can be significantly different from those of the constituent single surfactants. Consequently, understanding specific interactions between the various surfactant species present in the solution is of central importance to the surfactant tech- nologist. Indeed, in order to tailor surfactant mixtures to a particular application, the surfactant technologist has to be able to predict and manipulate (i) the tendency of surfactant mixtures to form monolayers, micelles, and other self-assembling aggregates in solution, (ii) the properties of the formed aggregates such as their shape and size, (iii) the distribution of the various surfactant 'Present address: Center for Bio/Molecular Science and Engineering, *To whom correspondence should be addressed. Naval Research Laboratory, Washington, D.C. 20375-5000. 0022-3654/92/2096-5561~03.~Q/O species between monomers and aggregates, and (iv) the phase behavior and phase equilibria of solutions containing surfactant mixtures. In spite of their considerable practical importance, as well as the challenging theoretical issues associated with the description of these complex fluids, solutions of surfactant mixtures have not received the full attention that they deserve. Specifically, previous theoretical studies of mixed micellar solutions have evolved along two very different, seemingly unrelated, fronts. On the one hand, significant efforts have been devoted to understand the mixture cmc,"6 as well as the micellar size and composition distribution.7J On the other hand, very little effort has been devoted to understand the solution behavior at higher surfactant concentrations where intermicellar interactions become increasingly important and control the phase behavior and phase separation phenomena? In view of this, it is quite clear that there is an immediate need to develop a theoretical description of mixed micellar solutions ca- pable of unifying the previously disconnected treatments of micellization and phase behavior, including phase separation, into a single coherent computational framework. The central goal of this paper is to contribute to this much needed theoretical unification. The theoretical approach that we p r o p , inspired by our recent work on single-surfactant solutions,I0 consists of blending a thermodynamic theory of mixed micellar solutions, which captures the salient features of these systems at the macroscopic level, with a molecular model of mixed micellization, which captures the essential physical factors at the micellar level. The resulting molecular-thermodynamic approach provides a valuable tool to predict solution properties of mixed surfactant systems using molecular information that reflects (i) the nature of the surfactant molecules involved in the micellization process and (ii) solution conditions such as temperature and the presence of additives such as salts and urea. As such, the molecular-thermodynamic ap- 0 1992 American Chemical Society

Thermodynamic description of micellization, phase behavior, and phase separation of aqueous solutions of surfactant mixtures

Embed Size (px)

Citation preview

J. Phys. Chem. 1992, 96, 5561-5519 5567

Thermodynamic Description of Micellization, Phase Behavior, and Phase Separatlon of Aqueous Solutions of Surfactant Mixtures

Sudhakar Puvvadat and Daniel Blankschtein* Department of Chemical Engineering and Center for Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 021 39 (Received: December 9, 1991; In Final Form: March 24, 1992)

We present a thermodynamic framework to describe micellization, phase behavior, and phase separation of aqueous solutions of surfactant mixtures. The theoretical framework consists of evaluating the Gibbs free energy of the solution which is modeled as the sum of three contributions: the free energy of formation of the various solution species, the free energy of mixing micelles, monomers, and water molecules, and the free energy of interactions between the various solution species. By utilizing the conditions of multiple chemical equilibria and thermodynamic stability, all relevant micellar and thermodynamic properties of the solution can be predicted. The predicted properties include (i) the critical micellar concentration (cmc), (ii) the micellar size and composition distribution including its moments, (iii) thermodynamic properties such as the osmotic pressure and compressibility, (iv) the critical line signaling the onset of phase separation, and (v) the coexisting surface bounding the two-phase region of the phase diagram. We find that the predicted properties i-v can be evaluated from knowledge of two molecular contributions: the free energy of mixed micellization, gd, reflecting intramicellar interactions, and a mean-field intermicellar interaction parameter, C,. In this paper, we formulate a simplified phenomenological model for g,, and provide a physical rationalization of the typical values that gdc and Cen can attain for various types of binary surfactant mixtures. Using these typical values of gdc and Ccn, we are able to predict qualitative trends of micellar and phase behavior properties as a function of surfactant type and composition. Specifically, we have derived an expression for the mixture cmc which is identical to the well-known expression derived by Rubingh in the context of the pseudophase separation model using the regular solution theory with an empirical interaction parameter c. Therefore, in our formulation, the molecular basis of the parameter c is clarified. We have also derived expressions for the variations of the optimum micellar composition with solution monomer composition and the weight-average mixed micelle aggregation number with total surfactant composition. In addition, utilizing the conditions of thermodynamic stability, the effects of adding small amounts of a second surfactant on the phase separation characteristics of a single-surfactant solution have been investigated. Specifically, we have derived expressions for the change in the critical temperature with total surfactant composition, as well as for the distribution of the added surfactant between the two coexisting micellar phases. We fmd that all our qualitatiue theoreticalpredictions reproduce very well the experimentally observed trends in aqueous solutions containing nonionic-nonionic, nonionic-ionic, zwitterionic-ionic, and anionic-cationic surfactant mixtures. In addition, we predict some new interesting trends which have not yet been observed experimentally.

I. Introduction Surface-active compounds used in commercial applications

typically consist of a mixture of surfactants because they can be produced at a relatively lower cost than that of isomerically pure surfactants. In addition, in many surfactant applications, mixtures of dissimilar surfactants often exhibit properties superior to those of the constituent single surfactants due to synergistic interactions between the surfactant molecules.' Indeed, in solutions containing mixtures of surfactants, the tendency to form aggregated structures (mixed micelles) can be substantially different than in solutions containing only the constituent single surfactants. For example, the critical micellar concentration (cmc) of a mixture of anionic and cationic surfactants in aqueous solution is considerably lower than the cmc's of each individual surfactant.* On the other hand, antagonistic interactions, in mixtures of hydrocarbon-based and fluorocarbon-based surfactants in aqueous solution, result in mixture cmc's that can be considerably higher than the cmc's of the constituent single surfactants.j In general, specific interactions (synergistic or antagonistic) between surfactants result in solutions of surfactant mixtures having micellar and phase behavior properties which can be significantly different from those of the constituent single surfactants. Consequently, understanding specific interactions between the various surfactant species present in the solution is of central importance to the surfactant tech- nologist. Indeed, in order to tailor surfactant mixtures to a particular application, the surfactant technologist has to be able to predict and manipulate (i) the tendency of surfactant mixtures to form monolayers, micelles, and other self-assembling aggregates in solution, (ii) the properties of the formed aggregates such as their shape and size, (iii) the distribution of the various surfactant

'Present address: Center for Bio/Molecular Science and Engineering,

*To whom correspondence should be addressed. Naval Research Laboratory, Washington, D.C. 20375-5000.

0022-3654/92/2096-5561~03.~Q/O

species between monomers and aggregates, and (iv) the phase behavior and phase equilibria of solutions containing surfactant mixtures.

In spite of their considerable practical importance, as well as the challenging theoretical issues associated with the description of these complex fluids, solutions of surfactant mixtures have not received the full attention that they deserve. Specifically, previous theoretical studies of mixed micellar solutions have evolved along two very different, seemingly unrelated, fronts. On the one hand, significant efforts have been devoted to understand the mixture cmc,"6 as well as the micellar size and composition distribution.7J On the other hand, very little effort has been devoted to understand the solution behavior at higher surfactant concentrations where intermicellar interactions become increasingly important and control the phase behavior and phase separation phenomena? In view of this, it is quite clear that there is an immediate need to develop a theoretical description of mixed micellar solutions ca- pable of unifying the previously disconnected treatments of micellization and phase behavior, including phase separation, into a single coherent computational framework. The central goal of this paper is to contribute to this much needed theoretical unification.

The theoretical approach that we p r o p , inspired by our recent work on single-surfactant solutions,I0 consists of blending a thermodynamic theory of mixed micellar solutions, which captures the salient features of these systems at the macroscopic level, with a molecular model of mixed micellization, which captures the essential physical factors at the micellar level. The resulting molecular-thermodynamic approach provides a valuable tool to predict solution properties of mixed surfactant systems using molecular information that reflects (i) the nature of the surfactant molecules involved in the micellization process and (ii) solution conditions such as temperature and the presence of additives such as salts and urea. As such, the molecular-thermodynamic ap-

0 1992 American Chemical Society

5568 The Journal of Physical Chemistry, Vol. 96, No. 13, 1992 Puwada and Blankschtein

in the critical temperature as well as the distribution of the two surfactant species between the coexisting micellar-rich and mi- cellar-poor phases. Finally, in section IV we present some con- cluding remarks.

11. Thermodynamic Framework A. Cibbs Free Energy and Chemical Potentials. The ther-

modynamic formulation used to describe the free energy of a mixed surfactant solution constitutes a generalization of the one developed to describe single-surfactant s01utions.l~ For the sake of clarity, in this paper we present a theoretical description of aqueous solutions of a mixture of two surfactants. Extension of the for- malism to describe solutions containing additional surfactant species is conceptually similar and therefore will not be discussed.

Consider a solution of N , water molecules, NA surfactant A molecules, and NB surfactant B molecules in thermodynamic equilibrium at temperature T and pressure P. If the concentration of the surfactant mixture exceeds its cmc, the surfactant molecules will self-assemble to form a distribution of mixed micelles {N,,,], where N,,, is the number of mixed micelles having aggregation number n and composition a. Note that in such a mixed micelle there are na surfactant A molecules and n( 1 - a) surfactant B molecules. Note also that NA = C,,paN,,, and NB = &n(l - a)~,, , . In the spirit of the multiple-chemical equilibrium de- script~on,'~ mixed micelles of different sizes and compositions are treated as distinct species in chemical equilibrium with each other as well as with the free monomers in the solution.

The Gibbs free energy of the mixed surfactant solution C is modeled as the sum of three contributions: the free energy of formation Cr, the free energy of mixing G,, and the free energy of interaction Ci. These contributions, as in the single-surfactant case,14 are chosen to provide a heuristically appealing identification of the various factors responsible for micelle formation and growth, on the one hand, and for phase behavior and phase equilibria, on the other.

The free energy of formation is expressed as

= Nwk'w + NAF'A + NBP'B + CnNnagmic(sh,n,a) (1) n,a

where pow(T,P), woA(T,P), and poB(T,P) are the standard-state chemical potentials of water, surfactant A monomers, and sur- factant B monomers, respectively, at the solution temperature T and pressure P; gmi,(sh,n,a) is the free energy of mixed micelli- zation, which represents the free energy change per monomer associated with transferring na surfactant A monomers and n( 1 - a) surfactant B monomers from water into a mixed micelle of shape sh, aggregation number n, and composition a. The nu- merical magnitude of g&, which reflects the propensity of a mixed micelle to form and subsequently grow, summarizes the many complex physicochemical factors responsible for mixed micelle formation such as the hydrophobic effect, hydrogen bonding, conformational changes associated with hydrophobic-tail packing in the micellar core, steric and electrostatic interactions between the hydrophilic head groups, and the entropy of mixing the two surfactant species in the mixed micelle.lOJ1

The free energy of mixing the formed mixed micelles, free monomers, and water is modeled by an expression of the form

(2)

where X, = N,/(N, + NA + NB), X, = N,/(N, + NA + NB), k is the Boltzmann constant, and Tis the absolute temperature. -Gm/T is an entropic contribution which reflects the number of ways in which the distribution of mixed micelles, the free mo- nomers, and the water molecules can be positioned in the solution as a function of the solution concentration and composition. The free energy of mixing, as expressed in eq 2, opposes the tendency to form micelles because monomer aggregation reduces the total number of available spatial configurations. The entropy of mixing also opposes the tendency of the micellar solution to phase separate, because of the loss in available spatial configurations associated with this phenomenon. Note that eq 2 is a generalization of the

G, = kT[N, In X, + EN,,, In X,,,] n,a

proach may be utilized to design surfactant mixtures for a par- ticular application, as well as to modify and control the resulting micellar solution properties.

In this paper, we formulate a thermodynamic free energy model of mixed micellar solutions which captures the essential physical factors governing the behavior of these complex fluids in terms of two molecular contributions: the free energy of mixed mi- cellization, gdc, reflecting intramicellar contributions (see section IID for details), and the mean-field interaction parameter, C,, reflecting the magnitude of intermicellar attractions which are responsible for phase separation (see section IIA and Appendix A for details). It is noteworthy that all the thermodynamic properties of the mixed micellar solution can be computed solely from a knowledge of gmic and C,fp

The value of Ceff can be estimated as explained in section IIIA(b). The value of gmic can be estimated using a molecular model of mixed micellization which accounts for the detailed molecular structure of the various surfactant species, as well as for the effect of solution conditions. Such a molecular model for gmic has been developed1'J2 recently and has been successfully utilized to predict quantitatively a broad spectrum of micellar and phase behavior properties of aqueous solutions containing binary nonionic surfactant mixtures. Furthermore, a simplified version of this model has also been successfully utilized to predict quantitatively cmc's of aqueous solutions containing anionic- nonionic, cationic-nonionic, and anionic-cationic surfactant mixtures.I2J3

The calculation of giC, in the context of the detailed molecular model of micellization, is quite involved and needs to be perjiormed numerically. As a result, it is not possible to obtain analytical expressions for micellar and phase behavior properties of interest. Furthermore, in view of the numerical nature of the resulting quantitative predictions, it is not straightforward to gain a clear appreciation of the relative importance of the various molecular factors contributing to the predicted properties. In view of this, in this paper, we decided to present a simplified model for gmi, (see section IID(c) for details). Although we recognize that with this simplified description of gd our predictions will be qualitative in nature, the proposed approach offers a number of useful fea- tures. These include the following: (i) Analytical expressions for many of the micellar solution properties of interest can be derived (see, for example, eqs 41,46, and 48). (ii) Qualitative trends can be predicted without the need of performing rather complex and time-consuming numerical calculations of the type presented in refs 10-12. In other words, the simplified approach proposed here can serve as a useful preliminary screening test in the design and selection of surfactant mixtures of practical importance. (iii) It is possible to rationalize the molecular basis of the predicted trends, without the complicating effects of the numerical calculations.

The remainder of the paper is organized as follows. In section I1 we formulate the general thermodynamic framework to describe mixed micellar solutions. Specifically, in section IIA we present the mathematical structure of the various contributions to the Gibbs free energy of the mixed micellar solution and derive ex- pressions for the chemical potentials of the various solution components. In section IIB we derive expressions for the micellar size and composition distribution and its moments, and in section IIC we analyze the phase behavior and phase separation phe- nomena. An important contribution to the micellar solution Gibbs free energy is the free energy of mixed micellization, gmic, and a simplified model for gdc is developed in section IID. In section I11 we present and discuss qualitative predictions of the ther- modynamic framework for aqueous solutions containing binary mixtures of nonionic-nonionic, nonionic-ionic, zwitterionic-ionic, anionic-cationic, and nonionic hydrocarbon-nonionic fluorocarbon surfactants. The predictions include (i) the cmc variation with surfactant monomer composition, (ii) the variation of micellar composition with surfactant monomer composition at the cmc, (iii) the variation of the weight-average mixed micelle aggregation number with total surfactant composition, and (iv) the effects of adding small quantities of a second surfactant on the phase sep- aration characteristics of the micellar solution, including the shift

Aqueous Solutions of Surfactant Mixtures

expression used with considerable success to d e ~ c r i b e ~ ~ * ' ~ * ~ ~ J ' micellization, phase behavior, and phase separation in single- surfactant solutions.

The free energy of interaction reflects interactions between mixed micelles, water molecules, and free monomers in the so- lution. At the level of a mean-field type quadratic expansion, in Appendix A we show that this free energy contribution takes the following form

where 4 = 'A + is the sum of the volume fractions, 'A and &,, of surfactants A and B, respectively, = NA/(NA + NE) is the composition of the surfactant mixture, and C,~(CY,~,,) is an effective mean-field interaction parameter for the mixture which is related to the single-surfactant interaction parameters CAW and CBw and a specific interaction parameter CAB through the fol- lowing expression (see Appendix A)

Ccfdasoln) = CAWasoln + cBW(1 - asoh) - CABasoln(l - a m l n ) ( G / Y e f f )

(4)

In eq 4, yA = n A / n w and yB = nB/fl,, where n,, nA, and nB are the effective molecular volumes of water, surfactant A, and surfactant B, respectively, and Teff = asoln~A + (1 - asoln)yB.

Note that, as stressed in the Introduction, the Gibbs free energy expressions given in eqs 1-3, as well as all the thermodynamic properties derived from it, are uniquely determined by the two molecular contributions gmic and Ccrp A detailed discussion of gmic is presented in section IID, and a physical rationalization of the typical values that gmic and Ccfr can attain for various types of binary surfactant mixtures is presented in sections IIIA(a) and IIIA(b), respectively.

The thermodynamic consequences of the proposed free energy formulation are described below. The properties examined include the cmc, the micellar size and composition distribution and its moments, and the phase behavior and phase equilibria. All these properties are governed by the proposed Gibbs free energy model through the chemical potential of water, p,, and the chemical potential of a mixed micelle, of aggregation number n and com- position a, pn,, which are obtained by differentiating the Gibbs free energy, eqs 1-3, with respect to N, and Nn,, respectively. The resulting expressions are given by

kT) + kT(1n X,,, + n(X - 1 - EX,,,)) + nap'A + n(l - a)&

( 6 ) nu

where X = X A + XB, with X A = NA/(Nw + NA + NE) and X B = NB/(Nw + NA + N E ) , is the total mole fraction of surfactant in the solution, and the interaction contributions to the monomer chemical potentials are given by

p'A = -3: '[CAW + $[asolncAW + ( l - asoln)cBWl(l - 4) -

CAB(1 - 4 A ) (7) (1 - asod- 1 6 Ycrr

and

The Journal of Physical Chemistry, Vol. 96, No. 13, 1992 5569

r

The chemical potentials of surfactant A and surfactant B monomers can be obtained from eq 6 by substituting n = 1 and a = 1 (for A) or 0 (for B), respectively, that is

where X1A and X I B are the mole fractions of free surfactant A and surfactant B monomers, respectively. Using eqs 5-10 for the chemical potentials, below we predict various properties of the mixed micellar solution. B. Micellar Si and Composition Distribution and IQ M o m &

When the mixed micellar solution is in thermodynamic equilib- rium, the chemical potential p, of a mixed micelle of aggregation number n and composition a is related to the chemical potentials of the free monomers through the constraints imposed by the conditions of multiple chemical equilibrium,'* that is

(11) Equation 11 implies that the chemical potential of a mixed micelle having aggregation number n and composition a is equal to the sum of the chemical potentials of its constituent na surfactant A and n(1 - a) surfactant B molecules. Substituting eqs 6, 9, and 10 in eq 11, we obtain the following expression for the equilibrium micellar size and composition distribution

pna = ~ W A + 4 1 - ~ ) P B

where fl = l/kT, flg, = [fig~c - 1 - a In a1 - (1 - a) In (1 - a l ) ] is a modified dimensionless free energy of mixed micellization per monomer, X 1 = X1A + X I B is the total mole fraction of free monomers in solution, and al = X I A / X I is the composition of free monomers in solution. Equation 12 indicates that a delicate balance between two opposing factors determines the nature of the micellar size and composition distribution {X,]. The first is the Boltzmann factor, e-'@g*, representing the energetic advantage (recall that gmic < 0) of assembling the various surfactant mol- ecules in a mixed micelle, which favors micelle formation. The second factor, XIAna XIBn(l-,), represents the large entropic dis- advantage associated with localizing na surfactant A molecules (each with probability X I A ) and n( 1 - a) surfactant B molecules (each with probability X I B ) in a single mixed micelle and opposes micelle formation. It is noteworthy that, with the choice of mean-field interaction potentials adopted in eq 3 (see also Ap- pendix A), the interaction free energy does not affect the micellar size and composition distribution.

The composition a*(n), at which X,, exhibits a maximum for a given micellar aggregation number n, is referred to hereafter as the optimum composition. Note that, in general, a*(n) is a function of the aggregation number n and can be obtained by setting the derivative of X,,, with respect to a equal to zero. Specifically, implementing this procedure with eq 12 leads to the following implicit equation

Using eq 13, the optimum compition for all aggregation numbers can be determined from a knowledge of a, and gmic(n,a). In addition, note that, for large n, X, will exhibit a sharp maximum at a = a* because n appears as a multiplicative factor in the exponential of eq 12, and consequently any small deviations of a from a* will be magnified significantly. This implies that, to

5570 Tke Journal of Physical Chemistry, Vol. 96, No. 13,. 1992 Puwada and Blankschtein

leading order in a, the mole fraction of micelles having compo- sitions a # a* will be negligible, and accordingly, E&,, can be approximated by Xna..

Equation 12 for the micellar size and composition distribution is applicable to mixed micelles of all shapes, sizes, and compo- sitions. However, as shown clearly in eq 12, to determine the distribution one needs to know (i) the free energy of micellization g,,&,a) as a function of n and a or, equivalently, g, as a function of n, a, and aI , (ii) the equilibrium solution monomer mole fraction X I , and (iii) the equilibrium solution monomer composition aI . Note that conditions ii and iii are equivalent to knowing X I A = a l X l and X I B = (1 - a l ) X I . Note also that X I and a, (or equivalently X I , and XIB) can be found by using eq 12 in the two constraints imposed by the conservation of the total number of surfactant A and surfactant B molecules in solution, that is, NA = CnanaNna and N B = Enan( 1 - a)Nna, or equivalently

XA = asolJ = a1Xl + End,,, (14a) n.a

X , = (1 - aSoln)X = (1 - a l ) X 1 + En( 1 - a)Xna (14b)

Given gmiC (or equivalently g,,,), and on inserting eq 12 into eqs 14a and 14b, one obtains two implicit equations for X I and a1 as a function of X and awl,,. Solving these two equations si- multaneously one can, in principle, obtain X,(X,a,,,,T,P) and al(X,asoln,T,P), which can then be inserted back into eq 12 to calculate the entire micellar size and composition distribution w,) as a function of X , asoh, T, P, and other solution conditions. Note that, in general, the free energy of micellization gmiC(n,a) will depend on the type and molecular structure of the two surfactant species present in the mixture, as well as on solution conditions such as temperature, pH, and ionic strength. A detailed molecular model to evaluate gmiC has been developed and will be presented elsewhere." However, as explained earlier, to illustrate some general features and prediction capabilities of the thermodynamic framework, in the present paper we develop a simplified analytical phenomenological model for gmiC (see section IID).

Having derived an expression for the micellar size and com- position distribution {Xnal, we describe next how the various moments of the distribution are related to one another. We define the moment of order k by Mk = EnankXna. Note that the first moment is given by MI = X , and the zeroth moment, Mo = E,,,&, is proportional to the total number of mixed micelles and free monomers in the solution. Note also that the zeroth moment plays an important role in determining the value of colligative properties such as the osmotic pressure (for example, see eq 22). In general, using eq 12, the kth moment can be expressed as

n,a

Mk = Ce-lnk(X,)ne-nSg,(a,ai) (15) na

The total derivative of Mk with respect to X I is given by

- = - - d M k M k + l Znk+lXna[ -1- f f - - 1 1 (16) HI XI na 1 -a1 X I

Since, as stated above, we assume that, for values of a # a*, X , is negligible, it can be shown (see Appendix B) that to a very good approximation

By applying the chain rule to eq 17 we obtain

(18) dMk Mk+l dMj Mj+l

z - -

and by setting k = 1 in eq 17 we obtain

Equations 18 and 19 show that with our choice of Gibbs free energy, eqs 1-3, all the moments of the distribution can be ex-

pressed solely in terms of the second moment M2 of the distri- bution. Note that M2 can be related to the weight-average ag- gregation number ( t ~ ) ~ of the mixed micelles through

( n ) w(Xyasoln, T,P) = M2(X,asoln, T,P) /X (20) In addition, the relative variance of the distribution is given by the following expression

It is noteworthy that both the weight-average mixed micelle ag- gregation number and the relative variance can be measured experimentally and through a comparison with eqs 20 and 21 can therefore serve as useful indicators of the applicability of our thermodynamic framework to specific solutions of surfactant mixtures. Note also that eqs 18-21, describing the various mo- ments in the mixed micellar case, are very similar to those de- veloped for single-surfactant solution^.'^

C. Phase Behavior and Phase Separation. Having derived expressions for the micellar size and composition distribution and its moments, the thermodynamic framework is used next to de- scribe the phase behavior and phase equilibria of mixed micellar solutions. In particular, the osmotic pressure T can be related1* to the water chemical potential by T = (pow - pw)/flw, where as stated earlier nW is the effective volume of a water molecule. Using eq 5 we obtain -@Tilw =

In (1 - X ) + X - MO(X,asoln,T,P) + Cedasod(@/2~ea) (22)

The osmotic compressibility of the solution (a?r/dx);$,,. can be obtained by differentiating eq 22 with respect to X , at constant T, P, and aSoln. This yields

(23) It is interesting to point out that the mathematical structures of eqs 22 and 23 for the osmotic pressure and the osmotic com- pressibility are identical to those obtained earlier for single-sur- factant solutions.14

When solution conditions such as temperature, pressure, or ionic strength are altered in a ternary solution consisting of two sur- factant species and water, stability criteria could be violated, and the solution may separate into two or more phases in thermody- namic equilibrium with each other.ls In this paper, we restrict our discussions to two-phase equilibria.

The boundary between the stable and unstable regions of a ternary solution is known as the spinodal surface, while that between the two-phase and one-phase regions is known as the coexistence surface. Note that a binary system at fixed pressure will have a single temperature versus concentration spinodal line and coexistence curve. However, a ternary system at fmed pressure will have a family of spinodal lines constituting a spinodal surface and a family of coexistence curves contstituting a coexistence surface in the three-dimensional temperature, total concentration, composition coordinate system.

In general, the spinodal surface in a ternary system is describedlg by the following stability condition

where G is the Gibbs free energy of the system. On the spinodal surface, the locus of points where the two coexisting phases become indistinguishable is the line of critical points. Note that, at a fixed pressure, there is a single critical point in a binary solution, whereas in a ternary solution there is a line of critical points. Note also that the critical line is the line of tangency between the spinodal surface and the coexistence surface. Thus, at the critical line,

Aqueous Solutions of Surfactant Mixtures The Journal of Physical Chemistry, Vol. 96, No. 13, 1992 5571

in addition to $ = 0, a second stability condition

needs to be satisfied.Ig A simultaneous solution of the two stability conditions, + = 0 and C#I = 0, yields the entire critical line. As described earlier, the coexistence surface bounds the unstable

two-phase region within which a solution spontaneously separates into two isotropic phases. The lines connecting the two coexisting phases in the phase diagram are called tie lines, and the critical line corresponds to the locus of points at which the length of the tie lines vanishes. The conditions of phase equilibria'* require that the temperature, the pressure, and the chemical potentials of each of the components in the solution be the same in the two coexisting phases. Specifically, for the ternary water, surfactant A and surfactant B system, p,( T,P,Y,ay) = pw(T,P,Z,az), pA- (T,P,Y,ay) PA(T,P,ZPZ), and P B ( T , P , Y , ~ Y ) = PB(T,P,Z,~Z), where Y and Z are the total surfactant mole fractions in each coexisting phase, respectively, and ay and az are the corresponding surfactant compositions. The three chemical potentials h, p A , and pe can be expressed as derivatives of the Gibbs free energy per particle, g = G / ( N , + NA + NB), and the resulting expressions are given byI9

ag p w = g - x - ax

Using eqs 26a-26c, the three conditions of phase equilibria can be reexpressed in the following mathematically useful forms

1 ag( T,P, y,ay) 1 ad T , P z , ~ , ) Y aay z aaz (27a) - I-

Using the Gibbs phase rule,'* a two-phase ternary solution in thermodynamic equilibrium has three independent intensive variables. Therefore, fixing temperature, pressure, and one ad- ditional intensive variable the remaining intensive variables are uniquely specified. Accordingly, by fixing one of the four intensive variables Y, 2, ay, and az in eqs 27a-27c, the remaining three can be calculated. In other words, at fixed pressure, the entire family of coexistence curves in the T-X-amIn coordinate system can be generated.

D. Modeling the Free Energy of Micellization. ( a ) General Considerations. As described earlier, the free energy of micel- lization g,,&h,n,a) summarizes the many physicochemical factors responsible for mixed micelle formation and depends on the molecular structure of the surfactants, as well as on solution conditions such as temperature, pH, and the presence of any additives. To illustrate some general features and qualitative prediction capabilities of the thermodynamic framework, below we present a simplified analytical phenomenological model for g&(sh,n,a). Although the predictions based on such a model will be qualitative in nature, major advantages of utilizing a simplified model for g& are that it enables us to obtain analytical expressions for many of the useful micellar solution properties as well as to shed light on the physical basis of some of the observed experi- mental trends. For more detailed and accurate quantitative predictions of micellar solution properties, a molecular model of mixed micellization, which takes into account the detailed mo- lecular structure of the surfactants as well as the effect of solution

conditions, is required. Such a model has recently been developed" and has been successfully utilized to make quantitative predictions of a broad spectrum of micellar and phase behavior properties of aqueous solutions containing mixtures of nonionic-nonionic surfactants."J2

The free energy of micellization is a function of the shape of the micelle sh, the aggregation number n, and the composition a. It is well-known that n can range from very small to very large values. Since gmic(sh,n,a) needs to be evaluated for all n and a, the problem becomes computationally intensive. For the purpose of illustration, we can simplify the calculations by making the following two physically reasonable assumptions. First, as ex- plained in section IIB, we approximate E&, by X,*, where a* is the optimum composition of a micelle having aggregation number n. Second, we evaluate gmic only for the three regular shapes of spheres, infinitesized cylinders, and infinite-sized disks or bilayers. For the nonregular finite-sized mixed micelles, gmic(n,a*) and a*(n) are estimated by linearly interpolating be- tween the g,,&,a*) and a*(n) values corresponding to the limiting regular shapes.

(b) Illustrative Example: Finite-Sized Cylindrical Mixed Micelles. To describe a finite-sized cylindrical mixed micelle exhibiting one-dimensional growth, gmic(n,a*) and a*(n) are estimated by interpolating between the optimum values of these quantities for a spherical micelle and an infinite-sized cylindrical micelle, that is

gmic(nra*) = &%(a*cyl) + (nsph/n) [&%a*sph) - &y~c(a*cyl)l (28)

(29) where hpb and a*,ph are the aggregation number and composition of the optimum spherical micele, is the composition of the optimum infinite-sized cylindrical micelle, and g"ik and &;iC are the free energies of micellization of the optimum spherical and the optimum infinite-sized cylindrical micelles, respectively. Substituting eqs 28 and 29 in the expression for X , given in eq 12, we obtain

and a*(n) = a*cyl + (nsph/n)[a*sph - a*cyll

where K = e+, with Ap = nsph[g"$h - g",] + kT, and X,, = &Em*.

Recall that /3gm = /3gmic - 1 - a In aI - (1 - a) In (1 - a,). Note that the parameter Ap is a growth parameter analogous to the one introducedZ0J4 to describe the one-dimensional growth of single-surfactant cylindrical micelles. Clearly, in order to have growth, one expects A p > 0, where an increase in the value of Ap will result in an increase in micellar size. Note also that the concentration Xcll reflects the propensity to form cylindrical micelles and that, 111 the limit of considerable growth, one has Xcyl = cmc.

It is possible to show, following an analysis similar to the one utilized in the case of single-surfactant cylindrical m i c e l l e ~ , ~ ~ J ~ that, in the limit ( n ) , >> naph, the following remarkably simple result for X,,,, is obtained

x,,,. = (1 /K)e-n("-"' (31) Equation 30 shows that, in the limit of significant micellar growth, the micellar size distribution, corresponding to the optimum composition a*, is a monotonically decreasing exponential function of n (for n > h p h ) whose width is directly proportional to ("I2. As stated earlier, if we approximate Ea,,, by X,,,., the second moment M2 of the distribution (see eq 15 with k = 2) is given by

(32) Similarly, the weight-average mixed micelle aggregation number (see eq 20) is given by

(33)

M2 = nsph + 2K'/2312

( n ) , = M2/X e nsph + 2(KX)'l2

5572 The Journal of Physical Chemistry, Vol. 96, No. 13, I992 Puwada and Blankschtein

Note that the analysis of Ben-Shaul et al.' suggests that when the entire size and composition distribution w,] is utilized, instead of only that corresponding to the optimum composition wm*}, (n), varies as P4 instead of 2,5 as predicted by eq 33. This, of course, does not affect the qualitative conclusions presented in this paper.

In addition, the relative variance of the distribution (see eq 21) is given by

u '/2 (34) (c) Simplified Model for gMc. Below we introduce a simplified

model to estimate the free energy of mixed micellization gdc(sh,a) corresponding to the three regular shapes of spheres, infinite-sized cylinders, and infinite-sized disks or bilayers. As illustrated in (b) above, for nonregular finite-sized micelles, gmiC is estimated by linearly interpolating between the optimum gmiC values cor- responding to the limiting regular shapes.

We model gmiC as the sum of four primary contributions:l03" (i) Hydrophobic free energy gw,fic(sh,a), which represents the

free energy gain in transferring the hydrophobic tails from water to the hydrophobic interior of a mixed micelle characterized by a shape sh and composition a. This contribution can be further expressed as gwlfic(sh,a) = gw/hc(a) + ghc/fic(sh,a). The quantity gw/hc reflects the free energy gain associated with transferring the hydrophobic tails from water to a bulk phase, having composition a, made from the tails of the two surfactant species (note that, as expected, this contribution is independent of micellar shape), and can be expressed as

g w / h c = a d j h c + (1 - a)g /hc + k T [ a In a + (1 - a ) In (1 - a)] + a(l - a)&!

where &/hc and &/hc are the hydrophobic free energy contribu- tions associated with pure surfactant A tails and B tails, re- spectively, kT[a In a + (1 - a) In (1 - a)] reflects the free energy of mixing the two tails of the two surfactant species in the mixed micelle, and &! reflects the strength of the binary interactions between surfactant A and B tails based on the regular solution theory. Note that &/hc and &/hc can be estimated using available solubility data of pure hydrocarbons (or fluorocarbons) in water.10," Note also that d! is typically equal to zero for a mixture of hydrocarbon (or fluorocarbon)-based surfactants1 but is greater than zero3 for mixtures of hydrocarbon- and fluoro- carbon-based surfactants which exhibit repulsive antagonistic interactions. The quantity ghc/mic(sh,a), which reflects the free energy loss associated with the reduction in conformational degrees of freedom of the two types of surfactant tails (at composition a) in the constrained environment provided by the micellar core, depends on micellar shape. For computational simplicity, we assume that this contribution is linear in composition, that is, gkImic

formational free energy contributions associated with pure sur- factant A tails and B tails, respectively. Note that &c/mic and dC m,c can be computed numerically using a single-chain mean- fierb 'approach.IoJ1

(ii) Interfacial free energy g,,(sh,a), which represents the free energy per monomer associated with creating an interface sepa- rating the micellar core from the bulk solution. This contribution can be approximated as g, = a d + (1 - a ) z , where d = UAUA

and = gag. Note that the interfacial tensions uA and uB associated with the tails of surfactants A and B, respectively, are available experimentally. Furthermore, for a given micellar shape, the areas per monomer uA and aB can be determinedlo." from the known chemical structures of surfactant A and B tails, respectively.

(iii) Steric free energy g,,(sh,a), which represents steric in- teractions between the surfactant headgroups at the micellar surface. These interactions have been shown1° to play an important role in the micellization of nonionic surfactants. By treating the headgroups at the micellar surface as an adsorbed localized monolayer, it can be shown" that g,, can be approximated by a d + (1 - a)& where = -kT In (1 - q,A/a) and & = -kT 1n (1 - ahB/a) are the steric contributions associated with the headgroups of pure surfactants A and B, respectively. Note that the average headgroup cross-sectional areas ahA and a h B can be

= a&c/mic + (1 - a)dc/mic, where d c / m i c and d c / m i c are the con-

estimated from the known chemical structures of pure surfactant A and B headgroups, respectively, and for a given micellar shape, the available area per monomer, a, can be evaluated from the known chemical structures of pure surfactant A and B tails.

(iv) Electrostatic free energy gel,(sh,a), required to describe charged surfactants, which represents the free energy associated with creating a charged interface in a sea of counterions. From simple electrostatic considerations2'J3 one expects that, to leading order, gel, should be proportional to q2, where q is the net charge per monomer at the micellar surface. Consequently, for a binary surfactant mixture, q = aezA + (1 - a)ezB, where zA and zB are the valencies of the two surfactants, and e is the electronic charge. Thus, gel, * a& + (1 - + 4 1 - a)dL where C, = K e l d b 2 and d, = Ke,dB2 are the electrostatic contributions associated with pure surfactants A and B, respectively, and A,", = -Kels(zA - z ~ ) ~ reflects binary electrostatic interactions between surfactants A and B. Note that Kk is a numerical constant which can be evaluated from electrostatic t h e o r i e ~ ' ~ . ~ ~ and depends on micellar shape. In addition, Kdoc is expected to be a strong function of salt concentration and should decrease (screening effect) with increasing salt concentration. Note that although in general K,,, may also be a function of surfactant type and composition, for illustrative purposes, in this paper we assume that K,,, is a con- stant.

Combining the four contributions described above, gmiC can be expressed using the following simple relation

gmic(sh,a) = agA,ic(sh) + (1 - a)&ic(sh) + a(1 - a)& + k T [ a In a+ (1 - a) In (1 - a ) ] (35)

energy of micellization of pure surfactant A, Zic(sh) = (&/hc + &c/mic + 8 + & + #,=) is the free energy of micellization of pure surfactant B, and &,(sh) = &! + reflects contributions due to specific intramicellar interactions. Note that, as emphasized earlier, $mic, g",,, and $,E can be computed using readily available experimental information about surfactants A and B as well as some numerical calculations.1"'2 A detailed rationalization of typical values that these contributions can attain for various binary surfactant mixtures is presented in section IIIA(a). The modified free energy of micellization g, at the optimum composition a*, namely, g,(sh,a*,al), can be evaluated by inserting eq 35 into eq 13 and using the resulting relation in the definition of g,. Carrying out this procedure, we find that g,(sh,a*,ai) = I \

where $,ic(sh) = (d/hc + d c / m i c + d + + d w ) is the free

&i,(sh) + $m:(sh)a*' + kT In ( - ; 1 ;;) - kT (36)

where a* can be obtained by solving eq 13 with gmiC given in eq 35.

Since the growth parameter Ak = nSph(g",ph - g$) + kT, then upon using eq 36 for an optimum spherical micelle and an op- timum infinite-sized cylindrical micelle in this definition we find that

Ab' = nsph (&ic)sph - (&ic)cyl + (&&)sph("*sph)' - I 111. Results and Discussions

A. Estimation of Molecular Contributions. ( a ) Zntramicellar Conrriburions gA. , gi ic , and git . As explained in section IID(c), dit, g",,, and 2: determine the value of g,, (see eq 35). Since the model for gmiC utilized in this paper is an approximate one, we can only predict qualitative trends of micellar solution prop- erties. For this urpose, below we discuss the range of typical values that $,,, AiC, and &E can attain for various types of binary surfactant mixtures.

In general, micellization requires that the free energy of mi- cellization be negative. In particular, it follows that both $,,,(sh)

Aqueous Solutions of Surfactant Mixtures The Journal of Physical Chemistry, Vol. 96, No. 13, I992 5513

TABLE I: Typical Values of the Intramicellar Contribution $2 and the Specific Interaction Parameter CAB

TABLE 11: Typical Values of the Interaction Parameter CAW and dCAw/dT for Various Types of Surfactants

surfactant mixture k ! ? l k T C n d k T surfactant type Cnwlk (K) d(CawIk)IaT - . . .-. ._ .. . . . . . . . . . . . ,

nonionic-nonionic -0 -0 monovalent anionic-monovalent anionic -0 50 nonionic-monovalent ionic -2 to-6 <O monovalent anionic-monovalent cationic -10 to -25 <<0 hydrocarbon-fluorocarbon (nonionics) >O >O

and &ic(sh) should be negative for single-surfactant micelles of that particular shape to form. In fact, as shown below, kiC and diC are related to the cmc's of pure surfactants A and B by In cmcA = odic - 1 and In cmcB = pZiC - 1 , respectively. Using typical cmc values for nonionic, ionic, and zwitterionic surfac-

one finds that the values of diC and die range between -6kT and -20kT.

= &: + &: is a strong function of the type of surfactant mixture. In mixtures of hydrocarbon-based (or fluorocabon-based) surfactants the hydrocarbon (or fluoro- carbon) tails mix ideally, and consequently &: = 0, which results in dt = &E = -KeiS(zA - z ~ ) ~ . Therefore, for nonionienonionic surfactant mixtures (zA = zB = 0), as well as for ionic-ionic surfactant mixtures having equal charges (zA = re), d: = 0. However, in nonionic-ionic surfactant mixtures where the ionic surfactant (B) is monovalent, that is, zA = 0 and zB = f l , &: = -KCIN. Simple electrostatic calculations, with no added salt, yield a value of KCIN in the range of 3-6kT,13*21 and the addition of salt further decreases the value of KCIN. Thus, for these non- ionic-ionic mixtures, has a value of approximately -3kT to -6kT at zero salt concentration, which decreases upon the addition of salt due to electrostatic screening effects. For monovalent anioniemonovalent cationic surfactant mixtures (zA = -1, zB = +l) , = -4KelS, which is 4 times the value for nonionic-mo- novalent ionic mixtures. For mixtures of hydrocarbon-based and fluorocarbon-based nonionic surfactants (4: = 0), mixing of the surfactant tails is nonideal due to repulsive interactions, and, therefore, &: > 0, suggesting that $,! > 0. Thus, depending on the nature of the surfactant mixture, g& can range from negative values of up to -25kT to positive values. Table I summarizes typical values of $,: for various types of binary surfactant mix- tures.

( b ) Intermicellar Molecular Contribution C,, As explained in section IIA, the mean-field intermicellar interaction parameter C,, depends on the three intermicellar interaction parameters CAW, CBW, and CAB (see eq 4 and Appendix A). The two interaction parameters, CAW and CBw, reflect interactions in aqueous solutions of surfactant A and surfactant B, respectively (see eqs A6 and A7 in Appendix A) and consequently can be obtained from ex- periments conducted in the corresponding single-surfactant so- lutions. Positive values of CAW (or CBw) indicate net attractive interactions between the surfactant molecules, while negative values indicate net repulsive interactions. Note that net attractive interactions can induce phase separation, while net repulsive in- teractions oppose phase separation.I* More specifically, note that, to leading order in surfactant concentration, YACAW/k and TBCBw/k are equal to the critical temperatures TcA and TcB of the corresponding single-surfactant so1~tions.l~ In particular, many nonionic surfactants in aqueous solution exhibit23 phase separation between 0 and 100 OC and thus have YACAW/k (or -yBCBw/k) values of approximately 300 K. More generally, any surfactant in aqueous solution which exhibits phase separation between 0 and 100 O C , for example, a zwitterionic surfactant such as C8- lecithin,I4 will have YACAW/k (or yBCBw/k) values of approxi- mately 300 K. On the other hand, aqueous solutions of ionic surfactants in the absence of salt typically do not exhibit phase separation because of the long-range electrostatic intermicellar repulsions. Therefore, we expect ionic surfactants to have large negative CAw (or CBw) values. In Table 11, we list typical values of CAW (or CBW) for nonionic, zwitterionic, and ionic surfactants.

As described in Appendix A, the interaction parameter CAB describes specific interactions that may exist between surfactants A and B and, therefore, constitutes a measure of the synergistic

The contribution

nonionic +0(10) +0(1) zwitterionic +0(10) =O ionic <<O =O

(attractive, for CAB < 0) or antagonistic (repulsive, for CAB > 0) interactions between the two surfactant molecules. Note that mixtures of nonionic surfactants usually form ideal mixtures and, therefore, are expected to have CAB = 0. On the other hand, mixtures of dissimilar surfactants (for example, ionic-nonionic or anionic-cationic surfactants) usually form nonideal mixtures with expected nonzero values of CAB. From the discussion in Appendix A, we expect CAB to become increasingly negative (increased synergism) in the sequence nonionienonionic, anion- ic-anionic, nonionic-ionic, and anionic-cationic surfactant mix- tures. Mixtures of hydrocarbon-based and fluorocarbon-based surfactants are expected to have CAB values greater than zero because the two surfactants exhibit repulsive (antagonistic) in- te rac t ion~.~ In Table I, we list typical values of CAB for various surfactant mixtures.

To summarize section IIIA, we have presented typical values of (1) the three intramicellar molecular contributions kit, dit, and Zit, which determine the magnitude of gdc, and (2) the three intermicellar interaction parameters CAW, CBw, and CAB, which determine the magnitude of Ceff (see Tables I and 11). Using typical values of gmiC and Cen in the next sections we predict qualitative trends of (i) the mixture cmc, (ii) characteristics of the micellar size and composition distribution, and (iii) phase separation characteristics for various types of binary surfactant mixtures.

B. Critical Micellar Concentration (cmc). At very low sur- factant concentrations most of the surfactant molecules exist as free monomers. However, as the total concentration of surfactant is increased, keeping its composition constant, micelles begin to form beyond a certain threshold concentration known as the critical micellar concentration (cmc). Beyond the cmc most of the added surfactant remains in the micellar form, and the total monomer concentration remains practically constant. Furthermore, the first micelles that form will have a composition close to the optimum value a*, because at a = a* the free energy of micellization exhibits a minimum. The mole fraction of these micelles can therefore be expressed, using a = a* in eq 12, as

From eq 38, it follows that X,. is vanishingly small for XI << t?gm

and infinitely large for X1 > &gm. However, for XI = &gm, Xnu, is finite but small. That is, to a very good approximation, the micelles first form when the monomer concentration XI = &gm.

In other words, one can approximate the cmc of the mixed micellar solution by &gm. Using the expression for g, given in eq 36, one finds that In (cmc) = Pg,(sh,a*,al) = , \

where sh corresponds to the shape of the optimum micelle, and a* is obtained by inserting eq 35 for gdc into eq 13. The resulting implicit equation for CY* is given by

(40) a* a1 P(g",, - &,) + &&E(l - 2a*) + In - = In -

1 -a* 1 - a l

Combining eqs 39 and 40 and performing some algebraic ma- nipulations, we obtain

5514 The Journal of Physical Chemistry, Vol. 96, No. 13,

2E-05

1992 Puwada and Blankschtein

* U

Figure 1. Predicted optimum composition of the mixed micelle a*, at the cmc, as a function of monomer composition al for solutions of (i) nonionic-nonionic (-), (ii) nonionic-monovalent ionic (---), (iii) mo- novalent anionic-monovalent cationic (- - -), and (iv) nonionic hydro- carbon-nonionic fluorocarbon (- - -) surfactant mixtures.

where In fA = p ( 1 - a*)', In fB = @$,:(a*)2, In (cmc,) = @ m,c - 1, and In (cmce) = a&,.-, 1. The variables fA and fB are equivalent to the micellar activity coefficients of each surfactant, and the parameter &$,E is equal to the empirical interaction parameter t used in the pseudophase separation model5 for mixed micelles. Our approach, therefore, enables us to rationalize the physical basis behind eq 41, an expression which has been utilized extensively to analyze and predict cmc's of mixed surfactant solutions. Furthermore, our approach also allows us to uantitatively predict mixture cmc's from a knowledge of $mic, Aic, and A:, which can be computed quite accurately utilizing a detailed molecular model"J2 or es- timated using the simplified model presented in this paper.13 Thus, the mixture cmc can be predicted as a function of surfactant type, surfactant composition, and solution conditions. As stated earlier, in this paper, we will only make general qualitative observations about the cmc behavior based on the simplified model for gmiC presented in the previous section. A detailed quantitative analysis of cmc's of aqueous nonionic-nonionic surfactant mixtures' I and of aqueous nonionic-anionic, nonionic-cationic, and anionic- cationic surfactant mixturesI3 will be presented elsewhere.

We begin by noting that since t = /3$,;, in view of the dis- cussions in section IIIA(a), we would predict that monovalent anionic-monovalent cationic mixtures have a value of t which is 4 times that of nonionic-monovalent ionic mixtures. Indeed, this is observed experimentally, where e values range between -2.5 and -5 for nonionic-monovalent ionic mixtures and between -10 and -25 for monovalent anionic-monovalent cationic mixtures.24

Our approach also indicates that, at the cmc, the mixed micelles have compositions which are significantly different from the so- lution composition. Indeed, Figure l shows the predicted optimum micellar composition a*, at the mixture cmc, as a function of the solution monomer composition aI. The predicted mixture cmc is also plotted as a function of ayl in Figure 2. These calculations were performed for $. - -10kT and &,, = -12kT and for the following four typical Zkialues: (i) &E = 0 kT, (ii) = -2kT, (iii) &: = -15kT, and (iv) &: = 2kT. Based on the values of g"m: (see Table I ) , cases i-iv correspond to mixtures of nonion- ic-nonionic, nonionic-monovalent ionic, monovalent anionic- monovalent cationic, and nonionic hydrocarbon-nonionic fluo- rocarbon surfactants, respectively. Since, as shown earlier, In (cmcA) = @di, - 1 and In (cmc,) = @gemic - 1, this selection implies that in all four cases pure surfactant B has a lower cmc than pure surfactant A (see Figure 2, where al = 0 ( a l = 1 ) corresponds to pure surfactant B (surfactant A)). Below we analyze cases i-iv separately.

I E - 0 5 c ,' /

5E -06 c I E - 0 6

5E-07

I E-07 1 I I I I 0 0.2 0.4 0.6 0.8 I

Q I

Figure 2. Predicted critical micellar concentration (cmc) as a function of monomer composition aI for various surfactant mixtures. The notation is the same as in Figure 1. Note that for illustrative purposes the cmc's of the pure surfactants (a, = 0 and a, = 1) are the same in all four cases.

Case i, corresponding to an aqueous solution of a nonionic- nonionic surfactant mixture (full lines in Figures 1 and 2) rep- resents an ideal mixture at the micellar level (&: = 0). For this system, Figure 1 shows that the solution monomer composition aI is always higher than the optimum micellar composition a*, indicating that the mixed micelles are enriched with surfactant B having the lower cmc. Case ii, corresponding to a nonionic- monovalent ionic surfactant mixture (dashed lines in Figures 1 and 2) represents a nonideal mixture at the micellar level (weak synergism, &E = -2kT). For this system, Figure 2 shows that the mixture cmc is lower than in case i, indicating negative de- viations from ideality due to synergistic interactions. Case iii, corresponding to a monovalent anionic-monovalent cationic surfactant mixture (dashed-dotted lines in Figures 1 and 2) , represents a highly nonideal mixture at the micellar level (strong synergism, &: = -15kT). For this system, Figure 2 shows that the mixture cmc is significantly lower than in case i, indicating large negative deviations from ideality. Furthermore, it is note- worthy that the mixture cmc is considerably lower than the cmc's of the pure surfactants over a broad aI range (see Figure 2). Over this al range, Figure 1 shows that the optimum micellar com- position a* is close to 0.5, a composition at which the two opposite surfactant charges would completely neutralize each other. In other words, this suggests that the distribution of the two oppositely charged surfactants between monomers and micelles reflects the tendency to minimize electrostatic repulsions within the micelles, thus leading to strong synergism and associated low cmc values (as compared to the ideal case). It has recently been brought to our attention that, under such strong synergistic conditions, many surfactants self-assemble to form vesicles rather than micelles.25 However, our theory is applicable only to solutions containing micelles and would therefore be strictly valid for those surfactant mixtures which form micelles rather than vesicles. Case iv, corresponding to nonionic hydrocarbon-nonionic fluorocarbon surfactant mixtures (dotted lines in Figures 1 and 2), represents a nonideal mixture at the micellar level (antagonism, $,: = 2k7"). For this system, Figure 2 shows that the mixture cmc is larger than in case i, indicating positive deviations from ideality due to antagonistic interactions. As noted in Figure 1, it is noteworthy that over a broad aI range, one has a* << al , indicating that the mixed micelles are highly enriched in surfactant B (more than 90%) reflecting the antagonistic interactions between surfactants A and B.

Aqueous Solutions of Surfactant Mixtures

4,000 2 5

\

-I..

aso ln

Figure 3. Predicted weight-average mixed micelle aggregation number ( n ) , (at X = and growth parameter AN as a function of total surfactant composition a,ln for solutions of (i) nonionic-nonionic (-) and (ii) nonionic-monovalent ionic (---) surfactant mixtures. Note that in both cases Ap is the same for the pure surfactants (aroln = 0 and amln = 1).

C. Characteristics of the Micellar Size and Composition Dis- tribution. For illustrative purposes we consider the case of cyl- indrical mixed micelles which exhibit significant one-dimensional growth, a case of particular importance for certain aqueous so- lutions of mixed nonionic-nonionic surfactants.” In section IID(b) we saw that the weight-average mixed micelle aggregation number is given by eq 3 3 , which we rewrite here in terms of the growth parameter Ap, namely

( n ) , = nsph + 2(eflA~X)1/z (42)

Equation 42 indicates a square-root dependence of ( n ) , on X as well as an exponential dependence on Ap. Using eq 37 for Ap and eq 42 for ( n ) w , in Figure 3 we predict variations of ( n ) , (at X = 1 0 - j ) and & with solution composition Q for two illustrative cases: (i) ($,E),,, = (&z)cyl = 0 and (ii) ($,E),,, = -1.7kT, (&E),,, = -2kT. Based on these values (see Table I), cases i and ii would correspond to nonionic-nonionic and nonionic-monovalent ionic surfactant mixtures, respectively. Note that the parameter valucs in (i) and (ii) were chosen to illustrate the effect of different types of interactions (electrostatic vs steric) on the mixed micellar size. In addition, for both cases, the free energy values (&ic)cyl

- 1 1.6kT were selected. Using these values and assuming an aggregation number of 50 for the spherical micelle (typicallo values of &ph range between 30 and loo) , the growth parameters for pure surfactants A and B can be computed using Ap = &ph(g‘,ph - g”,y) + kT and are found to be equal to 6kT and 21 kT, respectively. That is, in cases i and ii, pure surfactant B (aJoln = 0) exhibits significant growth (Ap = 21 kT), while pure surfactant A (a,,,, = 1) exhibits moderate growth (4. = 6kT); see Figure 3 . In both cases, Ap decreases with increasing amln from 21kT to 6kT. However, for small values of aWln, Ap decreases at a much slower rate in case ii (dashed line) than in case i (full line). On the other hand, for larger awl,, values, Ap decreases more rapidly in case ii, reaching a value of 6kT at a,h = 1 . This decrease in Ap with aWln is also reflected in the decrease of the weight-average mixed micelle aggregation number with awl,,, with ( n ) , decreasing more rapidly in case i (full line) than in case ii (dashed line). Figure 3 clearly shows that ( n ) , is larger in case ii than in case i, in- dicating synergistic interactions in the nonionic-monovalent ionic surfactant mixture. This prediction is consistent with experimental

-1OkTs (&ic)sph = -9-9kT9 (&ic)cyi = -12kT, and ($,ic)sph =

The Journal of Physical Chemistry, Vol. 96, No. 13, 1992 5575

measurementsz6 of micellar sizes in the system ClZE5-SDS-DzO, where CI2E5 is a nonionic surfactant and SDS is a monovalent anionic surfactant. Specifically, it was found that aqueous so- lutions of pure ClzE5 contain large micelles and that the addition of small quantities of SDS (up to approximately 20 wt %) further increases micellar size (due to the synergistic interactions between the ionic and nonionic surfactants). However, as the amount of SDS is further increased the micellar size decreases.

D. Phase Separation. In this section, we study the influence of adding small amounts of surfactant A (asoln << 1) on several aspects of the phase separation of an aqueous solution containing surfactant B. Specifically, we study the influence on the critical temperature, as well as analyze how the added surfactant partitions between the two coexisting micellar-rich and micellar-poor phases.

At the binary critical point of a single-surfactant solution, (dzG/aNWz) = (d3G/aNw3) = 0, suggesting that (W/dN,) = OI9 (see eqs 24 and 25). Consequently, very close to the binary (surfactant B + water) critical point, the condition 0 = 0 in eq 25 can be expressed as

dZC/dN,* = 0 (43) Substituting the value of the Gibbs free energy, given in eqs 1-3, into eq 43 we find, to leading order in X , that the critical tem- perature Tc can be expressed as

Inserting eq 4 into eq 44, and expanding the resulting expression to linear order in asoln ( < < 1 ) yields kTc a YBCBW +

Tc =s YcnCeff/k (44)

[(YA - 7B)CBW + YBICAW - CBW] - f i C A B l a s o l n ( 4 5 )

Note that the critical temperature of an aqueous solution of pure surfactant B (corresponding to ah = 0 in eq 45) can be expressed as T 2 = yBCBw(T2)/k, where we have assumed that CBw can depend on temperature. To estimate the change in Tc of an aqueous solution of surfactant B upon the addition of small quantities of surfactant A (amln << l ) , namely, ATc = Tc -. T 2 , we expand eq 45 to leading order in AT, and aSoln. This yields kATc -

[(YA/YB - 1)cBW + (CAW - CBW - ~ ~ c A B ) l ~ B ~ m l n

(46) If solution conditions are altered such that the solution enters

into the unstable region of the phase diagram, the solution will spontaneously separate into two coexisting phases, one micellar poor and the other micellar rich. Substituting the free energy contributions given in eqs 1-3 into g = G/(N, + NA + NB), we obtain g = ( l - X)pow + X[asolnboA + ( 1 - asoln)poBI +

kT[ (1 - X ) In ( 1 - X) + X - Mo + X In XI] +

To derive an expression for the composition difference az - ay, when small amounts of surfactant A are added to aqueous solutions of pure surfactant B, we insert eq 47 in eq 21a. Since the coexisting mole fractions Y and 2 are much smaller than 1 (about

and amln << 1 , we expand the resulting expression in powers of Y, 2, a,,, and az. To linear order in az - ay and Z - Y we obtain

kTX[asoln In + ( 1 - a s o d In (1 - - XCefdX (47)

z - Y “Z - =s ?rBaflB[(YA/YB - 1 ) c B W +

- C B W - dxcAB)I (48) Combining eqs 48 and 46, we find that the difference in com- positions (aZ - a,,) between the two coexisting phases is pro- portional to the change in critical temperature ATc when small quantities of surfactant A are added to aqueous solutions of surfactant B.

5576 The Journal of Physical Chemistry, Vol. 96, No. 13, 1992 Puwada and Blankschtein

Depending on the values of CBW, CAW - CBW, CAB, and YA/YB, appearing in eqs 46 and 48, below we consider the following four illustrative cases.

(i) 0 < C A W I CBW, CAB * 0, and Y A = 7 6 . Since C A B = 0, this case would represent a mixture of nonionic or zwitterionic surfactants, and since CAW and CBw are both greater than zero, it would represent the addition of a nonionic or zwitterionic surfactant to another nonionic or zwitterionic surfactant (see Table 11). In this case, the critical temperature of the mixture (see eq 46) does not change dramatically because the values of CAW and CBW are comparable. If surfactant B is a nonionic surfactant, (yB/k)(dCBw/dT) is typically greater than 1 at the lower critical point (see Table 11), and the addition of surfactant A raises the lower critical temperature (note that kAT, 0: (CBW - CAW)/ [(yB/k)(dcBw/dT) - 11 > 0 in this case). However, at the upper critical point of a nonionic or a zwitterionic surfactant solution, dCBW/dT= 0 (see Table 11), and consequently since CBw 1 CAW, the upper critical temperature will decrease upon the addition of small amounts of surfactant A. In other words, the addition of small quantities of a second surfactant (nonionic or zwitterionic) raises the lower critical temperature of a nonionic surfactant solution and lowers the upper critical temperature of both zwit- terionic and nonionic surfactant solutions. In addition, since Z - Y is t y p i ~ a l l y l ~ , ~ ~ of the order of the difference az - ay is of order loe5. This indicates that the composition of surfactant in the two phases is almost identical and that the added surfactant partitions almost equally between the two phases. Indeed, recent phase separation measurements in aqueous solutions of mixtures of two nonionic surfactants belonging to the alkyl polyethylene oxide family (CiEj) appear consistent with this prediction."

(ii) CAW << 0, CBW = O(lOkT), CAB < 0, and Y~ = 79. Since CAW << 0 and CBw = O(1OkT) this case would represent (see Table I) the addition of an ionic surfactant (species A) to a solution containing a zwitterionic or nonionic surfactant (species B). If surfactant B is zwitterionic (dCBw/aT = 0), then the upper critical temperature decreases more rapidly than in case i with the addition of surfactant A (see eq 46). However, if surfactant B is nonionic, then (yB/k)(dCBw/dT) is greater than 1 at the lower critical point (see Table I), indicating that the lower critical temperature in- creases with the addition of surfactant A. In addition, the upper critical point of a nonionic surfactant solution, where dCBw/dT = 0, shifts to lower temperatures, thus predicting that the entire closed-loop coexistence curve should shrink upon the addition of an ionic surfactant. All these results, for the addition of an ionic surfactant to both nonionic and zwitterionic surfactants in aqueous solutions, are consistent with experimental data.27 Since in case ii the difference CAW - CBw is large and negative, this suggests that - ay is finite and negative. That is, we predict that the added ionic surfactant preferentially partitions into the dilute phase where the electrostatic repulsions are smaller. Note that in the dilute phase the total surfactant concentration is smaller than in the concentrated phase, and consequently the surfactants are, on average, farther apart from one another thus reducing the net electrostatic repulsions.

(iii) C, << 0, CAW = CBw and yA = yB which represents strong specific (synergistic because CAB < 0) interactions between sur- factants A and B, for example, in anionic-cationic surfactant mixtures. In this case, since aCBw/aT z 0, eq 46 indicates that the upper critical point moves to higher temperatures. That is, when there are strong specific attractions between the two sur- factants, the addition of the second surfactant increases net in- termicellar attractions and thus increases the size of the unstable two-phase region. In this case, eq 48 indicates that the added surfactant partitions into the concentrated phase. This enhances electrostatic attractions between the two surfactants, and since, on average, surfactants are closer to each other in the concentrated phase than in the dilute phase, the overall electrostatic repulsion decreases. These interesting predictions remain to be tested ex- perimentally.

(iv) Y~ > YB, CAB = 0 and CAW = CBw > 0, which describes mixing two surfactants having similar interactions and different sizes. In this case, we predict that when a larger surfactant (species

A) is added to a solution of the smaller surfactant (species B), the upper critical temperature (associated with dCBw/aT = 0) increases, while the lower critical temperature (associated with (yB/k)(dCBw/dT) > 1) decreases, thus increasing the size of the unstable two-phase region (see eq 46). Although we are unaware of available experimental data in mixed surfactant solutions to test these predictions, it is noteworthy that a behavior similar to that predicted here has been observed in solutions of polymer mixtures. Specifically, the addition of a high molecular weight polystyrene (M, = 50 X lo4) to a solution of a low molecular weight polystyrene (M, = 4.5 X lo4) in cyclohexane raised the upper critical temperature of the solution. In addition, a com- position analysis of the two coexisting phases indicated that the high molecular weight polystyrene partitions preferentially into the concentrated

IV. Concluding Remarks Mixed micellar solutions are currently a subject of considerable

practical importance because they can exhibit properties which are superior to those of solutions containing the individual con- stituent surfactants. Consequently, detailed information is be- coming available on the critical micellar concentration, synergistic and antagonistic interactions between the surfactants, and micellar solution phase behavior including phase separation. In view of these experimental developments, it becomes increasingly necesary to construct a theoretical approach capable of unifying the rich variety of seemingly unrelated experimental findings into a single coherent computational framework. It has been the purpose of this paper to contribute to this much needed theoretical unification.

Accordingly, in this paper, we have presented a thermodynamic theory of mixed micellar solutions which can be utilized to predict a broad spectrum of micellar solution properties. Our approach represents a generalization of a recently developed thermodynamic theory for single-surfactant solutions, which has been successfully utilized to predict a wide range of properties of aqueous solutions of pure nonionic surfactants and zwitterionic surfactants with and without added solution modifiers such as salts and

In this paper, the theory has been utilized to make qualitative predictions of the critical micellar concentration, the micellar size and composition distribution, and the phase behavior, including phase separation, of solutions containing mixtures of nonionic- nonionic, nonionic-ionic, zwitterionic-ionic, anionimtionic, and hydrocarbon-fluorocarbon based surfactants. These qualitative predictions are consistent with experimentally observed trends. The theory has also clarified the molecular basis of some of the synergistic and antagonistic interactions which were described in earlier works using empirical parameters. For example, the mixture cmc was previously described using an empirical inter- action parameter t and the cmc's of the constituent surfactants. In this work, we have derived a similar expression for the mixture cmc (se eq 41), in the context of our Gib& free energy formulation, and thus have provided a clearer understanding of the molecular basis for t. In addition, we have computed the monomer and micellar compositions, including their dependence on the nature of the surfactants present in the mixture.

The theory has also been used to make some qualitative pre- dictions about the variation of micellar size with surfactant composition. The present work predicts that the average micellar size of nonionic micelles in aqueous solution should decrease upon the addition of an ionic surfactant. Interestingly, we also predict that synergistic interactions between nonionic and ionic surfactants (Ax < 0) result in larger (n), values as compared to those in aqueous solutions of two nonionic surfactants (Ax = 0).

In addition to the cmc, the micellar composition, and the av- erage mixed micellar size, the theory is also able to describe the phase behavior of mixed micellar solutions as a function of the composition of the surfactant mixture. Indeed, it successfully predicts that in aqueous nonionic surfactant solutions the lower consolute (critical) temperature will increase, and the upper consolute (critical) temperature will decrease upon the addition of small quantities of an ionic surfactant. On the other hand, it predicts that the upper consolute (critical) temperature of aqueous

Aqueous Solutions of Surfactant Mixtures

zwitterionic surfactant solutions will decrease with the addition of small quantities of an ionic surfactant. These predictions are consistent with experimental observations.6 In addition, the theory can be used to describe the entire two-phase region, where an isotropic mixed surfactant solution spontaneously separates into two isotropic phases having different surfactant concentrations and compositions, We find that the compositions in the two coexisting phases can be significantly different depending on the nature of the two surfactants. Specifically, in nonionic-ionic surfactant mixtures, the ionic surfactant preferentially partitions to the dilute micellar-poor phase in order to minimize electrostatic repulsions. This interesting prediction remains to be tested ex- perimentally. On the other hand, in nonionic-nonionic surfactant mixtures, the compositions of the two coexisting phases are almost identical.” We hope that the results presented in this paper will stimulate new experiments aimed at probing mixed micellar morphology, as well as the phase behavior and phase separation of mixed micellar solutions.

The availability of a detailed molecular model to evaluate the free energy of micellization of mixed surfactant solutions will further enhance the predictive capabilities of the theoretical framework presented in this paper. In particular, the theory can then be used to make quantitative predictions of useful micellar solution properties as a function of (i) the molecular architectures of the surfactant species present in the mixture and (ii) solution conditions, including the presence of additives such as salts. Progress in this direction will be reported in three following pa- p e r ~ . ” - ’ ~

Acknowledgment. This research was supported in part by the National Science Foundation (NSF) Presidential Young Inves- tigator (PYI) Award to Daniel Blankschtein, and an NSF Grant No. DMR-84- 187 18 administered by the Center for Materials Science and Engineering at MIT. Daniel Blankschtein is grateful for the support of the Texaco-Mangelsdorf Career Development Professorship at MIT. He is also grateful to the following com- panies for providing PYI matching funds: BASF, British Pe- troleum America, Exxon, Kodak, and Unilever.

Appendix A: Evaluation of the Free Energy of Interaction The free energy of interaction reflects interactions between

mixed micelles, water molecules, and free monomers in the so- lution. We adopt a mean-field type approximation to describe these interactions. At the level of a quadratic expansion in the number density, the free energy of interaction takes the following form Gi =

!hNwUww~w + CNwuw(na)Pna + YZ Nnlalu(nla,)(n2a3~n2a2

(AI)

Equation A1 regards the ith particle as interacting with an average local potential Uij p r o d u d by other j th particles. For example, UW(M) represents the average interaction potential between a water molecule (ith particle) and a mixed micelle composed of na surfactant A molecules and n( 1 - a) surfactant B molecules (ith particle). In the spirit of a mean-field type approximation, this potential is assumed to be proportional to the number density of thejth particles p . in the solution. The number densities are given by Nj/Q, where d is the total volume of the solution. For dilute solutions, where mixing volume effects are negligible, the total volume of the solution can be modeled as Q = (NwQw + NAQA + N B Q B ) , where Ow, QA, and QB are the effective volumes of a water, surfactant A, and surfactant B molecule, respectively, and are assumed to be constant. More importantly, we model the interactions contributing to Qj as the sum of painvise interactions between individual monomers in different mixed micelles, thus neglecting three- and higher-order n-body interactions. That is

n,a n l , q n 2 4 2

U(nlal)(n2a2) = ~ I ~ Z I ~ I ~ Z U A A + (1 - a1)(1 - ~ U B B + - + (1 - aI)a21UABj (A2) Similarly

The Journal of Physical Chemistry, Vo1. 96, No. 13, 1992 5511

Uw(na) ~ ~ U W A + n(1 - ~ ) U W B (A31

In eqs A2 and A3, U,, UAB, and UBB denote interaction potentials between two surfactant A molecules, between surfactant A and surfactant B molecules, and between two surfactant B molecules, respectively. Similarly, UwA, UwB, and Uww denote interaction potentials between water and surfactant A, water and surfactant B, and two water molecules, respectively. Using eqs A2 and A3 in eq A1 we obtain

Gi = Q[!!z~wwP~’ + Y Z ~ A A P A ~ + WBBPB’ + UWAP~PA + U W f i W P B + U A f i A p B l (A4)

Eliminating pw = Nw/Q from eq A4, using QJ?, + QANA + n$v~ = 52, we obtain

I \

where, $A = nANA/n and 4B = Q$vB/Q are the volume fractions of surfactants A and B, respectively, YA = QA/nW, Y? = QB/fzw, and ycrf = a,ln~A + (1 - aWln)~B. The mean-field interaction parameters CAW, CBw, and CAB are related to the various inter- action potentials by the following expressions

C A W = Q A [ ----- 2uwA z “ A ] (A6) QWQA QA’

From eq A6, it is evident that CAW reflects interactions between water molecules and surfactant A only. Similarly, eq A7 indicates that CBw reflects interactions between water molecules and surfactant B only. However, specific interactions between sur- factants A and B (captured in U,), which are not present between two surfactant A molecules (U,) or between two surfactant B molecules (UBB), are captured by CAB (see eq A8). Note, however, that, in general, the interaction potential Vi, is expected to be proportional to the product of the number of interaction sites on the two molecules i and j . Thus, if the two molecules interact through electrostatic forces, Uij is proportional to the product of thc number of charged sites on each molecule. Similarly, if the two molecules interact through van der Waals forces, U,j is pro- portional to the product of the volumes of the two molecules.30 Therefore, if surfactants A and B are nonionic and they interact primarily through van der Waals forces, then uAB/QAQB may be approximated by (U,/QAz + UBB/nB2)/2. In that case, eq A8 indicates that CAB = 0, suggesting that nonionic surfactants form ideal mixtures. However, if there are synergistic (attractive) interactions between surfactants A and B, CAB < 0. Similarly, antagonistic (repulsive) interactions between the two surfactants leads to CAB > 0. The magnitude of CAB for various surfactant mixtures can be estimated using similar arguments. In anionic- anionic or cationic-cationic surfactant mixtures, the molecules interact primarily through electrostatic repulsions, and if the two surfactants have the same valency, then UAB = UAA = UBB > 0. In that case, eq A8 indicates that CAB a -UAA(nA - flB)‘! (QAQB)3/2, which results in negative C A B values. Since C A B is proportional to (Q, - ne)’, ionic surfactants having very dissimilar sizes have larger values of CAB than those having similar sizes.

5578

In nonionic-ionic mixtures, electrostatic repulsions between the ionic (species A) surfactants leads to very large and positive values of U A A . However, interactions between ionic and nonionic sur- factants (present in UAB) or between two nonionic surfactants (present in UBB) are much smaller. Thus, nonionic-ionic mixtures will typically have CAB < 0. In monovalent anionic-monovalent cationic mixtures, both U A A and UBB are large and positive due to electrostatic repulsions. However, attractive interactions be- tween the oppositely charged surfactant A and B molecules leads to a large but negative value of UAB whose absolute magnitude is of the same order as that of U , or UBB. Thus, CAB << 0, and its value is approximately 4 times that for nonionic-monovalent ionic surfactant mixtures. For both ionic-nonionic and anionic- cationic mixtures, the presence of salts screens electrostatic in- teractions and hence the absolute magnitude of CAB becomes smaller. In hydrocarbon-fluorocarbon mixtures, repulsions be- tween the two surfactants leads to positive CAB values. Table I summarizes expected typical values of CAB for the various sur- factant mixtures described above.

The free energy of interaction given in eq A1 reflects inter- actions between all the components present in the solution. However, we are actually interested only in those interactions which are not already accounted for at the level of the free energy of formation Gf given in eq 1. In particular, the first three terms in eq A5, which are proportional to Nw, NA, and NE, respectively, reflect interactions between water molecules, surfactant A and water molecules, and surfactant B and water molecules, respec- tively. Since the standard state chosen in our derivation is pure water, these three contributions have already been captured at the level of GI. Accordingly, only the last term in eq A5, which is proportional to (NA + NE)$, reflects the excess interactions which should appear in the free energy of interaction Gi. In other words

The Journal of Physical Chemistry, Vol. 96, No. 13, I99

Gi = -f/zCefdasoln)(NA + ('49)

where

Cedasoln) = CAWffsoln + CBW(l - %oln) - ( 6 / ? e f f ) c A B f f s d n ( 1 - %In)

(A 10) reflects the magnitude of the effective mean-field interaction potential.

Appendix B: Moments of the Micellar Size and Composition Distribution

The kth moment of the micellar size and composition distri- bution is defined by Mk = xn,nkXna. Using eq 12 for X,,, the kth moment can be expressed as

Mk Ce-lnkX,ne-nBgm (B1) na

where X I is the mole fraction of free monomers in solution, and g,,,(cY,a,) is a modified free energy of micellization. Accordingly, Mk IS a function of the micellar composition a and the monomer solution composition al. Using eq B1, the total derivative of Mk with respect to X I is given by

2 Puwada and Blankschtein

- = - - dMk Mk+ I Znk+lXna[ -1- f f - f f 1 1 (B2) XI na I - a1 X I

As discussed in section IIB, X,, exhibits a sharp maximum at a = a*(n), and, therefore, to leading order in a, CJ,,, can be approximated by Xn,.. Thus, eq B2 can be rewritten as

To perform the summation in eq B3 one needs to know how a* depends on the aggregation number n. Below, we consider two important different limiting cases: (i) monodisperse micelles in equilibrium with monomers and (ii) polydisperse micelles which

exhibit significant one-dimensional growth. For micelles which do not exhibit growth, one can describe the

solution as one containing monodisperse micelles, having an ag- gregation number m, in equilibrium with the monomers. In that case, the two material balance equations can be expressed as X = X I + mX,,,,. and aso,,J = a l X 1 + ma*X,,,,.. A simultaneous solution of these equations yields a* = (awls - a , X l ) / ( X - X I ) . Substituting this value of a* in eq B3, we obtain (for the ex- perimentally relevant range X >> X I )

For micelles which exhibit one-dimensional growth into large cylindrical micelles, the optimum composition a* can be estimated by modeling the micelles as spherocylinders and linearly inter- polating the optimum compositions a*sph and ascYl associated with the corresponding sphere and infinite-sized cylinder, respectively. That is

(B5)

where nsph is the aggregation number of the sphere. Substituting eq B5 in eq B3, we obtain

f f*(n) = f f * c y l + (nsph/n)("*sph - f f * c y l )

d M k - - 1 - f f * c y l nsphMk a*sph - ff*cyl

dX, -e[,]-,[ 1 - a , ] (B6)

In surfactant solutions containing micelles having large average aggregation numbers, it has been shown that Mk+l >> Mk.I4 Consequently, to a very good approximation, the second term in eq B6 can be neglected. In addition, using the expression for a*(n) given in eq B5 in the material balance equation for surfactant A, asol,J = Cnna*Xna., we obtain

f f s o l J = fflxl + " * ~ y l ( ~ - + nsph(MO - Xl)(ff*sph - ff*cyl)

(B7) Since Mk << Mk+l, it follows that Mo << X . Therefore for X >> X I , eq B7 indicates that a*cyl = a,,,,, and therefore eq B6 reduces to eq B4.

Thus, in both limiting cases, one where micelles remain small and monodisperse, and the other where micelles exhibit one-di- mensional growth into large polydisperse cylindrical structures, one finds that

References and Notes (1) Scamehorn, J. F. Phenomena in Mixed Surfactant Systems; ACS

Symposium Series 31 I ; ACS Press: Washington, DC, 1986, and references therein.

(2) Rosen, M. J.; Hua, X. Y. J . Am. Oil Chem. SOC. 1982, 59, 584. (3) Mukerjee, P.; Handa, T. J. Phys. Chem. 1981,85,2298. Handa, T.;

Mukerjee, P. J . Phys. Chem. 1981, 85, 3916. (4) Lange, V. H. Kolloid Z . 1953, 96, 131. Clint, J. H. J. Chem. SOC.,

Faraday Trans. 1 1975,71,1327. Shinoda, K. J. Phys. Chem. 1954,58,541. (5) Rubingh, D. N. In Solution Chemistry of Surfactants; Vol. 1; Mittal,

K. L., Ed.; Plenum: New York, 1979; p 337. Holland, P. M.; Rubingh, D. N. J. Phys. Chem. 1983.87, 1984.

( 6 ) Nagarajan, R. Langmuir 1985, I , 331. (7) Ben-Shaul, A.; Rorman, D. H.; Hartland, G. V.; Gelbart, W. M. J .

Phys. Chem. 1986, 90, 5277. (8) Stecker, M. M.; Benedek, G. M. J. Phys. Chem. 1984, 88, 6519. (9) Meroni, A,; Pimpinelli, A.; Reato, L. Chem. Phys. Lett. 1987, 135, 137. (IO) Puwada, S.; Blankschtein, D. J. Chem. Phys. 1990, 92, 3710 and

references therein. Blankschtein, D.; Puwada, S. MRS Symp. Proc. 1990. 177, 129. Puwada, S.; Blankschtein, D. In Proceedings of the 8th Interna- tional Symposium on Surfactants in Solution; Mittal, K. L., Shah, D. O., Eds.; Plenum: New York, 1991; Vol. 11, p 95.

( 1 1) Puwada, S.; Blankschtein, D. J. Phys. Chem., following paper in this issue.

(12) Puwada, S.; Blankschtein. D. In Proceedings of thesymposium on Mixed Surfactant Systems; Rubingh, D. N., Holland, P. M., Eds.; ACS Symposium Series: New York. in press.

(13) Sarmoria, C.; Puwada, S.; Blankschtein, D. To be published. (14) Blankschtein, D.; Thurston, G. M.; Benedek, G. B. Phys. Reu. Lett.

1985,54, 955. Blankschtein, D.; Thurston, G. M.; Benedek, G. B. J . Chem.

J. Phys. Chem. 1992'96, 5519-5592 5579

Phys. 1986, 85, 7268 and references therein. (15) Corkill, J . M.; Goodman, J. F.; Walker, T.; Wyer, J. Proc. R. SOC.

London, A 1969, 312, 243. (16) Briganti, G.; Puwada, S.; Blankschtein, D. J . Phys. Chem. 1991,95,

8989. (17) Puwada, S.; Blankschtein, D. J . Phys. Chem. 1989,93,7753. Carale,

T. R.; Blankschtein, D. J . Phys. Chem. 1992, 96, 459. Carvalho, B. L.; Briganti, G.; Chen, S.-H. J . Phys. Chem. 1989, 93, 4282.

(18) Guggenheim, E. A. Mixtures: The Theory of the Equilibrium Properties of Some Simple Classes of Mixrures, Solutions and Alloys; Clarendon: Oxford, 1952.

(19) Lupis, C. H. P. Chemical Thermodynamics of Materials; North- Holland: New York, 1983; Chapter XI.

(20) Missel, P. J.; Mazer, N . A.; Benedek, G. B.; Young, C. Y.; Carey, M. C. J . Phys. Chem. 1980,84, 1044.

(21) Bockris, J. OM.; Reddy, A. K. N . Modern Electrochemistry; Plenum: New York, 1977.

(22) Mukerjee, P.; Mysels, K. J. In Critical Micelle Concentration of Aqueous Surjactanr Sysrems; Natl. Stand. Ref. Data Ser.-Natl. Bur. Stand.

No. 36; US. Department of Commerce: Washington, DC, 1971. (23) Mitchell, D. J.; Tiddy, G. J. T.; Waring, L.; Bostock, T.; McDonald,

M. P. J. Chem. Soc., Faraday Trans. 1 1983, 79,975 and referenccs therein. (24) Rosen, M. J. In Phenomena in Mixed Surfactanr Systems; Scame-

horn, J. F., Ed.; ACS Symposium Series 31 1; ACS: Washington, DC, 1986; p 144. Rosen, M. J. Surfacrants and Interfacial Phenomena, 2nd ed.; John Wiley: New York, 1989.

(25) Kaler, E. W. Personal communication. Kaler, E. W.; Murthy, A. K.; Rodriguez, B. E.; Zasadzinski, J . A. N . Science 1989, 245, 1371.

(26) Nilsson, P.-G.; Lindman, B. J. Phys. Chem. 1984, 88, 5391. (27) Souza, L. D. S.; Corti, M.; Cantu, L.; Degiorgio, V. Chem. Phys. Lett.

1986, 131, 160. Valaulikar, B. S.; Manohar, C. J. Colloid Interface Sci. 1985, 108, 403. Sadaghiania, A. S.; Khan, A. J . Colloid Interface Sci. 1991, 144, 191.

(28) Fujita, H. In Polymer Solutions; Elsevier: Amsterdam, 1990; p 311. (29) Hashimoto, T.; Sasaki, K.; Kawai, H. Macromolecules 1984, 17,

(30) Israelachvili, J . Infermolecular and Surface Forces, Academic: 28 12.

London, 1985; p 78.

Theoretical and Experimental Investigations of Micellar Properties of Aqueous Solutions Containing Binary Mixtures of Nonionic Surfactants

Sudhakar Puvvada and Daniel Blankschtein* Department of Chemical Engineering and Center for Materials Science and Engineering, Massachusetts Institute of Technology, Cambridge, Massachusetts 021 39 (Received: January 13, 1992)

We present a detailed description of a molecular model of mixed micellization for aqueous solutions containing binary surfactant mixtures. The molecular model incorporates the effects of surfactant molecular structure, surfactant composition, and solution conditions on the salient physicochemical factors which control mixed micelle formation and growth. These factors include (i) hydrophobic interactions between surfactant hydrocarbon chains and water, (ii) curvature-dependent interfacial effects associated with the creation of the micellar core-water interface, (iii) conformational effects associated with hydrocarbon chain packing in the micellar core, (iv) repulsive steric interactions between the surfactant hydrophilic moieties, (v) electrostatic interactions between charged surfactant hydrophilic moieties, and (vi) entropic effects associated with mixing the two surfactant species in a mixed micelle. The free energy of mixed micellization gmi, is computed for various micellar shapes sh, micellar core minor radii I, , and micellar compositions a. The optimum equilibrium values sh*, IC*, and a* are then obtained by a minimization procedure. The deduced optimum micellar shape sh* determines whether the mixed micelles exhibit two-dimensional, one-dimensional, or no growth. These results are then used in the context of a thermodynamic theory of mixed micellar solutions, and the resulting molecular-thermodynamic formulation is utilized to predict a broad spectrum of mixed micellar solution properties as a function of surfactant molecular structure, surfactant composition and concentration, and solution conditions such as temperature. The predicted properties include the critical micellar concentration (cmc), the average mixed micellar composition at the cmc, the weight-average mixed micelle aggregation number, and the liquid-liquid phase separation coexistence curve, including the compositions of the two coexisting micellar-rich and micellar-poor phases. The theoretical predictions are compared with experimentally measured micellar properties of aqueous solutions of three binary mixtures of nonionic surfactants belonging to the alkyl poly(ethy1ene oxide) (CiEj) family, including C&6-C12Esr CI2E6-C&E6, and C12E6-c10E4. The broad spectrum of theoretical predictions compares very favorably with the experimental data.

I. Introduction Solutions containing mixtures of surfactants (mixed micellar

solutions) are currently a subject of considerable practical im- portance because they can exhibit properties which are superior to those of solutions containing the constituent single surfactants.' For example, it is well-known that in aqueous solutions of binary surfactant mixtures synergistic (attractive) interactions between the two surfactant species result in critical micellar concentrations (cmc's) which can be substantially lower than those in solutions containing the constituent single surfactants. Moreover, in general, synergistic interactions between different surfactant species can be, and have been, exploited by the surfactant technologist in designing solutions of surfactant mixtures which display unique desirable properties. Accordingly, developing a fundamental understanding of the behavior of mixed micellar solutions, in- cluding a rationalization of the nature of synergistic interactions, constitutes a problem of great practical importance. Indeed, such

'To whom correspondence should be addressed.

0022-3654/92/2096-5579%03.~0~0

an understanding can assist the surfactant technologist in the rational design and property control of solutions containing sur- factant mixtures. Specifically, in order to tailor solutions of surfactant mixtures to a particular application, the surfactant technologist has to be able to predict and manipulate (i) the tendency of surfactant mixtures to form mixed micelles and other self-assembling microstructures, (ii) the distribution of surfactant species between mixed micelles and monomers, (iii) the nature of the mixed micelles such as their shape, size, composition, and size and composition distribution, and (iv) the phase behavior and phase equilibria of mixed micellar solutions.

In spite of their considerable practical relevance, as well as the challenging theoretical issues associated with the description of these complex fluids, solutions of surfactant mixtures have not been studied sufficiently. In particular, previous theoretical studies of mixed micellar solutions have evolved along two very different, seemingly unrelated, fronts. On the one hand, significant efforts have been devoted to understand the mixture cmc24 as well as the micellar size and composition distrib~tion.~ On the other hand,

0 1992 American Chemical Society