You are on page 1of 14

The Multiple Equilibrium Model of Micelle Formation Author(s): J. M. Corkill, J. F. Goodman, T. Walker and J.

Wyer Reviewed work(s): Source: Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, Vol. 312, No. 1509 (Sep. 2, 1969), pp. 243-255 Published by: The Royal Society Stable URL: http://www.jstor.org/stable/2416220 . Accessed: 03/03/2013 20:29
Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at . http://www.jstor.org/page/info/about/policies/terms.jsp

.
JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact support@jstor.org.

The Royal Society is collaborating with JSTOR to digitize, preserve and extend access to Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences.

http://www.jstor.org

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

Proc. Roy. Soc. A. 312, 243-255 (1969) Printed in Great Britain

The multiple equilibrium model of micelle formation


BY J. M.
CORKILL,

J. F.

GOODMAN,

T.

WALKER

and J.

WYER

Proctor and Gamble Limited, Basic Research Department, Newcastle upon Tyne (Communicated by D. Tabor, F.R.S.-Received 28 November 1968Revised 14 January 1969)
In dilute aqueous solution surface-active molecules aggregate to form micelles. This process has been described by a multiple equilibrium model which considers changes in the distribution of micelle aggregation numbers with concentration. Analysis of the colligative properties of dilute aqueous solutions of non-ionic surface-active agents has enabled the concentration of monomer to be determined as a function of the total solute concentration. The validity of the treatment has been confirmed by showing that these monomer concentrations are in good agreement with those deduced from nuclear magnetic resonance experiments, which do not depend on the assumptions employed in the thermodynamic method. Only one colligative property need be determined as a function of concentration in order to describe adequately the average aggregation numbers and free energies of micelle formation. INTRODUCTION

The abrupt change in the colligative and other solution properties of surface-active molecules in dilute aqueous solution at a critical concentration, known as the critical micelle concentration (c.m.c.), is attributed to the formation of molecular aggregates known as micelles (see McBain I950). The description of this process has been approached from two viewpoints, as a single kinetic equilibrium (Jones & Bury I927; Hartley I936) and as a phase separation (Stainsby & Alexander I950). Both models are limited by an oversimplification in the specification of the micellar species. Hall & Pethica (i 967) have applied Hill's (i 963, I964) small system thermodynamics to overcome this objection and have obtained exact thermodynamic relations for micellar systems. The multiple equilibrium model provides an equally rigorous approach for the thermodynamic treatment of micellization. This model is developed here in a manner analogous to treatments of association in liquid mixtures (Prigogine & Defay I954) to obtain relationships between the average thermodynamic properties of a micellar system; some of which are similar to those derived for protein aggregation (Steiner I952; Adams & Williams I964). The theory is used to obtain the concentration of monomeric solute as a function of the total solute concentration from thermodynamic measurements, and the validity of the treatment confirmed by determining this quantity using a spectroscopic method which is independent of thermodynamic assumptions. The variation of the surface tension of a micellar solution with solute concentration calculated on the basis of this treatment agrees with values determined experimentally. Finally, the change in the standard free energy of micelle formation with aggregation number has been calculated. [ 243

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

244

J. M. Corkill, J. F. Goodman, T. Walker and J. Wyer


THEORY

1. Colligative propertiesof micellarsolutions


It is well known that the colligative properties of ideal solutions are determined by the mole fraction of the solvent, xl. In a system in which the solute exists in both the monomeric and associated states, we may define a solute colligative mole fraction xc, by = I xi, = (1.1) where x2 and xr are the mole fractions of the monomeric and rth associated species respectively and the summation extends over all aggregated species. The total solute concentration, C (g/ml), is related to these mole fractions by

C = (M2/Vm) (X2+ E

nrXr),

(1.2)

where M2 is the solute (monomer) molecular weight, Vmis the molar volume of the solution and nr is the aggregation number of the rth micellar species. For dilute solutions, Vmwill be indistinguishable from the solvent molar volume, Vl. In the subsequent analysis, we shall find it useful to employ a 'stoichiometric' solute mole fraction, xt, rather than C, where
Xt = X2+ EfnrXr.

(1.3)

We can now show that for systems in which the solute species are in chemical equilibrium it is possible to determine x2 from the dependence of xc upon xt (or C). If ar and t2 are the chemical potentials of the rth micellar species and the monomer respectively, we have for the equilibrium between the monomeric and any associated species = (1.4) nr/2 1br We choose the hypothetical standard states in which the solvated species have the partial enthalpies and heat capacities corresponding to dilute real solutions and free energies and entropies corresponding to unit activity (Lewis & Randall I96I). For solutions sufficiently dilute to behave ideally, we have, from (1.4), ln xr =-(#0-nr?)/RT +nrln x2 (1.5)

where ,ar and 4a?are the standard chemical potentials of the associated and monomeric species respectively, R is the gas constant and T the absolute temperature. By differentiation of (1.5) at constant temperature and pressure we have dlnx2 We multiply each side of this equation by with respect to xr to obtain d E, nrkx now xrn dlnX di 2
-

nrxr
nrk where
k +1(1

(1.6) k is a positive integer and sum (1.7)

If we now differentiate x, (equation (1.1)) with respect to In X2, substitute for the

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

Micelle formation

245

derivative of the sum from (1.7) (k = 0) and introduce the definition of xt from (1.3) we have dxc/xt = dlnx2. (1.8) Integration of (1.8), using an experimentally determined relation between xc and xt, enables the change in lnx2 to be calculated. By choosing for the lower limits of integration solutions sufficiently dilute for association to be negligible (xc -- xt), x2 may be determined as a function of xc and hence xt (or C). The colligative mole fraction may be determined from any suitable colligative property; thus in ideal dilute solution from (1.1) (1.9) pi/P1 = xi = 1 -Xe, where pi and p? are the equilibrium vapour pressures (strictly fugacities) of the solvent in the solution and pure liquid states respectively. 2. Light scattering Although we are dealing with a multi-component system, the condition of chemical equilibrium between the solute species means that there is only one compositional variable. The solute specific refractive increment is independent of the concentration above the c.m.c. in these systems and so we may assume that the specific refractive increments of all solute species are the same. For these conditions, it is permissible to employ the two component light scattering equation - HCRTV (Debye I947) (2.1) where T1 is the excess turbidity of the solution due to the solute, H is an optical constant containing the specific solute refractive increment and ir the solution osmotic pressure. For ideal dilute solutions 7V1/RT = -ln x
e. x,

(2.2)

By differentiation of this expression with respect to C and substitution into (2.1) we have from (1.2), after rearrangement HC
T1

1 dxc 1dxc . 14d0 M2 M2dx~~~~~~~~~ VdC &t V,

(2.3)

This equation enables either xc or x2 to be determined from the experimental dependence of HC/I1 upon xt (or C). By choosing the lower limits of integration to correspond to solutions with negligible association (IICM2/1l -- 1) integration with respect to xt enables xc to be calculated. By multiplying each side of (2.3) by dlnxt and M2 we obtain from (1.8)

HCM2dlnxt = dxc =
T

dInX2.

(2.4)

Xt

Integration of this expression enables x2 to be determined. Since the same information may be obtained from light scattering or colligative property measurements, the choice of method is determined by the relative

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

246

J. M. Corkill, J. F. Goodman, T. Walker and J. Wyer

experimental accuracy. It has been found that vapour pressuredepressionmeasurements are the more sensitive when the degree of association is low, light scattering when it is high. 3. Aggregation numbers The average association numbers are a set of parameters characterizing the distribution of the solute between the monomeric and associated species. With respect to all solute species we may define number (N.) and weight (Nw)average association numbers by
N and
- X2
X2+

-rXr=--T
>Xr Xc

(from (1.1), (1.3))

(3.1)
(3.2)

x2+nr(nrxr)
X2 + E nrXr

By differentiating xt with respect to lnx2 and substituting from (1.7) (k = 1) we obtain dxt/dlnx2 = x2+En2xr. (3.3) From (1.8) we have and hence, by division, dx,/d lnx2
=

x2+ EnrXr,

(3.4) (3.5)

N, = dxt/dx,.

We note than Nw may be obtained from either the dependence of xt upon xc or directly from light scattering data (equation (2.3)). Average association numbers may also be defined with respect to the associated solute species. Thus we have for the number (<n>)and weight (<n>w) averages

<n> =

rxr
EXr

XtX2
Xc-X2

(3.6)

N-Xtx2 <n>w= >2nr(nrxr)

(3.7)

If we differentiate (3.6) with respect to lnx2 we obtain, with the aid of (1.7) for the derivative of the sums, after rearrangement, d<n>/d lnX2
=

(<n>w-<n>)

<n>.

(3.8)

Since for a system in which there is more than one associated species <n>w > <n>, it follows that <n> must increase with concentration since d lnx2 has the same sign as dxt. 4. Nuclear magnetic resonance chemical shifts If the resonance corresponding to a particular chemical group within a molecule has a chemical shift Vjfor a particular environment of the molecule, and if there are several possible environments with rapid exchange between them, then the observed chemical shift, P, is given by (4.1) _ E <g01E0, where Cj is the concentration of the molecules in the jth environment and the

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

Micelle formation

247

summation extends over all sites. If we assume that in solution the unassociated molecules have a shift of v2 and all associated species have the same shift vm, then
v X2 v2 +(EnrXr)vm Xt

(4.2)

Rearrangement of this equation leads to


x2=v2-v

Xt.

(4.3)

Thus, if Pm) P2 can be determined, the dependence of v upon xt leads to x2 as a function of xt. 5. Surface tensions For a plane interface at constant temperature and pressure the Gibbs adsorption isotherm may be written in the form -dy = Fj d,j, (5.1)

where y is the interfacial tension, F- and ,ajthe surface excess and chemical potential of the jth solute species and the summation extends over all solute species. For an associating solution, we have from (1.4) -dy = (F2 +
E Frfnr)d/a2.

(5.2)

As the surface excess Fr involves nr Fr moles of solute, the term in brackets in (5.2) is the total surface exceess (Ft) in terms of monomeric solute. We can express d,a2 1 dlnxt in terms of dlnxt since dlnxt dln2 =R1Tdlnx2' (5.3)

we have

and hence by substitution from (3.3) and the introduction of Nw (equation (3.2)) - dy RTFt dlnxt
Nw(54

If Ft can be determined, then y can be calculated from the relationship between


Nw and xt.

6. Standardfree energiesof association


In the development of the thermodynamic theory it has proved useful to ignore the individual micelles and consider only their average properties. If the average chemical potential <,a> of the micellar species is defined as the mean Gibbs free energy <G>per mole of the associated species then since
<G> = ,arXr/ EXr,

(6.1)
(6.2)

it follows that

<a> =

<MO> +

RT<lnxr>.

If we consider a mixture of the individual micellar species in their respective standard states in proportions corresponding to the distributions in a real solution,

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

248

J. M. Corkill, J. F. Goodman, T. Walker and J. Wyer

the free energy change of this mixing process will be DmG =RT E >Xr Xr-n(\X2r/ Exr)
= RT[<ln xr>-ln (E xr)]. (6.3)

The standard chemical potential of the associated species in the mixed state (<00>) is given by = iMG, (6.4) Kv0>-Kito> and thus from equations (6.2) to (6.4) the free energy change for association per mole of monomer referred to the mixed standard state is given by

<0

>2

<n>

R>-T

(lnx2n

<n>x))

(6.5)

Thus <K vO> is related to quantities which may be determined experimentally. The converse problem, that of calculating the standard free energies for particular micelles entirely on the basis of experimental results may in principle be solved by the method given below. From (1.5), by summation with respect to the associated species, we have
a)

EXr

= E>f(nr)exp(-nrs),
0

(6.6)

wheref(O),f(l)

are zero and for n > 2 (integer)


f(nr) = exp (-[?
- nr/,tO]/RT)

(6.7)

and

s=-lnx2.

(6.8)

The sum Exr is (XI - x2) and hence, from the experimental data, this may be obtained as a function of s, say g(s). Equation (6.6) then becomes
g (s)
=

co >f(nr)
0

expHnr

8).

(6.9)

The right hand side of this equation is related to the Laplace transform and from the Fourier integral theorem (Sneddon I95I), it can be shown that
1
Ifd+ir1

f(nr)

2ni)j

g(s)exp(nrs)ds

(nr

>

0, integral),

(6.10)

where a is a constant such that all singularities of g(s) in the complex plane lie to the left of the path of integration. Equation (6.10) provides a formal solution to the problem of obtaining the free energies of formation of the individual micelles from experimental data. In principle f(nr) may be obtained by fitting a suitable function of s to the relationship between

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

Micelle formation

249

s and g(s) obtained from experimental results and applying (6.10). The approxima-

tion of replacingthe series summation (6.9) by an integration and using the normal Laplace inversion to obtain a continuous approximation to f(nr) leads to large errors (- 20 %) when <n>is small (- 5) for the simple trial function we investigated (f(nr) cx I/n +,8In- + A where a, /3,Z and A are constants).

EXPERIMENTAL

Materials 1The zwitterionic surface-active agent 3(dimethyloctylammonio)-propanesulphonate (C8H17N Me2(CH2)3SO ) was chosen for this study. The change in the

nuclear magnetic resonance chemical shift between the monomeric and micellar states is much larger for a fluorine nucleus than for a proton (Muller & Birkhahn I967), and to enable 19Fnuclear magnetic resonance experiments to be conducted the seven terminal protons were replaced by fluorineatoms to give the heptafluoro . Both compoundswere preparedfrom the compound C3F7(CH2)5N+Me2(CH2)3SO correspondingbromides by the following synthetic route.
CH2CH2CH2

I
HNMe2

0~~S02

RBr-

2RNMe2

-RN+Me2(CH2)3SOj

The products were recrystallized from acetone and ethanol and gave the correct elemental analyses. The bromide used to prepare the heptafluoro compound was synthesized from 4-pentene-1-ol and heptafluoropropyliodide as follows:
C3F71

CH2=CH(CH2)30H

C3F7CH2-CHI(CH2)30H
PBr3 PBr3 C3F7(CH2)50H
Zn/HCI

C3F7(CH2)5Br

measurements Thermodynamic (i) Light scattering Aqueous solutions of the fluorinatedsurface-active agent were clarifiedby filtration through 'Millipore' filters (100nm pore size) and the intensity of scattered light at 25 ?Cmeasured at angles of 45, 90 and 1350 to the incident beam by means of a Sofica model 4200 P.G.D. instrument. The disymmetry (Z45) of the scattered light was less than 1.05 in all cases and the turbidities, T, were therefore calculated from the scattering at 900 on the basis of a benzene calibration (Coumou I960). Correctionsfor depolarization of the scattered light were found to be negligible. Refractive index increments were determined with a Rayleigh interferometer.

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

250

J. M. Corkill, J. F. Goodman, T. Walker and J. Wyer

(ii) Vapourpressureosmometry The differencein temperature between a drop of solution and a drop of solvent contained in a constant temperature enclosure saturated with solvent vapour is proportional to the vapour pressure depression of the solution (Huff, McBain & Brady I95I). Thermistor beads are used to measure the temperature difference and the apparatus calibrated with aqueous sucrose solution of known activity (Robinson & Stokes 1959). The Mechrolab 301 and Hitachi-Perkin-Elmer instruments were used for these measurements. The latter is more sensitive and was requiredto detect the small activity change in micellar solutions of the fluorinated compound. (iii) Surface tension Surface tensions of aqueous solutions of the fluorinated solute were measured using a drop volume apparatus which was immersed in a water thermostat (25.00 + 0.05?C). The drops were formed in a Teflon tip because difficulty was experienced in wetting correctly the glass tips normally used. The tip was attached to a micrometersyringe held in an atmosphere saturated with solvent vapour and the surfacetension was calculated by the proceduredescribedby Harkins & Brown
(I 9I9).

(iv) Nuclear magneticresonance(n.m.r.) measuremnents N.m.r. measurementsfor the 19Fnucleus were carried out using a Varian DA 60 system at a frequency of 56.4 MHz at 25 ?C.The external referencesolution (40 % trifluoroaceticacid in water) was contained in a 1 mm capillary and centred in the sample tube with Teflon bushes. The chemical shift of the terminal CF3group of with reference to trifluoroacetic acid was measured C3F7(CH2)5N+Me2(CH2)3SO by using a side band calibration with a Hewlett Packard 200 CD audiofrequency oscillator monitored on a Venner frequency counter. Chemicalshifts were corrected for bulk susceptibility effects, which were found to be small.

RESULTS

The turbidities, T, of aqueous solutions of C3F7(CH2)5N+Me2(CH2)3SO3 and the colligative mole fraction (xc)fromosmometry measurementsare shown as a function of the total solute concentrationin figures1 and 2 respectively. Values of xc obtained by integrating the light scattering data using equation (3.5) are in good agreement with the values obtained directly by osmometry (figure 2), thus verifying the equivalence of the two methods. The chemical shift of the terminal CF3 group is plotted as a function of the reciprocal concentration (x-y1)in figure 3, and the variation of the surface tension with the logarithm of the solute concentration (lgxt) in figure 4. The surface tension decreases with increasing concentration

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

Micelle formation
62!;4 -

251

5
102 C/gml-1

10

2
103Xt

4 FIGURE2

FIGURE1

solutions, turbidity against FIGURE 1. Light scattering, C3F7(CH2)5N+(CH3)2(CH2)3SOconcentration. solutions, xc against xt, FIGURE 2. Colligative properties, C3F7(CH2)5N+(CH3)2(CH2)3SO, integration of light scattering results (calculated by using (3.5)). 0, Osmometry;

300 -

250-

0.5
10-3x-1

1.0

1.5

FIGURE 3. Chemical shifts of the CF3 group for C3F7(CH2)5N+(CH3)2(CH2)3SO- solutions against x-1.

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

252

J. M. Corkill, J. F. Goodman, T. Walker and J. Wyer

at all concentrations, but the decline is relatively small in micellar solutions


Precise turbidity data could not be obtained for aqueous solutions (Xt > 1.5 x 10-3). of C8H17N+Me2(CH2)3S0- because of the small aggregation numbers, but the variation of xc with xt from osmometry measurements is shown in figure 5. The

36 -

34 0 cn 32 -3.0 lgxt solutions FIGURE 4. Surface tension of C3F7(CH2)5N+(CH3)2(CH2)3SO- -, lg xt. 0, Experimental; calculated (by using (5.4)). against -0_Q -2.5 -0

8-

-~~~~~~~~~~~~~

10
103 Xt

20

FIGURE 5. 0,

solutions. Vapour pressure osmometry, C8Hl7N+-(CH3)2(CH2)3SO


x. against

xt;

, x2 against xt (calculated

by using (1.8)).

absence of any clearly defined discontinuity illustrates the difficulty in defining a c.m.c. precisely for surface-active agents with low aggregation numbers. The plot of xc against xt shows deviations from ideal behaviour at concentrations well below the c.m.c. region. A verification of the interpretation of the deviations from ideal behaviour in terms of solute aggregation was obtained by examining the homologous compound C6Hl3N+Me2(CH2)3SO, which was found to behave almost ideally over the same concentration range.

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

Micelle formation
DISCUSSION

253

Multiple equilibriummodel The large change in the chemical shift for the 19Fnucleus in the monomeric and micellar states enables x2 to be determined as a function of xt by a method that is independent of thermodynamic assumptions. By extrapolation of the linear portions in figure 3, V2 and vm were found to be 241.0 and 315.0 Hz respectively relative to trifluoroacetic acid. The effect of micelle size on vm is assumed to be negligible. Values of x2as a function of xt calculated from (4.3) are in good agreement with those obtained by light scattering measurementsby means of (2.4), as shown in figure 6. We conclude that the multiple equilibrium model adequately describes the aggregation of surface-active molecules in dilute aqueous solution.

20

2 10wxt

of monomeric detergent, FIGURE 6. The concentration C3F7(CH2)5N+(CH3)2(CH2)3SO3, , from light scatterX2 against x,. 0, From n.m.r. results (calculated by using (4.3)); ing results (calculated by using (2.4)).

Surface tension It has been shown for a number of surface-active agents that a plot of surface tension against the logarithm of the concentration is linear well below the c.m.c. (van Voorst Vader I960) which is consistent with the experimental observation made by direct radiotracermeasurementsthat the adsorbedmonolayeris essentially complete at concentrations well below the c.m.c. (Corkill, Goodman, Ogden & Tate I963). Assuming that Ft is defined by the linear region of the relation of y against lg xt below the c.m.c. (figure 4), the variation of y with xt over the whole concentration range may be calculated by using (5.4) and values of NRobtained from light scattering experiments. The agreement with the experimentally detervalues of y against lg xt is good, showing that the abrupt change in y against *mined xt observed at the c.m.c. is simply a consequence of the abrupt change in NR (Elworthy & Mysels I966).
I7

Vol. 312.

A.

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

254

J. M. Corkill, J. F. Goodman, T. Walker and J. Wyer


Aggregation numbers and free energies of micellization

For C8Hl7N+Me2(CH2)3SO-, x2 does not change abruptly at any well defined concentration (figure 5) and values of <n> and <n>wcalculated from (3.6) and (3.7) are small and increase markedly with increasing concentration (figure 7). The large differencle between <n> and <n>w at any particular concentration shows that the micelles are polydisperse with respect to size.

10
3>5

<n>w

<n>

5-

10
1O3xt

20

FIGURE7. Aggregation numbers for C8H17N+(CH3)2(CH2)3SO3, (n>, and (n> against xt (calculated by using (3.6) and (3.7)).

As <n>is small in the concentration range in which it can be accurately determined the integral approximation to the series inversion formula cannot be applied to obtain the dependence of the individual standard free energies of micelle formation (,u?- nr/O) upon n,. However, the average free energy of micelle formation, with respect to the mixed standard states can be obtained. The value of <(ziqO>for C8H17N+Me2(CH2)3SO- decreases sharply with increasing <n> from -2RT at <n> = 2.5 at small values of <n>. For <n> = 8, <zJ00> has the value - 4.5RT and is much less dependent upon <n>. Thus as the micelle size increases the free energy for the addition of a further monomer molecule tends to become constant. For these small micelles a considerable alkyl chain-water interaction in the core of the micelle is to be expected. As the association number increases the extent of alkyl chainwater interaction in the core region will decrease. The limiting value, approached at high association numbers, represents the difference between the transfer of the hydrophobic chain from an aqueous to a non-polar environment and the free energy term associated with the packing of the head groups. The variation of x2 with xt for C3F7(CH2)5N+Me2(CH2)3SOT is small at the higher concentration (figure 6) and the concept of a critical micelle concentration calculated from equation (3.7) increases from (x* = 1.50 x 10-3) is reasonable. <n>w 20 to 25 over the concentration range studied above the c.m.c. Values calculated using the Debye approximation (Debye I949), where the micelles are considered to

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

Micelle formation

255

be in a solvent defined by the solution composition at the c.m.c., are the same as those calculated from (3.7) within the experimental error. For this compound, the standard free energy of micellization with respect to the mixed standard state is = RT ln x (Corkill, Goodman & Tate I968), and is well approximated by AGO - 6.50RT at 25 ?C. Conclusions
The use of the multiple equilibrium model in the treatment of the properties of micellar solutions has shown how the various colligative properties may be related to the distribution of the solute between the monomeric and associated states. The connexion between the colligative and other (spectroscopic, surface) properties of the system has been demonstrated. Although a formally exact relationship enabling the free energy of formation of a particular species to be calculated from the colligative properties has been deduced, mathematical difficulties have prevented us from applying it. It is, however, possible to obtain the average free energy of association as a function of the average association number. The behaviour of this quantity is consistent with the currently accepted model of the micellar state.
REFERENCES

Adams, E. T., Jun. & Williams J. W. I964 J Am. Chem. Soc. 86, 3454. Corkill, J. M., Goodman, J. F., Ogden, C. P. & Tate, J. R. I963 Proc. Roy. Soc. A 273, 84. Corkill, J. M., Goodman, J. F. & Tate, J. R. I968 Symposium on hydrogen bonded solvent systems (University of Newcastle upon Tyne) (eds. A. K. Covington and P. Jones). London: Taylor and Francis. Coumou, D. T. I960 J. Colloid Sci. 15, 408. Debye, P. 1947 J. Phys. Coll. Chem. 51, 18. Debye, P. 1949 J. Phys. Coll. Chem. 53, 1. Elworthy, P. H. & Mysels, K. J. I966 J. Colloid Sci. 21, 331. Hall, D. G. & Pethica, B. A. I 967 Nonionic surfactants (ed. M. J. Schick), ch. 16. London: Arnold. Harkins, W. D. & Brown, F. E. I919 J. Am. Chem. Soc. 41, 499. Hartley, G. S. I936 Aqueous solutions of paraffinic chain salts. Paris: Hermann. Hill, T. L. I963, I964 Thermodynamics of small systems, vols. 1, 2. New York: Benjamin. Huff, H., McBain, J. W. & Brady, A. P. 1951 J. Phys. Chem. 55, 311. Jones, E. R. & Bury, C. R. 1927 Phil. Mag. 4, 841. Lewis, G. N. & Randall, M. I96I Thermodynamics, ch. 20. London: McGraw-Hill. McBain, J. W. 1950 Colloid science, ch. 17. New York: Heath. Muller, N. & Birkhahn, R. H. I967 J. Phys. Chem. 71, 957. Prigogine, I. & Defay, R. 1954 Chemical thermodynamics (trans. by D. H. Everett), ch. 26. London: Longmans, Green and Co. Robinson, R. A. & Stokes, R. H. 1959 Electrolyte solutions, app. 8.6. London: Butterworths. Sneddon, I. N. 195I Fourier transforms, ch. 1. New York: McGraw-Hill. Stainsby, G. & Alexander, A. E. 1950 Trans. Faraday Soc. 46, 587. Steiner, R. F. 1952 Arch. Biochem. Biophys. 39, 333. van Voorst Vader, F. I960 Trans. Faraday Soc. 56, 1067.

I 7-2

This content downloaded on Sun, 3 Mar 2013 20:29:54 PM All use subject to JSTOR Terms and Conditions

You might also like