• Aucun résultat trouvé

Ginzburg-Landau theory of aqueous surfactant solutions

N/A
N/A
Protected

Academic year: 2021

Partager "Ginzburg-Landau theory of aqueous surfactant solutions"

Copied!
21
0
0

Texte intégral

(1)

HAL Id: jpa-00247762

https://hal.archives-ouvertes.fr/jpa-00247762

Submitted on 1 Jan 1992

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Ginzburg-Landau theory of aqueous surfactant solutions

G. Gompper, Stefan Klein

To cite this version:

G. Gompper, Stefan Klein. Ginzburg-Landau theory of aqueous surfactant solutions. Journal de Physique II, EDP Sciences, 1992, 2 (9), pp.1725-1744. �10.1051/jp2:1992230�. �jpa-00247762�

(2)

Classification Physics Abstracts

61.20G 61.25E 64.60C

Ginzburg-Landau theory of aqueous surfactant solutions

G. Gompper and Stefan Klein

Sektion Physik der Ludwig-Maximilians Universitit Mfinchen, 8000 Minchen 2, Germany (Received 26 May1992, accepted is June 1992)

Abstract. A simple two order-paraIneter Ginzburg-Landau theory for binary mixtures of

water and amphiphile is introduced. The amphiphile concentration is described by a scalar, the

orientational degrees of freedom of the amphiphile by a vector order parameter. Phase diagrams

are calculated by minimizing the free energy functional. Several ordered lyotropic phases of lamellar, hexagonal, inverse hexagonal, and cubic structure exist. Phase diagrams show the

sequence of ordered phases with increasing amphiphile concentration commonly observed in

such systems. Several scattering intensities characterizing the disordered phase are calculated.

They clearly show the tendency for bilayer or micelle formation. Finally, the elastic bending

constants of surfactant bilayers are determined. The bending modulus K is found to depend

only on the orientational profile of a flat bilayer.

1. Introduction.

Aqueous surfactant solutions, I-e- mixtures of water and amphiphiles, show a large variety of unusual phenomena [1-4]. For example, for low surfactant concentrations a disordered phase

is observed, in which the amphiphiles self-assemble into spherical and cylindrical micelles.

Amphiphiles in aqueous solution can also assemble in bilayers and vesicles. At higher surfactant concentrations and low temperatures, several ordered lyotropic phases have been found, with

Iamellar, hexagonal and cubic symmetries. Here, the cubic phase with bicontinuous structure [5, 6] is of particular interest. At higher temperature, another disordered phase exits with a

characteristic scattering behavior which is dominated by a broad peak at finite wavevector.

Due to this variety of interesting phenomena, and the importance of aqueous surfactant solutions in biology, these systems have been studied intensively over the last years, both

experimentally and theoretically. On the theoretical side, various models have been proposed,

each focusing on a different aspect of the amphiphilic system. The underlying mechanism, which is responsible for all these effects, is clear: polar molecules like water and polar head groups of amphiphiles want to avoid contact with the hydrocarbon chains of the amphiphile

tails. This leads to the formation of micelles and bilayers in aqueous surfactant solutions. With

increasing surfactant concentration these aggregates pack together to form the ordered lyotropic

(3)

phases. Due to the complexity of the system under consideration, drastic simplifications have to be made. "Lattice "models [7-10] simplify the two-particle interactions and discretize space, in order to facilitate a statistical analysis. "Interfacial"models [11-14] use bilayers as the

basic building block. All the properties of the molecules and their interactions enter the model via the elastic bending energy [it] of bilayers. Finally, "thermodynamic"models [15] consider

micelles as the basic units of the theory. Micelles of different sizes are considered to be in chemical equilibrium with each other and with amphiphiles in solution.

Each of these models has its disadvantages. In the "lattice" models, a realistic description of the ordered lyotropic phases is difficult. "Interfacial "models fail for weak amphiphiles. Fur-

thermore, they cannot easily describe the coexistence of a phase made of bilayers, like a lamellar phase, and another one with a different structure, like the hexagonal phase. 'Thermodynamic'

models are restricted to micelles.

In this paper, we introduce a Landau-Ginzburg model for aqueous surfactant solutions, which is inspired by Landau models for ternary oil-water-surfactant mixtures [16-19]. We will demonstrate that this model can account for a large part of the spectrum of phenomena in aqueous surfactant solutions. In section 2, the model is defined in detail. It is used in section 3 to calculate the scattering intensity of the homogeneous phases. In section 4, the phase

diagrams and the structure of the ordered lyotropic phases are discussed. Section 5 is devoted to the study of micelles and bilayers. Finally, the elastic properties of bilayers are calculated in section 6.

2. Ginzburg"Landau model.

In order to construct a Ginzburg-Landau model for water-surfactant mixtures, one has to decide first which degrees offreedom of the system are considered to be essential and therefore have to be included in the model. Obviously, a scalar order parameter field, #(r), for the local

density difference of the two components is necessary, as in any other binary fluid mixture.

To describe the anisotropy of the surfactant molecules, we introduce a vector order parameter field, T(r), which models the orientation of the amphiphiles [7-10, 18, 19].

Let us consider first a binary fluid mixture oftwo ordinary fluids, with isotropic interactions between all molecules. Its free energy is given by

~f~in = f d~rwi(vi)~

+ ii (i) »« (1)

where p is the chemical potential difference between the two components. Ii(#) is the free energy density of ho mogeneovs order parameter configurations; its form will be specified below.

Here, the gradient term, with pi > 0, disfavors spatially varying order parameter configura-

tions. We consider next a pure amphiphile system. In this case spatially modulated phases,

like a smectic liquid crystalline phase, do exist. This can be modelled in a Ginzburg-Landau theory by a negative gradient term, which favors spatial modulations. Higher order derivatives

are then necessary to make the model thermodynamically stable. Thus, for pure amphiphile

we have the free energy functional

~famph =

/ d~rltxi(v

T)~ + tx~(v~T)~ + tx3(v x T)~ + a4(v(v T))~ + f~(T~)1 (2)

with ai < 0 to describe the tendency of the amphiphilic molecules to put their heads or their tails together, and a2 > 0, a4 > 0 to insure thermodynamic stability. Here, we have included

(4)

all terms with first and second derivatives, which are rotationally invariant and are quadratic in the order parameter. Note that a term (VT)~ = £; ;(0;T;)(0;T;) does not appear, because by partial integration it can be shown that this term is $quivalent to a linear combination of other

terms in (2) (in case the boundary contributions can be neglected). Finally, the amphiphiles in

aqueous solution are oriented in such a way that the polar heads point towards the water. This

anisotropic interaction is described by terIns containing both order parameters. We include

again all terms with first and second derivatives, which are quadratic in the order parameters.

This is, in fact, only a single term, so that we finally arrive at J~ = / d~r[txi(V

T)~ + a2(V~T)~ + tx3(V X T)~

(3)

+ °4(i7(i7 T))~ + fli(i7#)~ + 7(T 'V#) + U(#,T~) P#]

To complete the definition of our free energy functional, we need to specify the free en- ergy density U(#,T~) of homogeneous order parameter configurations. We consider here two different forIns for this function, which are derived from different principles. The first is an expansion in powers of the order parameters, which makes our model a true Ginzburg-Landau theory, where all powers allowed by symmetry appear, so that

U(~,T~) # fl(~) + b2T~ + b4T~ + Cl~T~ + C2~~T~ (4)

with

fl(#) " £ ~n#" (5)

n=2,6

The coefficients in (5) are chosen in such a way that two minima exist; we have to go to a

sixth order polynomial because we want to break the symmetry between the two phases at coexistence. To ensure thermodynamic stability, a6 > 0, b4 > 0 and e2 > 0. We take $2 > 0

so that no homogeneous phases with non-zero orientation ("nematic ") can exist. The term

diT~# acts like a local chemical potential, which favors locally large surfactant concentrations

# where T~ is large for di < 0.

The second possibility is to obtain the free energy functional as

a sum of interaction energy and entropy. With pairwise interactions, the energy contributions to the free energy density U(#,T~) are terms proportional to #~ and T~. The entropy is calculated in the random mixing approximation. It is shown in Appendix A that this gives the generalized Flory-Huggins form

S(~,T~) # d~r (~in~ + M(1 ~)ln(1 ~) + ~ (C2f~ + C4f~ + )j (6)

where f

= T/# is the average local orientation of the amphiphiles at fixed concentration, and M is the relative molecular volume of the amphiphile (compared to water). In this case we have

U(~,T~) # fl(~) + b2T~ + T~ [C2f~ + C4f~j (7)

with

Ii(#)

= a2#~ + Tl#In# + M(I #)'n(I #)1 (8)

For T = 0, U(#, T~

= 0) = Ii(') is identical with the usual Flory-Huggins form [20] of the free energy.

In equations (7) and (8), we have introduced a temperature T for the case of the Flory potential. For the Landau potential (4), a temperature can be defined analogously, I-e- by

JOURNAL DE PHYSIOUE II T 2, N' 9, SEPTEMBER 1992 65

(5)

writing the potential U(#,T~) in the form [a2#~ + b2T~ TS(#, T~)j. Thus, in equations (4),

(5) we set d2 " a2 + d2T, $2 " b2 +12T, and a;

= d;T, $; = b;T for j > 2, as well as

d; = d;T for j = 1,2. We will calculate phase diagrams as a function of the parameter T in sections 3 and 4 below. However, it has to be kept in mind that T cannot be easily compared

with the temperature of a real system. The reason is that we have not taken into account all

degrees of freedom in our model. In particular the rotational degrees of freedom of the water molecules have been shown to be important [21]. These degrees of freedom are considered to be

integrated out in our model. The price we have to pay is that the interaction constants in our model become themselves temperature dependent. We will ignore this additional temperature dependence here. Thus, our results can be compared with experimental data only at constant

temperature, as a function of the amphiphile concentration.

3. Scattering intensity of disordered phases.

We calculate the phase diagram of our model by minimizing the free energy functional (3)

with respect to all order parameter configurations. The simplest phase is a homogeneous

and isotropic disordered phase, in which the order parameter field #(r) is constant, and T(r)

vanishes identically. Information about the order parameter correlations in this phase can be obtained from neutron or X-ray scattering experiments. We calculate the structure functions here in the Ornstein-Zernike approximation. To do so, we expand the free energy functional to second order in the deviations of the order parameters from their average values T

= 0 and

# = lo- This is most easily done in Fourier space, with the result

~F = FMF(#o) + /d3k (( L(k) ((

+.. (9)

where the matrix L is given by

I<(k) ~7k L(k) =

~

2 (lo)

-7k p~l(k)

2 with

K;i = (b;i + e~'~' j, I e (1,2,3)

i~ (II)

P ~ " flik~ + j0(U(#,T~)(j=j~,T=o

and

"~~~ ~ "~~~ ~ ~~~~~'~~~~~~~°'~~°

(12)

e = (ai a3)k~ + a~k~

The inverse of the matrix L contains all structure functions. In particular, we find

< #(k)#(-k) >

=

~~~ ~ ~

(13a) ( + e

~ pk~

4

~ ~~~~ ~~ ~~ ~ ~

~ ~ ~~~~

~~~~~

~ ~ ~~~~ ~~ ~

~ ~ ~~~~

~

~

~~ ~ ~ ~~~~

~~~~

(6)

where cos~

= k)/k~, so that is the angle between k and the z-axis.

We begin the discussion of this result with the case of the pure amphiphile system, for which

p(k) = 0. For the Flory free energy density (7), (8) thin is the case for lo " I. It is instructive to study the correlation function < Tz(k)Tz(-k) > as a function of the angle b, see figure I:

for = K/2, this correlation function is the conditional probability to find two amphiphiles

oriented in the z-direction, when the vector joining the two is perpendicular to the z-axis. It

can be shown easily with equation (13c) that for a3 > 0 (and a4 > 0), this function has its maximum always at k

= 0. This reflects the tendency of the amphiphiles to self-assemble in

monolayers: nearest neighbor molecules within a monolayer are oriented in the same direction.

The correlation function for

= 0, on the other hand, describes the correlation along the axis of orientation. In this case, we observe an "antiferromagnetic "ordering on short length scales, corresponding to a tendency for the formation of surfactant multi-layers. The structure

function (13c) has a peak at non-zero wave-vectors, as long as ai < 0.

The real space correlations can be obtained from (13) by Fourier transform. They are shown in figure 2, and illustrate again the physics of self-assembly discussed above. The correlation

Z

) ','

' ,

* '

~

Q~ '

'

M ' z X

f '

,

~

'

M ,' I

~ V ,

,

X '

0 2 k

Fig. I. Correlation funcfiions in Fourier space, < Tz(kz)Tz (-kz) > (solid line) and

< Tz(kz)rz(-kz) > (dashed line), for ai

" -2, a2 " 0.5, a3 " 1.1, a4 " 0.2, and 31U(~=o " 1.

z j

~

~

£ x

~_~

_~~ ', X

/ ' ,----,

v

, ,

2, , 6 r

, ,'

'-'Z

Fig. 2. Correlation functions in real space, < Tz(0)Tz(z) > (solid line) and < Tz(0)rz(z) > (dashed line). The parameters are the same as in figure 1.

(7)

function in the z-direction, integrated over the z- and y-coordinates, has the form

< Tz(z)Tz(°) >= Aze~~'~"+Sm(fi + ~lz (14)

~~~~~

fz+ " ~4(tX2 + tY4)/(~Z ~ ~~~ ~~~~

with

Dz = /2(tY~ + a4)oiu (16)

and cos ~fiz = fz- Ii. It decays with exponentially damped oscillations, as long

as ai < Dz, whereas it decays monotonically for ai > Dz. The line ai

# Dz is called a disorder line [22, 23]. We are interested here only in the case ai < Dz, because in this case the

amphiphile-amphiphile interaction is such

as to favor bilayer formation. The wave-length of the oscillations, fz-, can be interpreted as the average thickness of a bilayer. The correlation

function in the z-direction, integrated over the y- and z-coordinates, is found to be

< Tz(z)Tz(0) >= Aze~~'~"+sinh( ~ + tlz) (17) ix-

(18)

Whe~e

f~+ = /4°1/(°3 + ~~~

with

D~ = @@ (19)

and cosh~fi~ = f~- Ii. This correlation function decays exponentially, with no os- cillations for a3 > D~, whereas it starts to oscillate for a3 < D~. Here, another disorder line

occurs. For a3 < D~ the correlations become quite complicated. We will always stay on the

"monotonic" side of this disorder line. Note that the length scales in the

z- and z-directions are different, and their ratio can be varied over a wide range by an appropriate choice of interaction

parameters.

When water is added to the system, and the interaction between water and amphiphile is

weak, the phase diagram is usually dominated by a miscibility gap between a water-rich and a surfactant-rich phase, both of which are disordered phases. The phase-diagram for such a case is shown in figure 3. We can now compare the scattering intensities within the two phases at

coexistence. They are shown in figure 4 for T

= 0.5. We find that the Li-phase behaves like an ordinary binary liquid mixture: the scattering intensity < #(k)#(-k) > has a peak at k

= 0,

and < Tz(k)Tz(-k) > is essentially independent of the angle @. We conclude that in this phase the amphiphiles do not form aggregates, the fluid is disordered on a molecular scale. In the

L2-phase, on the other hand, the fluid is strongly structured, as indicated by the scattering peaks at non-zero wavevector of all correlation functions, except < Tz(k)Tz(-k) > for near

K/2. This demonstrates that self-assembly has set in, although the phase is still homogeneous

on a macroscopic scale.

The scattering behavior above the critical temperature Tc is also interesting. The position

of the scattering peak of the correlation functions < #(k)#(-k) > and < Tz(kz) Tz(-kz) >

as a function of the amphiphile concentration is shown in figure 5. It can be seen that the amphiphiles are always oriented antiferromagnetically along their axis of orientation, leading

to a peak at finite wavevector for all surfactant concentrations. The peak position is essentially independent of surfactant concentration. A peak at finite k in the density-density correlation

(8)

.o

A

T

L~

o.5

~ ~'~

4 ~'~

Fig. 3. Phase diagram for the Flory free energy ~vith the parameters ai " -6, a2 " 9, a3 " 5,

a4 " 5, fli " 10, ~ = 10, M

= 20, a2 " 1, b2 " -10, c2 " 17/12. The dot (.) marks the position of

the critical point.

<wi-k) wik)> <w(-kj wikj>

o i k o ~

<i(-ki iiki> <i(-ki iiki>

' '

' '

"

"

," "

, ' "

' "

o o ~

Fig. 4. Correlation functions in Fourier space at two-phase coexistence, for the temperature T

= 0.5

in figure 3. The curves on the left show the correlations in the water-rich Li-phase, the curves on the right in the surfactant-rich L2-Phase. The dashed lines are the < Tzik)Tz(-k) > correlations for e

= 0 (strongest peak at finite k), e

= K/6, e = 2x/6, and e

= x/2 (peak at k

= o). The parameters

are the same as in figure 3.

(9)

function, on the other hand, appears only at large surfactant concentrations. It does not move out continuously from k

= 0, but appears suddenly at k ci 0A. Furthermore, there is

a range of amphiphile concentrations, where < #(k)#(-k) > has two peaks, one at k

= 0,

the other at finite k. This indicates a structure with three different length scales: (I) there is a bilayer-type structure with a repeat distance I, which determines the peak position at

non-zero k ci kbi " 2K/1 of both correlation functions (it) this bilayer structure decays with

a correlation length fbi, which is inverse proportinal to the width of the peak at kbii (iii)

there are density fluctuations, which decay with another correlation length fj, which is inverse

proportional to the width of the peak at k

= 0. The density-density correlation function looses its peak at k

= 0 at a Lifshitz-line, which is given implicitly by

7~ = 2fl101U(#,T~)li=io,T=o. (2°)

At this line, the peak at k m kbi already dominates the scattering intensity, so that the Lifshitz-line is of minor relevance here. Thus, at very large amphiphile concentrations, the

density fluctuations become neglibible compared to the 'bilayer'fluctuations.

6

~max <Ii>jj

,

4 , <jj>

z

o

~ ~

lo ~'~

Fig. 5. Concentration dependence of the peak positions of the scattering intensities

< rz(kz)rz(-kz) > (full line) and < #(k)#(-k) > (dashed line) along the path A in figure 3.

4. Phase diagrams Ordered phases.

When the strength of the water-amphiphile and the amphipbile-amphiphile interaction is large enough, ordered phases can become stable in our model. We look again for configurations

which minimize the free energy functional (3). Only crystalline phases with mirror symmetry will be considered here. In this case, it can be shown easily that V x T n 0, and the vector field T(r) can be obtained as the gradient [24] of a scalar field, E(r). Both order parameter fields are expanded in a Fourier series,

N

#(r) = lo + £ £ #kcos(k r)

~

'~~ ~~~' (21)

E(r) = £ £ Ek cos(k r)

I=I kEB,

Références

Documents relatifs

At the critical point [4J (the point wher e a line of second order ph ase tran sitions becomes a line of first orde r phose transitions) it is necessary toluwc Acr(H, T) = 0

In this chapter, we are going to use Landau theory to investigate the R3m -+ C2/m structural phase transition in CuFe0 2. First, we derive a Landau free energy compat- ible

In two cases, it is known that this system admits bounded solutions coming from the invariance of the Ginzburg-Landau equation by translations and rotations.. The specific

For sake of completeness, we give also the limit behaviour of the previous Ginzburg-Landau equation with homogeneous Dirichlet boundary condition (for scalar problems with

In conclusion, we investigate families of fundamental and vortical spatiotemporal dissipative solitons in the framework of the 3D CGL equation with the CQ nonlinearity,

We would like to draw the attention of the reader that in the scalar case considered here the exact form of the optimal profile plays a central role in the analysis. We will

For a sufficiently large applied magnetic field and for all sufficiently large values of the Ginzburg–Landau parameter κ = 1/ε, we show that minimizers have nontrivial

The aim of this final section is to use the structure of the minimizers of the limiting functional F 0,g given by Theorem 4.3 to prove that minimizers of F ε,g 0 have the same