• Aucun résultat trouvé

The interdependence of intra-aggregate and inter-aggregate forces

N/A
N/A
Protected

Academic year: 2021

Partager "The interdependence of intra-aggregate and inter-aggregate forces"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00210161

https://hal.archives-ouvertes.fr/jpa-00210161

Submitted on 1 Jan 1985

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.

The interdependence of intra-aggregate and inter-aggregate forces

J.N. Israelachvili, D. Sornette

To cite this version:

J.N. Israelachvili, D. Sornette. The interdependence of intra-aggregate and inter-aggregate forces.

Journal de Physique, 1985, 46 (12), pp.2125-2136. �10.1051/jphys:0198500460120212500�. �jpa-

00210161�

(2)

The interdependence of intra-aggregate and inter-aggregate forces

J. N. Israelachvili

Department of Applied Mathematics, Institute of Advanced Studies, Research School of Physical Sciences, GPO Box 4, Canberra Act 2601, Australia

and D. Sornette

Laboratoire de Physique de la Matière Condensée (*), Université des Sciences, Parc Valrose, 06034 Nice Cedex, France

(Reçu le 6 mai 1985, accepté le 20 août 1985)

Résumé. 2014 Nous examinons l’interdépendance des forces responsables de l’association d’amphiphiles pour former des agrégats tels que micelles, microémulsions et bicouches et des forces s’exerçant entre de telles structures à courte distance. Nous présentons d’abord une analyse de l’origine et de la nature des forces d’hydratation entre têtes polaires qui prédominent à très courte portée. Un développement Landau-Ginzburg permet de relier naturellement la force d’hydratation

au

sein d’un même agrégat à celle s’exerçant entre deux bicouches. Le potentiel d’interaction unique obtenu permet de démontrer la corrélation entre l’aire par tête polaire d’une bicouche isolée et la force de

répulsion entre bicouches. L’application de cette approche à d’autres systèmes est discutée.

Abstract

2014

We examine the interdependence between forces responsible for the self-association of amphiphiles in

aggregates such as micelles, microemulsions and bilayers and the forces occurring between such structures when

they

come

close together. We first present

a

theoretical analysis of the origin and nature of the hydration force which

appears as the most important interaction occurring between head-groups at very short distances. A Landau-

Ginzburg approach allows us to link naturally the hydration force between the head-groups within an aggregate with that between two opposing bilayers. The single interaction potential obtained between amphiphilic head-

groups allows demonstration of

a

quantitative correlation between the measured head-groups

areas

of isolated

bilayers and the repulsive pressures between bilayers. The relevance of this interdependence to various other

phenomena is discussed.

Classification

Physics Abstracts

82. 70K - 82.65D

1. Introduction

The forces acting between amphiphilic structures (aggregates) such as micelles, bilayers, vesicles and microemulsion droplets in water determine many

important phenomena, for example, their stability to aggregation. in dilute solution, their phase behaviour

in concentrated solution, the thickness of soap films,

etc. By contrast, the forces acting between the same

amphiphilic molecules within these structures deter- mine the type of structure they self-assemble into, their shape, size, polydispersity and phase state, and in particular how these parameters are sensitive to changes in the solution conditions. For unlike solid

(i.e. rigid) colloidal particles, the forces holding amphiphiles together within aggregates are weak

(*) Laboratoire LA 190 Associe

au

C.N.R.S.

(viz. van der Waals, screened electrostatic, and sol-

vation forces), which usually results in fluid-like struc- tures that change their size and shape when the solu- tion conditions, temperature, or surfactant concen-

tration is changed. These factors, however, are the same

as those that affect the forces between these structures

(or any colloidal particles in general).

The purpose of this paper is to investigate how the

intermolecular forces within amphiphilic aggregates

are correlated with the forces betweel aggregates, how this interdependence manifests itself in practical terms, and how it may be quantified.

While such correlations have been noticed before [1],

no attempts appear to have been made to consider their generality or their theoretical basis. The reasons

for this are twofold. First, the short-range electro- static, hydration and hydrophobic forces involved

are still not sufficiently well understood. Second,

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198500460120212500

(3)

theories of interparticle and intersurface forces are

often formulated in terms of some smeared-out surface property (rather than between the molecules consti-

tuting the surfaces). For example, a uniform surface

charge density in the case of double-layer forces [2],

or a uniform surface polarization density in the case

of hydration (solvation) forces [3, 4].

Further, the interface itself is usually modelled as a

discontinuous dielectric boundary. By so treating

surfaces as homogeneous planar discontinuities rather than made up of discrete molecular groups (amphi- philic head-groups, for example) it is not readily

apparent how the forces .between these groups can be related to the forces between the surfaces.

Note however, that these formulations are the result of mean field approximations and are not due

to an inherent assumption where one disregards the

direct interaction between the ionic head groups at the surface (1).

In this paper we adopt another approach where right from the start we assume that the forces between individual head-groups within aggregates are the same

as those that give rise to the net force between the aggregates. In other words, a single interaction

potential is postulated to describe both.

This work can be considered as an attempt to generalize to neutral systems what has been established for charged systems.

In section 2 we proceed with a theoretical discussion of the problem, and in particular we derive a simple expression for the repulsive hydration potential

between two head-groups. In section 3 we apply this potential in a detailed quantitative analysis of lecithin

bilayers, while in the final section 4 we discuss the broader implications of the results and show that there are many other systems to which the « principle

of interdependence » applies.

2. Nature of the interactions within and between aggregates.

2.1 THEORETICAL BACKGROUND. - First we consi- der the interaction forces between amphiphiles (sur- factants, lipids) within aggregates that determine the type of structure that they self-assemble into. This has been well described and reviewed in the literature

[ 1, 5] and only the essential features will be given here.

For amphiphiles whose hydrocarbon chains are in

the fluid state (i.e. surfactants above their Krafft point)

their optimum head-.group area ao is determined by a

balance of two « opposing forces » (6) acting mainly

in the interfacial head-group region. The attractive

(hydrophobic) energy contribution per molecule may be expressed as ya where a is the head-group area and

where y, the hydrocarbon-water interfacial (hydro- phobic) free energy, is in the range 20-50 mJ/m2

[1, 7-9]. The repulsive energy contribution per head-

(’) As pointed out by

a

referee.

group may be expressed approximately as cla, where c

is a constant. The total energy per molecule is therefore ya + cla. This has a minimum value at a = ao

=

ffi,

which gives the « optimal area » per head-group [1, 9].

For fluid hydrocarbon chains the optimal area should

be independent of the number of chains or the chain

length, as is indeed found experimentally [10, 12]

since the van der Waals attraction between chains

gives a very small contribution compared to these

two terms.

The optimal head-group area is a very important

parameter since (when it is taken together with the

chain volume and chain length) it determines the types of aggregates that are formed and many of their

physical and thermodynamic properties [1, 5, 9].

From the preceding paragraph it is also apparent that the head-group area ao is mainly determined by the repulsive forces between head-groups, since the attrac-

tive component y is more or less constant [7-9] while

the repulsive component c can vary greatly depending

on the nature of the head-group and the solution conditions. Let us therefore consider these repulsive

forces in more detail.

The repulsive forces between adjacent head-groups

within a fluid micelle, bilayer or monolayer, which

determine their head-group area, include electrostatic,

steric and solvation (hydration) interactions. Since all these interactions are occurring over very short distances (a few Angstroms, corresponding to the mean spacing between head-groups) their theoretical model-

ling is still at a rudimentary stage. Simple equations for

the lateral electrostatic (double-layer) pressure within the interfacial region of a single monolayer or bilayer

have been derived by Payens [ 13] and Forsyth et al. [ 14]

on the assumption of a smeared out surface charge,

and their interplay with the attractive interfacial force has been analysed by Parsegian [7] and by Wenner-

str6m and Jonsson [8]. The steric repulsion, which depends primarily on the sizes of molecules or mole- cular groups which may be modelled by hard spheres,

discs or cylinders [ 15] was first considered by Lang-

muir [16]. However, the important role of water of hydration in determining the effective head-group size, and/or its contribution to the inter head-group pair potential, does not appear to have been considered in the literature and is still unresolved.

Turning our attention now to inter-aggregate forces

we notice that the repulsive forces are fundamentally

the same as those between adjacent head-groups

discussed above, viz. repulsive electrostatic, steric and hydration forces. Consequently, one should expect a

correlation between intrabilayer and interbilayer for-

ces. For example, we might expect larger head-group

areas to be accompanied by larger repulsions between bilayers. This is borne out by experiments : the large hydration of the lecithin head-group results in a large

surface area of N 70 A2 as well as a large swelling in

fully hydrated lecithin multibilayers. By contrast the

head-group repulsion in phosphatidylethanolamines

(4)

is much less, which leads to a smaller head-group area

of 45-55 A2 and a much reduced swelling, viz. 10 A compared to about 25 A for lecithin [ 17,18]. We shall

consider other systems where such correlations occur

later. Here we proceed with a consideration of the three major types of repulsive inter-aggregate forces.

For simplicity we shall only talk about the forces between planar bilayers (while recognizing that the

same forces occur between micelles, vesicles, etc.).

The repulsive electrostatic double-layer force bet-

ween colloidal surfaces, lipid bilayers and across soap films is now well understood theoretically and it has

been measured in all three cases [19]. The importance

of steric repulsion between fluid bilayers is still

controversial [20], Huh [21] applied the theory of

Dolan and Edwards [22] to the interactions between oil in water microemulsion droplets where the pro-

truding hydrophilic head-groups on each surface

were considered to act as short polymer chains interacting via a repulsive osmotic (entropic) steric

interaction. Other steric contributions to the net

repulsion between bilayers would arise from local thermal fluctuations in bilayer thickness and head-

group area [23], and long wavelength elastic undu- lations of bilayers [24, 25]. Again, the effect of water of hydration in increasing the effective excluded volume of head-groups and in modulating the undu-

lation forces is to enhance their magnitude and

range [25].

Finally, we briefly mention the various theories that have been proposed to account for the strong exponentially repulsive « hydration >> force measured between lecithin bilayers [26]. The mean field theory

of Marcelja and Radic [3] later extended by Gruen and Marcelja [4] considers hydration forces in general as

due to the decaying polarization of water molecules

away from a surface, although a more realistic mole- cular dynamics simulation of this system [27] shows

that an oscillatory force is expected when the discrete molecular nature of water is taken into account Jonsson and Wennerstrom [28] proposed a highly specific theory (for lecithin bilayers alone) based on

the interactions of the zwitterionic (dipolar) head-

groups on one surface with their electrostatic images

« reflected >> by the other. Somette and Ostrowsky [24, 25] have extended a previous calculation by

Helfrich [29] and have derived general expressions

for the steric repulsion contribution arising from the

thermal undulation of elastic bilayers in the presence of other interactions. Note that these three very different theories each predict an exponential, or near- exponential short-range repulsion of magnitude and

range close to that measured, and we may also mention that the theory of Dolan and Edwards [22] also predicts

a short-range exponential steric repulsion between bilayers with freely mobile protruding head-groups.

(The one consolation for this state of affairs is that it is quite common in the history of science to find

completely different theories « explaining >> one obser-

vation

-

a phenomenon that is known as the « inva- riance of current theories to current experimental

results ».)

However, we can make the qualitative remark that this uniformity of exponential behaviour in all theories

only reflects the existence of a well defined correlation

length (not necessarily the same for all theories) leading unambiguously to exponential decaying cor-

relation functions.

We shall return to consider the likelihood of these interactions again at the end of this paper. Here,

we shall continue with a more detailed analysis of the origin and nature of the hydration force since (i) it is

the least understood and yet appears to be the most

important (at least in contributing to the short-range repulsion between hydrophilic head-groups), and (ii) we require some formulation that naturally links

up the hydration force between head-groups (within

a bilayer) with that between two opposing bilayers.

2.2 A SIMPLE MODEL FOR THE HYDRATION INTERACTION BETWEEN HEAD-GROUPS.

-

The physical origin of

this so-called hydration or structural force is not yet fully understood. It has long been recognized that

water is an associated liquid which has rather unique

solvent properties [30].

A qualitative concept, developed from the beginning

of colloid and surface science, underlines the impor-

tance of the modification of solvent structure near a

solute molecule. This perturbation propagates over

some distance from the solute molecule. When a

second solute molecule approaches, the total per- turbation of the solvent structure is changed, leading

to an indirect solvent mediated solute-solute inter- action.

A complete statistical mechanical treatment of

homogeneous water is extremely difficult (and all

the more so for inhomogeneous water) : indeed, the

structure of water is not determined by hard core

contacts and smooth attractive forces. Instead, water

constitutes an open structure akin to an imperfect

random tetrahedral network with oxygen atoms at the vertices and hydrogen atoms placed along the 0-0

bonds in configurations satisfying the ice rule, forming

the so-called hydrogen bonds. As a result, water forms

a structure where both position and orientation of

nearby water molecules are correlated. Theoretically,

one has thus to take into account both the transla- tional and orientational degrees of freedom and

analyse the way they are coupled

Some attempts have been made to derive general Landau-Ginzburg formalisms by expanding the

Ornstein-Zernike equation in the weak coupling limit

with the additional hypothesis that the interaction

potential between water molecules is short-ranged [31].

Marcelja et al. [31 have obtained an expression for the

free-energy density in terms of the density order

parameter. Realizing the incompleteness of this

approach, they have also proposed a generalized

(5)

Pople model for water focusing on an orientational order parameter, namely the average of the cosine of the angle measuring the deviation of either the lone

pair direction or of the OH bond from 0-0 line.

They also obtained in this case a Landau-Ginzburg expansion very similar to the density order parameter.

A similar formalism had been proposed previously by Marcelja and Radic [3] who considered the orientation effects of a planar membrane which polarizes the surrounding medium. This case may be simplified if

one assumes that the dominant effect of the pertur- bation of water structure is orientational ordering

of molecular dipoles and that the density variation

and dipolar polarization are decoupled.

The free energy expansion describing the solvent perturbation then reads

where Eo is the applied electric field, P

=

XE, D

=

sE and ~ is the corresponding correlation length.

This first approach has been subsequently put on a

firmer basis [4, 32] by showing that equation (1),

which implies a fixed local linear relation between the electric field and the dielectric polarization P, may be derived from a statistical ice-model of water with

Bjerrum defects.

The advantages of this type of approach are : a) The very difficult problem of determining the

translational and orientational correlation functions is replaced by the search for a continuous order parameter profile subjected to the condition that it minimizes the total free energy expressed as a Landau expansion. This LG expansion is extremely general

and has a wider validity than expected naively from

its derivation. It is essentially equivalent to the assumption of the Ornstein-Zernike form for the order parameter correlation function i.e.

where is the order parameter correlation length.

b) The unknown region where the interaction is strong (near a solute particle for example) which corresponds to a breakdown of the validity of the

weak coupling limit is subsumed in the boundary

conditions that one has to implement to the Euler- Lagrange equations derived from the minimization of the total free energy. Provided one can experi- mentally determine the boundary conditions (this

may be the difficult point) the problem is well-defined.

The continuous limit of the free energy density in

the LG expansion neglects the discrete nature of the solvent and should be ideally used only when relevant distances are larger than the molecular size. In this

limit, the variation of, say, an exponential sinusoidal

correlation function is averaged to an exponential.

However, even if the correlation length and molecular

separation are comparable we can still expect a qualitative description [31]. In the case of light solute particles or fluctuating fluid membranes, one expects

a smearing out of the molecular discrete structure of the solvent when averaging over time (or alternatively

over statistical configurations) with the result that the continuous description then becomes more rele- vant and appropriate.

We may then summarize the possible options for describing the solvation interaction between solute

particles as follows :

Conceptually, the notion of a disruption of water

structure by the presence of a solute molecule (or macroscopic body or surface) seems well established.

The difficulty remains in determining what is (are)

the relevant order parameter(s). The most likely

candidates may be the solvent density, the hydrogen

bonds density and the polarization density. The two

first order parameters are scalar whereas the last one

is a vector. This recognition already leads to an interesting conclusion : when applied to solute-solute

interactions, the LG formalism with a single scalar

order parameter can be shown to be unable to dis-

tinguish clearly between the case of non-polar and polar solutes, and always predicts an attraction [31].

One could argue that this results from the restriction to quadratic expansions of the free energy as a function of the order parameter as in equation (1). Indeed,

for large solute particles, it has been recognized that

the next quartic term, as well as a surface energy

density replacing the fixed boundary conditions usually imposed, are essential.

This surface free-energy which describes both the

change in coupling of the order parameter induced

by the presence of the surface and a possible surface

chemical potential, together with the bulk free energy

density, leads to the qualitatively correct description

of the very rich behaviour such as prewetting, first-

order and second-order wetting transitions [33].

However, it has been shown [34] that one always

expects an attractive contribution for the solvation interaction between like-solute particles in the case

of symmetric boundary conditions (i.e. identical on

the two solute particles). This result holds as long as

the scalar order parameters (say, molecular density

and hydrogen bonds density) are decoupled. This assumption is clearly questionable considering the highly correlated local orientational and translational order of liquid water. Indeed, it can be shown for the

case of two coupled, scalar, order parameters, that there exits suitable choices of the free energy density

and (symmetric) boundary conditions which lead to a

repulsive interaction.

Thus, within the hypothesis that the interaction between solute particles results from the disruption

of the density and hydrogen bonding of water mole- cules, we might conclude that the observed repulsion

can only emerge when fully taking into account of the

coupling between the two scalar order parameters.

(6)

The alternative description in terms of the vector polarization order parameter is richer since it embo- dies in a sense the density and hydrogen bonding description (to be precise, a full description of the

solvent is only attainable by defining all multipole

densities and not only the polarization profile) and

allows a simple treatment in terms of a single three-

dimensional order parameter.

In the following to be specific we use this repre- sentation. Let us consider a film of water bounded by

two infinite planes from which polar heads protrude

and occupy on the average the equilibrium area ao.

We assume that the planes are placed at the hydro- phobic-water interfaces and that the water molecules do not penetrate in the hydrocarbon core [35]. The

water disturbance can then be decomposed into two contributions, the hydrocarbon-water and the polar-

heads-water interactions. We assume for simplicity

and as a first approximation, that the structure of

water is mainly due to its interaction with the polar

heads. This hypothesis is clearly wrong when no polar

or ionic species are fixed on the surface as is the case

for hydrophobic surfaces immersed in water [36].

In this context, a new type of inter-aggregate force

has been measured, namely a rather long-range

attractive hydrophobic interaction which may be partly rationalized within the Marcelja’s type of models with a scalar order parameter and symmetric boundary conditions.

In the presence of polar or ionic species bounded

to the surfaces, we expect the electrostatic interaction to dominate.

We may now write the form of the generic total

Landau free-energy of the water solvent as :

where Vd is the water volume bounded by the two planes and the polar heads. Pj(r) is the j th component

of the polarization at r and f (P) is the free-energy density of a bulk solvent having a uniform polarization

P. The polarization profile actually carried by the

solvent minimizes Fd ~ P } and thus satisfies the Euler-

Lagrange equations :

This equation must be implemented by either a

surface energy term or surface boundary conditions.

Since we only want here to convey the main ideas,

we assume for simplicity that the polar-heads are spherical with radius R The more general case is

treated in the Appendix but the main qualitative

results are not changed by this hypothesis.

We further assume that the polar-head group- water interaction is modelled by a spherically sym- metric radial boundary condition (see Fig. 1). The polarization of a water molecule in contact with a

head-group is thus taken radial and of amplitude,

say Po (2).

Finally, as is often done in many-body problems,

we simplify the treatment by considering only two- body interactions and assume pairwise additivity of

forces. This last hypothesis is usually valid for large

distances or weak interactions but has been shown to

yield incorrect results for short distances, in a number

of cases [37].

Now, the interaction between two polar-heads is computed using a differential method. If the distance d between the two head-groups is increased by bd by introducing a fluid slice of thickness bd in the vicinity

of the medium plane between the head-groups, the

total free-energy of the solvent will change by the

Fig. 1.

-

Configuration of two spherical polar heads sepa- rated by

a

distance d + bd. The azimuthal angle T, not

represented,

measures

the deviation of the plane Sl rS2

with respect to the reference xOy.

(2) This approximation is valid

as

long

as

the « extra=

polation

»

length A is shorter than the correlation length ~.

(The extrapolation length is defined

as

the ratio between the coefficient of the gradient term and the coefficient of the

quadratic term of the surface free energy [44].) The physical

status of the extrapolation length

can

be clarified by relating

it to the range and anisotropy of the microscopic interactions in the bulk and

near

the surfaces [44].

When A is not small compared to ~, the physical behaviour

at surfaces is governed by the interplay of these two charac-

teristic lengths. One can then show that expression (14) for

the interaction energy between spherical head groups

(derived in the limit h - 0), is still valid if ~ is replaced by j

function of ~ and ,1,. For d - ~, ~ takes

a

simple form :

(7)

amount

which can be expressed as :

The notation Vd, Pd ... illustrates the fact that the volume of integration and the polarization profile

of the solvent are functions of

As shown on figure 1, we divide the volume into three parts, Vj, V2 and the slab of thickness 03B4d at x = 0

Then equation (6) reads

where x

=

0 denotes the position of the central plane. We have used Vd

=

V 1 + V2. The first integral can be

transformed using the fact that Pd +,5d(r) is only slightly different from Pd(r) :

The first term in the r.h.s. of equation (8) is nothing more than the functional derivative of Fd { P } with respect

to P at fixed volume Vd, and is zero as seen on inspection of the Euler-Lagrange equation (4).

The second term arises from the integration by part and the sign derives from the choice of the sense of the x-axis and recalling that the slab (3) is not in Yd. Equation (7) can then be written as :

Developing

allows to express 6F to first order in bd as :

The first term in the r.h.s. of equation (10) does not

contribute to equation (11) since it cancels out with the two contributions from x

=

± 6dl2.

The pressure resulting from the interaction between the head-groups is simply

and F is computed up to an additive constant by integrating equation (11).

The calculation now requires the form of the polari-

zation profile P(x, r 1-) and of its first derivative. This is done in the appendix taking a specific model for the

free energy density [31 ] :

and using a superposition approximation which

amounts to saying that P(x, r.1) and its first derivatives

are given by the sum of their values computed for

isolated head-groups. We finally obtain

where d is the distance between the two head-groups

(8)

and where A is given by

R is the radius of a polar head (Fig.1 ).

We can now use this formulation equation (14), for

the hydration force, to link up the solvation force between head-groups (within a bilayer) with that

between two opposing bilayers.

3. Lecithin multilayer interactions.

Let us consider the hydration interaction between two

parallel bilayers. Within our model, an important contribution comes from the interaction between

pairs of head-groups, one sited in one bilayer the

second on the other one. Using the assumption of pairwise additivity, the total interaction energy per unit surface can be estimated by summing all pair

interactions between a given head-group on one

membrane and all head-groups on the other one :

ao is the equilibrium head-group area and p and r are

defined in figure 2. Using d 2 + p2

=

r2, we obtain,

for the hydration interaction free energy per unit area

between two bilayers.

Now, we can estimate consistently ao by generalizing

the thermodynamic approach briefly recalled in § 2. l.

Arguing that the optimal head-group area ao is

determined by the balance between an attractive

hydrophobic energy and a repulsive energy which

Fig. 2.

-

Schematic drawing of two opposing bilayers

a

distance d apart whose head-groups occupy

a

mean (optimall

area

ao. Each head-group is assumed to interact with all

other head-groups via the

same

potential function of equa- tion (14) regardless of whether these head-groups

are

in the

same

bilayer or the opposite bilayer.

origin is mainly of hydration nature, we write the free energy per head-group as :

where a is the hard-core diameter of a head-group.

The integral in equation (18) gives the sum of all pair hydration interactions between all head-groups in a bilayer and the head-group at the origin. Its expression

in equation (18) corresponds to the zeroth order term in a development of the free energy in powers of the

fluctuating part of the density [31]. ao is then given by

which yields

we can now replace expression (20) for ao in equation (17) and obtain :

where

The corresponding pressure is

with

Equations (22) and (23) constitute our main result Note that the repulsion between bilayers does not depend explicitely on the amplitude A of the hydration

forces between head-groups, but implicitely through

the dependence of the range ~ on A (or conversely through the dependence of A on ~ as seen for example

in Eq. (15)). Indeed, since G ~ A (Eq. (17)) and ao

ao - (Eq. (20)), it is clear that G should not

depend explicitely on A. In other words, this result

qualitatively says that when A increases (due to ion binding for example), ao increases so that the surface

density of head-groups decreases leading to an unchanged inter-bilayer interaction energy, if we

neglect the change of ~. Note however that for our

approach to be consistent, we should have added the energy given by equation (17) to the free-

energy per polar heads. This gives a correction of

d-(1

amplitude (2 e ~ ) times the hydration force comput- ed in equation (18) which is negligeable as long as

-r is larger than a few units.

(9)

However for smaller separation, the coupling

between bilayers as well as the correction terms involv-

ing the higher order contributions in power of the

fluctuating part of the density should become relevant and equation (22) should no more be valid. Let us now

examine quantitatively equations (22)-(23). We can

extract an order of magnitude for the pressure ampli-

tude no by taking y

=

50

which yields 1011 dynes/cm’.

This is well within the experimental range [18].

Moreover, our model predicts that the pressure

amplitude is related to the hydration range ~ as shown

in equation (23). To test this feature, we have plotted

in figure 3, Log Ho as a function of 1 /~ which should

obey the following scaling form :

The grouping of the experimental points taken

from reference [18] on a straight line is rather remar-

kable considering the crudeness and level of approxi-

mation of the model.

The prediction (23) verified quite accurately in figure 3 thus presents strong evidence for the quanti-

tative correlation between the forces determining the

structure within a bilayer and those occurring between adjacent bilayers. This can be stated alternatively as

follows :

The decay length ~ varies from system to system in a way that correlates with the area per polar group.

However, in the Landau-Ginzburg theory presented

Fig. 3. - Plot of log 77o against I/j (where 1Io is in

dynes/cm2 and ~ in A) for eight different lecithins, egg PE, and lecithin-cholesterol mixtures, each in the fluid-state,

at 25 °C. The experimental data is from reference [18],

table I, which gives the repulsive hydration pressures

J7(J)

=

no e-"/4 for these lipid bilayer systems (as well

as

for three other lipids, two in the frozen state and one at 50 °C - not included in this plot). The slope of the straight line yields

an

estimate of the hard-core radius

(1 ’"

8 A.

in § 2.2, ~ is identified as the correlation length and

should be constant for a given solvent at a given temperature (at constant pressure). This apparent

inconsistency is clarified in note (2) where the ~ used in equation (14) is argued to be related to the true correla- tion length and also to another characteristic « extra-

polation » length, function of surface quantities.

However, explaining in details the precise origin

of the change of the decay length ~ with each lipid system, would take us outside the scope of the paper and does not seem feasible without a careful analysis

of the direct water correlation functions in the bulk and near the disturbing surfaces.

An idea, analogous to the one discussed in this

paper, has been put forward very recently in the

context of phase separation of micellar solutions [38].

Regarding intermicellar interactions as resulting from pairwise interactions between amphiphiles of different micelles the authors of reference [38] have been able to demonstrate the existence of a link between self- association and phase separation via low-order mo-

ments of the micelle distribution using a classical thermodynamic approach.

Let us finally mention that our development in

section 3 can be applied also in the presence of other types of interactions (and not only hydration) as long

as they are short-ranged (i.e. showing exponential decay) and that the pairwise additivity approxima-

tion holds.

4. Discussion.

In the preceding sections we have argued that there are

good theoretical reasons for expecting the properties

of individual amphiphilic aggregates to be correlated with the forces between the aggregates. We have also shown that for lecithin and mixed lecithin bilayers

this manifests itself as a correlation between the head- group areas of isolated bilayers and the repulsive

pressure between two bilayers, and further that a single

interaction potential between amphiphilic head-

groups can account for both. In this case the correlation

was surprisingly good (Fig. 3) and it is unlikely that

such quantitative agreement would occur for other systems. Since the possible presence of other relevant interactions of different and longer range (double- layers forces...) may break down the simplicity of the previous analysis. However, we should expect at least qualitative correlations between various aggregate

properties and inter-aggregate forces and we now

consider some of these.

(i) We have already noted (Sect. 2) the correlation between the head-group area and amount of swelling

of lecithins (PC) and phosphatidylethanolamines (PE). We should also expect that the adhesion of two PE bilayers is greater than that of PC bilayers. This

indeed is the case [23].

(ii) All the examples so far have involved zwitte-

rionic head-groups. There are many other examples

(10)

involving both cationic, anionic and non-ionic head- groups where a reduction in head-group area results in increased inter-aggregate adhesion (due to a reduction

in inter-aggregate repulsion). Thus the head-group

area of the nonionic lipid DGDG is greater than that of MGDG, which correlates with the shorter-range repulsion and stronger adhesion of MGDG bilayers [39]. Analogous correlations occur in anionic lipid bilayers where in general a decrease in pH or addition

of divalent cations reduces the electrostatic head- group repulsion and hence the surface area per lipid,

and also leads to reduced bilayers swelling in multilayer phases and to increased adhesion of vesicles.

(iii) Note that any general correlation between

head-group area and inter-aggregate adhesion is

particularly interesting for self-assembly considerations since smaller head-group areas usually lead to larger

vesicles or inverted micellar structures (for doubled-

chained lipids), or larger micelles (for single-chained surfactants) [1]. Thus, aliphatic alcohols which are

often used as co-surfactants in microemulsion systems have a very small head-area (ao 20 A2). This causes

oil in water microemulsion droplets to increase in size when alcohol is added. Recently, Pashley and Mc Guiggan (private communication) have measured the

adhesion forces between adsorbed surfactant bilayers

and found that the adhesion increases significantly as

more pentanol is incorporated in the bilayers.

(iv) As a final example, the aggregation number of

anionic alkyl sulphate micelles increases as the elec-

trolyte is changed from LiCl -+ NaCl -> CsCI [40].

This implies a reduced head-group area which probably

arises from the reduced hydration repulsion of the

bound counterions as we go from Li + to Cs + . This correlates with the increasing adhesion between oil-in- water emulsion droplets, stabilized by alkyl sulphate monolayers, as monitored by the contact angles

between them [41]. Similar correlations occur for CTABr and CTACI micelles.

We conclude that many apparently unrelated properties of amphiphilic systems are correlated (interdependent), and that these correlations have

some sound theoretical basis. In the absence of a more

rigorous general theory to describe these correlations

quantitatively we propose that this phenomenon be

called the Principle of Interdependence of amphiphilic

systems.

The Principle of Interdependence should help in testing the validity of theories of hydration forces

between surfaces, since any such theory should be

at least consistent with interdependence. We may note in passing that the theory of double-layer forces is inherently consistent in this regard [13]. However, it

is not yet clear whether the theory of Jonsson and Wennerstrom [28] or that of Somette and Ostrowsky [25] satisfy the Principle of Interdependence since

neither tell us anything about the lateral interactions between amphiphiles. On the other hand any theory

of the entropic (steric) interaction between head-groups

must obviously satisfy the Principle, as does the

extension of Marcelja’s theory discussed in section 2 and applied in section 3.

It is also possible that the hydration force between

head-groups involves a combination of some of these effects. Thus, it is very likely that the intrinsic hydration

force is oscillatory, possibly superimposed on a

monotonic repulsion [42], and that the thermal fluctuations of head-groups and bilayers smear out the oscillatory nature of the force-law leaving a purely

monotonic repulsion that is enhanced by the thermal fluctuations (i.e. a combination of hydration and steric

effects [20, 23, 25]. However, these interesting develop-

ments take us outside the scope of this paper.

Appendix A.

In this appendix, we give the main steps leading to equation (14) in the text. As seen from equation (11),

in order to compute 6F/64 one has to estimate first the

polarization profile P(x, rl) and its first derivatives for the problem of two spherical head-groups distant

from d imposing a radial uniform water polarization Po on their boundary.

We first solve the case of a single sphere for the particular choice f(p) = 1 P 1. This form for the free energy density yields a linear differential equation

which allows in a second step to solve the two-spheres problem using a superposition approximation.

The Euler-Lagrange equation (4) then reads :

which must be solved with the boundary conditions

where R is the radius of the head-group.

The problem being spherically symmetric, it is

convenient to use spherical coordinates.

In this basis, equation (A .1) reduces to :

where Pr is the radial component of P, the other two angular component being zero.

Integration of (A. 2) is straightforward and gives :

Using the superposition principle, the two sphere

problem is solved immediately and within this approxi-

(11)

mation the polarization profile is given by :

The notations are defined in figure 1.

We are now interested in computing surface inte-

grals of P2 and its derivatives in the mid-plane x

=

0.

For x

=

0, 81 = 02 and ri = r2, expression (A. 4)

reduces to

Let us first estimate the integral

Using

and

we obtain

I I I I

where we have kept only the leading term in power of

j /d

Evaluation of equation (11) also requires computing

surface integrals of derivatives of P.

As an illustration, let us compute for example

Using the equalities valid at constant p

we have

Equation (A. 9) allows to express in terms of r, p.

Proceeding as before, we end up with the result

Comparison of the results given by (A. 10) and (A. 9)

shows that a large distances (~ld 1) the contribution of the terms involving derivations are dominant. This is reminiscent of the planar problem [3] where indeed

only the gradient term contribute to the repulsive

interaction. Examination of equation (11) shows that

the third negative term under the integral in the r.h.s.

dominates over the first and gives an overall negative contribution which can be easely checked to yield amplitude of the form given by (A. 10).

Therefore, as expected from the continuation of the

planar case we obtain a repulsive interaction which

strenght scales as

For large

Références

Documents relatifs

Polyaromatic hydrocarbons (PAHs) Concentrations of PAHs (pulp deposit, 25.1 mg kg −1 of 15 PAHs in total; pulp deposit pore water, not measured) were above the effect concentrations

Interpretation of the results: In the numerical calculation of the stress coefficient concentration for the different species studied, that presented in table 3, we obtained

The search for a transient neutrino counterpart has been extended to the full sky with different energy thresholds for events originating from be- low and above the Antares horizon,

The doorway-mediated mechanism for dynamical processes represents the first step beyond statistical dynamics toward an explicit mechanism. A bright state →doorway state→dark

State tomography on the single-ion system of atomic and motional qubits requires a nontrivial set of operations, since a single qubit rotation on the motional qubit cannot be

Precipitation rate spectra as dependent on dynamic forcing: application to probabilistic

/ La version de cette publication peut être l’une des suivantes : la version prépublication de l’auteur, la version acceptée du manuscrit ou la version de l’éditeur.. For

Consistent with the Knudson model of tumor suppressor genes (Knudson, 1971), our work suggests that the loss of both Nfl alleles is a step toward tumorigenesis in