• Aucun résultat trouvé

Bending moduli of polymeric surfactant interfaces

N/A
N/A
Protected

Academic year: 2021

Partager "Bending moduli of polymeric surfactant interfaces"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00210874

https://hal.archives-ouvertes.fr/jpa-00210874

Submitted on 1 Jan 1988

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.

Bending moduli of polymeric surfactant interfaces

S.T. Milner, T.A. Witten

To cite this version:

S.T. Milner, T.A. Witten. Bending moduli of polymeric surfactant interfaces. Journal de Physique,

1988, 49 (11), pp.1951-1962. �10.1051/jphys:0198800490110195100�. �jpa-00210874�

(2)

Bending moduli of polymeric surfactant interfaces

S. T. Milner and T. A. Witten

Exxon Research and Engineering, Corporate Research Science Laboratories, Annandale, NJ 08801, U.S.A.

(Reçu le 24 mai 1988, accept6 sous forme définitive le 26 juillet 1988)

Résumé.

2014

Nous étendons notre théorie récente des « brosses » polymériques greffées en bout de chaînes à des polymères attachés sur des interfaces courbées. Plusieurs systèmes importants, par exemple les surfactants

polymériques ou les copolymères formés de deux blocs fortement ségrégués, peuvent être décrits comme des brosses. Par un développement de l’énergie libre d’une brosse sur une surface courbée en puissances de la courbure, on obtient des expressions analytiques pour les modules de courbure moyenne et de courbure

gaussienne. Les valeurs K et K pour des brosses monodisperses sont en accord avec les arguments d’échelle qui donnent, K, K ~ N303C35 pour les conditions de « fondu » et ~ N 3 03C37/3 pour des brosses de densité modérée. On traite également de manière analytique le cas important d’une brosse formée d’un mélange de chaînes longues

et courtes. On montre aussi qu’en remplaçant une faible fraction des chaînes longues par des chaînes courtes,

on réduit de manière spectaculaire les modules de courbure.

Abstract.

2014

Our recent theory of the free energy and conformations of end-grafted polymer « brushes » is

extended to polymers attached to curved surfaces. Several important systems, e.g., layers of polymeric

surfactants or of strongly segregated diblock copolymers, can be well described as brushes. By expanding in

powers of the curvature the free energy of a brush on a curved surface, the mean and Gaussian bending moduli

may be obtained analytically. Results for K and K of monodisperse brushes are consistent with scaling arguments, which imply K, K ~ N3 03C3 5 for melt conditions and ~ N3 03C3 7/3 for moderate-density brushes with solvent. The important case of a brush composed of a mixture of long- and short-chain molecules is also treated

analytically. The replacement of a small fraction of long-chain molecules in a brush by short chains is shown to

dramatically reduce the bending moduli.

Classification

Physics Abstracts

36.20E

-

81.60J

-

87.15D - 82.70D

Introduction.

Among the most basic and important physical properties of an adsorbed surfactant layer at a liquid-liquid interface, or of a bilayer composed of amphiphilic molecules, are the mean and Gaussian bending moduli. Much progress has been made in

predicting the phase behavior [1-5] and fluctuations

[6-9] in systems of self-associating amphiphiles, by parameterizing the monolayers or bilayers in these systems in terms of bending moduli and spontaneous

curvature. However, progress in understanding the origins [10-14] of these fundamental parameters has been somewhat less complete.

One origin of the elastic constants of amphiphilic layers is the interaction of the long hydrocarbon

« tails » of the molecules. These tails may be de- scribed with the language of polymer physics, a description which becomes increasingly good as the

molecular weight of the tails increases, i. e. , in the

limite of a « polymeric surfactant ».

Recently, progress has been made both exper-

imentally [15, 16] and theoretically [17, 18] is understanding the conformations and elastic proper- ties of a layer of polymer molecules attached by one

end to a surface at relatively high surface coverage.

The resulting structure, called a « brush » because the polymer chains stretch away from the grafting

surface to avoid high monomer concentrations, has

been studied both for the monodisperse case, [17, 19, 20] i. e. , when all the chains are of the same

molecular weight N, and for the case of arbitrary

molecular weight distributions [18].

A strong analogy may be drawn between the system of end-grafted polymer chains, and a surfac-

tant monolayer ; a bilayer may be regarded as two

such monolayers back to back. In this paper, we shall exploit this analogy, and extend our results on

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198800490110195100

(3)

polymer brushes to the case of curved grafting ’ I surfaces, in order to calculate the elastic properties

of the amphiphilic layers. The paper is organized as

follows. In section 1, a brief review of methods and results for monodisperse brushes is presented. In section 2, the bending moduli for a monodisperse

brush are calculated by comparing the free energies

of the brush in flat, cylindrical, and spherical geomet- ries. In section 3, we treat the important case of an amphiphilic layer composed of chains of two differ-

ent molecular weights

-

a model for the effects of a

short-chain « cosurfactant » on the properties of a layer of long-chain surfactant molecules. Recent exact enumerations of configurations of very short- chain molecules [14] observe that a small admixture

of shorter-chain molecules to a surfactant layer

reduces the bending stiffness with remarkable effi-

ciency. Using our analytic methods, we calculate a dramatic nonlinear dependence of the bending mod-

uli on the amount of cosurfactant added : a relatively

small amount of short chains serves to make the

layer nearly as flexible as a layer made only of short

chains. We conclude with a short discussion of

physical features left out of the present treatment, including interactions between the hydrophilic

« heads » of the amphiphiles, and the effects of the

(very) finite tail molecular weight. Details of the

mixed-brush calculations are relegated to Appendix A ; a prescription for converting our expressions for bending moduli from « theorists’ units » to physical

units is contained in Appendix B.

1. Review of « Brushes ».

Consider the problem of a polymer chains per unit

area attached by one end to a surface, and exposed

to a not-too-good solvent (such that the monomers

interact with one another through a mean monomer

density). A description of this brush would consist of the conformations of the chains, and thus such

quantities as the monomer density 0 (z ) at a height z

above the surface, the local monomer chemical potential V (z ), and the density of free chain ends (z).

Energy-balance [19] and blob [20] arguments have been used to show that, as a function of molecular

weight and surface coverage a, the thickness of this brush h scales as NQ ll3 and the free energy per unit

area F scales as No- 513

More recent work [17, 21] has shown that these

simple scaling arguments miss several important

features of the brush ; namely, 1) the conformations of different chains in the brush are not necessarily similar, nor is a particular chain uniformly stretched ; 2) the density profile, rather than being nearly a step-function as was suggested, [19, 20] is instead parabolic ; and 3) because the density profile goes

continuously to zero at the outer extremity of the

brush, the force to compress the brush slightly is

weaker (by one power of the strain) than calculated

using a step-function ansatz profile.

These results [17] were obtained by a solution of the one-dimensional self-consistent field (SCF) equations [22], which give a mean-field description [23] of the system valid for chains at high coverage

interacting through a sufficiently weak two-body repulsion. In practice this description should be valid for chains at moderate concentration in a not-too-

good solvent [24] ; for brevity this limit is referred to below as the « moderate density » regime. A closely analogous treatment was given earlier by Semenov [25] to describe the statistics of chains in a molten

block-copolymer system having a lamellar mesoph-

ase geometry. This situation corresponds formally to

a polymer brush in which no solvent whatever is present ; for brevity we below refer to this as the

« melt » regime. In both the moderate density and

the melt regimes, the solution of the SCF equations

makes use of the fact that the chains in a brush are

strongly stretched. This strong-stretching assumption

allows a « classical limit » to be taken, in which fluctuations of chain conformation about the most

probable path between its endpoints can properly be neglected [17, 25]. What remains is the many-chain problem of finding the self-consistent set of chain

paths, and chain-end positions with density E (z), in

a monomer chemical potential V (z). This V(z) is proportional to the average monomer density 0 (z) in the moderate density regime. In a melt, V(z) is fixed by the requirement that 0 = 1. The self-consistent solution is one in which 1) all chains

are in equilibrium in the potential V (z ), with configurations of the correct number of monomers ; 2) the monomer density cP (z) is reproduced from

the sum over the end density E (z ) of the contribution to the density d 0 of each chain.

This many-chain problem was solved exactly, for

both the melt and moderate density cases, by employing an « equal-time » argument [17, 25],

which we now summarize.

First we assume that the polymer chains are all of equal molecular weight, and that their free ends are

distributed with nonzero density at all distances from

the grafting surface, up to the « brush height » h, beyond which 0 (z > h ) = 0. (Detailed stability arguments are given in reference [17] to demonstrate that the chain end density is indeed nonzero

everywhere in the brush.)

Then the self-consistent chemical potential V (z)

must be such that a chain of molecular weight N may be in equilibrium

-

with no force applied to its free

end

-

with that end located anywhere in the brush

(at zo, say). There is then a precise analogy [23]

between the most probable path of the chain be- tween z = zo and z = 0, and the trajectory of a

certain classical particle. Under this analogy, discus-

(4)

sed in detail in reference [17], the arc-length of the

chain (roughly, the monomer index) corresponds to

« time », and the degree of local chain stretching to velocity. The fictitious particles move with equation

of motion

where the potential energy U(z) of the fictitious particle is related to the monomer chemical potential V (z ) by U (z ) = const - V (z ). (We define U(O) = 0, and assign V (h ) = 0 for equilibrium brushes.) Since there is no tension at the free end of

a chain, dz/dn = 0 there ; i.e., the particle starts (at

« time » zero, at point zo) from rest.

Then it is evident that for monodisperse chains,

the potential U (z ) must be an equal-time potential, i. e. , it must give the same « time of flight » (chain length) for a particle starting from rest at any distance away from the grafting surface. That is, ’

I

U(z ) must be a harmonic-oscillator potential, so that V (z ) = A - Bz. (See Fig. 1). The coefficient B then determines the period of the « oscillator » ; one quarter period is the length N of the chain. The

coefficient A may be determined e.g., by 1) insisting on a brush profile which accomodates the required

number of monomers per unit area ; and 2) requir- ing the monomer chemical potential to be zero at the

outer extremity of the brush, V (h ) = 0.

Finally, we may check our assumption that a positive end density c(z) exists which reproduces the

monomer density 0 according to the second self-

consistency condition. The contribution dO (z ; zo)

of a chain with free end at zo to 0 at a point z is the

inverse of its velocity [17]

The density is then the sum of contributions from all the chains, and may be written

This equation may be solved explicitly for e (z) in

both the moderate density and melt cases (see Fig. 1) ; in addition, general arguments may be

given which establish that E (z) > 0 in a brush grown

on a flat surface [17]. (This point is discussed further in Sect. 2).

In a subsequent paper, we explored the ever-

present effects of polydispersity in molecular weight

on the properties of end-grafted polymer brushes [18]. An understanding of these effects requires an

extension of the equal-time arguments of reference

[17] to a case where many molecular weights and

thus « transit times » coexist within a single brush.

Fig. 1.

-

The self-consistent potential V (z ) (solid curve),

and density of free chain ends s (z) for the case of a melt

brush (dashed curve) and a moderate density brush (dotted curve). The height z is given in units of the total brush height ho ; the y-axis units are arbitrary.

One characteristic feature of the polydisperse brush

is that the « equal-time » requirement forces the free ends of chains of different molecular weight to segregate in the z-direction : the self-consistent po- tential V (z ) has an unique time of flight (molecular weight) associated with each height z above the grafting surface. All shorter chains will have their free ends closer to the grafting surface than that of

any longer chain. This segregation is a general

feature of the « classical » long-chain limit for the

polydisperse brush [26].

We were able in reference [18] to construct a complete extension of our description of the mono- disperse brush to the case of arbitrary polydispersi- ty ; analytical expressions in terms of the molecular

weight distribution were obtained for the potential V (z) (and thus the density, via the equation of state), as well as the force required to compress a

polydisperse brush. In particular, explicit formulas

for the potential V (z ) were obtained for a « bimod- al » brush, composed of a mixture of two molecular

weights M and N. These results, used in the calcu- lation of the bending moduli of a bimodal brush, are quoted in section 3.

2. Monodisperse case.

A formal expansion of the free energy per unit area

of a bent surface (brush, membrane, or other thin sheet) may be written as [27, 28]

Here K and K are the mean and Gaussian rigidity

moduli respectively, cl, C2 = I/Ri, 1/R2 are the

local curvatures (inverse radii of curvature), and

co is the preferred or « spontaneous » curvature.

If we can compute the free energy of a brush bent

into the inside or outside of a cylinder or sphere, we

(5)

can extract K, K, and co. We write F, (R) (Fc (R)) as

the free energy per unit area of a brush assembled on

the surface of a large sphere (cylinder). (Throughout

this section, subscripts c ans s will refer to cylindrical

and spherical bends respectively). We write R posi-

tive (negative) if the brush is on the outside (inside)

of the bend (see Fig. 2). Fo is the free energy per unit area of the corresponding flat brush. Then

equation (3) implies

We may now compute the bending rigidity moduli

of a « monodisperse » brush, i. e. , one consisting of

chains of a single molecular weight. We consider a

bend performed with the constraint of fixed cover-

age, i.e., a fixed number per unit area of chain

attachments, denoted by cr. The grafting surface will taken to be the « surface of constant area » of the bend (see Fig. 2). (Either of these two assumptions

may be later relaxed, by adjusting the coverage and/

or area of the brush prior to the bend, to satisfy

some other physical requirement than fixed coverage and a grafting surface of constant area during the bend).

Fig. 2.

-

A brush bent with « positive » radius of

curvature (in this paper, R > 0 means the chains are on the

« outside » of the bend).

To compute the quantities F, (R ) and Fc(R), we imagine assembling the brush on a slightly curved substrate, grafting chains to this substrate at the

same coverage (chains per unit area) a present in the flat brush. We assume for the moment (this will be

examined below) that in the bent configuration, the

local monomer chemical potential V (z ) is still para-

bolic, V (z, R) = A (R) - Bz 2. This is equivalent (because of the « equal time » arguments of the introduction and Ref. [17]) to the assumption that

some free ends of the grafted chains may be found

anywhere inside the brush. The coefficient B must still be given by the « equal-time » requirement on

the chain conformations, i. e. , B = ir 2/ (8 N 2) so

that the chains of molecular weight N may reside in

equilibrim in V (z, R). The values of A (R) and h (R) must depend on the radius (and type) of bend,

as we will now show.

Consider first the case of a « melt brush », i. e. , on

in which the density is forced to be uniform. (This is perhaps a reasonable model for the hydrocarbon

tails of a surfactant layer). The height of the brush is

now given trivially by counting monomers ; in the flat geometry we have h = Na , while in a cylindrical geometry we have

with ho - No- and E = ho/R.

Because we require the monomer chemical poten- tial to vanish at the outer extremity of the brush, V (h (R), R ) = 0, we have A (R ) = Bh 2(R), so V is

now determined. We may now compute the free energy to assemble the melt brush on a cylindrical

surface by the same technique used in the flat case,

developed in reference [17]. The method, which calculates the free energy to assemble the brush by adding chains to an existing brush of lower coverage, may be briefly described as follows. The free energy increment to add a chain to a brush can be shown to be independent of where the free end of the new

chain is placed. There are chains with their free ends

arbitrarily close to the grafting surface, for which the free energy increment is purely due to the potential

V and is simply NV (o ) = NA. Hence we may write the free energy per unit area as

Application of equation (6) in the above case of a

melt brush grown on a cylinder gives

The analogous calculations in a spherical geometry give

Then equation (4) can be applied to extract K and

K for the brush as

(6)

A bilayer, composed of two such brushes back to

back, would have bending moduli twice as large, and

co = 0 by symmetry. (Appendix B outlines the

conversion of Eq. (10) into physical units).

We may also consider a moderate density brush, i. e. , grafted chains at moderately high coverage in the presence of a not-too-good solvent. As above,

we assume the monomer chemical potential of the

bent brush remains parabolic,

V (z, R) = A (R) - BZ2. In the moderate density

case, the potential V is proportional to the density,

V = 0 (we shall work in units such that the interac- tion parameter U is unity). The height h,,(R) of such

a brush when bent into a cylinder of radius R is again given by integrating the density profile in the bent

geometry,

With the explicit expression for 0 above, this implies

where in the moderate density case, we have

and E - ho/ R as before.

Now equation (6) may be used to give, in the

moderate density case,

The analogous calculations for the moderate den-

sity case in a spherical geometry give

Then equation (4) may once again be applied to

extract K and R for the moderate case as

Our results for the bending moduli of monodis- perse brushes are naturally consistent with a simple scaling argument for their magnitude ; namely, since

K and 9 have dimensions of energy, they should

scale as the free energy of a characteristic volume

ho of the brush [29], or as Fo h2 Recalling the scaling arguments given by de Gennes [19] (or examining equations (7, 9, 12, 14)) we observe F 0 ’" NO’ 5/3, ho - No-’13 in the moderate density case and Fo - Nu3, ho - No, in the melt case. These imply K and K scale as N 30,7/3 in the moderate density case

and N3 uS in the melt case.

Several previous authors, including references [1, 10, 14], have obtained estimates of the bending

stiffness of surfactant layers which are also consist- ent with these scaling results. Analogous scaling

results are contained in reference [1] for the case of a good solvent ; reference [14] presents a simple model

which results in K - N 3 cr 5for the melt case.

Other authors derive results which are equivalent

to the above scaling when combined with some

model of how the coverage a depends on N. For instance, in reference [10] the coverage cr is op- timized within a model in which a « bare surface tension » y is assigned to the surfactant interface, so

that a minimizes (F - y )/ o-. (This is not necessarily

a good model for determining a (N ) ; see Ref. [30]).

This implies 0’ -- N -315 for the moderate density

case. Together with the above scaling result for K,

K, this gives K, k - N 8/5 for the moderate density

case, which agrees with the scaling of the results

reported in reference [10].

The sign of the Gaussian modulus 9 is such that

bending into a « saddle surface » (with Cl = - C2)

costs free energy. This feature may be anticipated.

quite generally within a polymer model [31] of bending energy, by the following argument. Con- sider the volume V of a thin shell of height h above a

surface of area A bent with local curvatures cl and C2. The ratio VI(AH) is the ratio of volume available for chains in a layer of height h and grafting

area A in the bent and unbent geometries, and may be expanded as [32]

Then for a saddle surface, ci = - c2 = h/R and VI(AH) -- / 1 - 3 3 (hlR )2 the bent thin shells have

less room for monomers than the unbent shells, and

so the chains must stretch upon bending, which must

cost free energy.

The spontaneous curvature co may also be ex- tracted from equations (4, 7, 12) for the moderate

density and melt cases ; co is of order 1 /ho in each

case, as a scaling argument would suggest. This is, of

course, a strong hint that a flat monolayer of grafted

chains would prefer to break up into micelles.

However, the calculated value of co cannot be

identified with the actual preferred curvature, be-

(7)

cause the value was obtained in an expansion in

powers of hc about c = 0, the flat surface. When a

brush is bent a finite amount, two things occur which

cause the present calculation to break down.

First and most fundamentally, the strong-stretch- ing assumption fails when R - h ; this is because chains find so much free space far away from the

grafting surface, they no longer stretch to avoid high

monomer density. To treat properly the corrections to a scaling picture of polymeric micelles or « stars » [33] requires a major extension of the strong-stretch- ing methods.

Second, a difficult technical problem arises when

we consider a brush bent by any nonzero amount ;

namely, it is no longer true that the density of free

chain ends E(z) is nonzero everywhere inside the brush. A « dead zone » (region of z where no free

chain ends are found) opens up near the grafting

surface. We may estimate the size of this region by assuming that the potential V is still parabolic in a

bent brush, and computing in the bent geometry the

E (z, R ) required to reproduce the monomer density (0 = V for the moderate density case). This c (z, R ) will be negative in a small region near

z = 0, which will be approximately the region of the

dead zone.

For convenience, we define the end density e(z) in the bent (cylindrical) geometry to be the number of ends per unit (height x projected area).

Then, in the melt case, the requirement that E (z) reproduce the (uniform) density is, by analogy

to the discussion in section 1,

Assuming that V is a parabola amounts to writing U (z ) = BZ2; using this and A = Bh 2, equation (16)

may be expanded. It is convenient [18] to rewrite the

equations in « U space », defining E (U) d U =

= e(z) dz, as

This may be solved perturbatively in powers of 5 -= h/R for e (U), yielding the end density of a monodisperse brush bent into a cylinder as

The second term in equation (18) diverges at U = 0, i.e., at z = 0. However, if we ask for the root of E (z), we find z- h exp(2013 1/6 ) ; the dead zone is exponentially small in 8, and will not contaminate an

expansion in h/R. We need only contend with the

9

Fig. 3.

-

From reference [18], the self-consistent potential U(z ) (V = A - U ) for a bimodal brush, with M/N = 3/2 (solid curve). Below h, = 1, the potential is that of a

monodisperse brush of N-chains (dashed curve) ; above hl, the potential approaches that of a brush of M-chains

(dotted curve).

complications of a dead zone (which e.g., makes it

impossible to guess the form of V) when considering

« large » (nonperturbative) curvatures.

3. Mixing short and long chains.

The methods of treating strongly stretched polymers developed in reference [17] were extended in refer-

ence [18] to brushes of arbitrary molecular weight

distribution. Using some of the results of reference

[18], we can compute the bending moduli K and

K for the interesting and important case of a melt

brush composed of a mixture of short and long polymer chains. This may be regarded as a model for

the effectiveness of short-chain cosurfactants in

reducing the bending constant of a layer composed

of long-chain surfactants.

We begin with a brief summary of results for the unbent mixed melt brush. In reference [18], it was

shown that in a brush composed of a mixture of chains of molecular weights N and M::-. N, the free

ends of chains segregate in the z-direction (normal to

the grafting surface). The segregation results from the « equal-time » requirement : chains of two differ-

ent molecular weights, experiencing the same poten- tial V, cannot both be in equilibrium with their free ends at the same distance from the grafting surface.

The free ends of the shorter (longer) chains will be found at distances z -- z 1 (z :-.. z from the grafting

surface.

The self-consistent potential V (z ) for this mixed

brush, with a partial coverage a 1 of short chains and

o- 2013 o-i 1 of long chains, was shown in reference 18 to

be

(8)

where Ul = U(zl) is determined below. This may be inverted to find U(z) for z > zl as

with a = M/N -1 and BN M lr2/(8 N 2). (We have U(z) = BN Z 2for z : zl, just as for a brush made of

pure N-chains). This bimodal brush potential is displayed in figure 3.

It was also shown in reference [17] that the end density, converted to «t/-space» by c(z)dz=

E ( U ) dC/ (as proved convenient in the calculation of

E in a bent geometry in sect. 1), could be determined independently of polydispersity as

Using equation (21), the value of U(zl ), where

zl is the location of the boundary between the regions where short and long chain ends are found,

was determined by integrating the end density until

all the short chain ends are accomodated :

The value of A can be obtained in the same way

(replacing U1 by A and o-i by a in Eq. (22)). The

results [18] are

The relation between cr 1 and Ul may be conveniently given in terms of A as

Solving for Ul, we obtain

The free energy of the brush in a bent geometry

can be calculated in two equivalent ways : by directly assembling the brush on the bent surface, as was

done for the monodisperse brush in section 1 ; or by evaluating the work required to bend a flat brush,

discussed by Helfrich [34]. To extract bending con-

stants, one must work to 0 (c 2) in the first approach

and only to 0 (c ) in the second. The second method offers a welcome advantage in the more tedious case

of the mixed brush.

Consider a unit area of unbent brush. To deform it

so that it may become part of the inside of a

cylindrical shell of radius R, we must reduce the

cross-section C of the brush at a height z from

C = 1 to Cc (z, R) = 1 - z/R. When we do this, the height of the brush grows from the unbent value

ho = M((7- -al)+Nal to h (R ) given by

JOURNAL DE PHYSIQUE. - T. 49, N 11, NOVEMBRE 1988

equation (5), i.e., to h (R ) = Ao(l - E/2 +

E2/2 ... ), with E = ho/ (- R ), in order that the same

number of (incompressible) monomers be accomo-

dated in the reduced cross-sections. Work is done

during this process only [35] against the osmotic

pressure nc(z, R ), which is a function both a height

z above the grafting surface and the radius of the bend. Writing c = 1/7?, we may integrate with respect to c to find the change in free energy per unit

area in a brush bent into a cylinder as

The analogous construction for compressing a

brush into a section of a sphere of a radius R, for

which C s (z, R) = (1 - zIR)’ and hs (R ) is given by equation (13), yields

(In Eqs (26, 27), the signs have been chosen so that R > 0 corresponds to a brush on the outside of a

bend).

If we expand the osmotic pressure in equations (26, 27) to 0 (c ), and compare to the formal bending

energy expansion equation (3), we may obtain gen- eral expressions for K, K, and co. What remains is to relate 8H/ac I c = 0 in the cylindrical and spherical geometries. The result,

may be understood in several ways. Physically, the

factor of two arises from the fact that the cross-

section at a height z in the spherical bend is reduced twice as much (to 0 (c )) as for a cylindrical bend of

the same radius ; thus, the first-order change of the

osmotic pressure II(z ) from its flat-geometry value

is doubled. More formally, we may write for a

general bend H = H(Z, Cl, C2) ; then Hc(z, c) = II (z, c, 0) = II (z, 0, c ), and IIS (z, c ) = II (z, c, c ).

Differentiating with respect to c, equation (28) fol-

lows immediately.

-

Now we carry out the expansion of equation (27) (noting that H(h (c), c) always vanishes, we need

not expand the dependence in the limit of integration

of h(c)). The final result is

(9)

We see from equations (29-31) that K and Kco are properties of the unbent configuration only,

while K depends on how the osmotic pressure begins

to change when the brush is bent slightly. Noting the signs of equations (30, 31) (and assuming K > 0), we

see that brushes generically prefer to bend so that

the chains are on the outside (not surprisingly) ; also, we see that the sign of k generically opposes a

« saddle surface » bend. These conclusions depend only on the assumption of a bend at fixed coverage, and not on the molecular weight distribution. Huse and Leibler [9] have noted the importance of the sign of K for interesting phase behavior in am- phiphilic systems. The present polymeric model for

the origin of bending moduli suggests that positive .K, which may give rise to « plumber’s nightmare » phases [9], may be difficult to achieve.

We now proceed to evaluate equations (29, 30) for

the case of the bimodal melt brush. For a melt brush,

the osmotic pressure 77(z) is the same as the

monomer chemical potential V (z ), since monomers

are incompressible. Hence equation (30) for K may be evaluated, using II = V = A - U, with A and U(z) from equation (20) and equation (23). After

some algebra, the results are

with a = (MIN - 1) as given after equation (20).

Writing n =- N IM and 0 = o, 1 / o,, the quantities

h= (1-0)+no and hI = n (1 - (1 - p )2 )112 are

respectively the total height of the flat bimodal

brush, and the height of region where short chain ends are found in the flat bimodal brush (both in

units of Me-). The quantity in curly brackets equals unity for 0 = 0 and n 3for 0 = 1, thus reproducing

the monodisperse result of equation (10) in the appropriate limits.

Before examining this (somewhat opaque) result,

we shall compute the mean bending modulus K. If

we write equation (29) in U-space and integrate by parts, we get

The bimodal melt profile, equation (19), remains

valid in the bent geometry, with A(c) and Ul(c) replacing their flat-geometry values ; hence, z ( U, c)

is a function of c only through A (c ) and U1 (c). Thus

we may expand the innermost integral in equation (33) to 0(c) as

where M and 5 U, are the 0 (c) changes in A and Ul ; and z(U), az/aUl and A are evaluated for an

unbent brush.

The changes SA and 5 Ul could be calculated from

equation (22), which counts the free ends of the

chains, if we knew the end density E (z,.c) to 0 (c ) for the bimodal melt brush in the cylindrical geometry. The details of this procedure are con-

tained in Appendix A. The results are

with h again the total height of the flat bimodal

brush, as given after equation (32).

Combining equation (35, 36) with equation (34)

and inserting the result into equation (33) gives (after more algebra)

where we have defined n - N 1M, cp == 0’ II 0’, and

k - (1 - cp ) + n4,. (Appendix B describes the con-

version of equations (32, 37) into physical units).

The somewhat formidable-looking expressions for

K is plotted in figure 4 as a function of .0 for several values of n. (It turns out that the ratio K/K is

constant to within a few percent for all values of 0

and a wide range of values of M/N ; hence we have

not included a plot of K(cp) for various MIN).

Notice that the dependence of the modulus on the fraction of short chains present is far from an

interpolation between the bending moduli of a pure short-chain brush and a pure long-chain brush (dashed line). We may understand this without

appealing to the complete formula by considering

two limits : a long-chain brush with a few short chains replacing long chains, and the opposite case

of a short-chain brush with a few long chains

substituted for short chains.

In the case of a few short chains, if M > N, we can crudely think of the effect of shortening a few chains

as removing them, since the space occupied and free

energy per chain are both proportional to chain

length. Hence the largest effect of shortening a small

(10)

Fig. 4.

-

The change in mean bending modulus AK(O’)

upon substituting by chains of length M a fraction 0’ of the chains in a brush of N-chains (At > lV ), in units

of AK (1 for several values of n = N/M. Solid curve,

n = 1/10 ; dotted curve, n = 1/2 ; dashed curve, n = 9/10.

(The analogous curves for the Gaussian bending modulus

look essentially identical).

fraction 0 ’ == 1 - cp of the chains is the same as

reducing the coverage to o- (1 - 0’) in the monodis- perse long-chain brush. Recall that the monodisperse

melt brush bending constants scaled as o- 5 ; then in

the reduced units of figure 4, the slope of the curve near 0 = 1 should be five (rather than unity, as the interpolation would suggest). Indeed, if we expand equation (32, 37) around .0’ = 0, we obtain the first correction to the monodisperse M-chain value of

equation (10) as

For a small fraction 0 of long chains substituting

for short chains, we observe that the conformations of the long chains are very much like the most- stretched short chains (those extending to the full height ho), with an extra unstretched piece of length

M - N. These extra pieces are unstretched because

they are in a region of vanishingly small chemical potential V. Thus they make only a very small contribution to an integral of the osmotic pressure such as appear in equations (29, 30) for K and K. In fact, expanding equations (32, 37) around 0 = 0 gives

Any smooth curve connecting the slower-than- linear behavior at small 0 with the steep slope at

0 = 1 must deviate strongly from the linear interpo- lation, as does the exact result.

Recent studies of mixed short and long chains in a

melt brush [13], though only very small (N -- 20)

molecular weights are examined, show several of the

same features as the above analytic calculation. In

particular, reference [13] notes the qualitative impor-

tance of an admixture of shorter chains in lowering

the stretching energy of a brush made of long chains (see especially their sketch in Fig. 1). The less-

stretched long chains interact less violently when the

brush is bent; the reduction of K and 9 which results is expressed in equations (38, 39).

Conclusions.

We have employed an analogy between surfactant

monolayers and bilayers, and the well-characterized

polymer « brush », to obtain an analytic calculation of the bending moduli of amphiphilic bilayers. The

method becomes increasingly valid in the limit of long hydrocarbon tails on the surfactant molecules.

Our results for a monodisperse brush, correspond- ing to an amphiphilic layer of a single molecular weight, are consistent with simple scaling arguments for the dependence of bending moduli on coverage

(interfacial area per molecule) and tail molecular

weight.

In our polymeric model of the origin of bending stiffness, the Gaussian modulus in always of a sign (negative) which opposes bending into « saddle

surfaces ». This means that a special molecular interaction, perhaps between the surfactant

« heads » (see below), is required to produce a positive Gaussian modulus, which may give rise to interesting « plumber’s nightmare » phases [9].

Our methods break down when applied to sharply

curved interfaces, such as occur in wormlike or

spherical micelles. This results from the breakdown of the « strong stretching » approximation when the

bend radius is of the order of the layer thickness.

Thus, we cannot perform a reliable calculation of the spontaneous radius of curvature for surfactant mono-

layers.

The present model can be applied to the interest-

ing case of mixed long- and short-chain amphiphiles ;

the bending moduli for this system are found to depend nonlinearly on the fraction of short chains.

In particular, replacing a relatively small fraction of long chains with shorter ones serves to make the bilayer nearly as flexible as the pure short-chain

bilayer. This phenomenon is observed in preparation

of surfactant-cosurfactant mixtures used in the study

of fluctuating flexible bilayers [36]. We have given

no consideration in this work to interactions of

amphiphile « head groups » (which, by their aversion

to organic solvents, bind the amphiphile to the

interface). However, any specific model of head-

(11)

head interactions can be combined with the present work to give a complete description of a flexible amphiphilic layer. If the head-head interaction en-

tails a change in the area per molecule as the layer is bent, this can be incorporated by calculating the free

energy change upon bending in two stages : first change the area per molecule in the flat layer, and

then bend the layer at fixed coverage.

An important question (discussed in Ref. [17, 18])

is the size and nature of finite molecular weight

corrections to the fluctuation-free « classical limit »,

upon which our methods rely. Briefly, we may say that all quantities have corrections of relative order

Rglh - O(N-1/2); features of the local monomer

chemical potential V (z ), the positions of chains, etc.

will all be smeared out by this amount.

However, a direct investigation of small-N poly-

mer models of amphiphilic layers seems to indicate

that these fluctuation effects are not severe. Re-

cently, Szleifer et al. have performed exact enumer-

ations of configurations of short random walks attached at one end to a curved surface, and interacting with the mean monomer density. From

these calculations [14], the bending stiffness of

monolayers and bilayers were extracted. Even for such short random walks as five to 20 steps, the scaling dependence of the mooduli on molecular

weight, and the dependence of the bending moduli

in mixed bilayers on the fraction of short chains, are

in qualitative agreement with our analytic calcu-

lations. (Szleifer et al. observe several features, including a slight minimum in the bending moduli of

the mixed bilayer as a function of the fraction of

short chains, which they attribute to interdigitation

of the chains in the two monolayers forming the bilayer. Our work indicates [17] that interdigitation

effects should disappear as the molecular weight is increased).

Acknowledgments.

We thank S. Alexander, M. E. Cates, W. Gelbart,

P. Pincus, and S. A. Safran for helpful discussions,

and I. Szleifer for a stimulating discussion of his work prior to publication.

Appendix A.

Computing E (z, c).

To reproduce the uniform density in a bent (cylindri- cal) geometry, we use equation (16) together with

the bimodal potential profile of equation (19).

Writing the equation in U-space and expanding as in

section 1, we obtain for C/> U1

where 6 = Nac. A set of solutions of this integral equation for various left-hand sides was obtained

[25] and give

We would like to obtain two equations for the two

unknowns A (c) and U, (c) by counting chain free ends, as in equation (22). That is, we write

Rather than compute E (U, c) for U : Ul, we may argue as follows. The first term in equation (43) for 8 (U, c) for U:--. U, satisfies equation (16) for

U : Ul, i.e., with U = BN z2. The contribution from the second term (proportional to 8 (Af/W - 1) of e

in equation (43) spoils the equation, and must be

cancelled by a small negative contribution to E (U, c )

for U ,- Ui. The number of monomers cancelled out

by this negative contribution must equal the contri- bution of the second term in equation (43) to the

left-hand side of equation (16). Each chain removed

has N monomers, which makes it possible to count

the free chain ends inside z, as

where 81 I and E2 are the first and second terms in

equation (43).

Références

Documents relatifs

Bacchanales antiques L’horizon enflammé annonce dans le ciel L’ardente promesse d’une nuit sans sommeil.. La clairière cernée d'obscures frondaisons S’anime peu à peu

Keywords: salicylic acid, transport, phytohormone, defense response, systemic acquired resistance, plant-microbe interactions.. DISCOVERY OF

Diurnal variations of the Enph/VGLUT1interaction Using a GST fusion protein of the SH3 domain of EnphA1 (EnphSH3) we confirm the described interaction with VGLUT1 ( De Gois et al.,

Although these studies are valuable in assessing nontarget effects of Bt proteins expressed in plant tissues, they have limitations in their use for laboratory assessments on

ers of such weakly adsorbing homopolymers carrying a strongly adsorbing terminal monomer, in order to investigate the adsorption isotherm of these polymers and the possible existence

amphiphile going into discs (say, the 2-4fb of GM3 of the present case), is inversely proportional to the disc radius. Therefore « infinite » lamellae could be in equilibrium

For frustrated membranes with the intermediate surface, saddle bending moduli of finite values, but much smaller thon the cyhndrical modulus m magnitude, were obtained, the sign

1 J.C alais -Auloy, Droit de la Consommation, Dalloz, 2 éme éd, 1986,n° 25, p.36. 2 دﻘﻌﻟا" كﻼﻬﺗﺳﻹا دﻘﻌﺑ دﺻﻘﻳ تﺎﻣدﺧ وأ ﻊﻠﺳ نﺄﺷﺑ كﻠﻬﺗﺳﻣﻟاو