• Aucun résultat trouvé

Covariant elasticity for smectics A

N/A
N/A
Protected

Academic year: 2021

Partager "Covariant elasticity for smectics A"

Copied!
12
0
0

Texte intégral

(1)

HAL Id: jpa-00208301

https://hal.archives-ouvertes.fr/jpa-00208301

Submitted on 1 Jan 1975

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.

Covariant elasticity for smectics A

M. Kleman, O. Parodi

To cite this version:

M. Kleman, O. Parodi. Covariant elasticity for smectics A. Journal de Physique, 1975, 36 (7-8),

pp.671-681. �10.1051/jphys:01975003607-8067100�. �jpa-00208301�

(2)

COVARIANT ELASTICITY FOR SMECTICS A

M. KLEMAN

Laboratoire de

Physique

des Solides

(*),

Bât.

510,

Université de

Paris-Sud,

91405

Orsay,

France

and O. PARODI

Groupe

de

Dynamique

des Phases

Condensées,

Laboratoire de

Cristallographie,

Université des Sciences et

Techniques

du

Languedoc,

Place

Eugène-Bataillon,

34060

Montpellier,

France

(Reçu

le

13 janvier 1975, accepté

le 5 mars

1975)

Résumé. 2014 On présente une théorie des déformations statiques d’un smectique A, dans l’hypo-

thèse où les variations en

épaisseur

des couches sont faibles, mais en incluant la possibilité de larges déformations de courbure. Les variables

indépendantes

utilisées sont le directeur n(r) et la phase des couches 03A6(r)

(03A6(r)

est constant sur une couche

donnée).

Ces variables sont liées par une relation,

ce

qui

introduit un

multiplicateur

de Lagrange. Le calcul tient compte de la

perméation.

Les

équations

d’équilibre en volume et sur les surfaces s’obtiennent très simplement à 1’aide des variables choisies,

et on peut donner un sens

physique

au multiplicateur de Lagrange. On étudie en

application

la

stabilité d’une structure circulaire cylindrique. On trouve que les courbes cylindriques sont instables

vis-à-vis de certaines déformations hélicoïdales et sinusoidales.

Abstract. 2014 A

theory

of static deformations of smectics A is

presented,

taking into account the large curvature deformations of the layers, but assuming that the variation in the thickness of the

layers is small. The director n(r) and the layer

phase

function 03A6(r) (03A6 is constant on a given layer)

are chosen as independent variables, but they have to be related by a relation, to which corresponds

a Lagrange multiplyer. Permeation is taken into account. With the

help

of these variables, the bulk

and boundaries equilibrium

equations

can be expressed

simply,

and the

physical

meaning of the Lagrange

multiplyer

is

straightforward.

The formulation is

applied

to the case of the stability of a

cylindrical

structure. It is found that the cylinders are unstable versus certain helical and sinusoidal deformations.

Classification

Physics Abstracts

7.130

1. Introduction

(1).

- A number of attempts have been made to describe the

elasticity

of smectic

phases

on a sound basis. The difficulties are of various kinds.

l.l Smectic

phases

are lamellar

phases. They

can suffer very

large

deformations of the

lamellae,

while the lamella thickness remains very little disturb- ed. In that sense one must reach a

description

in

which the radü of curvature can show very

large variations,

while the local distortions of the thickness

can still be described in the framework of linear

elasticity,

as

long

as the local radii of curvature are

large compared

to the

layer

thickness.

(*) Laboratoire associé au C.N.R.S.

(1) This work has been partly done in order to satisfy the requi-

rements of a D.G.R.S.T. contract on defects in smectic phases.

1.2 Smectic

phases

are

mesophases,

and may be described

by

a director field

n(r) (n2 = 1).

In the

ground

state the directors which represent the direc- tion

of

the local molecule are

perpendicular

to the

layers (inside

the

layers),

and the

layers

are

planar.

The molecular

length

is

equal

to the

layer

thickness

do.

For any deformation of the

layer,

a

displacement

of

the molecules

through

the

layer

can be

superimposed, independently.

This process has been called

per-

meation

by

Helfrich

[1].

As will be shown

later,

the

only

molecular

displacements through

the

layers

which are to be considered are those for which n remains

perpendicular

to the

layer.

A first attempt, due to Oseen

[2],

insists on the

analogy

between nematic and smectic

phases :

his

theory

uses the director as the fundamental variable.

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01975003607-8067100

(3)

Oseen considers that the

layer

thickness cannot be

modified in any

displacement

of the

layers,

and this

leads to the condition

The

only

term

surviving

in the energy is therefore the

splay

term

The

Lagrange

condition curl n = 0

implies

that

Dupin cyclides (the shape

taken

by

the

layers

in the

classical confocal domains first described

by

G. Friedel

[3])

are solutions of the minimization

equations. Any

distorted medium has

only splay

energy and core energy

(along

the dislocations and

disclinations).

Permeation is of course forbidden in such a

theory

and molecules are restricted to stay inside the

layers.

The first

description

which includes a small varia- tion in the

layers’

thickness is due to de Gennes

[4].

De Gennes restricts his

theory

to the distortions in the

vicinity

of

planar

structures. The free energy now

reads

where u is the

layer displacement

and z the direction of the normal to the

layers

in their

ground

state.

Since n is

along

the normal to the disturbed

layers,

one

has,

in a linear

theory

This free energy has been used to discuss the distor- tions related to the energy of

edge

and screw disloca-

tions in

planar specimens [5, 6],

the free energy of

symmetrical low-angle grain

boundaries

[7, 8]

and

of

low-angle

conical deformations of

layers [9]

similar to those which appear

along

the focal lines in the confocal domains. All these calculations assume

that the distortions of the

layers

are small with

respect

to the

planar

situation. In this

theory,

permea- tion does not appear

explicitly,

but it is

implicit

that

permeation

relaxes the stresses exerted on the fluid

itself

by

the

displacement

of the

layers.

More

recently,

a new

theory

of smectic

phases [10]

(referred

there after as

MPP)

has

appeared

based on

a

general analysis

of

hydrodynamics

in terms of

conserved

quantities. Apart

from the classical

hydro- dynamic

variables

(mass,

energy and

momentum)

a new variable is

introduced,

which characterizes the local state of symmetry. A

general property

of conserved variables is that their fluctuation relaxa- tion time tends to

infinity

when the fluctuation wave- vector tends to zero. In that sense the new variables

can be considered as

quasi-conserved variables,

and

are treated as the other classical

hydrodynamic

variables. In the case of smectics

A,

the new variable is the

displacement

of the

layers.

Quite simultaneously,

de Gennes

[11] and

MacMil-

lan

[12]

have

given

a

phenomenological theory

for

the smectic A-nematic

phase

transition based upon

Landau-Ginsburg equations.

In these theories one uses

a complex

order

parameter Y’

related to the

modulation of the

density :

The

amplitude of §

vanishes in the nematic

phase

and characterizes the

degree

of smectic order. The

phase

describes small distortions of the

planar configuration.

Up

to now, except

perhaps

for the Oseen

model,

all these

approaches

considered

only

small deviations of the

planar

structure.

They

succeeded in

explaining

some

experimental

results such as

Rayleigh [13]

or Brillouin

scattering

in

planar configurations.

However non

planar configurations

are of outstand-

ing importance

in smectics A. Confocal domains can

be formed with

only

small variations of the

layers interspacing. Hence,

even for

largely

curved

layers,

linear

elasticity qualifies

for the

study of layer

thickness

variation. In other words it is necessary to have a

covariant

representation

of linear

elasticity. This

is the aim of the

present

paper.

In section 2 a covariant

expression

for the free- energy based on the

previous

considerations is

presented.

This

expression

has

already

been used

by

one of us

[14]

without any detailed

justification.

In

section 3

equilibrium

conditions are obtained

through

a minimization

procedure. Though

the director may appear as an

independent variable,

it is related to the

gradient

of the

phase

ç of the order parameter

through

a constraint which is taken into account

by

a

Lagrange multiplyer.

This

procedure,

which has

already

been used

by

one of us

[15],

allows for a

clear distinction between

permeation

pressure

(which

is the

thermodynamic conjugate

of

cp)

and torques.

In section

4,

this

general theory

is

applied

to the

expression

of the

free-energy

derived in section 2.

In this framework we discuss

planar

and

homeotropic boundary conditions,

the

possible

role of surface energy terms in thin

films,

the existence of

Dupin cyclides,

and

finally

the order of

magnitude

of the

variation of the

layer

thickness in confocal structures.

In section 5 we show that

cylindrical configuration

are unstable versus an helical deformation.

2. A covariant

expression

for the

free-energy.

-

In the former

descriptions

of smectics

A, only

small

deviations from the

planar

structure

(parallel

and

equidistant layers)

were studied. The

displacement

of the

layers

was then the

good

variable. If we want now to describe more

complex

structures, such as dislocations on

Dupin cyclides,

these

displacements

are no

longer

small. One has therefore to find another

(4)

starting point, and,

as in the case of

nematics,

the

good

one is the

phase

of the order parameter. Let us define it now.

As was

pointed

out

by

de Gennes the existence of

layers

results in a modulation of the

density

where

is the order parameter. Here qs = 2

nldo

where

do

is the lattice parameter of the

planar

structure.

At a

given

temperature and far from the N --+ A ,

transition, tf 1

is constant, and the

configuration

is

completely

described

by

the

phase (f’(r) :

the

layers

are surfaces

and their

interspacing

is d =

do/ grad

(p

.

The

planar

structure, with Z-axis normal to the

layers

corres-

ponds to

In smectics A a local minimum energy is obtained when the director n is normal to the

layers,

and when

the

layers spacing

is

equal

to

do,

i.e. when

Vep

= n.

Therefore one

expects

in the

free-energy

per unit- volume a term due to the

layers elasticity

where the tensor B has an axial symmetry around n :

Hence

To this energy must be added Frank’s curvature

free-energy :

In the above we have taken ç and n as

independent

variables.

However,

the

free-energy

is

changed

in

an uniform rotation of n - which means that the initial

configuration

is restored in a finite relaxation time. n is not an

hydrodynamic

variable

(a complete

discussion of this

point

is

given

in

MPP).

This does

not mean

by

itself that an

equilibrium configuration

could not be achieved with a tilted n

(n. grad

p *

0),

the increase

of fl being

balanced

by

a decrease of

fF.

Let us now discuss this

point.

Let v be the unit-

vector of

Vp

and

One would then have

Vôn is of the order of

1 /R

where R is a

typical

curvature

radius. Hence

Now,

far from the N --+ A and A - C transition temperatures, B ~

10’

cgs, K ~

10-’

cgs, which

means that

;’1. 10-’

cm, which is a molecular

length.

As a consequence, we can expect tilted n

only

in dislocation or disclination cores. In

fact,

as

BI,

and

Bl

are of the same

order,

one should

rather

expect

a nematic structure in these cores.

Since we are

dealing

with a

macroscopic

continuum

theory,

we can set

Here e is the relative dilatation of the

layers.

It is

easily

seen,

using (2.5)

that

The director is bound to be normal to the

layers,

and twist deformations are therefore

strictly

forbidden.

Let us now look at bend deformations.

Using (2. 5),

one obtains

Or

This term does not involve the director curvatures.

B 2 It must be

compared

to the

layers free-energy, B 2 II le

Ve is of the order of

elR,

where R is

again

a

typical

curvature radius.

Then,

since R is much greater than

A3

=

(K3/BII)1/2 (~ 10-’ cm),

one can

neglect

the bend term.

The

expression

for the elastic

free-energy

per unit- volume is now very

simple

This

expression

is very similar to that introduced

by

de Gennes

[4]

in the case of the deformation of a

planar

structure.

(5)

where u is the

displacement

of the

layers.

With our

notations,

we have for a distorted

planar

structure

and,

from eq.

(2.5),

On another hand this

expression

is valid

only

for

small layer

dilatations : the relative dilatation of the

layers is

As

long

as we deal with the case of linear

elasticity,

we can

neglect

terms in

e 3

in the

free-energy,

and

put e = y. For stronger

deformations,

one should have taken into account

higher

order terms in e

in the

free-energy expansion.

This means that the

form of the

free-energy

introduced

by

Bidaux et al.

[7]

for low

angle grain

boundaries is

strictly equivalent

to eq.

(2.6)

and

is,

as

pointed

out

by

the

authors,

valid

only

for low

angles.

The formal derivation of the

equilibrium

condi-

tions will be done in the next section. Two

slightly

different

approaches

could be used :

(i)

The

only

variable is

cp(r).

Then

and, using

the notation (p,i -

a (P ~xi

and the usual

summation convention over

repeated indices,

the

free-energy

per unit-volume is

given by :

In this

expression, pf

is not a

quadratic

function of

~,i and (pji, and this leads to non linear

equations.

The second

approach

is

simpler.

(ii)

One

keeps n(r)

and

({J(r)

as

non-independent

variables : the

free-energy

per unit-volume is

again

with the constraints

We now need to use

Langrange multiplyers

when

formulating

the

equilibrium conditions,

but these

equilibrium

conditions will take a much

simpler

form than in the first

approach.

Before

deriving

the

equilibrium conditions,

it is worthwhile to compare this

expression

of the free- energy with that used

by

de Gennes

[ 11 ] and

McMil-

lan

[12]

near the N H A transition temperature.

De Gennes introduced a

Landau-Ginsburg

free-

energy per unit-volume

for small deformations of the

planar

structure. Here

The z-axis is taken normal to the

layers

of the unper- turbed

configurations,

and ôn is the director

projec-

tion on the

X,

Y

plane. ( 1 j M)

is a tensor with axial

symmetry :

It is

easily

seen

that, using (2.11),

eq.

(2.10)

can

be rewritten as

This

expression,

which was first introduced

by McMillan,

is covariant. Far from the transition

temperature 1 If¡ 1

can be considered as a constant.

We then find (2.12) :

which is

exactly

eq.

(2. 2a),

with the

correspondence

3.

Equilibrium

conditions. - 3 .1 MINIMIZATION OF THE FREE-ENERGY. - We will use the second

approach

defined in section 2. The

free-energy

per

unit-volume, pf,

is a function

with the constraints

given by

eq.

(2.9)

which can be

rewritten as

(6)

where

Pi

is the

projection

operator on the

layers :

The

equilibrium

conditions will be obtained

by

a

minimisation of the total

free-energy

with

Here ,ui and 1

are

Lagrange multipliers.

Let us now

apply

the usual

procedure.

The

posi-

tion r of an

elementary

volume of the

fluid,

the

phase cp(r)

and the director

n(r)

are assumed to be inde-

pendent variables,

which is correct as far as we use

f instead of f

as a

specific free-energy. Starting

from

an

equilibrium configuration,

let us

perform

the

following

infinitesimal transformation :

Then the

specific free-energy f

becomes

And the new total

free-energy

is

given by :

From eq.

(3. 5),

one obtains

A standard calculation

gives

where

Then,

with the

help

of eq.

(3. 9)

and after a

partial integration,

eq.

(3. 7)

can be rewritten as

Here

bfs

are surface terms, uii is the stress tensor, h the molecular field

acting

on the

director, and g

the

permeation

force

(thermodynamic conjugate

to

the

layers displacement). Using

the

following

notations

One finds

Here v is the outer

normal, p

is the pressure with the usual definition

and the vector S has components

The bulk

equilibrium

conditions are therefore

Note that the

Lagrange multiplier A.

has

disappeared

from eq.

(3.14).

From eq.

(3.14c)

one

gets

0 :

3.2 THERMODYNAMIC POTENTIAL AND EQUILIBRIUM.

- We will now show that eq.

(3.14a),

which allows for the pressure

determination,

can be rewritten as

where y

is the usual

specific thermodynamic potential

Let us start from eq.

(3.12a).

Then

using

eq.

(3.13)

one derives

(7)

Using

now eq.

(2.9),

one has

since 0 is normal to the director n. Then eq.

(3.13)

can be rewritten as

Or, using

eq.

(3.16)

Taking

into account eq.

(3.14b)

and

(3.14c),

one

has,

for the

equilibrium

conditions

Another

interesting point

is that the stress tensor,

as defined

by

eq.

(3.12a),

is not

symmetrical.

As was

pointed

out

by MPP,

one can

always

define an

equi-

valent

symmetrical

stress. However from this non-

symmetrical

stress-tensor, one can derive a torque balance

equation relating

the

antisymmetrical

part of the stress-tensor, the torque due to the molecular field and the surface torques. In order to do that

we need

i)

to derive a space

isotropy relationship relating

the

partial

derivatives of the

free-energy

and

expressing

the invariance of the local

free-energy

in an overall infinitesimal rotation of the

physical

system and

ii)

to

give

a

physical interpretation

of the

surface term

bfs

defined in eq.

(3.12d).

3.3 SPACE ISOTROPY RELATION. - Let us

perform

an infinitesimal rotation.

Then,

in eq.

(3.6),

we have

where : eijk is the

antisymmetrical

rank-3 tensor

and 80 the infinitesimal rotation vector. In such a

rotation

pf

must be

invariant,

if there is no external torque

coming

for

example

from a

magnetic

field.

From eq.

(3.19)

and

(3.9),

we have

We therefore have

This

expression

must vanish for any value of 8w.

Hence we find the

space-isotropy

relation

It will be

easily

seen that this relation is satisfied

by

the

free-energy

defined

by

eq.

(2.8).

3.4 SURFACE TERMS. - Let us now look at the

physical meaning

of

bfs

defined

by

eq.

(3.12d).

There

are two terms. The first

is

clearly

related to

permeation.

From eq.

(3.13),

S has two components. The first component is

from eq.

(2.8).

It will be

important

for

layers parallel

to the surface. Let us assume n = v and e >

0,

which

means that we have a dilatation of the

layers.

Then

the total

free-energy

is lowered for

positive ’cp,

i.e.

by

nucleation of new

layers

on the

surface,

which tend to relax the

layers’

dilatation. On the other

hand,

for

negative

e

(contraction

of the

layers)

this relaxa- tion process is obtained

by

a surface destruction of

layers.

The second component

is dominant when the

layers

are normal to the surface.

It tends to make the

layers glide along

the surface in order to relax their

bending.

,

It must be

emphasized

here that S takes into account

only

the bulk

configuration

effects on the interface.

Depending

on the true nature of the surface there

are other terms that take into account surface effects.

There is a surface

permeation

force

S,

characteristic of the

surface,

and the

equilibrium

will be achieved for

Let us now look at the second term

It will be shown that this

corresponds

to a surface

(8)

torque. 8n is the variation of the unit-vector n in an

infinitesimal rotation defined

by

8w :

Then and

where

r

clearly

appears to be a surface torque exerted on the direction as a result of bulk effects. Here

again

we

must take into account the external surface torque r characteristic of the

physical

nature of the true

surface,

and the

equilibrium

conditions will then be

given by

The

interesting point

is that the

knowledge

of the

equilibrium configuration

allows the determination of S on

r,

and hence

of S

and

T.

3.5 TORQUE BALANCE EQUATION. - Let us start from eq.

(3.20)

which we can write

Using

eq.

(3.12a)

and

(3.13)

we have

From eq.

(2. 5),

Using

eq.

(3.12c)

and

(3.22),

we find from eq.

(3.23)

that :

This is the torque balance

equation.

We must

emphasize again

that this

equation

is valid

only

if

6, h and t are derived from a

purely

elastie free-

energy, .i.e.

including

no effect of external fields such as

magnetic

or electric fields.

From eq.

(3.23)

the

origin

of the asymetry of the

stress-tensor can be

easily

seen. In an

isotropic

medium

there is no elastic

body

torque. On the other

hand,

in a

liquid crystal,

the elastic

body

torque is the result of the torque due to the action of the molecular

field h on the director and of the surface torques

acting

on the director.

4.

Application

to Frank

free-energy.

- 4. 1 GENE-

RAL FORMULATION. - Let us now

apply

the

previous

results to the case when the curvature free energy

fF is,

as was discussed in section

2,

reduced to

We shall assume that

K,

is

independent

of the

density,

which is the standard

incompressibility hypothesis

made for smectics A. This

hypothesis

is

corroborated,

as far as we

know, by

the

high

velocities measured in first sound

experiments.

The pressure p however still appears in the

equations

we have obtained : p is now the

Lagrange multiplier

relative to the

constraint of

incompressibility (2)

As

explained above,

we have

dropped

in the Frank

force energy the term relative to

twist, strictly

for-

bidden on the basis of the existence of

layers,

and

the term relative to

bend,

since this latter term is

so small

compared

to the term of

splay.

We are then left with the

following equations

-- .

---- I,j -1 .. -1

There are no

simple

solutions of the

equation

for

equilibrium

div S = 0

and,

in

particular,

as we

shall see, the

Dupin cyclides

of the classical confocal textures are not, in

general,

solutions of this

equation.

However,

before

going

to

applications

in which we

shall take an

approximate equation

for

pf

and

div S =

0,

a few remarks on eq.

(4) (3

to

6)

should

be made.

4.2 HOMEOTROPIC BOUNDARY CONDITIONS. -

Consider a

sample

limited

by

a

boundary perpendi-

cular to the molecules

(homeotropic sample).

The

boundary

conditions

involving

the

stability

of the

layers

read

(see

section

3.4)

i.e.,

for v

along

n : e = 0.

(1) Let us notice that the condition div u = 0 is debatable in a

stratified médium ; div u = 0 is a condition which is natural in a true 3. D liquid, but the question arises whether a smectic phase is

from that point of view a true liquid, or a 2.D liquid, for which the

incompressibility condition reads rather

div( P. u) = 0.

A discussion of these topics is given in reference [16] for the case

of a membrane.

(9)

This represents a

layer

of constant thickness

equal

to the

equilibrium

thickness

do.

Also,

the

boundary

conditions

involving

the sta-

bility

of the fluid read

where F is an

applied force, necessarily

normal to the

boundary.

One notices that the

particular

form of

pfF

does not

play

a role in the

boundary

conditions in

homeotropic samples. According

to the set of

eq.

(4.3 to 6)

and div S =

0,

one can therefore fix at will the values of div n on the

boundary,

i.e. fix

the mean curvature

1 1

It is easy to show that if u1 + U2 is fixed on one

boundary

L, the distribution of the director n is

completely

determined. If one considers the

neighbor- ing layer

L’

(at

an infinitesimal distance

do), applica-

tion of Stokes theorem to the vector S

(on

a volume

limited

by

two small areas on L and

L’,

viz.

dSL

and

dSL,

and

by

the common normals to

dSL

and

dSL,)

enables us to calculate

E’,

since

(div n) L’

is known

by

the

knowledge

of L’. The process can be continued

on

adjacent layers L",

etc...

4. 3 THIN HOMEOTROPIC SAMPLES. - If the homeo-

tropic sample

is of

vanishing

thickness

(one

or a

few

layers thick),

and if one assumes that such a

thin film is

supported by

a

(closed) loop

of wire of

a

given shape,

the surface energy terms, up to now

completely disregarded,

will

play

a role in the deter- mination of the

shape

of the surface that the

layers

will take. It is reasonable to assume that the surface terms consist of the sum of a surface tension A

(inde- pendent

of the

curvature),

a term

proportional

to

((j 1

+

62)2,

viz.

B(ul

+

(j 2)2,

and a term

proportional

to the

gaussian

curvature ai 62

(3).

But one can show

that the term in (j 1 (j 2 is irrelevant in a

truly liquid layer, by

the

following

argument, taken from refe-

rence

[16] :

assume that Ul Q2 is an

hydrodynamic variable ;

one then has to consider as a

physical

fluctuation the

layer

fluctuations which

keep

Ul (J 2 constant. Under such a constraint the

layer

surface 1

is transformed to a surface E’

applicable point by point

on 1

(ex :

a

plane

transformed to a

developable).

Such a

peculiar

deformation can

always

be

relaxed,

if necessary, to a surface f of smaller

splay

energy than

l’,

in a

(short)

finite

time, by

viscous relaxation.

Hence Q1 Q2 is not an

hydrodynamic

variable.

If this is true, one is therefore

left,

for the

equation determining

the

shape

of the

surface,

with the

Laplace equation

(3) Such a term appears in the Frank and Oseen expressions of

the free energy of a mesophase (the K4 term in Frank’s original free energy).

If

Ap

=

0,

then 61 + U2 = 0 and therefore the

layer

is a minimal surface. Since e = 0

(4.7),

the bulk

energy of the

layer vanishes,

and the surface area

spanned by

the closed curve is that which minimizes the areas enclosed

by

the curve.

If

L1p ::/= 0,

ai + 02 takes a constant value over the surface area span

by

the curve.

4.4 SOLUTIONS WITH CONSTANT CURVATURE LAYERS.

- It seems

interesting, according

to the

previous considerations,

to look for solutions of div S = 0

consisting

of

layers,

of which all have constant cur- vature. Such solutions would

always imply

strong

singularities.

Let us

apply

Stokes theorem to the vector S for a volume

consisting

of two elements of

surface

dS,

and

dS2

on two different

layers,

linked

on their boundaries

by

the lines normal to the set of successive

layers

between

dSi

and

dSz.

One finds

or,

introducing

the

layer

thicknesses

dl and d2

This conservation law holds

along

the lines normal

to the set of

layers.

It means

than,

when d -

do,

dS becomes infinite.

The

simplest

solution to our

problem

is

evidently

when the surfaces of constant curvatures are

spheres

and

cylinders.

In that case, the

layers

are

parallel,

and one finds

according

to eq.

(4.10)

.for a

cylinder

e

= a/r

for a

sphere

e =

b/r2 ,

where a and b are constants.

Inspection

of the free energy shows

that,

in the

bulk, pf d V

is minimized for a = b = 0. But solutions with a and b different from zero must not be

disregarded.

4.5 PLANAR BOUNDARY CONDITIONS. - If the

sample

is limited

by planar boundaries,

i.e. such that the director is in the

plane

of the

boundaries,

which

implies

that the

layers

are

perpendicular

to

the

boundary,

the

boundary

conditions read :

- For the

layers

This

equation

means that the intersection of each

layer

with the

boundary

is a line

orthogonal

to the

set of lines div n = Const. in the considered

layer.

- For the

layers again

This

equation

expresses the fact that a torque has to be exerted in the direction

parallel

to the intersection

(10)

of the

layer

with the

boundary

to compensate for the surface torque.

- For the fluid

There is a force exerted on the surface

boundary

on

the molecules.

Except

for the case when the

boundary

is

planar,

this force does not reduce to a normal pressure.

4.6 DUPIN CYCLIDES. - Are

Dupin cyclides

solu-

tions of the

equations

of

equilibrium

div S = 0 ?

Dupin cyclides

in a confocal domain are

parallel

surfaces. One therefore has

which reads

If one now expresses div S = 0 as a function of

e

and g

one notices that one obtains an

equation

of

the

type

where 0 is a linear operator

acting

on g. Since a

is a function

of g only (and

not of r : e is a constant

on a

given layer),

eq.

(4.16)

has solutions

only

if

0(g)

is a function of g. This occurs

if g

is an

eigen-

function of the operator 0. We therefore are led to infer that

Dupin cyclides

are solutions of div S = 0

only

for a discrete set of values of g. Such a result has in fact been reached

by

Kleman

[16],

who has

shown

that,

for a confocal domain defined

by

its focal

curves,

only

one

Dupin cyclide obeys

the

equilibrium equation.

It is also shown in reference

[16]

that this

cyclide

has no cusp

points

and is of the first

species (in

the sense of

Bouligand). Hence,

in a

given

confocal

domain,

all the

layers

are

(slightly)

different from

Dupin cyclides,

except one.

4.7 ORDER OF MAGNITUDE OF e. - A crude eva-

luation of E can now be

given :

div n is of order

1/R,

the mean radial curvature ;

Pij Vj

div n is of order

1/RL

where L is a characteristic transverse

length

for the variation of

R,

and div 0 is therefore of order

Kl/ R 2

L. In the same way diB

(Ben)

is of order

Be/R.

Hence the

equilibrium

condition leads to

where

Ac

=

(K1/B)1/2

is the coherence

length.

For a smectic

A, Ac

is of the order of a molecular

length,

except near the upper transition temperature.

Near the focal

lines,

L can be very small

but,

except in the core, if is greater than

Â..

Then 8 gg

Âc/R.

Taking Â,,, - 10-7, R ~ 10-4,

we have 8 gg

10-’.

On the other

hand,

for

cholesterics, Ac is

of the

order of the

pitch

qo, which means that s is no

longer

small. This can

explain

the strong deformations of focal structures that have been observed on

cholesterics.

5. The limit of small s.

Stability

of

cylindrical

solutions. - 5.1

QUASI-PLANAR

LAYERS. - In the

limit of small a, the small distortions from a

planar

situation are described

by

the same

equations

as

those used

by

de Gennes

[4].

One considers a function (p

given by

lp.

= z -

u(r)

where u is a finite

quantity,

which varies

slowly.

Its

derivatives are small

quantities.

One

finds,

when

keeping

the second order terms in those derivatives

from which one obtains E and n.

hence one obtains the same force energy as de Gennes’

but one notices

that,

if one

keeps

third order terms in the B term

(as

has been done

by

Clark

and

Meyer [17]),

it is necessary to include terms of the

same order in the

K,

term.

5. 2 CYLINDRICAL SAMPLES. - Consider now a

stacking

of

layers

in concentric

cylinders.

We want

to

investigate

the

stability

of such a

system

in the

vicinity

of the solution e = 0

(layers

of constant

thickness).

Let us define

(11)

Calculation of e and n up to the second order in u

yields :

where

div

where

Building

now the free energy up to the second

order,

we notice a remarkable différence with eq.

(5.4).

The term in B is

basically

of the same form

but the term in

Ki

now contains

quadratic

terms of a

complete

different nature. We have

This

quantity

can be

negative

and

gives

rise to

instabilities of the

cylindrical solutions,

for some types of fluctuations of u. In order to

investigate them,

we first write the extra free energy

b(pf ), drop- ping

the terms which add up to surface terms since their presence would

just

shift the energy

by

a cons-

tant amount of little

physical meaning

as

long

as the

boundaries conditions are not taken into account.

We have

u

obeys

the

Euler-Lagrange equation

We

investigate

différent solutions of

(5.9),

whose

difference with

(5.4)

is

striking.

5.2 .1 Solutions

depending only

on 0 and z. -

They

are

necessarily

of the type

where

do

is the

layer

thickness at

equilibrium,

and S

is an

integer

in order for u to be consistent with the

periodicity

of the

layers

after 0 has

changed by

an

angle equal

to 2 n, u therefore

represents

a kind of dislocation

along

the axis of the

cylinder :

the

layers spiral

about the axis

(at

constant

z),

but if one now

considers the variation

along

z, the

spiral

rotates

about the axis with a

pitch equal

to p

= 2 n/q.

This

helical variation means that an extra

layer

of thickness

do

is added

along

the axis after a variation of z

equal

to p.

Eq. (5.10)

represents

only

the

beginning

of the

instability ;

but the

symmetry

which is

displayed by

this

equation clearly

indicates that a helical distortion of the core from the

cylindrical

disclination we have started from will take

place.

Let us calculate the variation

b(pf )

in energy due to

(5.10).

This variation is

always negative.

We

find

(per

unit

length

of

line)

where rc

is a core radius and R a

(large)

cut-off radius.

This

expression

is valid as

long

as the terms of third

order which we have

dropped

are

negligible.

This is

certainly

true if

do q « 1. q

is therefore limited.

Exp. (5.11)

would suggest that S is as

large

as

possible,

but the core

always gives

a

positive

contribution to the energy

which,

when balanced with

(5 .11 ),

should

fix the

optimal

values of

S. q

is related to that same

balance and also to the size of the

specimen.

Accord-

ing

to a calculation

by

Durand

[18] concerning

the

effects of fluctuations of size 2

nlq along

a

layer,

it is

reasonable to

relate q

to the

specimen

size

by

the

relationship q2 Â,,

R - 1.

5.2.2 Solutions

oscillating

with z. -

They

are

of the form

where m is an

integer.

We have

Références

Documents relatifs

- Phase diagrams of binary systems in which a significant lowering of the smectic phase stability is observed : (a) and (b) mixture of compounds yielding smectic

Abstract 2014 The effect is tested of the smectic layer spacing ratio, r, on the phase diagram for the binary systems consisting of 80CB (smectic Ad) and one of

qZ nevertheless dominates over the Bragg signal since the relative Bragg to small-angle intensity goes as d- 3/2 Nevertheless, that effect is much weaker than in

ing the smectic as incompressible and neglecting inertial effects. Consider for instance the vicinity of the upper plate : we have a vertical pressure drop of FIR* taking

- Proton magnetic resonance experiments [1] parallel and perpendicular self-diffu- sion coefficients were measured in smectics A-TBBA and EBBA (D~~ is related to

Then we use this decomposition technique and estimates to justify a nonlinear bending model for elastic shells for an elastic energy of order δ

Bamler to study the behavior of scalar cur- vature under continuous deformations of Riemannian metrics, we prove that if a sequence of smooth Riemannian metrics g i on a fixed

The stereo kernel [6] which compares the minimal stereo subgraphs of two ordered graphs has one drawback : graph information are reduced to a bag of subgraphs without taking