• Aucun résultat trouvé

Flow alignment of nematics under oscillatory shear

N/A
N/A
Protected

Academic year: 2021

Partager "Flow alignment of nematics under oscillatory shear"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00247914

https://hal.archives-ouvertes.fr/jpa-00247914

Submitted on 1 Jan 1993

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.

Flow alignment of nematics under oscillatory shear

A. Krekhov, L. Kramer, A. Buka, A. Chuvirov

To cite this version:

A. Krekhov, L. Kramer, A. Buka, A. Chuvirov. Flow alignment of nematics under oscillatory shear.

Journal de Physique II, EDP Sciences, 1993, 3 (9), pp.1387-1396. �10.1051/jp2:1993209�. �jpa-

00247914�

(2)

J. Phys. II £Fance 3

(1993)

1387-1396 SEPTEMBER1993, PAGE 1387

Classification Physics Abstracts

61.30G 47.20

Flow alignment of nematics under oscillatory shear

A.P. Krekhov

(~),

L. Kramer

(~),

A, Buka (~) and A-N- Chuvirov

(~) (~)Institute

of Physics, University of Bayreuth, D-95440 Bayreuth, Germany

(~)Physics

Department, Bashkirian Research Center, Russian Academy of Sciences, 450025 Ufa, Russia

(~) KFKI-Research Institute for Sohd State Physics, Hungarian Academy of Sciences, H-1525 Budapest, P.O.B.49. Hungary

(Received

1

Aprfl1993, accepted

2 June

1993)

Abstract. Using a simple averaging method we show that under plane oscillatory shear

a

net torque acts on the director of a nematic

liquid

crystal leading to unexpected flow alignment.

Under quite general conditions the usual flow-alignment orientations are unstable, and the ori-

entation parallel to the shear

(perpendicular

to the

velocity)

becomes most stable within the shear plane. The effect drops out exactly under

(unrealistic)

linear shear unless the director has

a nontrivial space dependence. An explicit application to plane Poiseuifle flow is

presented.

Pre- liminary experimental data are shown which demonstrate a transition under planar oscillatory

shear from a homeotropically aligned nematic to a spatially homogeneously deformed state.

1. Introduction.

The

coupling

between flow and director is one of the basic features of

liquid crystals.

Thus in a

steady velocity

field

u(z) along

the x axis

(plane

shear

flow)

the director of a nematic will tend to

align

in the x -z

plane

at an

angle

off = + tan~~

(a31a2)

with the x axis if a3

la2

> 0 and the shear

8u/8z

is

positive/negative

(a3> a2 are Leslie

viscosities) ill.

The

respective

orientation is the

only

stable

solution,

the one with the other

sign representing

an unstable

equilibrium.

The

only

other unstable

equilibrium corresponds

to the orientation

along

the y axis. A number of

interesting phenomena

result from this

flow-alignment

effect

(see

e-g- 11, 2] and references

therein).

In most nematics

a31a2

is small but

positive,

in some

materials,

however, one can have

a3/02

< 0

(especially

near the transition to a smectic

phase)

and then instead of flow

alignment

one has a

tumbling

motion which has also attracted much attention [3].

When the

velocity

field oscillates

periodically

and

symmetrically

around zero the situation is different. Then for the flow

alignment

case,

a31a2

>

0,

and a linear

ii-e- constant-velocity-

gradient)

shear

flow,

the director oscillates between two

positions

that

approach

the

alignment

angles

+

tan~~(a31a2)

as the oscillation

amplitude

becomes

large

[4].

However,

one should

(3)

note

that,

at least for not too

large amplitudes,

so that instabilities can be

excluded,

there exists in

principle

an infinite set of

possible (oscillatory) trajectories depending

on the initial director orientation. Thus there is no real flow

alignment.

Of course this can

apply only

when there is no other torque

acting

on the director

(e,g.

surface

effects)

and therefore the statement is difficult to

verify experimentally.

For the non-flow

alignment

case the situation is rather

similar,

except

possibly

at

large amplitudes

[4].

Actually

this

degeneracy

is

peculiar.

One has to

postulate

a

large degree

of

reversibility

in this

dissipative

system, since the director is

presumed

to return after each

period exactly

to its

previous position.

The purpose of this communication is to show that the above

picture

is

only

an

approximation

and that in

general

there does indeed exist a net torque so that there is in fact real flow

alignment

even under

oscillatory

shear. The effect is contained in the standard

hydrodynamic description

of nematics [5, 6]. It cancels

exactly

under linear

shear,

I.e. when au

/8z

is constant

(at

least when the director is assumed to be space

independent).

In a time-

dependent

situation a

strictly

linear shear is unrealistic.

Actually

a

time-averaged

net torque is

responsible

for the

pattern-forming

instabilities observed in thin nematic

layers subjected

to

an

elliptic oscillatory

shear

ii,

8], and was included in the theoretical treatments of this effect [9]. We mention in

passing

that the linear response of nematics to

small-amplitude oscillatory

shear flow with

rigid

surface

anchoring

of the director was considered

recently

[10].

Experiments

under the

plane oscillatory

shear flow considered here show

instabilities,

which up to now have remained

essentially unexplained.

A transition to a

spatially periodic

roll

pattern oriented

perpendicular

to the shear with

period

of order of the cell thickness was found

by

Scudieri in MBBA

ill].

The

frequency

was 7.5 kHz and the

alignment homeotropic.

Apparently

the roll

instability

was

preceded by

a

homogeneous

transition

leading

to "biaxial-

type"

behavior. With the same material and

alignment,

but at much lower

frequency (500 Hz),

a transition to rolls that are

parallel

to the shear was observed at

considerably higher

oscillation

amplitudes by

Guazzelli [8]. In this work the

plane

shear was

just

a

special

case of

elliptical

shear. With small deviations from

planarity

the threshold first increased, but then there was

apparently

a crossover to the usual

elliptic

shear

mechanism,

after which the threshold

decreased. At very low

frequencies (20-100 Hz)

and rather

large amplitudes

with

planar

and

homeotropic alignment

and materials that included non-flow

aligning

ones a

variety

of effects

were observed [4]. The

periods

of patterns were much

longer

than the cell thickness and the orientation

predominantly oblique. Finally,

at ultrasonic

frequencies

a transition from the

homeotropic prealignment

to a state with a

planar

component of the

director,

which rotates

slowly

and

develops approximately axisymmetric

wave

phenomena ("autowaves"),

has been

found [12].

In the next section we present some

preliminary experimental

observations of a

homoge-

neous transition under

plane oscillatory

shear which

similarly

to the roll patterns

presumably

appearing

for

higher

shear

amplitudes require

the action of net torques. Further on we calculate the net bulk torque, for

simplicity dealing only

with motion in the x z

plane.

In section 3 we present our

general averaging

method and

apply

it in section 4 to the

simplest

type of

prescribed

nonlinear

shear, namely small-amplitude plane

Poiseuille flow. In section 5

we comment our results and

give

an outlook for future work. Some detailed

analytic

formulas

are

presented

in the

appendix

A.

2.

Experiments.

In this section we want to demonstrate

qualitatively

a transition under

planar oscillatory

shear from a

homeotropically aligned

nematic to a

spatially homogeneously

deformed state.

Exper-

(4)

N°9 FLOW ALIGNMENT OF NEMATICS UNDER OSCILLATORY SHEAR 1389

~)

3

' ' ' '

', ', ',

' , ',

,', ' '

' , , ,

' ' ,'

Fig-I-

Time resolved measurements of the zero level of the optical

signal ii),

the acceleration of the upper glass plate

(2),

and the optical response signal in arbitrary units

(3).

iments have been carried out with two nematics, EBBA and 7CB. The cell was

horizontally placed

on a

thermostatting block,

the

height

and orientation of which could be

adjusted

with micrometer screws. The lower

glass plate

of the cell was fixed to the

heating

block while the upper one was

rigidly

attached to the membrane of a

loudspeaker.

Great care was taken to avoid

ellipticity

of the

oscillatory

motion. No spacers have been

used,

the

glass plates

were

adjusted parallel

to each other and the distance

(sample thickness)

has been measured. The

sample

was

placed

into a

polarizing microscope

combined with a

photodetector

and the inten-

sity

of the

light passing through

the

sample

was recorded. An accelerometer was attached to the upper

glass plate

in order to measure the mechanical vibrations.

In

figures

la-c we show time resolved low

frequency

measurements at three different vibration

amplitudes.

In each

picture

curve I is the "zero level" of the

optical signal

without

vibration;

curve 2 shows a

signal proportional

to the

oscillating

acceleration of the upper

plate

and 3 is

the

optical

response to the vibration. In

figure

la the vibration

amplitude

is

small,

the

optical signal

follows the vibration with double

frequency.

Zero acceleration which

corresponds

to zero

displacement

coincides with zero

intensity

of the

optical signal.

This shows that the director

(5)

(optical

axis of the

system)

reorients

periodically

with the external

frequency, symmetrically

around the initial

homeotropic alignment angle. (The

small asymmetry of the

optical signal

is

probably

not

relevant,

it could arise from an

imperfect sample adjustment). Figure

16 shows the situation after a transition to an

asymmetric

state occurred at

higher amplitude

of the shear. Director reorientations

corresponding

to

positive

and

negative

directions of the

displacement

are not identical any more. The

optical signal

has now the same

frequency

as the external shear and moreover a small

phase

shift appears between the two

signals. Figure

lc

corresponds

to a further increase of the vibration

amplitude.

The asymmetry and

phase

shift of the

optical signal

is even more

pronounced.

The threshold field for the onset of the transition was measured in a

frequency

range of10-600 Hz. For both EBBA and 7CB we found a linear

frequency dependence

of the critical

velocity,

which means there is a

frequency independent

critical

displacement amplitude

at which the transition occurs.

Typical birefringence

measurements where the

sample

is

placed

between crossed

polarizers

which make an

angle

of 45

degrees

with the shear

direction,

have also been carried out. Both substances showed similar behavior. The

birefringence

started at the critical vibration

ampli-

tude and reached a

frequency independent

saturation value. The saturation value

corresponds

to a director orientation

angle

which is

substantially

less than 90

degrees.

For EBBA the

experiment

was also carried out in an electric field in order to record the "total"

birefringence

curve, I-e- director reorientation from 0 to 90

degrees. Experiments presented

here do not allow to determine the direction of the symmetry

breaking

orientation.

3. Basic

equations.

We consider a nematic

layer

of thickness d confined between two infinite

parallel plates.

The z-axis is chosen normal to the

bounding plates,

the

oscillating

flow in the x-direction and the

director fi confined to the x z

plane.

With this choice one can write

n~ = cos@, ny = 0, nz =

sine,

U~ " U, fly " 0, uz " 0,

Ii)

where = @(t,

z),

u =

Hit, z)

with ~

< z <

~ The

hydrodynamic equations

[5,6] reduce to

2 2

~ii@,t

(a2 sin~

a3 cos~@)u

z =

(Kii

cos~ +

K33 sin~

@)@,zz +

(K33

Kii sine cos

@@)~,

(2)

pu,t " -P~ +

8z(-(a2 sin~

a3

cos~

@)@,t + [M(@) +

N(@)]u,z ), (3)

where P is the pressure,

2M(@)

= a4 +

(05 a2) sin~

@,

2N(@)

=

(a3

+ a6 + 2ai

sin~

@)

cos~

@,

the a; are

viscosity coefficients,

~ii = 03 a2

,

K;; are elastic constants, and p is the

density

of the

liquid crystal.

The notation

f,;

a

8f/8i

has been used

throughout.

A full

sylution

of the

coupled equations (2)

and

(3)

with

time-periodic boundary

conditions

at z = + " for Couette flow or

time-periodic

pressure

gradient

P~ for Poiseuille flow is

quite

a

2 '

formidable task. For

simplicity

we will here treat

Hit, z)

as a

prescribed 2ir/w-periodic

function in t with

Hit

+

ir/w, z)

=

-u(t, z),

so that the

time-average

<

u(t, z)

> is zero. This should

not

change

the

qualitative

features and should be a reasonable

quantitative approximation

for small oscillation

amplitudes

or low

frequencies

such that the viscous penetration

depth fi,

(6)

N°9 FLOW ALIGNMENT OF NEMATICS UNDER OSCILLATORY SHEAR 1391

where a is a

typical viscosity,

is

larger

than the thickness d. In this

spirit

we may even include

a g-

dependence

in @, which leads to an additional term K22@,yy on the

right-hand

side of

equation (2),

and which we will need for a very

specific

reason later on.

With the dimensionless variables

I

= cot, 2 = z

Id,

fi =

u/dw, (4)

the

equation

for can be rewritten as

@,t

K(@)u,z

= e~[P(@)@,zz + ~P'(@)@)~ + k2@,yy],

(5)

where the tildes have been omitted and

K(@)

m

(I

cos~

sin~

@)

Ill I), P(@)

=

cos~

+ k3

sin~

@, 1 =

a31a2, k,

=

K;; /Kii,

e~

=

I/(Tdw),

Td

='tid~/Kii,

h~(

f)

a

8h/8f. (6)

Note that for the

flow-alignment

case, I > o~

K(@)

vanishes at = +@fl. Since the

frequency

w is

usually

much

larger

than the inverse director relaxation time we may assume e~ « I,

Neglecting

the terms on the

right-hand

side of

equation (5)

the

remaining

first-order

ordinary

differential

equation

can be solved and the solution @o is obtained from

f°° (

= g,

gin, z)

=

f~

dtu

~it, z) ii)

so that

@o =

90(J/, g(t, z))

is a

periodic

function in t. For J/

= +@fl

ii)

cannot be used

(then

Y/ e +@fl).

Actually

the

o-integral

in

ii)

can be solved

analytically

and @o(11,g) can be

expressed

in terms of

elementary

functions [4]. For our purpose this is not needed.

Choosing

the

origin

of t such that

g(t

=

0)

= 0 one has @o(t =

0)

= Y/. Thus

@o oscillates around

Y/, but in

general

Y/ does not coincide with < @o >. We conclude that

neglecting

the elastic

coupling

in the bulk leads to a continuous

family

of periodic oscillations of that can be

parametrised by

the

"phase"

Y/.

In order to

investigate

the influence of the elastic

coupling

we use the method of

multiple-

scale

analysis

[13].

By introducing

a "slow" time T =

e~t,

that modulates

(slowly)

the

periodic

behavior on the fast time

scale,

so that

= @(t, y, z,

T)

and

at

- at +

e~8T,

one can formulate

a

systematic perturbation expansion

of the form

= @o + e@j + e~@2 +..

,

where all the @; are

periodic

in t. Then from equation

(5)

at order e° one has the solution

ii). However,

the

phase

Y/ is now allowed to vary on space and slow time T and is undetermined at this order. At first order in e one has

L91

= 0, L m

at K~(90)g,t, (8)

where L is the linear operator of the

perturbational equation

of

ii)

and we can choose

0j

= 0.

Finally,

at order e~ we find

L82

=

-80,T

+

P(80)80,zz

+

P'(80)8(,~

+ k280,~~.

(9)

This

inhomogeneous

linear

equation

for 82 is not

generally

solvable because L has the null

eigenvector

80,~. The

solvability

condition takes on the form:

<

V+80,T

> = <

V+[P(80)80,zz

+

jP'(80)8(,~

+ k280,~~] >

(10)

JOURN~L DE PHYS]QUE Ii -T 3 N'9 SEPTEMBER 1993

(7)

where we introduced the scalar

product

< V+

f

> =

/~~

dtv+

f (11)

27r o

and V+ is the null

eigenvector

of the

adjoint

operator L+V+

= 0, L+

= -at

K'(go )g,t. (12)

Using

the fact that go

depends

on y, z and T

only through

i~ and g

equation (10)

can be cast into the form of an evolution

equation

for the

phase

i~:

i~,T = B + Ci~,z + Di~,zz + Ei~)~ +

Fi~,m

+

Gi~§, (13)

where

B, C, D,.

G are scalar

products involving

the functions i~, go(~1,

g),

g,z and g,zz. The

expressions

are

given

in

appendix

A. The calculations are

quite simple

because 80,~ and V+

can be calculated

analytically.

Let us present more

explicitly

the

expressions

for the case of small

g(t, z),

which

essentially

amounts to A

Id

< 1 where A is the maximum

displacement

in the

oscillatory

flow. Then from

equation (7)

for go up to the second order in g one has:

80 " ~1+

K(~1)g

+

(K(~1)K'(i~)g~

+

(14)

and for the coefficients in

equation (13)

B =

P'K~

<

gg,zz >

+[PK~l'j

< g(z >, C =

2K[PK'

+

)P'KI'

< gg,z >, D = P +

K[P'KI')

< g~ >,

E =

)P'

+

(-K'[P'KI'

+

[K[PK'+ jP'Kl'l'))

< g~ >, F

= k2, G

=

k2(KK"l'~

< g~ >

(15)

Here P

=

P(i~)

and K

=

K(i~) everywhere.

Equation (13)

is the central result of this work. It describes the slow evolution of the

phase

i~

(the precise

definition of

i~ was

given

after

Eq. (7)).

We see from

(15),

and also from the

general expressions (see Appendix A),

that B is in

general

nonzero if g is

spatially varying,

I-e- when there is a shear

gradient (see Eq. (7)).

Nonzero B means that there is a net bulk torque

acting

on i~. If

spatial

variations of

i~

play

no role the

stationary

solutions of

(13)

are

simply given by

the zeros of B

(note

that B

= 0 for i~ =

~8fl).

As

pointed

out

before,

in

general

the

velocity

field u has to be determined

self-consistently

from

equation (3) together

with

i~ from

equation (13),

from which go follows. For that purpose it would be useful to also separate the fast and slow time

dependence

in

equation (3).

4.

Oscillatory

Poiseuille flow.

For

simplicity

we here choose a geometry where one has a shear

gradient

even at low

frequencies.

Thus we

apply

the

theory

to

plane

Poiseuille flow. The results are

expected

to carry over to Couette flow for

frequencies

such that the viscous penetration

depth fi$

is of order d.

(8)

N°9 FLOW ALIGNMENT OF NEMATICS UNDER OSCILLATORY SHEAR 1393

For

simple plane oscillatory

Poiseuille flow one has

u =

a(4z~ 1)

cost, g = 8az sin t.

(16)

Then

equation (13)

lvith

(15)

goes over into

YJ,T =

a~lllYJ)

+ a~z©lYJ)YJ,z + lPlYJ) + a~z~l5lYJ)lYJ,zz +

+[~ P'(i~)

+ a~z~fl(i~)]i~)~ + k2~1,yy +

a~z~d(i~)i~§, (17)

where the functions

fi

etc.

are

easily

calculated. In

particular

one has

tj~)

=

16jpj~)K2j~)ji j18)

In the

flow-alignment

case,

a3/02

> 0, the function

fi(i~)

has three zeros

corresponding

to two stable

equilibria

at i~ = 0 and

7r/2 (here

P'

= K'

=

0),

and an unstable

equilibrium

at the

flow-alignment angle

i~ = 8fl =

tan~~(a31a2) (here

K =

0).

When a3 goes to zero two

equilibria

coalesce and for a3 > 0 there remains an unstable

equilibrium

at i~ = 0

(no changes

at i~ =

7r/2).

Since for

typical

nematics 8fl is

small,

one expects the solution

i~ = 0 to be

only

very

weakly

stable

(small

domain of

attraction) compared

to i~ =

7r/2.

This is also borne out

by

the fact that the

quantity

U(il)

=

-16P(i~)K~(i~), (19)

which

plays

the role of a

potential (up

to a constant

factor)

for the motion of i~, I-e-

fi(i~)

=

-U'(i~),

is much lower at i~

=

7r/2

than at i~ = 0.

Our conclusion about the relative

stability

of the two minima of U can be substantiated further

by considering

the effect of fluctuations and the direction of motion of a domain wall in the y direction

connecting

the two states

(here

we need the y

dependence). Clearly

for a~ « the last term in

equation (17)

can be

neglected compared

to the one before it.

Discarting

z

dependence

as

before, equation (17)

represents a

simple Ginzburg-Landau equation.

It is well known that in this

theory

domain walls

always

move in the direction that lowers the

potential

and localized fluctuations and

perturbations

can

only

lead from the

higher

to the lower state.

When

oscillatory

Poiseuille flow is induced in a thin slab in the usual way

by application

of a pressure difference one must be aware of the fact that as a result of

compressibility

the pressure

decays along

the flow direction x on a

length

b

~

d/pc~ /(3wo))

where c is the sound

velocity

and a a

typical

shear

viscosity

[14]. Since b » d for all reasonable

frequencies, experiments

should nevertheless be

possible. Moreover,

the

decay

is avoided when Poiseuille flow is induced

by oscillating

in

parallel

both

plates

of an

open-ended

slab.

5. Discussion.

Our

theory

shows that under

oscillatory

Poiseuille flow there is a net bulk torque which renders the

time-averaged homeotropic

position

(<

8 >

=

7r/2) globally

stable within the shear

plane.

Actually

the

interpretation

of the effect is

fairly straightforward.

Let us consider a much

simpler

system,

namely

that of a

parametrically land oscillatory)

driven

highly-damped particle

described

by

an

equation

of the form

8,t

"

K(8)au(t)

where

u(t)

is

27r/w-periodic

and a is

some factor. This system does not loose the memory of its initial

conditions,

I-e- one has a continuum of solutions

including

the

(neutrally stable) equilibrium position.

The situation is

changed immediately

when an identical

particle

is

coupled

to the first one via discrete diffusion

(9)

and the second

particle

is not driven with the same

amplitude (but

with the same

frequency),

I.e. when one has

81,t

=

K(81)aiu(t) D(81 82),

82,t "

K(82)a2u(t) D(82 81) (20)

with al

#

a2. It can then be shown

quite generally

that the

equilibrium position, K(8)

= 0,

becomes

unstable,

which is the

analog

of the flow

alignment angle becoming

unstable as a result of

spatial coupling,

and that the

long-time

solutions are attractors

(in fact,

it is

easily

seen that the

time-averaged divergence

of the flow in

phase

space defined

by equation (20)

is

-2D,

I-e-

negative).

The additional

dissipation

comes from the fact that the

inhomogeneous driving always

activates the diffusive

coupling,

which attempts to restore

homogeneity.

Our

treatment takes

properly

care of the

singular

nature of the

perturbation.

Our result appears to hold for more

general plane

shear flow. A

simple

estimate of the

magnitude

of the effect shows

that,

in order for the flow-induced torque to overcome strong

(planar)

surface

anchoring,

the

displacement amplitude

of the fluid has to be of order of the thickness d.

Unfortunately

then the

small-amplitude approximation

invoked in section 3 breaks down. A more accurate treatment within the framework of the

theory

is in progress.

We point out that a net torque can also arise from the term

proportional

to i~~,z in

equation (13).

To activate this term, an asymmetry across the

sample

with respect to the z = 0

plane

has to be present which could be induced

externally by asymmetric boundary

conditions. This effect can destabilize the

homeotropic alignment

and would lead to a state where i~

changes monotonically

across the cell.

Let us now comment on the

possibility

of

out-of-plane

motion under

planar oscillatory

shear.

Neglecting

the elastic

coupling

one

again

has the situation that the system does not relax, and one now has a two-parameter

family

of

trajectories. Clearly

none of the out-of

plane trajectories

crosses the shear

plane.

From the theoretical treatments of the

elliptic

shear

instability

it is known that the threshold

diverges

in the

planar

limit [9]. This must mean

that, taking

into account elastic

coupling,

the

in-plane

motion of

flow-aligning

nematics under

planar

shear is indeed

linearly

stable for all

amplitudes.

Therefore transitions to

out-of-plane

motion can

presumably only

be

explained by assuming

the existence of stable

out-of-plane

orbits that coexist with the

in-plane

orbits but never bifurcate. Therefore the transitions should be

hysteretic.

A

study

of the out-of-

plane

motion

using

the method

presented

here is in progress. Whereas in

low-frequency experiments

the critical

amplitudes

of the oscillations

were found to be of order of the thickness d [8] or

larger

[4],

they

were much smaller

(below

0.1

~m)

for the

high-frequency

measurements

[11,

12]. The latter situation seems difficult to

explain by

our

theory,

except

by assuming

that the effective

(averaged)

surface

anchoring

becomes weak at

high frequencies.

At very

high frequencies

it may be necessary to include other effects like

compressibility

[15].

The

theory presented

here shows that in

plane oscillatory

shear flow a net

hydrodynamic

torque acts in

general

on the

director, opening

the

possibility

for reorientational transitions of the kind

presented

in section 2. More

detailed, quantitative experimental

studies of the

phenomenon

are in progress.

Acknowledgments.

We are

grateful

to W. Pesch for a critical

reading

of the

manuscript.

Two of us

(A.B.

and

A-K-)

wish to thank the

University

of

Bayreuth

for its

hospitality.

Financial support from Deutsche

Forschungsgemeinschaft (SFB

213,

Bayreuth)

as well as the

Hungarian Academy

of Sciences

(OTKA 2976)

are

gratefully acknowledged.

(10)

N°9 FLOW ALIGNMENT OF NEMATICS UNDER OSCILLATORY SHEAR 1395

Appendix

A.

From

equation (7)

we have go = go(i1,

g)

,

where i~ =

il(T,

y,

z),

g

=

g(t, z)

and

80,T

= i1,T80,~, 80,z = i1,z80,~ +

g,z80,g,

80,y = i1,Y80,~,

80,zz = i1,zz80,~ + il~z80,~~ + i1,zg,z280,~g +

g,zz80,g

+ g~z80,gg,

80,yy = ~l,vv80,~ + ~l(y80,~~.

II)

We can calculate

80,~, 80,g, 80,m,

80,gg, and 80,~g from our solution

(Eq.(7)):

a~ /°° ]I=

o, j2)

~

which

gives

80,~

"

fl@. Analogously,

80,g =

K(80), 80,m

=

fl@~'°#(l'~,

80,gg

=

K(80)K'(80)

and 80,~g =

@K'(80).

The null

eigenvector

V+ of the

adjoint

operator

(12)

can be

easily

found as V+

=

@

and

we choose

C(il)

=

K(il)

so that

V+80,~

=1. Then for the coefficients in

equation (13)

we will have:

B =

K(il)(< P(80)g,zz

> + <

[P(80)K'(80)

+

)P'(80)K(80))g)z >),

C = 2 <

[P(80)K'(80)

+

)P'(80)K(80))g,z

>, D = <

P(80)

>,

E "

)(-K'(i1)

<

P(80)

> + <

P(80)K'(80)

+

(P'(80)K(80) >),

~

~~'

~ ~~

)i1)

~'~~~~

~

~'~~~ ~~'

~~~

Note added in

proof

: We

recently

learned that experiments under the

plane oscillatory

shear flow in

homeotropically

oriented nematic

layers

considered here have also been

performed

by

Belova G-N- and Remizova

E-N-,

Akust. Zh. 31

(1985)

289

(Sov. Phys.

Acoust. 31

(1985) 171).

A theoretical

analysis

of the transition to rolls found there has been

presented by

Kozhemikov

E.N.,

Zh.

Eksp.

Teor. Fiz.

91(1986)1346 (Sov. Phys.

JETP 64

(1986) 793).

Similar theoretical and experimental results for low

frequencies if

= 20 200

Hz)

and

planar alignement

were

published

very

recently by Hogan S-J-,

Mullin T. and Woodford

P.,

Proc. R.

Sov. London A 441

(1993)

559.

References

[1] Leslie F-M-, Adv. Liq. Cryst. 4

(1979)

1.

[2] Leslie F-M-, Theory and Applications of Liquid Crystals, il. Ericksen, D. Kinderlehrer Eds.

(Springer-Verlag,

New York, 1987) p. 235.

[3] Zuniga I., Leslie F-M-, Liq. Cryst. 5

(1989)

725; Europhys. Lett. 9

(1989)

689.

(11)

[4] Clark M-G-, Saunders F-C-, Shanks I-A-, Leslie F-M-, Mol. Cryst. Liq. Cryst. To

(1981)

195.

[5] Ericksen J.L., bans. Soc. Rheol. 5

(1961)

23.

[6] Leslie F-M-, Arch. Mech. Anal. 28

(1968)

265.

[7] Pieranski P., Guyon E., Phys. Rev. Lett. 39

(1977)

1281;

Guazzeli E. and Guyon E., J. Phys. France 43

(1982)

985.

[8] Guazzeli E., Thbse de 3me cycle, Chap. III.7, Universit6 Paris- Sud

(Orsay, 1981).

[9] Dubois-Violette E., Rothen F., J. Phys. France 39

(1978)

1039;

Sadik J., Rothen F., Bestgen W., J. Phys. IFance 42

(1981)

915.

[10] Burghardt W-R-, J. Rheol. 35

(1991)

49.

[11] Scudieri F., Appl. Phys. Lett. 29

(1976)

398.

[12] Chuvyrov A-N-, Zh. Eksp. Teor. Fiz. 82

(1982)

761

(Engl.

Transl: Sov. Phys. JETP 55

(1982) 451);

Chuvyrov A-N-, Scaldin O-A-, Delev V-A-, Mol. Cryst. Liq. Cryst. 215

(1992)

187.

[13] see e-g- Bender C-M-, Orzag S-A-, Advanced Mathematical Methods for Scientists and Engineers

(McGraw-Hill,

New York,

1978).

[14] Blinov L-M-, Davidyan S-A-, Reshetov V-N-, Subachyus D-B-, Yablonskii S-V-, Zh. Eksp. Teor.

Fiz. 97

(1990)

1597

(Engl.

Transl: Sov. Phys. JETP To

(1990) 902).

[15] Helfrich W., Phys. Rev. Lett. 29

(1972)

1583.

Références

Documents relatifs

2014 The nature of the flow alignment of nematic liquid crystals is discussed within the hydrodynamic theory of Leslie, Ericksen and Parodi.. It is shown that there is a

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

From long-time strain auto- correlation data we next bring evidence that, in the New- tonian regime, thermal fluctuations trigger irreversible shear transformations giving rise

In the macroscopic geometric limit (that is the macroscopic, rigid and quasistatic limits), with either volume or normal stress controlled simulations, static and dynamic

Focusing on the specific case of resonantly driven models whose effective time-independent Hamiltonian in the high-frequency driving limit corresponds to noninteracting fermions, we

Generalizing a recently introduced approximation scheme valid when the director relaxation time is large compared to the inverse frequency of an oscillatory flow, we derive equa-

Steady poiseuille flow in nematics : theory of the uniform

in table I relative to MBBA at 25°C.. - Instability threshold Sc as a function of the magnetic field. 1) Homogeneous distortion : - Curve A corresponds to