• Aucun résultat trouvé

Surface-polarization effect in the alignment of nematic liquid crystals

N/A
N/A
Protected

Academic year: 2021

Partager "Surface-polarization effect in the alignment of nematic liquid crystals"

Copied!
10
0
0

Texte intégral

(1)

HAL Id: jpa-00247809

https://hal.archives-ouvertes.fr/jpa-00247809

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.

Surface-polarization effect in the alignment of nematic liquid crystals

G. Barbero, A. Chuvyrov, G. Kaniadakis, E. Miraldi, M. Rastello

To cite this version:

G. Barbero, A. Chuvyrov, G. Kaniadakis, E. Miraldi, M. Rastello. Surface-polarization effect in the alignment of nematic liquid crystals. Journal de Physique II, EDP Sciences, 1993, 3 (1), pp.165-173.

�10.1051/jp2:1993118�. �jpa-00247809�

(2)

Classification

Physics

Abstracts 6 1.30

Surface-polarization effect in the alignment of nematic liquid crystals

G. Barbero

('),

A. N.

Chuvyrov

(Z), G. Kaniadakis

(I),

E.

Miraldi(I)

and

M. L. Rastello

(3)

(J) Dip.

Fisica, Politecnico, C. Duca

degli

Abruzz124, l0l29Torino,

Italy

(2) Bashkir State

University,

Ufa, Russia

(3) 1-E-N-G- Ferraris, Str. Delle Cacce91, 10135 Torino, Italy

(Received 3 March 1992, revised 7 August 1992, accepted 28 September 1992)

Abstract. The

change

in the orientation of nematic

liquid crystals

due to the increase of the

sample

thickness is

experimentally analysed.

The data show that the

phenomenon

is a critical one,

typical

of a second order

phase

transition. The

change

in the surface

polarization

induced

by

a

selective

ion-adsorption

on the two

boundary

solid substrates is

proposed

as the cause of the observed facts and discussed. The considered model is correct for

angles

of tilt small with respect

to the initial homeotropic

alignment,

and the theoretical curves fit both the interferometric and the

voltage-threshold

data. A few comments on the obtained results are

reported.

1. Introduction.

It is well known that the mean molecular orientation of a nematic

liquid crystal,

the so-called

« director » n,

depends

on both the intrinsic symmetry of the

material,

and the structural

symmetry

of the substrate surface.

The director

profile imposed by

a solid substrate has been

analysed by

many groups I both

experimentally [2]

and

theoretically [3],

in order to find a connection between the observed orientation and the

physico-chemical properties

of surfaces and

liquid crystal.

A

complete understanding

of the

ordering properties

of a

surface,

due to the chemical and

physical interactions,

is far from

being achieved,

but

they

can be described in terms of an

«

anchoring

» energy.

From a

phenomenological point

of

view,

the

anchoring

energy is defined as a function of the director orientation close to the surfaces its minimum coincides with the mean molecular direction in the absence of bulk torques. This energy

depends

on both the interaction between the

liquid crystal

and

substrates,

and the interaction among the nematic molecules themselves.

In the absence of

long-range forces,

the surface energy is a local

property

in the sense that it does not

depend

on the

sample

thickness. On the contrary, it does

depend

on the thickness when there are

long-range forces,

like electrostatic

forces,

due to

ion-adsorption

on the

surfaces,

or viscous

forces,

due to the

velocity

field in the nematic volume

[4].

This energy can be

investigated by controlling

the

boundary conditions,

e. g.

by covering

the cell surfaces with suitable

conducting

and/or

insulating

materials and

by rubbing them,

as for the most

widely

used

displays.

(3)

166 JOURNAL DE PHYSIQUE II N°

In this paper, the average director

profile

of different nematic

cells,

filled

by capillarity,

has been determined

by measuring,

with standard interferometric

techniques,

their

birefringence

for different discrete thicknesses. In this manner,

varying

the

sample thickness,

one can vary its base and lateral

surfaces,

and

indirectly

the concentration of the

selectively

adsorbed

ions,

the

sample

volume

being

constant.

In the

following

the average

birefringence

will be defined as

usually

:

t12 n

(An )

= An

df

=

2Rt (sin~

4l

)

,

(1)

-t12

~

C

~~~~/~~~~~~~~~~~~ ~

_ ,

fl '

/ t9'

,1' Wit)

°

j

/

,

,11,,llllll/ till/

(llllll

Ill

a)

llzllllllllllll (jllllll/ j

( $ jj

j

~l

t)

Ii

,

j

(

,

rll/ / II

llllll11(/

II film

_y

z b)

t

«lllllljll/,llllfllllll/, (

f

~

/

l

,/

,

i ,~ ~~~~

~ 4~

_?

Ill II / /, / II Ill Ill IN llllllll Ill Z

C)

Fig.

I. The co-ordinate system used in text and the schematic director

profile

: a) in presence of surface

polarization

b) induced

by

a

velocity gradient

c) in the case of a Fr£edericks transition.

(4)

where R

= I

(n~/n~)~,

with n~n~ the

ordinary

and

extraordinary

refraction

indices,

and

(sin~ 4l)

is the average of the

squared

sinus of the

angle

between the

optical

axis and the

surface

normal,

and ( the coordinate normal to the

boundary surfaces, posed

at (

=

± t/2 as

reported

in

figure

I.

The

figure

also shows the

peculiar

differences of the director

profile

in three conditions,

Figure

la refers to the case when

only

surface forces are

present,

so that the

alignment

is

nearly uniformly

tilted from one surface to the other and 4l

(()

= const.

Figure

16 shows the director

profile

due to a kind of « memory » of the

filling

process : near

the

boundary surfaces,

at ( = ±

t/2,

the

hydrodynamic

forces and the

anisotropy

of the

liquid crystal viscosity impose

that

4l(()

reaches the maximum absolute value. The

profile

is

antisymmetrical,

that is 4l

((

=

4l

(-

(

),

In

figure

lc the

liquid crystal undergoes

the so-called Fredericks transition in the presence of

an extemal field:

4l(()

reaches the maximum value in the middle of the cell at

(

=

0. The

profile

now is

symmetrical.

The three cases can be

distinguished experimentally by

interferometric

techniques

and the measurements discussed in this paper refer to the first case shown.

Only

the Frdedericks transition with weak

anchoring

conditions

(in

the sense that 4l

t/2 can vary under the torque

imposed by

the extemal

field)

can

give

a result similar to the

uniformly

distorted

alignment

due to the

only change

in the surface energy.

Indeed for this reason the same

phenomena

were

discussed,

some time ago, as

« Frdedericks » effects due to

spontaneous polarization [5],

but a critical look at the accuracy of the former

experimental set-up (mainly

the surface treatments and the

samples preparation)

induced us to

repeat

the

experiments.

With a new

experimental apparatus, designed

to match the necessary

stability conditions,

and with

nearly

strong

anchoring

conditions

imposed by

the

boundary surfaces,

as

exposed

in the next

paragraph,

this

ambiguity

is resolved.

In section

3,

a model of surfaces

anchoring

energy that

depends

on the

selectively

adsorbed ions is

proposed.

The

corresponding

surface

polarization

is proper to

explain

the observed

abrupt change

in the director

alignment,

induced

by

the

substrates,

when a criticaI thickness of the

sample

is reached. The model

predictions

are

compared

with measurements of both mean

director

profile

and the electric field Fredericks

transition,

as a function of the cell thickness. In the last

paragraph

some comments are

reported

on the numerical results.

2. The

experimental

results.

The nematic

liquid crystal

used was

methoxy-benzylidene-butyl-aniline (MBBA),

both pure

and mixed

(3

:

1)

with

ethyl-hydroxy-benzylidene-butylaniline (EBBA).

Two cells were

prepared

with unmixed MBBA and the

corresponding

sets of

experimental

data

(points

and squares in

Fig. 2a)

refer to nematic

samples kept

between

glass surfaces, optically

flat within 100 nm, cleaned in sulfo-cromic acid and covered with

Cr~,

30 nm thick.

The two sets of data have been obtained in different

times,

with different

accuracies,

but with the same

experimental apparatus

: the

complete matching

between the data confirms that the cell thickness is a

good

control parameter for this

phenomenon,

The last set of

experimental

data

(Fig, 2b)

refers to a third nematic

sample

filled with mixed

MBBA and EBBA and

kept

between fused quartz

plates

covered with

ln~O

and

SnO~,

The chosen surface treatments

give

in any case an

homeotropic alignment,

I-e- the

starting

director

profile

is normal to the

walls,

with

nearly

strong

anchoring conditions,

and in that

position

it would continue to be while the cell thickness is

changing,

unless other causes, from inside or outside the

specimen,

are able to support a tilted director

alignment.

(5)

168 JOURNAL DE PHYSIQUE II N°

. , m m m mm . mm m

, m " "

0

30 ~

jp~j

60

a)

0.12

m

~ W

'$ W

~e

C

~

W

m

0

50

jp~j

loo

b)

Fig.

2.

Average

sinus of the tilt

angle (sin~

~P

)

~'~ versus the thickness t for

samples

filled with pure MBBA (a), or a nematic mixture MBBA + EBBA (b). Points and squares in al refer to two sets of

measurements

performed

in different times but with the same

experimental facility.

The theoretical

curves are drawn with the value of the parameters discussed in text.

The parameter under control in every run has been the cell thickness. This can be varied

mechanically

with the aid of micrometer screws, and a

great

care was delivered in the

design

of

the mechanical device

appropriate

to

modify

the

sample

dimension without

modifying

the

parallelism

of the

boundary

surfaces.

To the same purposes the metallic materials used for

supporting

the

liquid crystal cells,

and handled to

change

their

thickness,

were chosen with the smallest thermal

expansion

coefficient and

kept

in a thermostatic chamber.

The overall accuracy of these micrometric measurements has been of the order of

2 ~Lm for what concems the absolute

length determination,

but somewhat better if one looks

at the remake of the

experiment,

as

previously reported.

The

general properties

that the collected data show are the

following

ones :

I)

the director orientation is normal to the

boundary

surface if the

sample

thickness is lower than a critical

value,

I.e. if t w t~;

(6)

it)

for t

~ t~, the director

profile

is a tilted one, the average

birefringence (An )

# 0 and

depends

on the thickness ;

iii)

for t

~ co,

(An) approachs

a constant value that

depends

on the

liquid crystal

used.

One can summarize in this way the three above-mentioned statements : the trend of

(sin~ 4l)

~'~, 4l

being

the surface tilt

angle,

is

typical

of a second-order

phase

transition near t

= t~ and shows a limit for t

~ co.

For what concems

point

I the two sets of data

referring

to unmixed MBBA

give

the same

critical

thickness,

as can be seen in

figure 2a,

with a numerical value t~ m 32 ~Lm.

On the contrary the critical value measured with the nematic mixture was t~ m 60 ~Lm, as can be seen in

figure 2b, showing

that also the critical thickness

depends strongly

on the

liquid crystal properties.

3. The

physical

model.

As

already

mentioned in the introduction, the

change

in the director

alignment

with the

liquid crystal

cell thickness could be

interpreted,

in

principle,

in two different ways.

According

to the first

point

of

view,

the flow of the

liquid crystal during

the

filling

process

modifies the surface

anchoring

energy, and

imposes

an easy

direction, parallel

to the flow direction.

Only

a few papers consider the influence of the flow on the nematic orientation. Some time ago, Wahl and Fischer

[6]

observed that the nematic average orientation

drastically changed by

a shear

flow,

with velocities even so

negligible

as

10~~ mm/day.

Note that this mechanism

implies

the surface tilt

angle 4l(f)= 4l(- (),

I-e- an

antisymmetric

director

profile. Furthermore,

it

implies

also a

unique flow-direction,

whereas in our

experimental

set-up it is

only

a noise

effect,

without any well defined direction : the

flow-effects,

if any, average out to zero.

Looking

for a different way to

analyse

the

experimental data, namely

the observed re-

orientational

transition,

a

possible explanation

could be the selective ions

adsorption

phenomenon recently

discussed

[7].

According

to this

point

of view the effective

anchoring

energy

depends

on the adsorbed ions.

If the dielectric

anisotropy

of the

liquid crystal

is

negative,

and the easy axis due to the surfaces treatment is

homeotropic,

as

certainly they

are in our case, the

adsorption phenomena give

a

destabilizing

effect on the surface orientation.

In this framework let us suppose that the surface

anchoring

energy is

represented by

:

f~

= iii

cos~

4l

+ y

cos~

4l

,

2 4

(2)

where

cos~ (4l

is connected with the surface

polarization. Actually

the energy term associated

to the order-electric

polarization [8]

is

proportional

to

(rj cos~

4l

+r~)~,

where rj and

r~ are the order-electrical coefficients.

Straightforward

calculations show that a term

proportional

to

cos~

4l should appear in the surface energy, as

reported

in

equation (2),

where ik is the effective

anchoring

energy,

taking

into account the « ions » effect.

As it is

given by

reference

[9]

:

iii=w- ~~~~ ~

.~~, (3)

2

e~

where w is the

anchoring

energy

strength,

A is the

Debye screening length,

e is the average dielectric constant, e~ is the dielectric

anisotropy

and

finally

~r

= ~r

(t)

is the surface

charge

density

due to the selective ions

adsorption.

~r

depends

on the

thickness,

as discussed in

(7)

170 JOURNAL DE PHYSIQUE II N°

reference

[9],

and the second term in

equation (2)

can be

guessed

as a term

correcting

the usual

Rapini-Papoular' expression

and connected to the order-electric

polarization [8].

The nematic orientation in the cell can be obtained

minimizing

the total energy of the

sample,

defined as the sum of the bulk and surface contributions. The bulk term is due to the elastic deformation of the

nematic,

and it is

proportional

to the square of the

gradient

of the average molecular orientation. In the

hypothesis

of a uniform orientation of the nematic in the

bulk,

this term vanishes and the orientation near the surfaces is deduced

by minimizing only

the surface contribution

f~. Simple

calculations

give

an

homeotropic profile

stable if

ik ~ y, that is for :

«

(t)

~ «

(tc)

=

l~ )~j

~

(4)

~a Y

Equation (4)

defines the critical thickness t~. for t ~ t~, that is if

~r(t)

~

~r(t~),

the stable nematic orientation is characterized

by

a tilt

angle given by

:

sin~

4l

= a

i~r~(t> ~r~(t~)1

,

(5)

where

e~j

A

"

(2 ye~l'

~~~

Of course the tilted orientation is stable

only

if iii

~

0,

that

implies

~r ~ ~r*

=

fi, (7)

~a(

otherwise the stable orientation is the

planar

one, with 4l

=

ar/2,

not observed in our case.

Note that

according

to this mechanism the

dependence

of

(An )

versus t is due to a

change

of the surface energy. More

precisely

the

phenomenological

parameters

characterizing f~ depend

on t.

Consequently

also the casy direction deduced with

df~/d4l

=

0, depends

on t.

Hence the

equilibrium

condition

stating

that the total torque on the surface has to be zero,

gives

a surface tilt

angle

that

depends

on the thickness.

From

equation (5)

one can see, for t 5S t~.

sin4lmp /~, (8)

where

» =

la ~i~ ,~)"~ (9)

Equation (8)

shows that the

homeotropic

~ distorted transition is of the second

order,

as

experimentally

observed.

According

to the model

presented

in reference

[9]

~r

(t

is

given by

:

"~~~

"~

2

~~/r

'

~~~~

with r =

~r~/p~,

where p~ is the ions concentration in the

bulk,

~r~ is the

asymptotic

value of the surface

charge density,

and both concentrations refer to an infinite

sample.

(8)

By substituting equation (10)

into

(4)

one obtains :

sin~

4l

=

awl ~~~~~~

~

~~~~~~

~

l. (ll)

(

+ t/r

) (

+

t~/r)

It is

possible

to estimate the order of

magnitude

of the parameters

appearing

in the model.

For what concems the surface parameters

[8, 10]

one has

wm10~~erg/cm~

and y

m

10~ ~ =

10~~ erg/cm~

for the electrical

[9,

it ones one has :

e~(

m

0.5,

e m

5,

A m 0. I

=

I ~Lm, ~r~ m 10~ ~

C/cm~,

p~

m 10~ ~

= 10~ ~

C/cm~.

Consequently

: r

=

~r~/p~

m10 ÷

10~

~Lm,

a =

e~(

Al

(2 ye~)

m

10~

=

10~ (cm~/C)~, awl

m

10~

=

10~.

By using equations (4)

and

(10)

one can calculate the critical thickness t~.

tc

12 e~(w

y

)

"~

t~ + 2

r e~ y

From this

equation

and

using

the values of the

physical

parameters listed

above,

one has t~ m

(IN)

r, that is of the correct order of

magnitude.

In

figures

2a and b two theoretical curves,

computed

with

equation (11),

and the

corresponding

sets of

experimental data, (sin~ 4l)"~

versus t, are

reported.

It is

straightforward

to see that the agreement of the theoretical curves with the

experimental points

is

fairly good only

near the critical thickness. This is due to the very naive model

proposed

for the ion

adsorption,

described

by equation (10).

As is well

known,

this

expression

is similar to the

Langmuir isotherm,

and

consequently

it can run well

only

in the limit of small

~r values. A more

general expression

can be obtained

only by taking

into account the

electrostatic

repulsion

between the adsorbed

ions, neglected

in this

analysis.

The model used leads in fact to a

polynomial expression

that contains

only

three

parameters, namely

one is the critical thickness and the other two are related to the ion concentrations in the bulk and on the surfaces.

By

means of the above-mentioned

model,

nevertheless, one can deduce the critical

voltage

U for the Fr£edericks transition as a function of the thickness of the

sample.

To this aim

4

0

lo t

jp~j

32

Fig.

3. The theoretical curve and the

experimental

data collected for the threshold voltage in the Frdedericks transition U versus the

sample

thickness t. The

samples

used were the same as in

figure

2a.

(9)

172 jOURNAL DE PHYSIQUE II N°

measurements of the threshold

voltage

U for the Frdedericks transition have been

performed,

and the uniform

homeotropic alignment

of the director for

samples

with t

~ t~ was confirmed.

In

samples

filled with pure MBBA U was

falling

with

increasing

thickness of the

crystal,

till

a

complete vanishing

was reached for t m t~,

supporting

the existence of a permanent tilt of the

director,

as in the

optical

measurements.

By minimizing

the total free energy of the

sample,

obtained

by summing

up the bulk term

(including

the elastic term and the extemal field

interaction)

and the surface one, as shown in reference

[12],

one obtains that the critical

voltage depends

on the elastic

(k)

and electric

(e~) properties

of the

nematic,

the effective surface energy

(iP)

and the

sample

thickness

(t).

Its

analytical expression

is

~t= ~tg (i~ ), (12)

" k UC 2 UC

where U~ = ar

kleo

e~ is the critical

voltage

in the strong

anchoring

case, and U is the one in the weak

anchoring

case.

By using

also

equations (3)

and

(10),

it is

possible

to calculate

U/U~

versus t.

The

experimental

results

conceming

the critical

voltage

and the theoretical curve derived from the

physical

models are

reported

in

figure

3. The agreement between

points

and curve is

obtained with the same values of the

parameters

used above.

4. Conclusions.

The influence of the thickness of a

sample

on the average orientation of a nematic

crystal

has

been considered. The

experimental

data show that the

reorienting phenomenon

connected with the thickness of the

sample

is a critical one. More

precisely

it is

experimentally proved

that for t ~ t~, where t~ is a critical

thickness,

the

samples

are in a

homeotropic

orientation. For t m t~ the average orientation

changes according

to

(t t~)~'~, typical

of a second order

phase

transition.

The

experimental

data are

analyzed by using

a

simple model,

in which the surface energy

depends strongly

on the adsorbed ions.

If one assumes in the

expression

of the surface energy a term

proportional

to

cos~ 4l,

connected with the order-electric

polarization,

a transition from the

homeotropic

to the tilted director

configuration

must be

expected

if one suggests that the selective ions

adsorption

process follows the

Langmuir

law for the

perfect

gases.

In the same frame of the

Langmuir-adsorption

law the threshold

voltage

for the Frdedericks transition has been

analyzed

: the theoretical

predictions

are in

good

agreement with the

experimental

data for t

< t~.

In contrast, the theoretical

(sin~ 4l)~'~

curves and the

experimental

data are in

good

agreement

only

for t 5St~. This fact can be

interpreted

in the sense that the

Langmuir

law must be considered

only

as a very crude

approximation,

valid

just

in the

beginning

of the

transition,

for t

m t~. Work is in progress to

improve

the theoretical model.

References

[1] JEROME B.,

Rep. Frog.

Phys. 54 (1991) 391.

[2] See for instance COGNARD J., Mol. Cryst. Liq.

Cryst. Supp.

1(1987) 1.

[3] YOKOYAMA H., Mol.

Cryst. Liq. Cryst.

165 (1988) 265.

[4] LESLIE F. M., Arch. Rat. Mech. Anal. 28 (1968) 265.

(10)

[5] CHUVYROV A. N., Sov.

Phys. Crystallogr.

25 (1980) 188.

[6] WAHL J., FISCHER F., Mol. Cryst. Liq. Cryst. 22 (1973) 359.

[7] YOKOYAMA H., KOBAYASHI S., KAMEI H., J. Appt. Phys. 56 (1984) 2645.

[8] BARBERO G., DURAND G., J. Phys. France 47 (1986) 2129.

[9] BARBERO G., DURAND G., J. Appt.

Phys.

67 (1990) 2678.

[10] BLINOV L. M., KABAENKOV A. Yu., SONIN A. A., Liq. Cryst. 5 (1989) 645.

[11] THURSTON R. N., CHENG J., MAYER R. B., BOYD G. D., J.

Appt. Phys.

56 (1984) 263.

[12] RAPINI A., PAPOULAR M., J.

Phys. Colloq.

France 30 (1969) C 4-54.

Références

Documents relatifs

For a flow-aligning nematic, the HI shear rate threshold is studied as a function of surface tension anisotropy 03C3a and applied magnetic field strength.. It is shown

compounds : open circles, electrical measurements; filled circles, magnetic measurements. measurements of density or susceptibility have so far extended; we used

Non linearities close to the thermal threshold in a planar nematic liquid

2014 In this paper we report measurements of the angle 03B8t between the director and the vertical axis at the free surface of the nematic liquid crystals MBBA and

described by the molecular field theory, the result of the calculation cannot be expected to be identical with a more correct determination of the elastic constants

2014 Nous étudions le diagramme de phase d’un système dense de particules allongées avec une interaction Coulombienne, et montrons l’existence d’une transition

We have applied this approximation to the Maier-Saupe model of nematic liquid crystals and studied the effect of short range correlations on the.. nematic-isotropic

This is mainly for two reasons : First, this kind of viscosity is an original property of nematic liquid crystals not linked to any macro-.. scopic shear and,