• Aucun résultat trouvé

Threshold conditions for hydrodynamic instabilities in two-frequency nematics

N/A
N/A
Protected

Academic year: 2021

Partager "Threshold conditions for hydrodynamic instabilities in two-frequency nematics"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: jpa-00210137

https://hal.archives-ouvertes.fr/jpa-00210137

Submitted on 1 Jan 1985

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Threshold conditions for hydrodynamic instabilities in two-frequency nematics

P.R. Gerber

To cite this version:

P.R. Gerber. Threshold conditions for hydrodynamic instabilities in two-frequency nematics. Journal

de Physique, 1985, 46 (11), pp.1865-1872. �10.1051/jphys:0198500460110186500�. �jpa-00210137�

(2)

Threshold conditions for hydrodynamic instabilities in two-frequency nematics

P. R. Gerber

F. Hoffmann-La Roche & Co. Ltd., Central Research Units, CH-4002 Basel, Switzerland

(Reçu le 20 mars 1985, revise le 7 juin, accepti le 28 juin 1985)

Résumé.

2014

Les équations électrohydrodynamiques pour une couche de nématique en configuration planaire

sont traitées dans le cas de petites déformations. On étudie les conditions pour une instabilité hydrodynamique : premièrement dans le cas d’une anisotropie diélectrique positive sous la forme d’un développement en petits vec-

teurs d’onde (déformation de longue période) et deuxièmement dans le cas de nématiques à « deux fréquences ».

Pour une anisotropie positive de la conductibilité électrique, la déformation statique de Frederiks est supplantée

par une instabilité dynamique à période finie. Les instabilités ne se produisent que dans une bande de fréquence large d’à peine un facteur deux. Les conditions sont optimales pour un processus de relaxation de grande amplitude

et pour une anisotropie diélectrique négative forte à fréquence élevée.

Abstract

-

Electrohydrodynamic equations for a layer of nematic material with planar homogeneous alignment

are given for alternating applied voltage, for the case of small deformations. Conditions for hydrodynamic instabi-

lities are studied firstly for the case of a positive dielectric anisotropy in a small wavevector-expansion, and secondly numerically for two-frequency nematic materials in which the conductivity originates from orientational relaxation of polar molecules. It is found that, except for some cases of negative conductivity anisotropy the static Frederiks deformation can only be overtaken by a finite-(non-zero) wavevector dynamic instability. In two-frequency mate-

rials instabilities hardly occur over a frequency range of more than a factor of two. Best conditions are a high ampli-

tude of the relaxation process and large negative dielectric anisotropy at high frequencies.

Classification Physics Abstracts 61.30G201385.60P

1. Introduction.

Electrohydrodynamic instabilities in nematic layers

which are oriented homogeneously parallel to the plane of the substrates by surface forces have been studied for some time [1]. The observation by Wil-

liams [2] of regular patterns which occur in certain substances upon application of a perpendicular electric

field have found a transparent explanation through

the work of Helfrich [3]. Heilmeier et al. [4] found that

at higher applied voltages turbulent patterns develop

which owing to their strong light scattering properties appeared to be promising for display-application.

Helfrich’s interpretation required an anisotropic elec-

trical conductivity which enabled the generation

of space charges. In most experiments this conduction

was provided by ionic impurities in the nematic material. These impurities in turn hampered the application of the effect because they were a source

of electrochemical degradation. Later on de Jeu et al.

[5] discovered similar phenomena in a so called two- frequency material for which part of the parallel

ohmic conductivity was provided by orientational

relaxation processes of long molecules. This showed,

in principle, a way of avoiding degradation.

In the present work we aim at exploring the

necessary conditions on the dielectric parameters of ion-free nematic material for the generation of hydrodynamic instabilities, by studying theoretically

the threshold conditions of the electrohydrodynamic equations. To this end the existing theories had to be extended While the treatment of Dubois-Violette

et al. [6] allows for alternating applied voltages,

it is restricted to infinitely thick layers. On the other hand the finite-layer treatments [7, 8] are restricted

to dc-applied voltage in which case the orientational- induced conductivity vanishes.

A further, more fundamental question regards the changeover from static deformation to hydrodynamic instability which occurs in these substances upon

increasing the frequency of the applied voltage.

2. Basic equations.

There are three types of equations which govern the

electrohydrodynamic effects in nematics namely the

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198500460110186500

(3)

1866

torque equation describing the alignment of the

nematic director, the acceleration equation governing

the material flow and Maxwell’s equations which

determine the charge distribution. Derivations of these equations as valid in a nematic layer have been given in several places [1, 6, 7] whence we restrict

ourselves to writing them down for the case of a

layer of nematic substance which is homogeneously aligned in the substrate plane by rigid boundary

forces at both substrates. The director deformations

are assumed to be restricted to the plane given by

the layer normal and the substrate aligning direction.

Similarly we assume that the material and charge

flow directions are also restricted to this plane.

The z-axis of the coordinate system is normal to the substrates with the origin in the middle of the

layer. The substrate positions are thus at z

=

± d/2.

The x-direction corresponds to the alignment direc-

tions at the substrates.

The variable fields are 9, the angle between x-axis

and director, Vx and vZ, the components of the mass

velocity, p the pressure, q the charge density and

the electric field variables Px and P. which constitute the electric field as

With these definitions the torque balance equation

reads

where

where y, and Y2 are viscosity coefficients [1]. With the density p and the viscosity coefficients al to a6 [1] the

acceleration equations read

which have to be supplemented by the continuity equation of the fluid which is assumed to be incompressible :

With the conductivities all ~ and (J.1 Maxwell’s equations read

All these equations are derived under the assumption

that the variable fields are small which means that all terms quadratic or of higher order in the fields

can be neglected. This procedure is exact for the cal- culation of threshold conditions as long as the insta-

bilities under consideration develop continuously

from the undeformed homogeneous state.

These equations have to be supplemented by the

boundary conditions at the substrates, namely

(4)

3. Time dependence.

The externally applied field j2 Eo cos mo t oscillates

in time with frequency mo. In the present treatment

we will assume that a period io

=

2 nlcoo is short compared to characteristic times governing the chan-

ges in director orientation or material flow. However, the electric fields and charge field are assumed to respond with times comparable to To (conduction regime [6]). Thus, we set

Furthermore we restrict ourselves to stationary solu-

tions such that all fields 9, p, vj, qk, fli,k can be consi-

dered time independent To obtain equations for

these quantities, we take time-averages of equations (2.2) and (2.9) while equations (2.11) to (2.13)

each yield an in-phase equation, obtained by multi- plying with cos roo t and time-averaging, and an out-of-phase equation, obtained by multiplying with

sin wo t and averaging. We are thus left with ten

equations. Similarly the boundary condition requir- ing vanishing tangential electric field at the substrate

splits into two components for P.,, and lix,2.

There is little use in writing down all the equations

at this point However, it might be worthwhile men-

tioning that a treatment of the dielectric regime [6, 9]

would require to take q as time-independent (or slowly varying), while the director field qJ and the

velocity and pressure fields would have to be taken

as time dependent, in a form analogous to equa- tion (3.1).

4. Further treatment : characteristic equation and boundary conditions.

The treatment of the ten equations proceeds by elimi- nating variables and taking spatial Fourier trans-

forms. Some details can be found in the Appendix.

Taking the viscosity coefficients ai and a3

=

(11 + y2)/2 which seem to be small in the few known cases, equal to zero for the sake of simplifying the

arithmetic somewhat, one eventually arrives at the following three equations :

in which we have set

From the k-dependent variables t/J, v and the original variables are calculated through

An important equation relates P2 which occurs in the boundary conditions with Pl and

For a nontrivial solution of equations (4. 1) to (4. 3) one has to require a vanishing determinant of the coeffi-

cients of v, Pl and V/ which leads to the characteristic equation

(5)

1868

This fifth order polynomial in y has one solution yo

= -

1 (due to the choice ai

=

0). In comparison with the

static case [7] (lim coo -+ 0) the degree of the polynomial is augmented by one thus providing an additional solu- tion to allow to fulfil the additional boundary condition for the out-of-phase electrical field.

The k,,

=

0 solutions obtained by setting y

=

0 in equation (4.14) give the threshold condition

This is precisely the result obtained by Dubois-Violette et al. [6] for the conduction regime.

Each root yi of equation (4.14) has with it a corresponding small-amplitude solution, the coefficients vi flu and of which are calculated through equations (4.1) to (4. 3) up to a common factor.

To fulfil the boundary conditions, any linear combination of these solutions may be chosen. Actually, there

are ten conditions which, because of the mirror symmetry at the x-axis of the problem, reduce to five each for

symmetric and antisymmetric solutions. These conditions again have the form of five homogeneous linear equa-

tions, having in general no solution except for special values of the coefficients which give zero value of their

determinant. For the case of symmetric solutions which is most interesting in the present context, the determinant reads

where we have used the abbreviations

The values of the external field parameter U for which equation (4.16) is obeyed, give the required threshold

conditions.

5. Expansion for small wave numbers k,

For two-frequency nematics in which the dielectric

anisotropy changes sign to negative values with

increasing values of the frequency, a transition from

a static deformation to an electrohydrodynamic ins- tability can occur for increasing frequency. Since the

static deformation is a kx

=

0 solution, one might

expect to gain some insight into the transition from static to dynamic instability by attempting a small- kx expansion of the above threshold conditions (4.16).

To perform the calculation, we restrict ourselves to vanishing magpetic field (h

=

0) and use as an

expansion parameter the dimensionless quantity

where A is the spatial period of the deformation.

Remembering that the parameter U diverges when kx

goes to zero we set

where the parameter

is just the square of the ratio between the applied voltage V and a reference voltage PB which is related to the Frederiks threshold voltage [1] ] VF by

Inserting (5. 2) into the characteristic equation reveals

that one solution, Y4, is of order, - 1, while the remain-

ing four are of order (including yo = - 1). A small

calculation yields

(6)

Furthermore, we have

.

and for the solution y4

such that

Because of

the quantities C4 and S4 are at most of order çO.

Thus to order (° the determinant (4.16) reads

In this expression the two determinants contain only

the four solutions y°, ... Y3 of O(çO), and have in

general nonzero values. Because Y4 is of order equation (5.13) requires

This immediately yields from (4.17) and (5.12)

The threshold voltage for the lowest (n

=

0) mode is

thus given by

This expression shows that for the common nema-

tics which almost always obey Q II > (J.1 the static deformation with ç

=

0 always sets in before a dyna-

mic deformation with small ç. In other words, at the changeover from static to dynamic instability the i

latter always sets in with a finite, non zero wave-

vector kx as a first order transition. In section 6 this will also be illustrated by a numerical example.

In contrast to ordinary nematics composed of elongated molecules, discotic molecules which show

a nematic-type order, seem to obey the condition

(J.1 > all II [10]. Thus, for low enough frequencies and

k 1

-

(J II I (J .1 one would expect that the usual

Frederiks transition is prevented by the occurrence

of a hydrodynamic deformation mode in such mate- rials. In addition, for suitable parameter values one may observe a second order transition from static to dynamic deformation.

6. Numerical calculations for two-frequency nema-

tics.

We restrict ourselves here to the case of a single relaxa-

tion process at frequency Cùr in the dielectric constant

parallel to the nematic director and zero ionic conduc-

tivity

The crossover frequency uy is obtained from the con-

dition y

=

0 (Eq. (6.1))

Our aim is mainly to trace the conditions for dynamic

instabilities when the dielectric parameters yo and

Ay vary. For this reason, the elastic and viscous para- meters are kept fixed at the more or less representa- tive values

With these material parameters the following results

have been obtained.

In figure 1 the dependence of the instability voltage

is shown for varying wave number for several fre-

quency values of the applied voltage. Above the cross-

over frequency We the instability sets in at a finite

wavevector given by the voltage minimum. Below We a second minimum occurs at dl À.

=

0 which at

first lies above the other but at a frequency Wx drops

below. At this frequency the changeover between dynamic and static deformations takes place. Upon

further lowering the frequency, the now metastable dynamic minimum disappears’ altogether. The wave-

vector of the dynamic instability decreases steadily

with decreasing frequency to a finite non-zero value,

where the dynamic minimum disappears. This picture

as well as the curvature at dl À.

=

0 for co > we is in accordance with the result of the small wavevector

expansion of section 5.

The threshold voltage as a function of frequency

is shown in figure 2. The crossover between static and dynamic deformation occurs at the frequency Wx’

somewhat below Wc. Also shown is the wavevector of the deformation (as dl À). With this value, one can,

for comparison, calculate a threshold as obtained

(7)

1870

Fig. 1.

-

Wavevector dependence of threshold voltage for

various frequencies in a two frequency material. For low

frequencies the minimal voltage occurs at À.

=

oo, which represents the static Frederiks deformation. For increasing frequency a finite wavelength hydrodynamic instability

takes over while for A

=

oo a local minimum is still present.

For w > Wc’ only the dynamic minimum remains. The minimal voltage instability shows a pronounced wavelength

variation with varying frequency (dashed curve).

by the kz

=

0 approximation of reference [6], which

is given in equation (4.15). Clearly, this value is

expected to improve with increasing ratio dIA. For

the calculations of figure 2, this approximate threshold voltage is low by almost a factor of two at wx and by

some 30 % at wlw,,

=

1.3.

Also shown on figure 2 are threshold voltages for

a few nonzero values of the ionic conductance assum-

ing (Ji II I Ui.1

=

1.5. As found in experiments [5],

the range of frequencies for which the dynamic ins- tability is observed increases when the ionic relaxa- tion frequency

increases to values comparable to w,,.

For our purpose, the most relevant results are

shown in figures 3 and 4. For these calculations the ionic conduction was again set to zero and the dielec-

tric parameters yo and Ay were varied. Figure 3

shows contours of constant threshold voltage V x

as observed at the frequency Wx where static and

dynamic instabilities exchange. The lowest values are

encountered for large Ay and low yo values. The fre-

Fig. 2.

-

Frequency dependence of minimal threshold

voltages. The plott illustrates the fairly narrow frequency

interval in which hydrodynamic instabilities occur. Also shown (dotted curve) is the threshold voltage for the bulk

approximation (Eq. (4.15)) as calculated from the actual wavenumbers shown in the upper graph. The dashed curves

illustrate the enlargement in frequency range obtained by a

finite ionic conductivity.

quency OJx hardly goes below 0.9 oj,,, (see Eq. (6.3))

such that it appeared not necessary to present a

,co.,/coc in graphical form.

Of importance for possible applicability of such

materials is the range in frequency within which

hydrodynamic instabilities can be generated without

the addition of ionic dopands. Figure 4 gives a con-

tour plot for this range. For this graph, the range was defined by the limiting frequencies OJx and a)hl the

frequency at which the threshold voltage had increas- ed to 3 V x. It must however be kept in mind that coc may be a more practical lower range limit than cvx, because at the more elevated voltages used in dyna-

mic scattering the static deformation may again take

over for (o OJc. This is, however, a minor correc-

tion regarding the closeness of (ox and wc.

From figure 4 one sees that the ranges stay usually

below a factor of two in frequency. The largest ranges

occur for large Ay and low yo values, i.e. in the same

region where the threshold is low.

(8)

Fig. 3.

-

Contours of constant threshold voltage Vx at the changeover frequency between static and dynamic instabi- lity. The coordinates are the amplitude of the relaxation process Ay and the high frequency dielectric anisotropy yo.

Fig. 4.

-

Contours of constant frequency interval for the

occurrence of dynamic instabilities in the (Ay, yo)-plane.

The interval is defined by Wx and the frequency Wh at which the threshold voltage has reached 3 Vx.

In conclusion, nematic materials with large orien-

tational relaxation effects in the parallel dielectric

constants may possibly be used as substances for

dynamic scattering without showing the adverse degradation effects usually observed in ionically conducting substances.

However, the range of frequencies in which the phenomenon is encountered is fairly narrow. Fur- thermore, in order to obtain low driving frequencies,

the relaxation frequency coo must be kept low. This

in turn leads to rather strong temperature dependences

of coo [10] which would either limit the temperature

range of operation or otherwise require a suitable adjustment of the driving frequency to the display temperature. For materials with cvo/2 x = 1 kHz

1

is, in fact, very high.

The strong dependence of the wavelength of the

distortion on the value of the driving frequency

may possibly be used to produce gratings of variable, externally controlled period This would be in a range centred around a value given by the spacing of the liquid crystal cell.

Appendix.

Some of the tedious but straightforward arithmetical steps which lead from the basic equations of sec-

tion 2 to equation (4.1) to (4.3) are sketched here.

As mentioned in Section 4 we put ai

=

a3

=

0. The main steps include time averaging as described in Section 3 and spatial Fourier transformation by

means of equations (4.10) to (4.12). The choice (4 .11 ) leads to a single velocity transform vk due to equation (2.10). Similarly the number of transforms for the electric field (Eq. (4.12)) is reduced by (2.11).

Equation (2.2) yields directly (4.2). From (2. 8)

and (2.9) one eliminates p and arrives at

Similarly (2.12) and (2.13) yield four equations (in-

phase and out-of-phase, see Section 3) namely

(9)

1872

Elimination of q1 in (A. .1 ) with the help of (A . 3) leads

to (4.1). Elimination of 41 from (A. 3) and (A. 6)

and of 42 from (A. 4) and (A. 5) yields

From these equations (4. 3) is obtained by elimination of P2’ while (4.13) is identical with (A. 7).

References

[1] For a review see equation DE GENNES, P. G., The Phy-

sics of Liquid Crystals (Oxford, 1974) or HEL- FRICH, W., Mol. Cryst. Liq. Cryst. 21 (1973) 187.

[2] WILLIAMS, R., J. Chem. Phys. 39 (1963) 384.

[3] HELFRICH, W., J. Chem. Phys. 51 (1969) 4092.

[4] HEILMEIER, G. H., ZANONI, L. A. and BARTON, L. A.,

Proc. IEEE 56 (1968) 1162.

[5] DE JEU, W. H., GERRITSMA, C. J., VAN ZANTEN, P.

and GOSSENS, W. J. A., Phys. Lett. 39A (1972) 355.

[6] DUBOIS-VIOLETTE, E., DE GENNES, P. G. and PARODI, O.,

J. Physique 32 (1971) 305.

[7] PENZ, P. A. and FORD, G. W., Phys. Rev. A 6 (1972)

414.

[8] NANDY, P., Mol. Cryst. Liq. Cryst. 104 (1984) 281.

[9] HEILMEIER, G. H. and HELFRICH, W., Appl. Phys. Lett.

16 (1970) 155.

[10] DUBOIS, J. C., HARENG, M., LE BERRE, S. and PERBET, J. N., Appl. Phys. Lett. 38 (1981) 11.

[11] GERBER, P. R., Z. Naturforsch. 37A (1982) 266.

[12] SCHADT, M., Mol. Cryst. Liq. Cryst. 89 (1982) 77.

Références

Documents relatifs

In this Letter, we present results of electrical magnetotransport and X-ray absorption spectroscopy (XAS) experiments on LAO/STO samples with either bare LAO surfaces or a

The DFAP begins with an initial set of links, called the initial kernel. The first sub- problem is to assign frequencies to the paths constituting the kernel. Then, new links

Dependence of the relative rms fluctuations of the threshold voltage &V~/ (V~) on specimens length... II 8 JOURNAL DE PHYSIQUE

- Instability threshold reduced field a is plotted versus the reduced frequency p. Let us note that the diffusion currents are in

Fig 2: The black stars means the 10 reference stars mentioned in the paper (Pickering 1908), the red marks() means the reference stars found using our method to identify the same

La forte variabilité temporelle mais également spatiale des apports des sources de matière organique aura, par conséquent, un impact non négligeable sur la dynamique de la

Pour lui, le couple folk/art reflète une musique pure, authentique, échappant à la touche commerciale et dépendant du génie (d’un individu, d’un peuple) alors que la

In areas where maize is a dietary staple, future studies intended to re- veal any possible linkage between maternal fumonisin exposure and reproductive toxicity and growth