Download as pdf or txt
Download as pdf or txt
You are on page 1of 476

This is a new and greatly revised edition of Professor Chandrasekhar's

classic book Liquid crystals, first published in 1977. In this second edition
the author has brought the subject completely up to date.
The subject of liquid crystals has now grown to become an exciting
interdisciplinary field of research with important practical applications.
This book presents a systematic and self-contained treatment of the
physics of the different types of thermotropic liquid crystals - the three
classical types, nematic, cholesteric and smectic, composed of rod-shaped
molecules, and the newly discovered discotic type composed of disc-shaped
molecules. The coverage includes a description of the structures of these
four main types and their polymorphic modifications, their thermodynamical, optical and mechanical properties and their behaviour under
external fields. The basic principles underlying the major applications of
liquid crystals in display technology (for example, the twisted and
supertwisted nematic devices, the surface stabilized ferroelectric device,
etc.) and in thermography are also discussed.
This book will be of great value to advanced students and research workers
in condensed matter physics, chemical physics, materials science and
technology with an interest in the physics, chemistry and applications of
liquid crystals.

LIQUID CRYSTALS

LIQUID CRYSTALS
Second edition
S. CHANDRASEKHAR, F.R.S.
Centre for Liquid Crystal Research, Bangalore

CAMBRIDGE

UNIVERSITY PRESS

Published by the Press Syndicate of the University of Cambridge


The Pitt Building, Trumpington Street, Cambridge CB2 1RP
40 West 20th Street, New York, NY 10011-4211, USA
10 Stamford Road, Oakleigh, Victoria 3166, Australia
Cambridge University Press 1977, 1992
First published 1977
Second edition 1992
A catalogue record for this book is available from the British Library
Library of Congress cataloguing in publication data
Chandrasekhar, S. (Sivaramakrishna), 1930
Liquid crystals/S. Chandrasekhar. - 2nd ed.
p. cm.
Includes bibliographical references (p. ).
ISBN 0 521 41747 3 (hardback). - ISBN 0 521 42741 X (pbk.)
1. Liquid crystals. I. Title.
QD923.C46 1992
530.4'29 - dc20
91 38434 CIP
ISBN 0 521 41747 3 hardback
ISBN 0 521 42741 X paperback

Transferred to digital printing 2004

UP

Contents

Preface to the first edition


page
Preface to the second edition
Introduction
.1
Thermotropic liquid crystals: structure and classification
of the mesophases
.1.1
Liquid crystals of rod-like molecules
.1.2 Liquid crystals of disc-like molecules
.1.3 Polymer liquid crystals
.2
Lyotropic liquid crystals
.3
Polymorphism in thermotropic liquid crystals
2
Statistical theories of nematic order
2.1
Melting of molecular crystals: the Pople-Karasz model
2.1.1
Order-disorder in positions and orientations
2.1.2 Plastic crystals and liquid crystals
2.1.3 Pressure-induced mesomorphism
2.2
Phase transition in a fluid of hard rods
2.3
The Maier-Saupe theory and its applications
2.3.1 Definition of long-range orientational order
2.3.2 The mean field approximation
2.3.3 Evaluation of the order parameter
2.3.4 Theory of dielectric anisotropy
2.3.5
Relationship between elasticity and orientational order
2.4
Hybrid models: hard rods with a superposed attractive
potential
2.5
Short-range order effects in the isotropic phase
2.5.1 The Landau-de Gennes model
2.5.2 Magnetic and electric birefringence
2.5.3 Light scattering
vii

xiii
xv
1
1
1
8
10
12
14
17
17
17
22
27
29
38
38
41
43
51
57
60
61
61
63
66

viii
2.5.4
2.5.5
2.6
2.6.1
2.6.2
2.7
3
3.1
3.1.1
3.1.2
3.1.3
3.1.4
3.2
3.3
3.4
3.4.1
3.4.2
3.4.3
3.4.4
3.5
3.5.1
3.5.2
3.5.3
3.5.4
3.5.5
3.5.6
3.5.7
3.5.8
3.5.9
3.6
3.6.1
3.6.2
3.6.3
3.6.4
3.6.5
3.7
3.8

Contents
Flow birefringence
Comparison with the Maier-Saupe theory
Near-neighbour correlations: Bethe's method
The Krieger-James approximation
Antiferroelectric short-range order
The nematic liquid crystal free surface
Continuum theory of the nematic state
The Ericksen-Leslie theory
Conservation laws and the entropy inequality
Constitutive equations
Coefficients of viscosity
Parodi's relation
Curvature elasticity: the Oseen-Zocher-Frank equations
Summary of equations of the continuum theory
Distortions due to magnetic and electric fields: static
theory
The Freedericksz effect
The twisted nematic device
The Freedericksz effect in highly anisotropic nematics:
periodic distortions
The constant k13
Disclinations
Schlieren textures
Interaction between disclinations
Non-singular structures (s = 1): escape in the third
dimension
Twist disclinations
Singular points
Interaction between disclinations and surfaces
Defects in the presence of an external field
Some consequences of elastic anisotropy
The core structure
Flow properties
Miesowicz's experiment
Tsvetkov's experiment
Poiseuille
flow
Shear
flow
Transverse pressure and secondary
flow
Reflexion of shear waves
Dynamics of the Freedericksz effect

69
70
71
71
75
80
85
85
86
88
91
93
94
97
98
98
106
113
115
117
117
120
123
126
129
132
135
139
143
144
144
145
148
152
157
159
161

Contents
3.8.1
3.8.2
3.9
3.9.1
3.9.2
3.9.3
3.10
3.10.1
3.10.2
3.10.3
3.10.4
3.10.5
3.11
3.11.1
3.11.2
3.12
3.13
3.13.1
3.13.2
3.13.3
4
4.1
4.1.1
4.1.2

4.1.3
4.1.4
4.1.5
4.1.6
4.2
4.2.1
4.2.2
4.2.3
4.3
4.4

Twist deformation
Homeotropic to planar transition: backflow and kickback
effects
Light scattering
Orientational
fluctuations
Intensity and angular dependence of the scattering
Eigenmodes and the frequency spectrum of the scattered
light
Electrohydrodynamics
The experimental situation
Helfrich's theory
DC excitation
Square wave excitation
Sinusoidal excitation
Hydrodynamic instabilities
Homogeneous instability in shear
flow
Roll instability in shear
flow
Thermal instability: stationary convection
Flexoelectricity
Determination of the flexoelectric coefficients
Influence of flexoelectricity on electrohy drody namic
instability
Order electricity
Cholesteric liquid crystals
Optical properties
Propagation along the optic axis for wavelengths ^ pitch
Propagation along the optic axis for wavelengths ~pitch:
analogy with Darwin's dynamical theory of X-ray
diffraction
Exact solution of the wave equation for propagation along
the optic axis: the Mauguin-Oseen-de Vries model
Equivalence of the continuum and the dynamical theories
Oblique incidence
Propagation normal to the optic axis
Defects
/-disclinations
Lattice disclinations
Dislocations
Leslie's theory of thermomechanical coupling
The Lehmann rotation phenomenon

ix
161
162
167
167
170
173
177
177
183
186
187
192
19 5
195
197
202
205
205
210
211
213
213
213

222
237
241
245
247
248
248
252
254
258
262

x
4.5
4.5.1
4.5.2
4.6
4.6.1
4.6.2
4.6.3
4.7
4.8
4.9
4.10
5
5.1
5.2
5.3
5.3.1
5.3.2
5.3.3
5.3.4
5.3.5
5.3.6
5.3.7
5.4
5.4.1
5.4.2
5.4.3
5.4.4
5.5
5.5.1
5.5.2
5.6
5.6.1
5.6.2
5.6.3
5.7
5.8

Contents
Flow properties
Flow along the helical axis
Flow normal to the helical axis
Distortions of the structure by external fields
Magnetic field normal to the helical axis: the
cholesteric-nematic transition
Magnetic field along the helical axis: the square grid
pattern
Electric field along the helical axis
Anomalous optical rotation in the isotropic phase
The blue phases
Some factors influencing the pitch
Molecular models
Smectic liquid crystals
Classification of the smectic phases
Extension of the Maier-Saupe theory to smectic A:
McMillan's model
Continuum theory of smectic A
The basic equations
The Peierls-Landau instability
The Helfrich deformation
Fluctuations and Rayleigh scattering
Damping rate of the undulation mode
Ultrasonic propagation and Brillouin scattering
Breakdown of conventional hydrodynamics
Defects in smectic A
Focal conic textures
Edge dislocations
Screw dislocations
Disclinations
The smectic A-nematic transition
Phenomenological theory of the smectic A-nematic
transition
Pretransition effects in the nematic phase
Smectic A polymorphism
Smectic A phases of strongly polar molecules
The phenomenon of re-entrance
Smectic A* or the twist grain boundary phase
The hexatic phase
Smectic C

267
270
274
277
277
281
286
289
292
296
298
300
300
302
310
310
311
313
317
320
323
325
327
327
333
338
338
340
340
341
350
350
355
358
360
362

Contents
5.8.1
5.8.2
5.8.3
5.8.4
5.9
5.10
5.10.1
5.10.2
5.10.3
6
6.1
6.2

Description of the structure


Continuum theory of smectic C
Defects in smectic C
The smectic C-smectic A transition
The nematic-smectic A-smectic C multicritical point
Ferroelectric liquid crystals
The properties of smectic C*
The smectic C*-smectic A transition
Applications of ferroelectric liquid crystals
Discotic liquid crystals
Description of the liquid crystalline structures
Extension of McMillan's model of smectic A to discotic
liquid crystals
6.3
The columnar liquid crystal: applications of the
continuum theory
6.3.1 Fluctuations
6.3.2 X-ray scattering
6.3.3 Light scattering
6.3.4 Mechanical instabilities
6.3.5 Acoustic wave propagation
6.4
Defects in the columnar liquid crystal
6.4.1 Dislocations
6.4.2 Disclinations
6.4.3 Developable domains
6.5
The discotic nematic phase
6.6
The biaxial nematic phase
References
Index

xi
362
365
369
370
374
378
378
383
386
388
388
394
398
398
399
400
400
401
403
404
409
410
411
414
418
451

Preface to the first edition

' I am at a loss to give a distinct idea of the nature of this liquid, and
cannot do so without many words.'
The Narrative of Arthur Gordon Pym of Nantucket,

Edgar Allan Poe

The foundations of the physics of liquid crystals were laid in the 1920s but,
surprisingly, interest in these substances died down almost completely
during the next three decades. The situation was summarized by F. C.
Frank in his opening remarks at a Discussion of the Faraday Society in
1958: 'After the Society's successful Discussion on liquid crystals in 1933,
too many people, perhaps, drew the conclusion that the major puzzles were
eliminated, and too few the equally valid conclusion that quantitative
experimental work on liquid crystals offers powerfully direct information
about molecular interactions in condensed phases.' In the last few years
there has been a resurgence of activity in this field, owing partly to the
realization that liquid crystals have important uses in display technology.
An exposition of the physics of liquid crystals involves many disciplines:
continuum mechanics, optics of anisotropic media, statistical physics,
crystallography etc. In covering such a wide field it is difficult to define
what precisely the reader is expected to know already. An attempt is made
to present as far as possible a self-contained treatment of each of these
different aspects of the subject. Naturally, discussion of some topics has
had to be curtailed for reasons of space. For example, we have not dealt
with lyotropic systems, whose complex structures are only just beginning
to be elucidated; or the special applications of magnetic resonance
techniques, as these have been adequately reviewed elsewhere; or the very
recent results of neutron scattering experiments. The primary aim of this
monograph is to provide an insight into the variety of new phenomena
exhibited by these intermediate states of matter.
Xlll

xiv

Preface to the first edition

This book would probably never have been completed without the
unstinting cooperation of my young colleagues in Bangalore. I am
particularly indebted to N. V. Madhusudana, G. S. Ranganath, R.
Shashidhar, U. D. Kini, R. Nityananda and T. G. Ramesh, whose notes
and critical comments were of inestimable value at every stage of the
writing, and to K. A. Suresh and B. R. Ratna for their assistance in
preparing the diagrams and the list of references. Finally, it is a pleasure to
express my thanks to Michael Woolfson for his encouraging interest and
advice.
Raman Research Institute
August 1975

S. Chandrasekhar

Preface to the second edition

Since the publication of the first edition 15 years ago, the subject of liquid
crystals has grown enormously to become a fascinating interdisciplinary
field of study. A variety of new thermotropic phases have been discovered,
including over a dozen different smectic modifications, discotics, biaxial
nematics, etc., which have opened up a veritable treasure-house for the
theoretical condensed matter physicist. On the technological side, the
advances have been no less spectacular: portable computers and hand-held
TV sets using liquid crystal display devices are being sold in large numbers,
and high-definition LCD-TV would seem to be just round the corner. The
aim of the present edition is to bring the coverage up to date. The chapters
dealing with the classical nematic, cholesteric and smectic types of liquid
crystals have been revised substantially and a new chapter has been
included on discotics. However, mainly for reasons of space, special topics
like the applications of magnetic resonance techniques, non-linear optical
properties, etc., have not been discussed here as these have been
comprehensively reviewed elsewhere.
It is my privilege to express my thanks to my young colleagues who kept
me alive to the subject: in particular to V. N. Raja, S. Krishna Prasad,
D. S. Shankar Rao, S. M. Khened and Geetha Nair for their invaluable
help throughout the preparation of this book, and to Sriram Ramaswamy
and U. D. Kini for their advice on certain theoretical points. I am indebted
to the Council of Scientific and Industrial Research, New Delhi, for a
Bhatnagar Fellowship which made it easier for me to undertake this task.
Centre for Liquid Crystal Research
March 1992

S. Chandrasekhar

xv

1
Introduction

The term liquid crystal signifies a state of aggregation that is intermediate


between the crystalline solid and the amorphous liquid. As a rule, a
substance in this state is strongly anisotropic in some of its properties and
yet exhibits a certain degree of fluidity, which in some cases may be
comparable to that of an ordinary liquid. The first observations of liquid
crystalline or mesomorphic behaviour were made towards the end of the
last century by Reinitzer(1) and Lehmann.(2) Several thousands of organic
compounds are known now to form liquid crystals.(3) An essential
requirement for mesomorphism to occur is that the molecule must be
highly geometrically anisotropic in shape, like a rod or a disc. Depending
on the detailed molecular structure, the system may pass through one or
more mesophases before it is transformed into the isotropic liquid.
Transitions to these intermediate states may be brought about by purely
thermal processes {thermotropic mesomorphism) or by the influence of
solvents (lyotropic mesomorphism).

1.1 Thermotropic liquid crystals: structure and classification of the


mesophases
1.1 A Liquid crystals of rod-like molecules
The vast majority of thermotropic liquid crystals are composed of rod-like
molecules. Following the nomenclature proposed originally by Friedel,(4)
they are classified broadly into three types: nematic, cholesteric and
smectic.
The nematic liquid crystal has a high degree of long-range orientational
order of the molecules, but no long-range translational order (fig. 1.1.1).
Thus it differs from the isotropic liquid in that the molecules are
1

1. Introduction

Crystal

Nematic

Fig. 1.1.1. Schematic representation of molecular order in the crystalline, nematic


and isotropic phases.
spontaneously oriented with their long axes approximately parallel. The
preferred direction usually varies from point to point in the medium, but
a uniformly aligned specimen is optically uniaxial, positive and strongly
birefringent. The mesophase owes its fluidity to the ease with which the
molecules slide past one another while still retaining their parallelism.
X-ray studies (56) indicate that some nematics possess a lamellar type of
short-range order, i.e., they consist of clusters of molecules - called
cybotactic groups(7) - the molecular centres in each cluster arranged in
layers. In ordinary nematics the cybotactic groups, if they do exist, are
smaller than can be detected by X-ray methods (see figs. 1.1.2 and 1.1.3).

1.1 Thermotropic liquid crystals

(a)

(b)

Fig. 1.1.2. X-ray diffraction patterns from an 'ordinary' nematic liquid crystal:
4-n-propyloxy-benzylidene-4/-n-propylaniline: (a) an unaligned (though not quite
randomly oriented) sample; (b) a well aligned sample. (De Vries.(6))
A biaxial modification of the nematic has been discovered recently (see
6.6).
The cholesteric mesophase is also a nematic type of liquid crystal except

1. Introduction

(a)

(b)

Fig. 1.1.3. X-ray diffraction patterns from a 'cybotactic' nematic: bis-(4'-noctyloxybenzal)-2-chloro-l,4-phenylenediamine: (a) an unaligned (though not
quite randomly oriented) sample; (b) a well aligned sample. (De Vries.(6))
that it is composed of optically active molecules. As a consequence the
structure acquires a spontaneous twist about an axis normal to the
preferred molecular directions (fig. 1.1.4). The twist may be right-handed
or left-handed depending on the molecular conformation. Optically
inactive molecules or racemic mixtures result in a helix of infinite pitch
which corresponds to the true nematic. The energy of twist forms only a
minute part (~10~ 5 ) of the total energy associated with the parallel
alignment of the molecules,(8) so much so that when a small quantity of a

1.1 Thermotropic liquid crystals

Fig. 1.1.4. The cholesteric liquid crystal: schematic representation of the helical
structure.
cholesteric substance,(4) or even a non-mesomorphic optically active
substance,(9) is added to a nematic the mixture adopts a helical configuration. The spiral arrangement of the molecules in the cholesteric is
responsible for its unique optical properties, viz., selective reflexion of
circularly polarized light and a rotatory power about a thousand times
greater than that of an ordinary optically active substance. Cholesterics of
low pitch, less than about 5000 A, exhibit what are known as blue phases.

1. Introduction

(a)

Illllllllllllllllll
Illlllllllllllllll
Illllllllllllllllll

Fig. 1.1.5. Schematic representation of the molecular arrangement in (a) smectic A


and (b) smectic C.
These phases exist over a small temperature range (~ 1 C) between the
liquid crystal phase and the isotropic liquid. Their structures will be
discussed in 4.8.
Smectic liquid crystals have stratified structures but a variety of
molecular arrangements are possible within each stratification. In smectic
A the molecules are upright in each layer with their centres irregularly
spaced in a 'liquid-like' fashion (fig. I A.5 (a)). The interlayer attractions
are weak as compared with the lateral forces between molecules and in
consequence the layers are able to slide over one another relatively easily.
Hence this mesophase has fluid properties, though it is very much more
viscous than the nematic. Smectic C is a tilted form of smectic A, i.e., the
molecules are inclined with respect to the layer normal (fig. 1.1.5(6)).
Several polymorphic forms of smectics A and C exist, as we shall see later.
Over a dozen other distinct smectic modifications have been identified/ 10 ' U)
Some of them (e.g., SB, S E , SG, SH, ST and SK) have three-dimensional longrange positional order as in a crystal, though with weak interlayer forces
(and hence energetically weak interlayer ordering), while some others,
referred to as hexatic phases, have three-dimensional long-range 'bondorientational' order, but without any long-range positional order. (12) The

1.1 Thermotropic liquid crystals

Fig. 1.1.6. (a) Threads in a nematic liquid crystal. Crossed polarizers. Film
thickness ~ 100 jum. (b) Schlieren texture in a nematic film of thickness ~ 10 jum.
Crossed polarizers. (Sackmann and Demus. (14))

1. Introduction

D phase has a cubic structure(13) and would appear to be an exception to


the rule that smectics have layered structures.
The energy required to deform a liquid crystal is so small that even the
slightest perturbation, caused say by a dust particle or a surface
inhomogeneity, can distort the structure quite profoundly. Thus when a
liquid crystal is taken between glass plates and examined under a polarizing
microscope one rarely sees the familiar interference figures expected from
the equilibrium structures shown in figs. 1.1.1,1.1.4 and 1.1.5. Instead, one
usually obtains a rather complex optical pattern. For example, a nematic
film shows a characteristic threaded texture from which this mesophase
derives its name (fig. 1.1.6) and a smectic A film a focal conic texture (fig.
1.1.7). These textures are useful in the optical identification of the
mesophases (1516) and their nature is generally well understood, as we shall
see in later chapters. 'Single crystal' films with the molecules aligned
perpendicular to the plates (homeotropic structure) or parallel to them
(homogeneous or planar structure) can be prepared by suitable prior
treatment of the glass surfaces.
From purely geometrical arguments, Herrmann (17) concluded that there
should be 18 distinct mesomorphic groups between the perfectly ordered
crystalline arrangement and the truly amorphous one. Examples of some
of these groups have been found in plant virus preparations (18) and in
surfactant-water compositions,(19) but it is not yet clear whether all of
them can give rise to energetically feasible configurations. Thus Friedel's
nomenclature offers a convenient basis for the classification of thermotropic liquid crystals and is universally adopted at present. The term
calamitic has come into use in recent years to describe liquid crystals
composed of rod-like molecules and to distinguish them from discotic
systems, which will be discussed in the next section.

1.1.2 Liquid crystals of disc-like molecules


The first liquid crystals of disc-shaped molecules, now generally referred to
as discotic liquid crystals, were prepared and identified in 1977.(20) Since
then a large number of discotic compounds have been synthesized and a
variety of mesophases discovered.<21) Structurally, most of them fall into
two distinct categories, the columnar and the nematic. The columnar phase
in its simplest form consists of discs stacked one on top of the other
aperiodically to form liquid-like columns, the different columns constituting a two-dimensional lattice (fig. 1.1.8 (a)). The structure is somewhat
similar to that of the hexagonal phase of soap-water and other lyotropic

1.1 Thermo tropic liquid crystals

Fig. 1.1.7. Focal conic textures in smectic A. (a) The polygonal texture. Crossed
polarizers. (Friedel.(4)) (b) Simple fan-shaped texture. Crossed polarizers. (Sackmann and Demus.(14))
systems (see fig. 1.2.2), but a number of variants of this structure have been
identified: hexagonal, rectangular, tilted, etc. The nematic phase has an
orientationally ordered arrangement of the discs without any long-range
translational order (fig. 1.1.8(6)). Unlike the classical nematic of rod-like
molecules, this phase is optically negative. A cholesteric (or twisted
nematic) phase has also been identified. A smectic-like lamellar phase has
been reported(22) but the disposition of the molecules in the layers has not
yet been resolved.

10

1. Introduction

(a)

Fig. 1.1.8. The structure of (a) the columnar phase and (b) the nematic phase of
disc-like molecules.

(b)

Fig. 1.1.9. Polymers that exhibit liquid crystalline phases. The basic monomer units
are low molar mass rod-like or disc-like mesogens which may form part of the main
chain (a) or attached as side groups (b).
1.1.3 Polymer liquid crystals
The structures of polymers that form liquid crystals are illustrated
schematically in fig. 1.1.9. The basic monomer units are low molar mass
mesogens, rod-like or disc-like, which are attached to the polymer
backbone in the main chain itself (fig. 1.1.9 (a)), or as side groups (fig.
1.1.9(6)). The nature of the mesophase depends rather sensitively on the

1.1 Thermotropic liquid crystals

11

(a)

to

Fig. 1.1.10. Mesophases of discotic polymers: (a) the hexagonal columnar phase.
The diagram illustrates intercolumnar binding as well as intracolumnar backfolding of the main chain (Herrmann-Schonherr, Wendorff and Ringsdorf (24) );
(b) the' sanidic' nematic phase composed of boards stacked parallel to one another
(Herrmann-Schonherr et al.{25)); (c) the columnar nematic phase (Ringsdorf
et al (26) ).

backbone, the mesogenic unit and the spacers. With rod-shaped repeating
units, mesophases similar to the nematic, cholesteric and smectic types of
rod-like molecules are observed(23) whereas with disc-shaped repeating
units, some new kinds of mesophase structures have been found. For

12

1. Introduction

example, a polyester with (disc-shaped) triphenylene as the repeating unit


in the main chain separated by flexible spacers forms a hexagonal columnar
structure (24) (fig. 1.1.10 (a)). On the other hand, rigid aromatic polyamides
and polyesters with disc-shaped units in the main chain exhibit a different
type of mesophase which has been described as a ' sanidic' (or board-like)
nematic,(25) the boards stacked parallel to one another as shown in fig.
1.1.10(6). The addition of electron acceptor molecules to discotic polymers
results in the formation of charge transfer complexes which stabilize - or in
certain non-mesomorphic materials, induce-mesophases. (26) A new type
of induced mesophase identified in such a system is the' columnar nematic'
illustrated in fig. 1.1.10(c).

1.2 Lyotropic liquid crystals


Lyotropic liquid crystals are made up of two or more components. (27)
Generally, one of the components is an amphiphile (containing a polar
head group attached to one or more long hydrocarbon chains) and another
is water. A familiar example of such a system is soap (sodium dodecyl
sulphate) in water. As the water content is increased several mesophases
are obtained. The types of molecular packing in these mesophases are
represented schematically in figs. 1.2.1 and 1.2.2, but several modifications
of these structures exist.(28)
In the lamellar or neat phase, water is sandwiched between the polar
heads of adjacent layers, while the hydrocarbon tails, which are disordered
or in a liquid-like configuration, are in a non-polar environment (fig. 1.2.1).
In the cubic or viscous isotropic phase the layers are bent to form spherical
units, the polar heads lying on the surface of the sphere and the
hydrocarbon chains filling up the inside. The spherical units form a bodycentred cubic arrangement, (29) water taking up the space between the units.
In the hexagonal or middle phase, the layers are rolled up into cylinders.
The cylindrical units of indefinite length are arranged parallel to one
another in a hexagonal array (fig. 1.2.2) rather like the columnar phase of
disco tics. A nematic type of ordering has also been observed in some soap
systems.(30) It is believed that there exists a cylindrical superstructure
similar to the hexagonal phase, but this does not appear to have been
proved conclusively.(31)
In hydrophobe-dominated compositions, such as aerosol OT-water
systems,(32) inverted middle and inverted viscous isotropic phases can occur
in which the tails point outward towards the hydrophobic medium while
water is trapped inside.(33)

1.2 Lyotropic liquid crystals

13

Fig. 1.2.1. The lamellar or neat phase of soaps.

Fig. 1.2.2. The hexagonal or middle phase of soaps.


Two mesophases may coexist over small ranges of composition and
temperature. Phase diagrams have been constructed for a large number of
binary systems. Transitions may be brought about from any one of the
mesophases directly to the isotropic solution at appropriate temperatures.
Ternary and higher-component systems exhibit essentially the same types
of structures, the phase diagrams being of course much more complicated.
Cholesteric liquid crystals are formed by solutions of synthetic polypep tides, e.g., poly-y-benzyl-L-glutamate, in organic liquids when the
concentration exceeds a certain critical value. (34)

14

1. Introduction

Lyotropic liquid crystals occur abundantly in nature, being ubiquitous


in living systems.(35) Their structures are quite complex and are only just
beginning to be elucidated. However, in this monograph we shall be
confining our attention mainly to the physics of low molecular weight
thermotropic liquid crystals and do not propose to discuss polymer and
lyotropic systems in any further detail. In chapters 2-5, we deal with the
nematic, cholesteric and smectic mesophases of rod-like molecules and in
chapter 6 discotic systems.

1.3 Polymorphism in thermotropic liquid crystals


Extensive tables are now available listing practically all known mesogenic
compounds, their transition temperatures and latent heats. (336) We
therefore give just a few typical examples of rod-like mesogens to illustrate
the rich polymorphism exhibited by some of these thermotropic materials.
The latent heats (in kilojoules per mole) are also presented for those cases
for which data are available. Three nematogenic compounds (1.3.1)(1.3.3)
are included in the list for a special reason: the first has a permanent electric
dipole moment inclined at a large angle (~ 60) with respect to the long
molecular axis, the second a strong longitudinal dipole moment, while the
third is non-polar. This brings out the fact emphasized earlier that it is the
molecular shape that plays the major role in determining mesomorphic
behaviour. All the transitions listed here, except one, are enantiotropic, i.e.,
they take place reversibly on heating and cooling, though the reversal to
the solid phase is usually accompanied by supercooling. The one exception
is cholesteryl nonanoate (1.3.5) which shows a monotropic transition - the
smectic A phase occurs only on cooling. Examples of disc-like mesogens
will be given in chapter 6.
A considerable body of experimental data has been accumulated over
the years in regard to the relationship between mesomorphism and
chemical constitution, especially for rod-like molecules, but we shall not be
discussing this aspect of the problem. For a concise review of current
trends in the chemistry of mesogenic compounds, with emphasis on the
special requirements for technological applications, reference may be
made to a recent book edited by Gray.(37)

1.3 Polymorphism in thermotropic liquid crystals

15

4, 4'-Di-methoxyazoxy benzene (/?-azoxyanisole/ ^

CH.O

(1.3.1)

OCH,

...
118.2 C
solid -*

29.57 kJ

.
135.3 C
.
nematic -*
isotropic
0.57 kJ

4'-n-Pentyl-4-cyanobiphenyl'I (39)

(1.3.2)
111

solid

22.5 C

17.2 kJ

nematic -*

401 C

35 r

/ /

isotropic

445 C
nematic -*

isotropic

7?-Quinquephenylv

solid -*

o.8 kJ

(1.3.3)

4'-n-Octyl-4-cyanobiphenyl (39)

C8H17-

solid -*

24 C
25.3 kJ

(1.3.4)

t\

34 C
+.
A
smectic A -*

0.13 kJ

.
42.6 C
nematic -*

0.97 kJ

isotropic

16

1. Introduction
Cholesteryl nonanoate (monotropic smectic A) (41>42)

O
CH 3 (CH 2 ) 7

78.6 C
91 C
solid -
cholesteric -*
0.017 kJ

91.3 C
blue phase I -*
0.017 kJ

blue phase II

75.5 C 0.25 kJ

91.45 C 0.53 kJ

smectic A

isotropic

(1.3.5)

N-(4-n- Pentyloxybenzylidene)-4'- n- hexylaniline,(43)

C5HnO-

36 C
solid -*

15.2 kJ

-CH =

-C6H13

38 4 C
42.4 C
smectic G -*:
smectic F -*

0.39 kJ
0.15 kJ

smectic B

50 C 25 kJ

isotropic ^ 7 2 - 8 c
0.66 kJ

nematic
0.42 kJ

smectic A ~m smectic C
very
small

(1.3.6)

2
Statistical theories of nematic order

2.1 Melting of molecular crystals: the Pople-Karasz model

X-ray analyses of the crystal structures of typical nematogenic compounds(13) have established that the long narrow molecules are more or
less parallel and interleave one another to form what was described by
Bernal and Crowfoot(1) as an 'imbricated' arrangement (fig. 2.1.1). The
transformation from the solid to the nematic phase is characterized by the
breakdown of the positional order of the molecules but not of the
orientational order. The mesophase is fluid and at the same time
anisotropic because of the facility with which the molecules slide over one
another while still preserving their parallelism. The degree of orientational
order in the liquid crystal drops gradually on heating, whereas certain
other thermodynamic properties, such as the specific heat, thermal
expansion and isothermal compressibility, increase rapidly as the temperature approaches the nematic-isotropic point T^r At Tm a 'weak' first
order (discontinuous) phase transition takes place, accompanied by the
complete breakdown of the long-range orientational order. The changes of
entropy and volume attending this transition are typically only a few per
cent of the corresponding values for the solid-nematic transition.
In this section, we shall consider an elementary model(4"6) which
illustrates the principal mechanisms responsible for this unusual type of
melting phenomenon.
2.1.1 Order-disorder in positions and orientations

The simplest treatment of the melting of inert gas crystals is that due to
Lennard-Jones and Devonshire(7) (LJD) who regarded the mechanism of
fusion as a positional order-disorder phenomenon. They postulated that
17

18

2. Statistical theories ofnematic order


cJ2

Fig. 2.1.1. Molecular arrangement in the crystalline phase of anisaldazine: (a)


projection of the structure along [010]; [b) projection along [001]. Crystals are
monoclinic with a= 17.46 A, c = 8.45 A, /? = 113 48'. Space group Cc with
four molecules per unit cell. (After Galigne and Falgueirettes.(3))
the atoms may occupy sites on one of two equivalent interpenetrating
lattices, which we shall refer to as ^4-sites and 2?-sites. The lowest energy
configuration (corresponding to the state of perfect order) is that in which
all atoms occupy the same kind of sites, say the ^4-sites. With increasing
temperature, the number of atoms in the interstitial 2?-sites increases till a
critical stage is reached when there is complete collapse of the long-range
order and both kinds of sites become equally populated. The system in
which there is an equal number of occupied and unoccupied ^4-sites
corresponds to a liquid, for under such circumstances migration from one
site to another becomes easily possible. By assuming an appropriate
volume dependence for the energy of interaction between atoms in A and

2.1 Melting of molecular crystals

19

Table 2.1.1. Melting entropy of molecular crystals(8) (in entropy units,


AS/R)
Inert gas crystals

Ne
3.26
Kr
3.36
A
3.35
Xe
3.40
Crystals in which the rotational transition precedes the melting transition
O2
2.0
PH 3
1.92
SiH4
1.8
CD 4
2.42
CF 4
2.0
CH 4
2.47
Crystals in which the two transitions coalesce
CO 2
9.25
SiCl4
9.06
N2O
8.58
HCN
7.73
C2H4
7.70
CS 2
6.51
SO2
8.95
(CN) 2
7.90

2?-sites, Lennard-Jones and Devonshire showed that the melting parameters predicted by the theory agree well with the data for a number of
spherical or nearly spherical molecules.
However, the theory fails for anisotropic molecules where the effects of
orientational disorder become important. The thermodynamic data on
melting suggest that there are two classes of molecular crystals: those
which undergo phase transitions associated with rotational motions at
temperatures below the melting point and those in which the rotational
and melting transitions coalesce. The former have entropies of fusion
lower than the inert gas crystals, while the latter have much higher
entropies of fusion (table 2.1.1).
Pople and Karasz ( 4 ) proposed a simple extension of the L J D model
which at once provides an interpretation of these variations in the melting
properties. They assumed that the molecule can take u p one of two
orientations on any site, so that it now has four possibilities, Ax, A2, B1 and
B2. The state of perfect order (or the solid at zero temperature) may then
be regarded as one in which all molecules occupy sites and orientations of
the same type, say A19 and the state of complete disorder (or the liquid
phase) as one in which all four configurations are equally populated.
Clearly, there can also be states with positional order and no orientational
order and vice versa.
When a molecule is turned into an unfavourable orientation, local
strains are set up and consequently there is an increased tendency for the
molecules in the vicinity to move to interstitial sites. T o allow for this

20

2. Statistical theories ofnematic order

coupling between orientational and positional ordering, Pople and Karasz


made the simple assumption that the orientational component of the AB
interactions is negligible compared with that of the A A or BB interactions,
so that the B-sites near a misorientation in an ^4-lattice are favoured.
Accordingly the energy required for a molecule to diffuse to an interstitial
5-site in an ^4-lattice is determined only by the AB interactions regardless
of orientation, and that required for a molecule to assume an A2orientation in an ^-lattice (or a i?2-orientation in a ^-lattice) is
determined by the AXA2 (or BXB2) interactions only.
Let W be the energy of an AB interaction and W the orientational
energy of an AXA2 or B1B2 interaction. We neglect interactions between
more distant neighbours. If for any configuration of the system of N
molecules, NAB is the number of neighbouring AB pairs, NA A the number
of relative misorientations on neighbouring ^4-sites and NB B is similarly
defined, the partition function for the whole assembly is
N

z=f n,

(AU W+NAiA% W + NBxB% W')/kBTl


where the sum is over all configurations taking into account all possible
arrangements in positions and orientations;/is the partition function per
molecule in a state of perfect order which is treated as a function of volume
per molecule and temperature only, and kB is the Boltzmann constant.
Approximate solution of the cooperative order-disorder problem
We define the degree of positional order as q = 2J 1, where J is the
fraction of molecules in the ^4-sites, and the degree of orientational order
as s = 2f' 1, where Sf is the fraction of molecules in 1-orientations. If TV
is the total number of molecules in the system, there are evidently
\N(\ -Vq)(\ +s) molecules in Ax positions,
\N{\ +q)(\ s) molecules in A2 positions,
\N(\ q){\ +s) molecules in B1 positions,
\N(\ q)(ls) molecules in B2 positions.
Let there be z 5-sites adjacent to each ^4-site and z^-sites closest to each
^4-site. This implies, of course, that there are also z ^4-sites adjacent to each
B-site and z'5-sites closest to each i?-site.
Applying the Bragg-Williams (or zeroth order) approximation,(9)
the configurational partition function

2.1 Melting of molecular crystals

21

where
r(q,s) = \

la
1UJV(1 - ? ) ( 1 +^)]![17V(1 Using Stirling's approximation

n+s
2 )

W(\-q
KT\
4
B
J

kBr{

K1AA)

To evaluate the contribution of the disorder to the various thermodynamic


quantities it is necessary to specify the dependence of PFand W on volume.
It is found(5)6) that the empirical trends in the properties, especially for
large orientational barriers which is the main region of interest in the
present discussion, are reproduced well by the model when W
W0(V0/vy and W = WO(VO/V)Z where the suffix zero denotes the value
corresponding to the equilibrium intermolecular separation as defined by
the Lennard-Jones potential. When W = 0, the theory reduces identically
to the LJD model for spherical molecules. Applying the equilibrium
conditions,
8 In Q 8 In Q
=
= 0,
dq
ds
we obtain

-(H-FWr)*

<2U)

where v = z'W'JzWQ is a measure of the relative barriers for the rotation


of a molecule and for its diffusion to an interstitial site.

22

2. Statistical theories ofnematic order

The Helmholtz free energy may be conveniently expressed as F =


F' + F'\ where F' = NkBT\nfis the contribution due to the perfectly
ordered system and F" = kBT\nQ is that due to disorder. The pressure
due to disorder/' = -@F"/dV)T, so that from (2.1.1),
7w

NkBT

vx

kBT

where q and s are now the equilibrium values determined by (2.1.2)


and (2.1.3). If/?' is the ordered part of the pressure, the total pressure
p = pf +//'. Extensive calculations of p' V'/'NkB Tand other thermodynamic
parameters have been carried out as functions of V/ Vo and kB T/s for a
face-centred cubic lattice in terms of the Lennard-Jones 612 potential, e
being the minimum energy of interaction of a pair of particles.(10) Using
these values and (2.1.4) the complete isotherm can be obtained. In the
actual calculations the ratio WJe was adjusted empirically to give the
correct melting temperature of argon. The parameters used were z = 6 for
an interpenetrating face-centred cubic lattice, and W^/e = 0.977. The
melting properties of a range of materials can then be investigated as the
non-dimensional parameter v increases from zero to a high value.

2.1.2 Plastic crystals and liquid crystals


Hereafter q and s will be understood to refer to those values that satisfy
(2.1.2) and (2.1.3). For low zW/kB T, q = s = 0 is the only solution that
minimizes the free energy for any given kB T/e. The behaviour for higher
zW/kB T depends critically on the strength of the orientational barrier. The
variation of the equilibrium values of q and s with zW/kBT for three
typical values of v and the corresponding isotherms are shown in figs.
2.1.2-2.1.4. The sigmoid portions of the isotherms represent phase
transitions, i.e., the two phases will be in equilibrium at a given pressure
when the areas enclosed above and below the pressure line are equal. For
small v (less than about 0.3), the theory predicts two transitions, a solid
state rotational transition followed by a melting transition. The intermediate phase corresponds to the orientationally disordered crystal,
sometimes referred to as the plastic crystal For v between 0.3 and 0.975,
the two transitions coalesce and there occurs a single transition with a
much greater entropy of fusion. For v > 0.975 there are again two
transitions; the positional melting now precedes the rotational melting,
and there occurs an intermediate phase similar to the nematic liquid crystal

2.1 Melting of molecular crystals

23

-0.2
-0.4
15

zW/kBT

Fig. 2.1.2. (a) Variation of the equilibrium value of the positional order parameter
q and the orientational order parameter s with zW/kB rfor v = 0.3. (b) Theoretical
isotherm for v = 0.3 showing rotational transition in the solid state followed by the
melting transition. (The isotherm is drawn for the solid rotational transition
temperature kB T/e = 0.593.)
L (b)

Fig. 2.1.3. (a) Variation of the equilibrium values of q and s with zW/kB T for
v = 0.8. (b) Theoretical isotherm for v = 0.8 showing a single transition in which
both positional and orientational orders collapse simultaneously (k B T/e = 0.625).
with orientational order but no translational order. (Over a small range of
v, 0.975 < v < 1.047, the theory predicts a second order (continuous)
nematic-isotropic transition, whereas experimentally it is found that this
transition is always discontinuous, though weakly so. This limitation of the
model probably arises from the restriction that the molecule can take up
only two orientations. An extension of the treatment to more than two
discrete orientations has been given by Amzel and Becka (11) for low v.)

2. Statistical theories ofnematic order

24

1.0

q, s

Fig. 2.1.4. (a) Variation of the equilibrium values of q and s with zW/kB T for
v = 1.3. (b) Theoretical isotherm for v = 1.3 showing solid-nematic and nematicisotropic transitions. (The isotherm is drawn for the solid-nematic transition
temperature kB T/e = 0.678.) (After reference 5.)

0.6

0.4

0.2

0.4

0.8

1.2

Fig. 2.1.5. Reduced transition temperature versus v. S-S, solid-solid; S-I,


solid-isotropic liquid; S-N, solid-nematic; N-I, nematic-isotropic. (After
reference 6.)

2.1 Melting of molecular crystals

25

a-1

3 --

\S-N

2 -

S-I

'

1 -

s-s
0 -

0.2

N-I/
i

0.4

0.6

0.8

1.0

1.2

Fig. 2.1.6. Entropy of transition versus v (see legend of fig. 2.1.5).


The entropy S, specific heat cv and isothermal compressibility p can be
evaluated from the thermodynamic relations:
S" = -

ZT);
\

+ 2 T ( \ ]

)/2T{ aril*

Figs. 2.1.5-2.1.7 give the reduced temperatures of transition and the


corresponding entropy and volume changes as functions of v. For a certain
range of v, AS/R and AV/Vfor the nematic-isotropic transition are only
very small fractions of the values for the solid-nematic transition. This is
indeed a distinctive feature of such transitions in general (table 2.1.2).
Curves for the order parameter derived from the theory are presented in
fig. 2.1.8(a) and as can be seen they are closely similar to the experimental
results (fig. 2.1.8(6)). The theoretical curves(5) for the thermal expansion,
specific heat and isothermal compressibility are also found to reproduce
quite well the observed trends in the properties of the nematic phase/ 14 ' 15 ' 20)

2. Statistical theories ofnematic order

26

Table 2.1.2. Latent heat and volume change data for some nematogenic
compounds
Solid-nematic
transition
Latent
heat
(kJ mole-1)

Compound
4-(4-n-Propylmercapto
benzalamino)azobenzene(12)
/?-Azoxyanisole
/7-Azoxyphenetole
4- Methoxy benzy lidene4/-butylaniline

31.30

29.57(14)
26.87 (14)
18.06 (17)

Nematic-isotropic
transition

Volume
change

AV/V(%)

5.07
11.03 (13)

84d6)

Latent
heat
(kJ mole-1)

Volume
change

AV/V(%)

0.368
0.574(14)
1.366(14)

0.09
0.36(15)
0.60(16)

0.582(17)

0.40(18)

S-N

0.2

0.1

s-s/

n
1

N-I /
^
1

0.4

0.8

1.2

Fig. 2.1.7. Relative change of volume on transition versus v (see legend offig.2.1.5).

Thus while a simple model of this type cannot be expected to be applicable


in detail to any particular substance, it does serve to bring out the effects
of orientational disordering on the thermodynamics of the melting process.
However, as near-neighbour correlations are neglected completely in the
Bragg-Williams approximation, the theory is unable to account for the
specific heat and other anomalies in the isotropic phase. An attempt has
been made to extend the theory by using the quasi-chemical or first

27

2.1 Melting of molecular crystals

Fig. 2.1.8. Orientational order parameters versus temperature: (a) theoretical


curves for v = 1.15,1.18 and 1.20; (b) experimental curves of s denned by (2.3.1) for
/7-azoxyanisole (PAA), anisaldazine and /?-azoxyphenetole (PAP).(19) The dashed
portions of the curves represent supercooled regions. (After reference 5.)

(b)

0.2

()

Solid /Nematic/Isotropic

pVo

/ iNemat

0.2 -

0.1

"T^~

Solid /
/ Isotropic
/ /
0.1 -

/
/

0.67

0.68

0.69

kBT/e

0.70

0.6

/
i

0.7
kBT/e

0.8

Fig. 2.1.9. Theoretical variation of the transition temperatures with pressure:


(a)v= 1.15; (b)v = 0.95.
approximation.(21) This results in a very weak anomaly in the specific heat
above 7^I? but not in the thermal expansion or isothermal compressibility.
It is clear that the model is far too elementary for describing these and
other short-range order effects in the isotropic phase (see 2.5).
2.1.3 Pressure-induced mesomorphism
Theoretical phase diagrams for two representative values of v are illustrated
in fig. 2.1.9. In fig. 2.\.9{a) the orientational barrier is large enough for the
nematic phase to occur at zero pressure. As the pressure is raised, both the
solid-nematic and the nematic-isotropic transition temperatures increase,

2. Statistical theories ofnematic order

28

500

400

Solid
A

Nematic

/
0/

Isotropic

300

200

100

390

400

410
T(K)

430

Fig. 2.1.10. Experimental phase diagram for PAA. (After reference 22.)

the slope AT/dp for the latter being greater in accordance with the
experimental data (fig. 2.1.10). Of course, such a behaviour is also to be
expected from the Clausius-Clapeyron equation.
Fig. 2.1.9(6) represents a more interesting case. Here v is just below the
critical value for the nematic phase to occur at zero pressure. As the
pressure is raised, there is initially a single transition, viz, the solid-liquid
melting transition, but at higher pressures branching takes place and
there are two transitions. The branching point represents the solidnematic-isotropic triple point.
Experiments were done on the first two members of the p-nalkoxybenzoic acid series to verify this prediction. (22) These two compounds, methoxy- and ethoxybenzoic acids, do not form liquid crystals
at atmospheric pressure, whereas propoxybenzoic acid and the higher
homologues show at least one liquid crystalline phase. As the pressure is
raised, both compounds exhibit mesophases, initially a nematic phase and

2.2 Phase transition in a fluid of hard rods

200 -

Solid

1/

29

/isotropic

' Smectic / Nematic

7k

100 -

Triple point tS
^
7

/
/

Triple point
i

205

215

225

Temperature ( C)

Fig. 2.1.11. Experimental phase diagram for /?-ethoxybenzoic acid showing


the solid-nematic-isotropic and solid-smectic-nematic triple points. (After
reference 22.)

then, at higher pressures, a smectic phase as well (fig. 2.1.11). (The smectic
phase is not included in the framework of the theory in its present form. It
would require the introduction of another order parameter, viz, a lamellar
order parameter (see chapter 5).) The transitions were detected by
differential thermal analysis and the mesophases identified by microscopic
observations using a high pressure optical cell (fig. 2.1.12). These studies
also established for the first time the existence of the solid-nematicisotropic and the solid-smectic-nematic triple points in single component systems.
2.2 Phase transition in a fluid of hard rods

We have emphasized that the asymmetry of the molecular shape is an


important factor in determining whether or not a substance exhibits the
liquid crystalline phase. Consider a fluid of long thin rods without any
forces between them other than the one preventing their interpenetration.

30

2. Statistical theories ofnematic order

Fig. 2.1.12. Pressure-induced mesophases in /?-ethoxybenzoic acid: (a) nematic


schlieren textures; (b) focal conies and batonnets in the smectic phase.
(Reference 22.)

2.2 Phase transition in a fluid of hard rods

31

At sufficiently low densities the rods can assume all possible orientations
and the fluid will be isotropic. As the density increases, it becomes
increasingly difficult for the rods to point in random directions and
intuitively one may expect the fluid to undergo a transition to a more
ordered anisotropic phase having uniaxial symmetry. That this does indeed
happen was first proved by Onsager.(23) Other hard rod theories have since
been proposed, which we shall refer to briefly at the end of this section, but
Onsager's approach, based on an exact density expansion for the free
energy, is still probably the most satisfying. However, the mathematical
analysis involved in the theory is rather complex - even in the lowest order
it leads to a non-linear integral equation. Truncation of the series after the
linear term, as was done by Onsager, would appear to be satisfactory only
for very long rods and not for shorter ones, but attempts to evaluate the
coefficients in the expansion beyond that of the linear term have so far been
unsuccessful because of computational difficulties.
Zwanzig(24) proposed a simplified version of Onsager's theory which
enables the virial coefficients to be evaluated to a much higher order. The
simplification consists in the following assumptions: (a) the rods can take
up only discrete orientations along three mutually perpendicular axes,
whereas in the original theory they can point in any direction, and (b) the
length to breadth ratio of the rods is very large (tending to infinity).
Though these assumptions may be somewhat too drastic as far as real
liquid crystal systems are concerned(25) the model does serve to illustrate
the essential principles of Onsager's method and may prove to be useful as
a starting point for possible extensions and applications of the general
theory.
In Zwanzig's model the rods are rectangular parallelopipeds of length /
and square cross-sectional area d2. Assuming that the potential energy of
interaction of TV such hard rods is a sum of pair potentials, we may write

UN = \tuij9

(2.2.1)

where Utj is oo if the molecules / and j intersect and zero if they do not. The
intersection may occur either in a parallel or perpendicular configuration
as shown in fig. 2.2.1. Let the position vector of they'th rod be represented
by Rj and its orientation by up and let the set of all positions be abbreviated
by R and the set of all orientations by u. The configurational integral for
the system consisting of TV rods in the volume V can then be expressed as

QN(T, V) = ~ -L JdR exp ( - UN/kB T),

32

2. Statistical theories of nematic order

where TV! allows for the indistinguishability of the particles and 3^ for the
total number of orientational states. The potential energy of the system
depends on the sets R and u. Since the configurational space available for
each molecule is V,
I dR exp ( - UN/kB T) = VN exp [ - <pN(u)/kB T],
where the new function <pN depends only on the orientations. Thus
tT}.

Now for any given configuration of the system, there will be N(\) molecules
pointing in the direction u(l), N(2) in the direction u(2) and N(3) in the
direction u(3). Therefore q>N becomes a function of the occupation numbers
in the three allowed directions, i.e.,
<pN =

<pN[N(l),N(2),N(3)].

The indistinguishability of the molecules requires that for any given set of

occupation numbers N(l), N(2) and N(3) there are N\/N(l)]N(2)\NQ)\


equivalent configurations, so that
=

V*

J^_
N\lN

{?

N )=a N )=o

N\exp(-<pN/kBT)
N(\)\N{2)\N{?)\
'

where N(\) + N(2) + NQ) = N. Putting

) Nil) M3)l

we have

Q=

t[N(l),N(2),N(3)].

iV(l)=0 A^(2)=0 iV(3)=0

(2.2.2)

To evaluate this configurational integral, we make use of the ' maximum


term method' of statistical mechanics,(26) i.e., in the limit N^ oo, V^ oo,
but N/V = p = constant, QN can be approximated by the largest term fmax
in the sum. The problem therefore reduces to one of maximizing t with
respect to the occupation numbers. Using Stirling's approximation and
introducing the mole fraction x a = N(tx)/N of the various components,

N-1]nt=l-]np-]n3-txttlnxa-(<pN/NkBT),

(2.2.3)

2.2 Phase transition in a fluid of hard rods

33

Fig. 2.2.1. Intersection of rods in the parallel and perpendicular configurations.


subject to the condition

To proceed further, we need to know the dependence of the function cpN on


density, which can be obtained by resorting to the virial expansion. Now,
the molecules having different orientations may be regarded as belonging
to different species, so that in effect we have a multicomponent system. The
virial expansion of the free energy of a gaseous mixture in ascending
powers of the density is well known :(27) for our present purpose we use the
form
N(3)]
N(2) JV(3)

(2.2.4)
where the virial coefficient B[N(l), N(2), N(3)] is defined as

/ i s the Mayer function given by


y;, = e x p ( - ^ . / / : B r ) - i ,
{ 1 if / and y intersect,
0 otherwise.

34

2. Statistical theories ofnematic order

is t n e standard abbreviation in the cluster theory of imperfect gases


which signifies the integral over all positions of a sum of products of Mayer
/functions taken over all irreducible clusters (or graphs) involving N{\)
molecules of species 1, N(2) molecules of species 2 and N(3) molecules of
species 3.
The maximum value of t is obtained from the condition
^(N-1lnt)

= 0,

a =1,2,3.

The configurational free energy of the system is given by


F=-kBTlnQN
and the pressure by

Thus by determining fmax, the pressure can be derived as a function of the


density.
Calculation of the virial coefficients
The simplification introduced by restricting the rods to three discrete
orientations will now be apparent. The rods must intersect only in those
three directions, so that the cluster integrals in three dimensions can be
factorized into products of cluster integrals in one dimension. Consider for
example the second virial coefficient for parallel rods. By definition

' 0) =
The first volume integral is evaluated directly so that

We observe that the Mayer function/ 12 is zero everywhere except in the


range / to / in the x direction and d to d in the other two directions. In
the range where/is non-zero, its value is 1. Hence
5(2,0,0) = -4W 2 .

2.2 Phase transition in a fluid of hard rods

35

Similarly, the second virial coefficient for two rods in an orthogonal


configuration
B { 1 l 0 )

ni+d)/2
J-(l+d)/2

r(i+d)/2 ra
J-(l+d)/2

d3R2 = -

J -d

When the rods are very long and very thin, I P d,


2?(l,l,0)^-2/ 2 41+2(<///)],
(2,0,0)

~-4l2d(d/l).

Thus, the virial coefficient for parallel configuration is smaller than that for
orthogonal configuration by an order d/l. Higher order virial coefficients
exhibit the same behaviour. To simplify the calculation Zwanzig neglected
all terms of order d/l, i.e., he confined himself to the limit /-> oo, d-^0,
l2d = constant, thereby reducing considerably the number of coefficients to be considered. It is readily verified that only coefficients of the
type B(m,n,0), B(0,m,n) and B(m,0,n) survive, and that by symmetry
B(m, 0, n) = B(n, 0,ra)= B(0,ra,n) = B(0, n, ra).
Let us now choose the z axis (/ = 3) as a preferred direction. By
symmetry the numbers of molecules in the x and y directions are equal, so
that the mole fractions along the three directions can be expressed in terms
of a single order parameter x:
v v

x2 = x,
x3 = 1 2x.

Therefore, (2.2.4) reduces to


VN

_ sr ^r

D/,^

NkBT
(2.2.6)
The mole fraction x is then determined by
(AT1 In 0 = 0,
ax

i.e.,
2 In

x \
\-2x)

d /
dx\

cpN
NkBT)'

36

2. Statistical theories ofnematic order

For example, in the second virial approximation this yields

The solution x = | corresponding to the isotropic distribution is trivially


satisfied at all densities and in any approximation. Below a certain critical
density, (2.2.7) has only the isotropic root. Above this density two new
roots appear; the one which corresponds to low x gives / max and the others
are discarded. Using (2.2.5), the pressure can then be evaluated as a
function of density or volume. The transition predicted by Onsager's
theory in the second virial approximation is confirmed by this model even
when all virial coefficients up to the seventh are included but the properties
of the anisotropic phase depend rather sensitively on the order of the
approximation. Fig. 2.2.2 gives the isotherm obtained in the sixth
approximation, and it can be seen that the shape of the curve is
characteristic of a first order transition.
A Pade analysis of Zwanzig's model by Runnels and Colvin (28) has
shown that the character of the transition is stable against increase in the
order of the approximation. However, the model does not converge to the
Onsager limit with increasing number of discrete orientations and
Straley<25) has concluded that it does not adequately represent real liquid
crystal systems.
Other hard rod theories
Most statistical treatments of the hard rod system, including Onsager's
theory and Flory's well known lattice model, (29) are valid only for very long
rods with length-to-breadth (l/b) ratios of about 100, typical of polymeric
systems, and are therefore not applicable to low molecular weight nematics
for which l/b is usually between 3 and 5.(30) However, a formulation of the
'scaled particle theory' due to Cotter (31) has proved to be convenient for
evaluating the excess free energy for shorter rods at high densities. This
theory provides a means of deriving approximate expressions for the
chemical potential and pressure of the fluid by considering the reversible
work necessary to insert a scaled particle (i.e., a 'solute' particle which is
a scaled replica of the ' solvent' particles) at an arbitrary point in the fluid.
The method was introduced by Reiss, Frisch and Lebowitz (32) for hard
spheres, first applied to hard rods by Cotter and Martire (33) assuming
restricted orientations (as in Zwanzig's model) and extended by Lasher (34)
for continuous orientations. The theory was later presented in a more
rigorous fashion by Cotter. (31) The transition parameters and the properties

2.2 Phase transition in a fluid of hard rods

37

3.0

2.0

0.2

0.4

0.6

0.8

Volume

Fig. 2.2.2. Theoretical isotherm evaluated from Zwanzig's model in the


sixth approximation showing a first order nematic-i so tropic transition. (After
Zwanzig.(24))
of the anisotropic phase predicted by this model for a system of hard
spherocylinders, i.e., cylinders capped at either end by a hemisphere, are in
reasonable agreement with those of a real system, viz, PAA, for l/b =
2 5 (31,35) Another theory which leads to comparable results, is a method
proposed by Andrews(36) for hard spheres, extended to spherocylinders.(35)
It is based on the idea that the reciprocal of the thermodynamic ' activity'
is merely the probability of being able to insert a particle into the system
without overlapping with other particles. The method enables an extrapolation of the 'exact' results of Monte Carlo and molecular dynamic
simulations in the isotropic phase carried out by Vieillard-Baron(37) and
others(38) for l/b = 2 and 3, to the region of the N - I transition. Other
treatments include the y-variable expansion(39) (where y = P/{\ V0P))9
the application of the density functional theory,(40) the functional scaling
approach of Lee,(41) Monte Carlo simulation studies on prolate and oblate
spheroids,(42) etc. Hard rod models have also been used to study the effect
of molecular flexibility on the density change at the transition,(43) the
contribution of the alkyl end-chain,(44) the properties of mixtures(45) and
the influence of molecular biaxiality.(46)
However, there is a basic difficulty with all hard particle models. Their
properties are essentially 'athermaP and depend entirely on the density.
Thus if one defines a quantity

= (31nr/ain/>).

(2.2.8)

38

2. Statistical theories of nematic order

which is a measure of the relative sensitivity of the order parameter s to


changes in density versus changes in temperature, it follows that y = oo for
a system of hard particles, whereas experimentally y = 4 for PAA (see
(2.3.18)). Clearly, therefore, attractive interactions must also be included
to give a proper description of the nematic phase.
2.3 The Maier-Saupe theory and its applications
An approach that has proved to be extremely useful in developing a theory
of spontaneous long-range orientational order and the related properties
of the nematic phase is the molecular field method, closely analogous to
that introduced by Weiss in ferromagnetism. Each molecule is assumed to
be in an average orienting field due to its environment, but otherwise
uncorrelated with its neighbours. The first molecular field theory of the
nematic state was proposed in 1916 by Born (47) who treated the medium as
an assembly of permanent electric dipoles and demonstrated the possibility
of a transition from an isotropic phase to an anisotropic one as the
temperature is lowered. Though this result is important qualitatively we
need not discuss this particular theory because it is now well established
that permanent dipole moments are not necessary for the occurrence of the
liquid crystalline phase (see (1.3.3)). Moreover, the theory predicts that the
aligned phase should be ferroelectric, which does not appear to be the case
even when the molecules are polar. The most widely used treatment based
on the molecular field approximation is that due to Maier and Saupe. (48)

2.3.1 Definition of long-range orientational order


We begin by defining the long-range orientational order parameter in the
nematic phase. Suppose that the liquid crystal is composed of rod-like
molecules in which (i) the distribution function is cylindrically symmetric
about the axis of preferred orientation n and (ii) the directions n and n
are fully equivalent, i.e., the preferred axis is non-polar. Subject to these
two symmetry properties, and assuming the rods to be cylindrically
symmetric, the simplest way of defining the degree of alignment is by the
parameter s, first introduced by Tsvetkov,(49)
.s = K 3 c o s 2 0 - l > ,

(2.3.1)

where 6 is the angle which the long molecular axis makes with n, and the
angular brackets denote a statistical average. For perfectly parallel
alignment s = \9 while for random orientations s = 0. In the nematic

2.3 The Maier-Saupe theory and its applications

39

phase, s has an intermediate value which is strongly temperature


dependent.
The order parameter can be directly related to certain experimentally
determinable quantities - say, for example, the diamagnetic anisotropy of
the liquid crystal.(50) Let us choose a space-fixed cartesian coordinate
system xyz with z parallel to n. If rj1 and rj2 are the principal diamagnetic
susceptibilities of the molecule referred to its own principal axes, the
average z component of the susceptibility per unit volume in the nematic
phase is evidently
(2.3.2)

where fj = \{n1 + 2n2) and n the number of molecules per unit volume.
Similarly
Therefore
n2).

(2.3.4)

r/1 and n2 can be obtained from the susceptibilities of the solid crystal if the
structure is known. Thus a measurement of xz>Xz o r t n e anisotropy xz~Xx
in the nematic phase gives the absolute value of s. In a similar manner,
s can be related to the optical anisotropy (birefringence (51) and linear
dichroism(50)) though, as is well known, there is a difficulty in this case as
there is no rigorous way of correcting for the effects of the polarization field
in the medium. Approximate methods have been proposed (50>52) which
appear to yield satisfactory results/ 53 ' 54)
Another important tool for investigating molecular order is nuclear
magnetic resonance (NMR) spectroscopy. (55) The principle of this method
is that if there is a pair of protons in a molecule, each spin is coupled to the
external magnetic field H as well as to the dipole field created by its
neighbour. The latter gives rise to a small perturbation of the eigenstates of
the Zeeman Hamiltonian so that the signal appears as a doublet. The
doublet separation is easily seen to be proportional to (3cos 2 # l)/r 3 .
where 9 is the angle which the interproton vector rtj makes with H. In the
isotropic phase xtj can assume all possible orientations with respect to H
and <3cos 2 # 1> vanishes, but in the nematic phase there is a splitting
which is a measure of the order parameter s. The analysis is, of course, not
so straightforward for complex molecules as it requires precise knowledge
of the interproton distances and angles. The quadrupole splitting in the
magnetic resonance spectrum of 14N(56) or of deuterium(57) has also been
employed for studying molecular order. Other methods involve the use of

40

2. Statistical theories of nematic order


0.7 r

0.6

a 0.5
O

6
0.4

0.3

-40

-30

-20

-10

r-rNI(Q
Fig. 2.3.1. The temperature variation of the long-range orientational order
parameter s in PAA. Open circles from NMR measurements (McColl and Shih(66)),
filled circles from diamagnetic anisotropy (Gasparoux, Regaya and Prost(67));
triangles from refractive indices.(65)
NMR (58) or electron spin resonance spectroscopy (59) of geometrically
anisotropic molecules aligned in nematic solvents. We shall not discuss any
of these techniques here as a number of excellent reviews on the subject are
available. (605) In fig. 2.3.1 we present experimental values of the order
parameter of PAA determined by three different methods.
NMR and dielectric measurements (see 2.3.5) indicate that there is a
fair degree of rotational freedom of the molecules about their long axes.
Nevertheless the assumption that the molecules are cylindrically symmetric
is not valid in general and can sometimes lead to errors in the determination
of s.{6S) For a molecule of arbitrary shape, (2.3.1) may be written in the
generalized form(69)
where a,/? = x9y9z refer to the space-fixed axes, ij = x\y\z' refer to the
principal axes of the molecule, /a, jp denote the projection of the unit vectors
i, j along a,/?, and 3a/3, S(j are Kronecker deltas. $"/is symmetric in /, j and in
a,/?. It is also a traceless tensor with respect to either pair, i.e.,
where repeated tensor indices imply the usual summation convention. The
generalization of (2.3.2) and (2.3.3) is

2.3 The Maier-Saupe theory and its applications

41

where ntj is the molecular susceptibility tensor which can, in principle, be


determined from the crystal data.
s$ is a 3 x 3 matrix. When diagonalized the matrix has the form
slx

0
-(Sll

+ s2i.

In the uniaxial nematic case, s n = s22.


If the molecule is flexible more parameters are required to define the
degree of alignment of its different parts. (64) However, in a large part of the
discussion that follows on the theory of the nematic state we shall ignore
all these details and assume the molecules to be cylindrically symmetric
rods.
2.3.2 The mean field approximation
We assume that each molecule is subject to an average internal field which
is independent of any local variations or short-range ordering. Consistent
with the symmetry of the structure, viz, the cylindrical distribution about
the preferred axis and the absence of polarity, we may postulate that the
orientational energy of a molecule
ut oc |(3 cos2 6i\)s,
where 0i is the angle which the long molecular axis makes with the
preferred axis and s is given by (2.3.1).
The exact nature of the intermolecular forces need not be specified for
the development of the theory. However, in their original presentation
Maier and Saupe(48) assumed that the stability of the nematic phase arises
from the dipole-dipole part of the anisotropic dispersion forces. The
second order perturbed energy of the Coulomb interaction between a pair
of molecules 1 and 2 is given by

^oo = i

r>2

Y ( 1 ) jr < 2 )

(2.3.6)
where x$,x$
as

etc. are the components of the transition moments defined

f
_J

4? = f^2)*42)42Vv2)dr(2), etc.,
k J

42

2. Statistical theories ofnematic order

y/^\ y/[2) and EM,EV are respectively the eigenfunctions and eigenvalues of
the unperturbed molecules corresponding to states ju and v, EMV = EM + Ev9
^ 1) ,xj 1) ^ 1) zj 1) are the charges and their coordinates in the first molecule
measured in the space-fixed coordinate system with the origin at the centre
of mass of molecule 1, R = (X, Y,Z) is the position of the centre of mass
of molecule 2, e(*\ x^y^z are the charges and their coordinates in
molecule 2 measured in the coordinate system obtained by translating the
one defined above by XYZ. The second summation in (2.3.6) is over terms
with x9 y and z changed cyclically.
We assume that the rotational motions of the molecules are independent
of their translational motions. Let ^rja)C(1) and f<2y2)f<2) be the two
molecular coordinate systems with their origins at the centres of mass,
coinciding with the long axis of each molecule. We define the Eulerian
angles as follows: 6 = the angle between z and , 0 = angle between and
the normal to the z plane and 0 ' = angle between x and the normal to the
z plane. The components of the transition moments are

where (p is the eigenfunction in the r/C coordinate system. Suppose now


that the nematic medium has cylindrical symmetry about z so that all
values of 0 ' are equally probable, and further that the molecules rotate
freely about so that all values of 0 are equally probable. Setting
<cos#(1)cos#(2)> = 0, i.e., the molecules do not distinguish between 9 = 0
and n, and assuming a spherical distribution of the intermolecular radius
vectors, the orientation dependent part of the potential energy of the /th
molecule can be written in the mean field approximation as

(237)
where M is a function of the intermolecular vector Rik only and
is a measure of the anisotropy of the 0-// transition. Maier and Saupe
expressed (2.3.7) in the simple form
A

/3cos2^-l\

[ j

where V is the molar volume and A is taken to be a constant independent


of pressure, volume and temperature.

2.3 The Maier-Saupe theory and its applications

43

Cotter(70) has examined the postulates underlying the mean field


approximation in the light of Widom's analysis of this general problem and
has concluded that thermodynamic consistency requires that ut should be
proportional to V'1 regardless of the nature of the intermolecular pair
potential. However, in what follows we have assumed a V~2 dependence as
in the original formulation of the theory by Maier and Saupe.
2.3.3 Evaluation of the order parameter
The excess thermodynamic properties of the ordered system relative to
those of the disordered one can now readily be derived on the basis of
(2.3.8). The internal energy per mole

I K<exp(-w</fcBr)d(cos0<)
= -\NkBTBs\
- ^
Jo

(2.3.9)

exp(-ui/kBT)d(cos0J

where B = A/kB TV2 and N is the Avogadro number. The partition


function for a single molecule

ft = f exp ( - ut/kB T) d(cos 0,),


Jo

(2.3.10)

so that the entropy

= -NkBkBs(2s+I)-In

rexp(ffecos20,)d(cos0jj. (2.3.11)

The Helmholtz free energy

F=

U-TS
(2.3.12)

The condition for equilibrium is

IT)
\

or

U l )

-o/V,T

3<cos^)
OS

l=0.

This equation is satisfied when


<cos20,> = <cos20>,
which may be called the consistency relation.

(2.3.13)

2. Statistical theories ofnematic order

44

T>TNJ

0.05 -

yL.

0.0

-0.05

-0.1
i

0.2

0.4
s

0.6

0.8

Fig. 2.3.2. Variation of the free energy with the order parameter calculated from the
Maier-Saupe theory for different values of A/kB TV2. The minima in the curves
occur at values of 5 which fulfil the consistency relation (2.3.13).(71)
The plot of F versus s at different temperatures evaluated from (2.3.12)
is shown in fig. 2.3.2. The minimum of the free energy occurs at that value
of s which satisfies the consistency relation (2.3.13). When T < Tm, there is
only one minimum which corresponds to the stable ordered phase. When
Tis slightly greater than 7 ^ there are two minima, but the one at s = 0, i.e.,
the isotropic phase, represents the absolute minimum. At T = Tm, there
are again two minima, one at s = 0 and the other at s = sc, but the two
states now have equal free energies; at this temperature therefore a
discontinuous transition takes place with no change of volume but an
abrupt change in the order parameter. Putting F = 0 at the transition yields
Vl = 4.541,

(2.3.14)

0.4292,

(2.3.15)

where Vc is the molar volume of the nematic phase at TN1. As an example,


for PAA Tm = 408 K, Vc = 225 cm3, and A = 13.0 erg cm6.
Strictly speaking it is the Gibbs free energy G=U-TS+PV
which

2.3 The Maier-Saupe theory and its applications

45

(b)

0-7

2 0.6

"2
O

0.5

0.4

-40

-30

-10
-20
T-Tm C
Fig. 2.3.3. Order parameter s versus temperature predicted by the Maier-Saupe
theory, (a) ut oc V~2; (b) ut oc V\

must be equated to zero at the transition. This leads to the following


expression for the volume change at Tmm'72)
or

2 In

Bs*X

exp (|fts0 cos2 0) d(cos 6) - Bsc(sc + 1) , (2.3.16)

from which sc can be determined if A V is known from experiment.


However, AV is usually so small (see table 2.1.2) that sc derived by this
method differs hardly by 1-2 per cent from that given by (2.3.15). In effect,
therefore, the theory predicts a universal value of sc 0.44 for all nematic
materials. While this value is in fair agreement with the observed data for
a number of compounds, there are systematic deviations/ 53 ' 60) The
introduction of higher order terms in the Maier-Saupe potential function
to allow for contributions other than the purely dipole-dipole part of the
dispersion energy has been suggested, (71"3) which enables calculations to be
made for specific cases. Further, the long wavelength orientational
fluctuations in the medium (see 3.9) will diminish the effective order
parameter and this has to be taken into account in the theory. (62) The
assumption that the molecule is a rigid rod is also an oversimplification.
Realistic calculations of the role of the flexible end-chain in the ordering
process have been made by Marcelja (74) and by Luckhurst,(75) whose results
will be discussed presently.
Empirical data suggest that the volume dependence of ut may be

2. Statistical theories of nematic order

46

V (cm3 mole- 1 )
224

6.05

6.00

222

220

218

5.38

Fig. 2.3.4. Lines of constant order parameter s versus In Kand In T for PAA. The
slope (aIn T/ein V) = -4.0. (After McColl and Shih.(66))
different from that assumed in (2.3.8). In general, if ut ccVm, where m is a
number, then (2.3.14) gets modified to
TmV = 4.541

(2.3.17)

but sc still has the same value. The rate of variation of s with temperature
does not depend too critically on the exponent m, as the isobaric change of
volume with temperature over the entire nematic range is usually only of
the order of 1-2 per cent (fig. 2.3.3). However, its precise value becomes
rather important in accounting for the effect of pressure on the order
parameter/ 66 ' 76>77) Pressure studies (6676) have established the following
results for PAA:
(i) (8In r / d l n V)s = -4 (fig. 2.3.4).
= -y as defined in (2.2.8).
(2.3.18)
(ii) The thermal range of the nematic phase at constant volume is about
2.5 times that at constant pressure.
(iii) The order parameter sc at the transition point is very nearly
independent of the pressure.
Now, it is clear from (2.3.12) and (2.3.13) that B should have a constant
value if s is to be invariant. If B is to be kept constant while T and V are
varied, as is implied in (2.3.18), TVm should be constant, or
(d\nT/d\nV)s = -

(2.3.19)

Empirically, therefore, m = 4 for PAA. It turns out that with this value of
rn, result (ii) follows at once from the theory; on the other hand with m = 2,

2.3 The Maier-Saupe theory and its applications

47

the thermal range at constant volume is only 0.74 times the experimental
value.(78) The pressure invariance of sc is, of course, a direct consequence of
(2.3.14) or (2.3.17). Thus, the experimental data can be fitted satisfactorily
with m = 4. However, it is doubtful if this has any theoretical significance,
in view of Cotter's result,(70) referred to earlier, that m should be 1
irrespective of the nature of the intermolecular pair potential.
A drawback of the theory becomes apparent when we calculate the
latent heat of transition from the nematic to the isotropic phase. The heat
of transition is easily shown to be given by(48>72)
H = Txl[(a/P)AV-S(Ve,

Tm)],

(2.3.20)

where a and p are respectively the coefficients of thermal expansion and


isothermal compressibility of the isotropic liquid at Tm. (Although AFis
usually small, its contribution to H is not negligible since H itself is rather
small.) However, the theoretical values of//are found to be much too high,
usually by a factor of about 2 or 3.<72) Large discrepancies are also found in
the specific heat Cv and the isothermal compressibility /? in the nematic
phase. The failure of the theory in this respect is evidently because the
molecular field method neglects completely the effect of short-range order.
A naive way of accounting for this discrepancy is to assume that the
molecules form small clusters of 2-3 molecules each, and that the single
particle potential (2.3.8) applies to each cluster rather than to the individual
molecules/ 48 ' 72) The theory is then able to give reasonable values for //, Cv
and /?.(72) However, such a procedure cannot account for the remarkable
effects of short-range order in the isotropic phase (see 2.5).
As observed earlier, the assumption that the molecule is cylindrically
symmetric is clearly not valid for real systems, and consequently the use of
a single order parameter is not adequate. Most molecules are lath-shaped
and have a biaxial character. Therefore two order parameters are required
to describe the uniaxial nematic phase composed of biaxial molecules. (68) If
, rj, C are the principal axes of the molecule ( defining the molecular long
axis), it is necessary to introduce an additional order parameter
/) = f<sin 2 /9cos2^>,

(2.3.21)

where y/ is the Eulerean angle made by with the line of intersection of the
n plane and the xy plane (i.e., the plane normal to the director). A finite
value of D indicates that there is a difference in the tendency of the two
transverse molecular axes to project on the z axis. Theories have been
developed(79) by incorporating an additional term proportional to D in the
Maier-Saupe potential function. It emerges that molecular biaxiality

48

2. Statistical theories ofnematic order

decreases the value of s at the transition, and leads to better agreement with
experiment. Further, since different properties have, in general, different
biaxial anisotropies, the order parameters determined by different techniques (optical anisotropy, diamagnetic anisotropy, etc.) need not
necessarily agree with one another if the biaxial order parameter (2.3.21) is
included in the calculations.
Until the 1970s only s (P2(cos9)} (where P2 is the second order
Legendre polynomial) was accessible experimentally, but a technique is
now available for measuring both </>2) and <i^> (where P4 is the fourth
order Legendre polynomial). It involves polarized Raman scattering
measurements on aligned samples. Jen et al.m) who developed this
technique, made use of the C = N stretch vibration of a cyano-compound
dissolved in the nematic phase as a probe for the Raman measurements. A
number of pure nematogenic cyano-compounds have since been investigated by this method/81'82) It turns out that in every case studied so far,
<P4> is negative for at least part of the nematic range. While the observed
<P2> can be fitted quantitatively by including higher order terms in the
Maier-Saupe potential, the negative <P4> cannot be accounted for
theoretically. Using the experimental value of <P2> and <P4> one can
calculate an orientational distribution function for the molecule :(80)
f(cos6) =
I (even)

j <i>(cos0)>i>cos0

(2.3.22)

truncated after the third term 1 = 4. The results show that the molecule has
a strong tendency to be tipped away from the nematic axis - very much
more than predicted by the Maier-Saupe theory (fig. 2.3.5). In the three
pure cyanobiphenyl compounds studied/81'82) it is found that the greater
the chain length, the greater is the magnitude of the negative <,P4> (fig.
2.3.6). This may possibly provide a clue to the origin of the observed
angular distribution function. However, further studies need to be carried
out before drawing any definite conclusions.
The odd-even effect
From chemical studies it has long been known that the end-chains play a
significant part in the stability of the mesophase. The transition temperature r m (83) and a number of other properties (e.g., the order parameter,
the excess specific heat, the transition entropy, etc.) show a pronounced
alternation as the homologous series is ascended, i.e., as the number of
carbon atoms in the end-chain is increased. This is often referred to as the

23 The Maier-Saupe theory and its applications

49

T-T=2C

3.0

2.4

0.6

-0.3

1.0

0.6

0.8

0.4

0.2

,OQ 6

0.0

cos 6

Fig. 2.3.5. Orientational distribution function /(cos 6) as defined by (2.3.22) in the


nematic phase of MBBA. Circles represent values from Raman measurements and
the line gives the distribution function derived from a two-term Maier-Saupe
potential in which the parameters are adjusted to give a good fit with the observed
<P2(cos#)>. The negative values of/(cos 0) at high angles arise from truncation
errors. (After Jen et al.m))
odd-even effect. Qualitatively, the origin of this effect can be understood
from a consideration of the molecular structure (fig. 2.3.7). In the even
members of this series, the disposition of the end group is such as to
enhance the molecular anisotropy and hence the molecular order, whereas
in the odd members it has the opposite effect. As the chains become longer
their flexibility increases and the odd-even effect becomes less pronounced,
until for long chains it becomes negligible. A quantitative calculation of the
contribution of the end-chain to the ordering has been made by
Marcelja.(74)
The possible conformations of the alkyl chain and the corresponding
internal energies can be worked out accurately, as was shown by Flory. (84)
The valence angle between each successive bond is constant, while the
energy as a function of the azimuthal angle has three minima defining three
possible states for each successive bond, viz, the extended trans state and
two symmetrical gauche states. Marcelja incorporated the configurational

2. Statistical theories ofnematic order

50

-o.i

-0.2

Fig. 2.3.6. Orientational order parameter <P 4(cos#)> in three nematic liquid
crystals from Raman measurements: (a) 4 /-n-pentyl-4-cyanobiophenyl (5CB); (81)
(b) 4/-n-heptyl-4-cyanobiphenyl (7CB);(82) (c) 4/-n-octyloxy-4-cyanobiphenyl
(80CB).(82) (After reference 82.)

Fig. 2.3.7. Structure of 4,4/-di-n-alkoxyazoxybenzenes. The addition of an evennumbered carbon atom in the preferred trans conformation is along the major
molecular axis. This is not the case for an odd-numbered carbon atom.

2.3 The Maier-Saupe theory and its applications

51

statistics of the end-chains in the theory of the nematic phase. In addition


to the conformational energy, each CC bond is subject to a mean field
which depends on the orientational order of the rigid central part of the
molecule as well as that of the end-chain. The detailed calculations yield
results in good accord with the observed trends. An interesting consequence of the theory is that depending on the strength of the interaction
between the rigid part of the molecule and the end-chain, there is a rising
trend in T^ versus n for some homologous series of compounds, and a
decreasing trend for others. This theory has been presented in a more
refined form by Luckhurst, (75) who applied it successfully to molecules
composed of two cyanobiphenyl moieties linked by flexible spacers in
which the odd-even alternation in T^ is as high as 100 C for the first few
members of the homologous series. Some of Luckhurst's results are
presented in fig. 2.3.8.
Binary nematic mixtures
The Maier-Saupe theory has been extended by Humphries, James and
Luckhurst,(85) and more recently by Palffy-Muhoray et al.,m) to investigate
the properties of binary mixtures. The latter treatment (86) has led to some
significant conclusions:
(i) For certain values of the parameters, it is possible to have a
nematic-nematic coexistence region. Coexisting nematic phases have
in fact been observed in mixtures of polymeric and low molecular
weight nematogens,(87) as also in mixtures of rod-shaped and discshaped nematogens. (88)
(ii) The order parameter of the mixture and its variation with temperature
agree very closely with the Maier-Saupe universal curve. However, the
order parameters of the individual components of the mixture may
differ appreciably. (62) The fact that the two components can have
different order parameters is of practical importance in the design of
dye displays. The dye molecules are chosen to have a high anisotropy,
and hence a high order parameter which, in turn, improves the contrast
ratio.
2.3.4 Theory of dielectric anisotropy
As an example of an application of the mean field method we shall consider
the theory of the dielectric anisotropy of the nematic phase. (89>90) The low
frequency dielectric anisotropy of a molecule is determined by two factors:
(i) the polarizability anisotropy aa which for the elongated molecules of
nematogenic compounds always makes a positive contribution (i.e., a

52

2. Statistical theories of nematic order

260 r

220

200

160

120

12

Fig. 2.3.8. The variation of the nematic-isotropic transition 7 ^ with the number of
methylene groups in the flexible spacer for a,o>bis(4,4'-cyanobiphenyloxy)alkanes,
the molecular structure of which is shown at the top of the diagram. The dashed
lines represent the theoretically calculated values. (After Luckhurst.(75))

larger contribution for the measuring field parallel to the long molecular
axis) and (ii) the dipole orientation effect. The sign of the latter contribution
is positive if the net permanent dipole moment of the molecule makes only
a small angle with its long axis and is negative if the angle is large. In the
last case the sign of the net anisotropy depends on the relative magnitudes
of the two contributions. Thus different nematic materials can exhibit
widely different dielectric properties (fig. 2.3.9).
In contrast to diamagnetism where the magnetic interactions between
molecules can be neglected, the polarization field in the medium becomes
important when discussing dielectric anisotropy (see, e.g., Bottcher(92)).
Maier and Meier(89) took this into account by applying the Onsager
theory.(93) The effective induced dipole moments per molecule along and

2.3 The Maier-Saupe theory and its applications

53

5.9 ,- (a)

tn

5.8

I 5.7

s
5.6
140

130

120
7TQ

30

20

10

1
100

120

140

7TQ
Fig. 2.3.9. Principal dielectric constants of (a) PAA (after Maier and Meier(90)) and
(b) n(4/-ethoxybenzylidene) 4-aminobenzonitrile (after Schadt (91)). s1 and e2 refer to
the values along and perpendicular to the optic axis of the nematic medium and is
is the isotropic value.

perpendicular to the unique axis of the nematic liquid crystal are then given
by expressions essentially similar to (2.3.2) and (2.3.3) but with appropriate
correction factors for the polarization field:
(2.3.23)
(2.3.24)

where h = 3e/(2s+ 1) is the cavity field factor, e the mean dielectric

54

2. Statistical theories ofnematic order

constant, F = 1/(1 af) the reaction field factor, a the mean polarizability,
aa the polarizability anisotropy (the direction of maximum polarizability is
assumed to be the long molecular axis),/= 4nNp(2e 2)/3M(2s+ 1), p the
density, M the molecular weight, s the order parameter defined by (2.3.1),
and E the applied field. Strictly speaking F and h have to be corrected for
the anisotropy of the dielectric constant but we shall ignore these
corrections.
To calculate the effective permanent dipole moment, we choose XYZ as
the space-fixed coordinate system, Z being parallel to the unique axis of the
medium, and rj( as the molecule-fixed coordinate system, the axis
coinciding with the long axis of the molecule. Let v be the Eulerian angle
between the axis and the line of intersection of the XYand rj planes, and
V the angle between this line and X. Suppose that the permanent dipole
moment ju is inclined at an angle ft with respect to the long molecular axis.
In dielectric measurements, low fields are usually employed so that the
potential energy of the dipoles due to the external field is small. We can
therefore write the effective dipole moments along the field directions as
,T)]juzy/(0)sin0d0dvdv/

o Jo Jo

[l+(M,hE1/kBT)]yf(ff)sin9d0dvdv'
hE2/kB T)] jux y/{0) sin 0 d6 dv dv'
> (2.3.26)

Jo Jo
Jo Jo
Jo
Jo

[l+(MxhE2/kBT)](S)sm8dvdv'

where
fiz = ///[cos P cos 6 + sinfi sin v sin #],
jux = Fju[cos P sin V sin 6 + sin /?(cos v cos V sin v sin V cos 9)]
and y/(0) d6 is the probability of a molecule having an orientation between 6 and 6 + d6. If y/(0) is given by the Maier-Saupe distribution,
exp( ut/kB T) where ut is defined by (2.3.8), the integrals reduce to
^

ft = ^T"^[l+i(l-3cos a i8)5]AF 2 2 .

(2.3.27)
(2.3.28)

2.3 The Maier-Saupe theory and its applications

55

Since

^f[\-(I-3

cos2 p)s
(2.3.29)

and similarly

s2= i + 4 ^ ^ ^ | a - i a ^ + ^;[l+Kl-3cos 2 y 9)^J. (2.3.30)


Therefore

a = e1-e2 = ^^hFW~^-f(l

-3cos 2 )L.

(2.3.31)

Clearly if /? is small, the two terms in the square brackets of (2.3.31) add to
give rise to a strong positive ea, whereas if /? is sufficiently large a may be
negative. In PAA, for example, ju and /? are estimated to be 2.2 debyes and
62.5 respectively from Kerr constant measurements in dilute solutions
and dielectric measurements in the isotropic phase. Substituting for the
other known parameters for this compound and using s from theory, it
turns out that ea should be weakly negative. The absolute value of ea as well
as its temperature variation are in fair agreement with the experimental
data.(90)
The dielectric constants are, of course, frequency dependent.(94) The
dipole orientation part of the polarization parallel to the preferred
direction (1-direction) may be expected to be characterized by a relatively
long relaxation time. This arises because of the strong hindering of the
rotation of the longitudinal component of the dipole moment about a
transverse axis. On the other hand the orientational polarization along the
2-direction will have a much faster relaxation, comparable to the Debye
relaxation in normal liquids, as this involves rotation about the long axis
of the molecule. If there are additional dipoles in parts of the molecule,
with their own internal degrees of rotation, the corresponding relaxation
times will again be similar to that in a liquid. The expected form of the
dispersion curves for a compound like /7-azoxyanisole is illustrated in fig.
2.3.10. The trends in the curves have been confirmed experimentally.(95>96)
Meier and Saupe(97) have discussed the mechanism of dipole orientation in
PAA and have shown that the relaxation time for polarization parallel to

56

2. Statistical theories of nematic order

e 3

10 12

Fig. 2.3.10. Expected form of the dispersion of the principal dielectric constants of
4,4/-di-n-alkoxyazoxybenzenes. The suffix 0 refers to the static values and the suffix
oo to the optical values. ex shows the low frequency relaxation and both x and e2
show the normal Debye high frequency relaxation. (After Maier and Meier.(95))

A \\ \l/7 ///
II.

I
Splay

Twist

Bend

Fig. 2.3.11. The three principal types of deformation in a nematic liquid crystal.

2.3 The Maier-Saupe theory and its applications

57

the 1-direction should increase from its value in the isotropic phase by a
'retardation factor' which may amount to several orders of magnitude
depending on the strength of the nematic potential. However, the
Maier-Saupe potential yields a retardation factor smaller than the
experimental value. This is not surprising since short-range order may be
expected to play a dominant part in the relaxation process and the mean
field theory neglects this completely. No theory of the relaxation
mechanism has yet been proposed taking into account near-neighbour
correlations.
2.3.5 Relationship between elasticity and orientational order
As remarked in chapter 1, a uniformly oriented film of nematic liquid
crystal may be prepared by prior treatment of the surfaces with which it is
in contact. If the preferred orientation imposed by the surfaces is perturbed,
let us say by a magnetic field, a curvature strain will be introduced in the
medium. The theory of such a deformation will be discussed at length in
3.2; for the present it will suffice to state some of the important results.
The free energy per unit volume of the deformed medium relative to the
state of uniform orientation is

where n is a unit vector called the director representing the preferred


molecular direction, dnx/dx, dnx/dy and dnx/dz are respectively the splay,
twist and bend components of curvature at any point, the curvature being
defined with respect to a right-handed cartesian coordinate system XYZ
with Z parallel to the preferred direction at the origin. The three principal
types of deformations are illustrated schematically in fig. 2.3.11.
At the molecular level, it is obvious that curvature elasticity is a
consequence of the orientational order in the liquid crystal. A quantitative
relationship between them was established by Saupe (98) using the mean
field theory. From (2.3.6) is it seen that the dipole-dipole part of the
dispersion energy of interaction of a molecule / in the average field due to
its neighbours k is given by
1
k

*Kik

^
juv* 00

1
J

Y2

o f j f c i \ ay (fc)
l
D2
}XOtiXOv

+ 3^

58

2. Statistical theories ofnematic order

The internal energy per mole due to orientational order in the mean field
approximation is then
U = \NAV~2s\
where V is the molar volume. Now let us suppose that the director at the
site of the fcth molecule is turned through an angle ak with respect to that
at /. If we define a new system of axes X Y'Z' at the site of the /:th molecule,
Z ' making an angle (xk with respect to the Z axis,
*oT = ^OM
v'(fc) _
v(k)

C 0 S

<** ~

^O

Since we are dealing with small deformations, we assume that the director
continues to have cylindrical symmetry about the Z ' axis at the site of the
fcth molecule and that the order parameter s is unchanged in magnitude.
Substituting for x($ etc. in (2.3.33) in terms of x'0{*\ defining the Eulerian
angles between the molecular coordinate system <fcy*><*> and the X'Y'Z'
system and taking appropriate averages, the extra internal energy per mole
of the system in the deformed state turns out to be
1

JLsin,flJ
^ik

r/ z 2
3

L\

\2

z 2 Y21

^ _ ! - 9 ^ 4 (2.3.34)
*^ik

^ik

Now for a pure bend distortion da/dX = da/d Y = 0 and we may set
sin ock ak 7^ifc 3a/6Z. Assuming a spherically symmetric distribution,
Xik/Rik etc. may be taken to be constant (independent of temperature),
and therefore the second sum in (2.3.34) is proportional to F~4/3. The free
energy of deformation per unit volume can then be expressed as
SF
where C is a constant. But by definition

so that

k3S = CV~7/3s2.

(2.3.35)

Similar expressions are obtained for the splay and twist elastic constants.
The temperature dependence of the elastic constants of' simple' nematics
is represented well by (2.3.35).(99'100)
To calculate the absolute values of the elastic constants, one has to
evaluate the second summation in the right-hand side of (2.3.34). Saupe(98)
evaluated the mean value of this sum for an isotropic distribution of

2.3 The Maier-Saupe theory and its applications

59

molecules and found fcn:fc22:fc33 to be 17: 7:11, whereas for PAA at


120 C it is 1.6:1:3.2. Saupe and Nehring(101) have attributed this
discrepancy partly to the neglect of k13 in the expression for the free energy
of deformation (see (3.2.8)). When this coefficient is included, k'^'.k^-.k'^
turns out to be 5:11:5, where k'xl = kxl k13 and k'ss = kss kls. Of course,
these calculations neglect the anisotropy of molecular shape. Attempts
have been made to evaluate the elastic constants on the basis of hard rod
models(102~5) (strictly valid only for very long and thin rods) and hybrid
models incorporating both hard rod features and attractive forces.(106"9)
A corollary to (2.3.35) is that though the order parameter as well as the
elastic constants decrease rapidly with rise of temperature, the intensity of
light scattering should be practically constant throughout the nematic
range. Light scattering from nematic liquid crystals is governed by the
long-range orientational fluctuations as described by the continuum
theory. It is shown in 3.9 that the differential scattering cross-section may
be written approximately as

da

[ns\2kBT

where Q is the irradiated volume of the liquid crystal, a the dielectric


anisotropy at optical frequencies, X the vacuum wavelength, ken an
effective average elastic constant and q the magnitude of the scattering
vector. Now the variation of kB r o v e r the nematic range can be neglected.
Since a varies approximately as s (see (2.3.4)) and kett approximately as
s2, a should be nearly temperature independent. Experimentally, this
result was first established by Chatelain(110) in PAA and PAP. Some
measurements on 4-methoxybenzylidene-4/-butylaniline (MBBA) are
presented in fig. 2.3.12.
An important source of error in these calculations is the neglect of shortrange order. In particular, the theory fails for the bend and twist elastic
constants when smectic-like short-range order is present in the nematic
liquid crystal. Under such circumstances these two constants exhibit a
critical divergence as the temperature approaches the smectic-nematic
point and the light scattering also shows a marked temperature dependence. The present treatment is then inadequate and more elaborate
models have been proposed/ 111 ' 112) The phenomenological theory of this
aspect of the problem will be discussed in chapter 5.

60

2. Statistical theories of nematic order


CD

3 1.2

'S
Q
o
A

% io
0.9
0.8

290

300

310

320

T(K)

Fig. 2.3.12. The temperature dependence of the intensity (/) of light scattering from
4-methoxybenzylidene-4/-butylaniline (MBBA) in the nematic phase. The values
(plotted as kB T/I) are normalized to unity at 286 K. The squares, triangles and
circles correspond to three different experimental configurations. For one of the
configurations (squares) the data were not fully corrected for the effect of the
temperature dependence of the refractive indices on the variation of one of
the wave-vector components, which probably explains the slight increase near the
transition temperatures in this case. (After Haller and Litster.(100))

2.4 Hybrid models: hard rods with a superposed attractive potential


A realistic theory of nematics should, of course, incorporate the attractive
potential between the molecules as well as their hard rod features.(113)
There have been several attempts to develop such hybrid models.
Equations of state have been derived based on the Percus-Yevick and
BBGKY approximations for spherical molecules subject to an attractive
Maier-Saupe potential/114'115) However, a drawback with these models is
that they lead to y = 1 (see (2.3.18)).
Cotter(31> has extended her scaled particle theory to include an attractive
potential of the form
ut = vQp v2psP2(cos0i).

(2.4.1)

The resulting distribution function is similar to that in the MaierSaupe theory, except that the coefficient of the potential has the form
[(v2p/kBT) + A(p)], i.e., a temperature dependent attractive part and an
'athermal' part as given by the scaled particle theory. A similar result can
be obtained using the Andrews model as well.(35) These last two approaches
appear to be promising; for example, calculations show that y 4 for
l/b ~ 2 without violating Cotter's thermodynamic consistency condition
that the mean field potential should be proportional to /?. Further the
transition parameters and the properties of the nematic phase are in
reasonably good agreement with the experimental values for PAA. Gen-

2.5 Short-range order effects in the isotropic phase

61

100 r-

80

60

40

20

120

TNI 140

160

180

Temperature (C)

Fig. 2.5.1. Magnetic birefringence in the isotropic phase of PAA: the horizontal
dashed line gives the value for nitrobenzene at 22.5 C. (After Zadoc-Kahn.(119))

eralized van der Waals models have also been developed(116 18) which lead
to results that are essentially the same as those predicted by the scaled
particle theory with a superposed attractive potential.

2.5 Short-range order effects in the isotropic phase


2.5.1 The Landau-de Gennes model
We have noted in 2.1.2 that though the long-range order vanishes
abruptly at 7^I? certain anomalous effects in the isotropic phase reveal that
an appreciable degree of nematic-like short-range order persists above the
transition point. The most direct evidence of this is the very high value of
the magnetic birefringence, which in the neighbourhood of Tm may be as
much as 100 times greater than in ordinary organic liquids (119120) (fig.
2.5.1). Similar anomalies are seen in the flow birefringence,(121) Kerr
effect(122) and nuclear spin lattice relaxation,(123) confirming the existence of
strong orientational correlations between the molecules. Foex(124) observed
many years ago that the magnetic birefringence exhibits a (T T*)1
dependence and drew attention to the fact that the behaviour is closely
analogous to that of a ferromagnet above the Curie temperature. More

62

2. Statistical theories ofnematic order

recently, de Gennes(125) has proposed a phenomenological description of


these pretransition effects on the basis of the Landau theory of phase
transitions.(126)
Consider an expansion of the excess free energy of any ordered system in
powers of a scalar order parameter s in the following form:
F=|^J2-|&

+ JCJ 4 + ...,

(2.5.1)

where B > 0 and C > 0. We observe that such an expression predicts a


discontinuous transition, for putting F = 0 and dF/ds = 0 at the transition
point Tc we get
sc = 2B/3C
(2.5.2)
neglecting higher order terms. A plot of F versus s results in a family of
curves similar to that shown in fig. 2.3.2.
If B = 0, the transition is continuous and A vanishes at the transition
point.(126) This is because in the disordered phase s = 0 corresponds to a
minimum of F only if A > 0, while in the ordered phase s #= 0 corresponds
to a stable minimum only if A < 0. Thus, since A is positive on one side of
the transition point and negative on the other, it must vanish at the
transition point itself. In the vicinity of the transition, we may therefore
write
A = a(T-T*)
(2.5.3)
where T* is the second order transition temperature.
If B > 0, T* lies below Tc and
(Tc - T*) = 2B2/9aC.

(2.5.4)

For a' weak' first order transition, B is small and 2B2/9aC may be expected
to be a very small quantity.
In principle, a free energy expansion of this type should be valid for
nematic liquid crystals, with s denoting the usual orientational order
parameter defined by (2.3.1). The term of order ss is not precluded by
symmetry, for the states s and s represent two entirely different kinds of
molecular arrangement which are not symmetry related and do not have
equal free energies. In the former case, the molecules are more nearly
parallel to the unique axis, while in the latter they are more nearly
perpendicular to it. However, in the nematic phase s is usually quite large
(greater than about 0.4) so that very many more terms have to be included
in the expansion in order to draw any valid conclusions. Consequently, the

2.5 Short-range order effects in the isotropic phase

63

model is conveniently applied only to the weakly ordered isotropic phase.


We shall discuss some of these applications.

2.5.2 Magnetic and electric birefringence


The free energy per mole of the isotropic phase in the presence of an
external field (magnetic or electric) may be written as
F=\a(T-T*)s2-\Bsz + \Cs\.. + NW(s\

(2.5.5)

where W(s) is the average orientational potential energy of a molecule due


to the external field and TV the Avogadro number. If the external field is
magnetic
W{s) = -^H\
(2.5.6)
where / a = X\\% *s the anisotropy of diamagnetic susceptibility of the
molecule. The magnetically-induced order is extremely weak ( ~ 10~5) so
that we may neglect cubes and higher powers of s. The condition dF/ds = 0
then leads to the result
(2.5.7)
and the magnetic birefringence
(2.5.8)
where, assuming a Lorenz-Lorentz type of relationship for the refractive
index n in the absence of a field,
C - 2nN2x^n^H\n2 + 2Y/21Vn,

(2.5.9)

rja is the anisotropy of optical polarizability of the molecule, and V the


molar volume.(127) The magnetic birefringence varies essentially as
(T T*)'1 since the dependence of C on temperature is relatively small.
Experimentally(128) this is found to be the case, with r N I - r * ^ l K (fig.
2.5.2). As we shall see later a (T r * ) " 1 law implies a classical mean field.
The behaviour is slightly more complicated in the case of electric
birefringence because, as explained in 2.3.4, the orientational energy in an
electric field E arises from the anisotropy of low frequency polarizability aa
and also from the net permanent dipole moment ju. An interesting example
is that of PAA in which the Kerr constant actually changes sign at about
rn + 5K (122) (fig. 2.5.3). The sign reversal is easily understood.(129) The

2. Statistical theories of nematic order

64

I 4

0
44

40

48

52

56

60

Temperature (C)

Fig. 2.5.2. Reciprocal of the Cotton-Mouton coefficient in the isotropic phase of


MBBA for two samples, one slightly impure and having a lower transition point
Tm. Both yield the same value of Tm- T* = 1 C. (After Stinson and Litster.(128))
average orientational energy of the molecule due to the induced dipole
moment is
(2.5.10)
and that due to the permanent dipole moment is
W2(s) = -(F2h2ju2E2/6kB

T)(3cos2p-

\)s,

(2.5.11)

where the symbols have the same meanings as in (2.3.31). In obtaining


W2(s), the distribution function is supposed to involve only even powers of
cos 9. This is clearly valid in the present case in view of the assumed form
of the free energy expression. Putting W{s) =WX+W2,

F=a(T-T*)s2-NFh2E2[(xgL-(Fju2/2kBT)(l-3cos2l3)]s.

(2.5.12)

Proceeding as before
s

NFh2E2[ota-(Fju2/2kB

3a(T- T*)

T)(\-3cos2p)]

(2.5.13)

and

21nVa(T-T*)

(2.5.14)

AE can be positive or negative depending on the sign of the quantity in the


square brackets of (2.5.14). The polarizability anisotropy aa is always

2.5 Short-range order effects in the isotropic phase

65
20

5 -

15

10
<l

-10

-15 r
405 Ni

415

425
Temperature (K)

435

445

Fig. 2.5.3. The electric birefringence (Tsvetkov and Ryumtsev,(122) open circles) and
the reciprocal of the magnetic birefringence (Zadoc-Kahn,(119) full circles) in the
isotropic phase of PA A versus temperature. The lines represent the theoretical
variations.(129)
positive for the long molecules under consideration, but the sign of the
dipole contribution depends on the angle /?. If /? is small, AE will be
strongly positive, whereas if /3 is sufficiently large AnK may be negative.
Moreover, the second term in the square brackets of (2.5.14) is proportional to T~x, so that, in principle, there can occur a reversal of sign of
AE with temperature. Using the values of a and T* derived from the
magnetic birefringence of PAA and substituting for //,/?, etc. (see 2.3.4) it
is found(129) that there is in fact a reversal of sign of AE, though it occurs
at I NI + 9 K (fig. 2.5.3). Since there is a competition between the
polarizability and the permanent dipole contributions, even a small error
in any of the parameters will cause an appreciable shift in the temperature
at which AE = 0. Taking this into consideration, the agreement may be
regarded as satisfactory.
If p = 0, AE given by (2.5.14) varies essentially as (T 77*)"1. In such
materials, the electric and magnetic birefringence may be expected to
exhibit the same type of behaviour over a wide temperature range.
Measurements(130) on pure samples of hexylcyanobiphenyl, a nematogen

2. Statistical theories ofnematic order

66

- 3.2

3.2 -

2.4

2.4

Magnetic

s
<l

1.6

0.8

0.8

T* Tm 30

35

40
45
Temperature (C)

50

55

Fig. 2.5.4. Reciprocals of the magnetic and electric birefringence in the isotropic
phase of 4-hexyl-4/-cyanobiphenyl versus temperature. Both give the same value of
T*(T* = 28 C, r N I - T* = 1.1 C). (After reference 130.)

of strong positive dielectric anisotropy with the dipole (C=N) pointing


almost exactly along the major molecular axis, confirm this prediction (fig.
2.5.4).
2.5.3 Light scattering
Although in the absence of an externally applied field the equilibrium value
of s in the isotropic phase is zero, there can occur fluctuations in the order
parameter about the zero value.(125) This gives rise to an anomalous
scattering of light.
We write the free energy expression (2.5.1) in a more general form in
terms of the tensor order parameter (2.3.5), with / =j = 3 corresponding
to the long molecular axis:
(2.5.15)

= F0 +

\a(T-

(2.5.16)

neglecting higher terms. For a given scattering wave vector q, the


differential scattering cross-section per unit volume is given by (see 3.9)
(2.5.17)
where
(2.5.18)
e is the mean dielectric constant and As is the dielectric anisotropy when the

2.5 Short-range order effects in the isotropic phase

67

16

t 14

12
10

i
I

44

48

52

56

60

Temperature (C)
Fig. 2.5.5. Reciprocal of the intensity of light scattering in the isotropic phase of
MBBA versus temperature. (After Stinson and Litster. (128))

molecules are all exactly parallel to one another (see (2.3.2)). Both e and Ae
refer, of course, to optical frequencies. If the incident light is linearly
polarized along z and the scattered light polarized along x
(2.5.19)
From the equipartition theorem
so that
An2

(Ae)2kBT

dQ

(2.5.20)

If the incident and scattered radiations are both polarized along z,

(where we have made use of the condition sxx 4- syy + szz = 0) and
da^

dQ

_ 8TT2

(Ae)2kBT

(2.5.21)

68

2. Statistical theories ofnematic order

Thus the intensity of scattering should vary essentially as (T T*)' 1 (fig.


2.5.5) and the ratio of the scattered light polarized along z to that polarized
along x should be 4/3. Both these predictions have been verified
quantitatively for MBBA. (131)
If the order parameter varies gradually from point to point, the free
energy expression should include gradient terms as well, which can be
written in the form
\Li^Py^s^

+ \L^as^pspr

(2.5.22)

where 8a = 6/8xa, repeated indices being subject to the usual summation


convention, and L 1? L 2 are constants. To elucidate the consequences of
these additional terms let us suppose that szz = s,sxx = syy = \s,s xy =
s
yz = szx = 0- Also, let s be a function of z only. The free energy is then

(J]

(2.5.23)

where
{

a(T-T*)\

may be called the coherence length. The spatial correlation function has
the form
<>(0) s(R)} = const. kB Tcxp ( - R/Q

(R > Q,

and the scattering cross-section (2.5.21) for a given scattering wavevector


q is modified to (125)
(2-5.25)
Experimentally it is found that the angular variation of the intensity of
scattering is rather small (132133) proving that the coherence length is much
smaller than the wavelength of light (q < 0.1), but by careful measurements Stinson and Litster(133) have established that f oc (T- T*)~1/2, in
accordance with (2.5.24).

2.5 Short-range order effects in the isotropic phase

69

2.5.4 Flow birefringence


To discuss flow birefringence, we have to make use of some results of the
continuum theory developed in chapter 3. In analogy with (3.1.38) we write
for the isotropic phase(125)
(2.5.26)
as 'fluxes'; t^ is the

where t^^cp^ are treated as 'forces' and


viscous stress tensor;
dF
ds,'a/?
= -AsxP from (2.5.15);

For an incompressible fluid daa = 0. All four tensors are symmetric and
traceless. Further from Onsager's relations, ju = ju'.
Now consider shear flow along x with a velocity gradient d*;/dz. The
flow induces a birefringence proportional to the velocity gradient with
the principal axes of the index ellipsoid inclined at 45 to the x, z axes. In
the steady state R^ = 0, cpxz = \pi dv/dz. Therefore
Sxz =

ju
dv
~2a(T-T*)~dz'

(2.5.27)

No other components of s exist, and

Transforming to the x\ z' axes which are inclined at 45 to x, z,


sxz
0
0

0
0
0

0
0
-sv

(2.5.28)

70

2. Statistical theories of nematic order

Therefore x'yz represent the principal axes of the order parameter tensor.
The difference between the dielectric constants (at optical frequencies) for
polarizations along the x' and z' axes is

where, as in (2.5.18), As is the dielectric anisotropy when all the molecules


are exactly parallel to one another. Putting Ss = 2ndn and As = 2nAn, the
flow birefringence

where An is the birefringence when the molecules are all perfectly parallel.
The flow birefringence may therefore be expected to show an anomalous
increase as the temperature approaches the transition. This was observed
by Tolstoi and Fedotov(121) many years ago in PAA. The experiments of
Martinoty, Candau and Debeauvais(134) on MBBA have confirmed the
temperature dependence predicted by de Gennes's equation (2.5.29) (see
fig. 2.5.6).

2.5.5 Comparison with the Maier-Saupe theory


It is of interest to examine the relationship between the phenomenological
model that we have just discussed and the molecular statistical theory of
Maier and Saupe. The free energy of the weakly ordered isotropic phase in
the presence of an external magnetic field is, according to the mean field
theory,

V2

-kBT\n

I e x p | ^ r ^ ( 7 7 ^ + ^ a // 2 )cos 2 0|sin0d<9^. (2.5.30)

Expanding and integrating


jo2

(T-T*)-

0.0762-

where T* = A/5kB V\. This expression is identical in form to the free

2.6 Near-neighbour correlations: Bethe's method


5 n

71
-i 5

45 7\

NI

47

49

51

53

55

57

Temperature (C)

Fig. 2.5.6. Flow birefringence in the isotropic phase of MBBA. Crosses represent
the birefringence (3n) per unit shear rate (G) and circles the reciprocal of this
quantity. (After Martinoty, Candau and Debeauvais.(134))
energy expansion (2.5.1) of the Landau model. However, it does not yield
a satisfactory value of 7*. Since (A/kBTNI V2C) = 4.54, T*/Tm = 0.908.
For PAA, TNI = 408K so that 7NI - T* ~ 40 K, whereas empirically

r NI -r* - 1 K.

Clearly near-neighbour correlations have to be allowed for in the


molecular statistical approach to give a better description of the pretransition effects. A step in this direction was taken in 1973(135~7) which
we shall proceed to consider in the next section.
2.6 Near-neighbour correlations: Bethe's method
2.6.1 The Krieger-James approximation
The theory is based on a method developed originally by Bethe (138) for
treating order-disorder effects in binary alloys. We suppose that every
molecule is surrounded by z neighbours (z ^ 3) and that no two of the z
neighbours are nearest neighbours of each other. (This implies that in
writing down the interactions between the central molecule and its z

72

2. Statistical theories ofnematic order

neighbours, we neglect the interaction between any two of the z


neighbours.) Let the pair potential between the central molecule 0 and one
of its neighbours j be E(9Qj), where 6Qj is a function of the usual spherical
coordinates #0, q>0,6p cpp and let every outer shell molecule y be coupled with
the remaining (external) molecules of the uniaxial medium by an
interaction potential V(9j). The relative weight for a given configuration of
a cluster of z + 1 molecules is then

flf(00j)g(0^

(2.6.1)

3=1

where

The relative probability of the central molecule assuming a certain


orientation 80, cp0 is

J f Jfl/(
J-J

(2.6.2)

while that for an outer shell molecule, say 1, to assume an orientation 015
^i is

JJ/(#oi) S(0i) d(cos 0o) d^ o J... J f l / ( ^ ) g(03) d(cos 0,) d^. (2.6.3)
If we postulate that these two probabilities are identical when 0 and 1 have
the same orientation, we obtain the following consistency relation due to
Chang :(139)

eoj) g{6d) d(cos 0,) d J

d(cos 90) d<p0 (2.6A)

which has to be satisfied for all values of 6, (p.


This condition was expressed in a slightly different form by Krieger and
James.(140) The relative probability that the central molecule 0 and one of
its neighbours, say 1, are oriented along #0, q>0 and 019 cpx respectively is

V(009tp0;Ol9q>1)=fL001)g(eei)ll
i)ll

ff f/(0
f o ,)^,)d(cos0,)d^,. (2.6.5)

3=2 J JJ

2.6 Near-neighbour correlations: Bethe's method

73

Krieger and James postulated that this probability should be the same
irrespective of which molecule is regarded as the central one, i.e.,
; 0 1 5 cp^ = ^ ( 0 1 ? q>x; 00, (p0),

so that

g(00)

TZI = p (constant),

(2.6.6)

which again has to be satisfied for all values of 6, q>.


We shall suppose that
,
V(6}) = - Bin P2n(cos 6,).

(2.6.7)
(2.6.8)

The assumption here is that the energy is independent of cp, i.e., that the
distribution is cylindrically symmetric. We also ignore the volume
dependence of the potential function. For a given value of z, the values of
B2, i?4, etc., can be derived in terms of B* at every temperature such that the
consistency condition (and hence the thermodynamic equilibrium of the
system) is satisfied accurately. All the properties of the system can therefore
be deduced in terms of a single parameter B*. By retaining terms up to
P4(cos0) in the mean field potential (2.6.8), it is found (137141) that the
maximum error in fulfilling Chang's relation (2.6.4) is less than 0.08 per
cent for z = 8. Further it has been verified that when terms up to P12(cos 8)
are included, the error becomes negligibly small (~ 10"9).
The long-range order parameter s is given by

...

P 2 (cos 0O) y/(90, (pQ; 0j9 ^ ) d(cos 0O) dy0 d(cos 0 ; ) d ^

(2-6-9)

...

y(009 (p0; 9P (p3) d(cos 0O) d(p0 d(cos 6d) dtp.

The internal energy of the system is


U = -\NzB*(P2(cos

where

eoj)},

(2.6.10)

</>2(cos0o,)>

... P2(cos 60j) y/(60, <pQ; 6p ^ ) d(cos 0O) dcp0 d(cos 0^) d ^
. (2.6.11)
... y/(0o, <p0; 9p I*,) d(cos 0O) d(p0 d(cos 0) &q>}
J

2. Statistical theories ofnematic order

74

B*/kBT
0.60

0.62

0.15

0.20

0.25 -

1
1

phase
" Disordered
/
S

0.66

0.64

'i

__

2nd order
transition

- 1st order
transition

Ordered phase

<
0.30

Fig. 2.6.1. Plot of <P2(cos 0OJ)> (= -2U/NzB*) versus B*/kB Tfor z = 8.(137) At the
first order transition temperature the shaded areas are equal so that the Helmholtz
free energy of the ordered and disordered phases are the same. At the second order
transition temperature <P2(cos6Oj)} = c2 = l/(z1).
The plot of U versus \/T (at constant volume) shows the characteristic
sigmoid shape (fig. 2.6.1). The curves for the ordered and disordered
phases meet at the apparent second order transition point T*, which is also
the temperature at which the short-range order parameter (P 2(cos60j)) in
the isotropic phase = l/(z1). The first order transition point 7 ^ is the
temperature at which the shaded areas are equal, i.e., when the Helmholtz
free energies of the ordered and disordered phases are the same. The
calculation gives (Tm-T*)/TNI
= 0.062 for z = 8, 0.04 for z = 4 and 0.03
for z = 3. This is an improvement over the Maier-Saupe value of 0.092,
though still much higher than the experimental value of 0.003.
The curves for s, <P4(cos#)> and (P2(cos60j)} as functions of temperature are shown in fig. 2.6.2. All three parameters change discontinuously at Tm; the long-range order drops to zero at the transition, while
the short-range order persists even in the isotropic phase. At the transition,
s = 0.400 (for z = 8) which is slightly lower than the Maier-Saupe value of
0.4292.
Methods involving less stringent assumptions, which have the advantage
that the relevant equations are somewhat easier to solve, have since been
put forward.(142"4) These are based on approximations used earlier in
theories of magnetism. (145146) However, when higher order terms (up to
P4(cos0) or higher) are used in (2.6.8), all these methods yield results that
are practically identical to those given by the Krieger-James approxi-

2.6 Near-neighbour
correlations: Bethe's method

0.6

Order parameter

0.5

Nematic

- - -

Isotropic

=-

<P2 (cos 60j))


0.4
^
0.3
0.2
<P2 (cos 60j)
0.1

1
0.90

1
0.95

1.0

T/Tm

Fig. 2.6.2. Short-range order parameter <P2(cos #0;.)> and the long-range order
parameters s and <P4(cos#)> versus T/T^ calculated in the Krieger-James
approximation.(137) The long-range order disappears at the transition but the
short-range order persists even in the isotropic phase.
mation. (137141) Lekkerkerker et al.a*7) and Van der Haegen et al.am have
used three-particle and four-particle cluster approximations in analogy
with a method followed in magnetism. (149) This leads to a slight
improvement in (TNI r*)/7^ I ? the precise value depending on the type of
lattice considered. A Monte Carlo simulation study/ 150 ' 151) of a lattice
version of the Maier-Saupe theory gives (Tm-T*)/Tm
0.01.
2.6.2 Antiferroelectric short-range order
The vast majority of nematogens are polar compounds but the absence of
ferroelectricity in the nematic phase shows that there is equal probability
of the dipoles pointing in either direction. Because of this it is generally
assumed that the permanent dipolar contribution to the orientational
order is negligibly small. However, a simple calculation shows that the
interaction between neighbouring dipoles is by no means trivial compared
with dispersion forces, particularly in strongly polar materials. We shall
now consider a model which takes into account the influence of permanent
dipoles and is at the same time consistent with the non-polar character of
the medium. (136137)

2. Statistical theories of nematic order

76

o
Fig. 2.6.3. Preferred orientation of neighbouring dipoles in the end-on and
broadside-on positions. However, because of the anisotropic shape of the
molecules, situation I is much more important than II in the nematic structure and
there results a net antiparallel correlation between neighbouring dipoles.

0.6

"

Isotropic

Nematic

0.4

0.2

(P2 (cos 0O,)>

0 0

</\ (cos 60j)y

<p 1 (cos eoj):


'

-0.2

0.90

0.95

1.0

Fig. 2.6.4. Short-range order parameters (P^cos^.)), <P2(cos^0j)> and the longrange order parameter s versus T/Tm. The negative value of (P^cos^.)) signifies
antiparallel ordering. The curves are calculated for A*/B* = 0.5.(137)
Suppose as before that the molecules are cylindrically symmetric so that
the dipole moment is along the major molecular axis. Now if a dipole is
fixed at O as shown in fig. 2.6.3,1 and II represent situations of minimum
energy for a neighbouring dipole in the broadside-on and end-on positions

2.6 Near-neighbour correlations: Bethe's method

11

respectively. Evidently, by virtue of the anisotropic shape of the molecule,


situation I will be much more important, i.e., there will be a greater
tendency for the nearest neighbours to assume an antiparallel orientation.
However, the absence of long-range translational order in the nematic fluid
precludes the possibility of antiferroelectric long-range order. To express
this in a mathematically tractable form we shall resort again to the Bethe
approximation. We modify the pair potential (2.6.7) to
E(0oj) = y4*P1(cos 00j) - 5*P2(cos 00j),

(2.6.12)

which favours an antiparallel arrangement of the permanent dipoles, but


let the interaction between j and the rest of the medium continue to be
(2.6.8) as before. Here P1 is the Legendre polynomial of the first order. Fig.
2.6.4 gives the curves for the long-range parameter s and the short-range
parameters <P1(cos0o^)> and <P2(cos#0j)> calculated on the basis of the
modified potential, taking A*/B* = 0.5. </>1(cos 0Oj)} is negative signifying
antiparallel order. (Tm-T*)/Tm
is now 0.059 for z = 8.
The first important question that needs to be considered is whether such
an antiparallel correlation leads to the right sign and magnitude of the
dielectric anisotropy. Proceeding in the usual manner, the principal
dielectric constants turn out to be
a + | a a ^ + - ^ ( 2 ^ + l ) + ^-<cos6> 0 cos(9 ; .>
J/C B 1

KB 1

Fju2 , . _

(2.6.13)
J

and
1

4nNphF\

-j <cos (p0 cos q>i sin 0O sin 0^

(2.6.14)

where the symbols F9 h, etc., have the same meanings as in 2.3.5. Using the
theoretically derived s of fig. 2.6.4, e1, e2 as well as s = l(e + 2e2) calculated
from these equations are presented in fig. 2.6.5 for a strongly polar (nitrile)
compound of the type studied by Schadt (91) (see fig. 2.3.9(b)). The
parameters used in the calculations are: ju = 5 debyes along the major
molecular axis, a = 28 x 10~24 cm3 and aa = 15 x 10~24 cm 3 (evaluated
approximately by assuming addivity of bond polarizabilities extrapolated

2. Statistical theories of nematic order

78

25 -

Nematic

Isotropic

20 -

I
15 -

-^^^_
10 -

5 -

0.85

0.90

0.95

1.0

Fig. 2.6.5. Theoretical variation of the dielectric constants e19 s2 and e = |(e1
2
with T/Tm.a31) The theory predicts that in the nematic phase should be slightly
lower than the extrapolated value of els, the dielectric constant in the isotropic
phase. This is found to be the case experimentally (see fig.2.6.6).
to low frequency). Since thermal expansion is ignored, the rate of variation
with temperature is reduced especially near the transition, but apart from
that it is clear from a comparison with fig. 2.3.9(b) that the dielectric
anisotropy a is of the right order of magnitude.
An interesting consequence of the theory is that the mean dielectric
constant e should increase by a few per cent on going from the nematic to
the isotropic phase because of the diminution in (P^cos 00j)}. This is found
to be the case experimentally in a number of strongly positive materials^ 1 ' 130152) (fig. 2.6.6). (A similar increase is seen in some negatively
anisotropic materials also, e.g., PAA (90) (see fig. 2.3.9 (a)); this can probably
be explained as due to an antiparallel correlation between the longitudinal
components of the dipole moments. As far as the transverse components
are concerned there will not be on the average any orientational correlation
for position I of figure 2.6.3 because of the cylindrically symmetric
distribution about the optic axis, but there will be an antiparallel
correlation for position II which is, however, likely to be so weak that it can
probably be neglected.)
Direct X-ray evidence for such antiparallel local ordering in the nematic
and isotropic phases of 4/-n-pentyl- and 4/-n-heptyl-4-cyanobiphenyls

2.6 Near-neighbour correlations: Bethe's method

79

18 -

16

J4

a
I

12

5
10

25

30

35

Temperature (C)

Fig. 2.6.6. Principal dielectric constants of 4/-n-pentyl-4-cyanobiphenyl (5CB);


s = i(i-f2e2) is calculated from the measured values of e1 and e2. The dashed
line denotes the extrapolated value of eis. (After reference 130.)
(5CB and 7CB), both of which are strongly polar compounds, has been
reported by Leadbetter, Richardson and Colling.(153) They have found that
the meridional reflexions correspond to a repeat distance along the
preferred axis of about 1.4 times the molecular length, which they have
interpreted as due to an overlapping head-to-tail arrangement of the
neighbouring molecules (fig. 2.6.7). In MBBA, on the other hand, the
repeat distance is approximately equal to the molecular length. These
studies appear to lend strong support to the model of antiparallel

80

2. Statistical theories ofnematic order

25.7 A

Fig. 2.6.7. Schematic diagram of antiparallel local structure in 5CB resulting in a


repeat distance along the nematic axis of about 1.4 times the molecular length.
(Proposed by Leadbetter, Richardson and Colling.(153))

correlation that we have just discussed. As we shall see in chapter 5


antiparallel ordering has important implications for the phenomena of
reentrance and SA polymorphism in polar materials.

2.7 The nematic liquid crystal free surface

We now turn our attention very briefly to the nematic liquid crystal
surface. A variety of experimental studies have established conclusively
that orientationally ordered states, and in certain materials even density
modulations, develop in the vicinity of the free surface. We describe below
the salient features of these observations.

2.7 The nematic liquid crystal free surface

81

37

35

130

120

140

T(C)

38r

36

34

160

190

180

170

r(c)

29

28

-16

-12

-S

12

16

20

r-r NI (c)
Fig. 2.7.1. Experimental variation of the surface tension with temperature for three
nematic liquid crystals, (a) PAA,(156) (b) /?-anisaldazine(156) and (c) 5CB. (158159)
Open circles represent measurements using the pendant drop method. Filled circles
in (c) are values determined independently by Gannon and Faber(159) using the
Wilhelmy plate method.

82

2. Statistical theories ofnematic order

Fig. 2.7.2. Schematic form of the density profile p(z) and its gradient \dp{z)/dz\
across the nematic liquid crystal-vapour transition zone.
It is a well known result that the gradient of the surface tension (y) versus
temperature is directly related to the surface excess entropy per unit area as
follows:
%- = -(So-SX

(2.7.1)

where the suffixes o and /? refer to the surface and bulk states. Thus if there
is surface ordering SG may be less than Sfi, and y may actually show a
positive slope.(154) Such a trend was discovered by Ferguson and Kennedy(155) in PAA and some other compounds many years ago and has been
confirmed by precise measurements on a number of liquid crystals using
different techniques. (1569) Some experimental curves are shown in fig.
2.7.1. The suggestion has been made (154) that the orientational order near
the surface is determined by two competing effects: (1) the disordering
effect of the spatial delocalization across the liquid-vapour transition zone,
and (2) the ordering effect of a surface torque field. The former may be
assumed to be proportional to the density profile across the interface and
the latter to the gradient of this profile (fig. 2.7.2). The different trends
observed in the different materials (fig. 2.7.1) can then be interpreted as
arising from the relative strengths of these two opposing effects and their
variations with temperature. (160)
Light scattering and optical reflectivity studies (161) again reveal the
existence of orientational ordering at the surface. For MBBA it is found

2.7 The nematic liquid crystal free surface

83

io 4 -

101 -

io1

1.0

l.i

QJQo

Fig. 2.7.3. High resolution specular reflectivity data for 80CB near the peak due to
the formation of smectic-like layers near the free surface of the nematic phase. The
open circles refer to the scale on the left and the filled circles to the scale on the right.
The temperatures T- TNA are (a) 0.10 C, (b) 0.21 C, (c) 0.40 C and (d) 1.80 C.
It is seen that the peak becomes significantly sharper as the temperature decreases,
showing that the number of surface induced layers increases on approaching the
nematic-smectic A transition point 7^A. (After Pershan et al.a62))
that the molecular long axis is inclined at about 75 with respect to the free
surface, the angle being temperature dependent, while for PAA the angle is
zero and temperature independent.
Smectic and cybotactic nematic systems may be expected to show an
oscillatory density profile near the surface. That this is the case has been
strikingly demonstrated by the very fine X-ray reflectivity measurements
of Pershan et a/.(162163) (figs. 2.7.3 and 2.7.4). Surface-induced smectic
layering is observed in the nematic or isotropic phase that extends into the
bulk, the number of layers increasing as the temperature approaches the
transition to the smectic phase.

84

2. Statistical theories ofnematic order

10"

10"

= (T-T1A)/TlA

Fig. 2.7.4. The measured X-ray reflectivity from the free surface of the isotropic
phase of 12CB at temperatures very close to the isotropic-smectic A transition
point. The step-like form of the intensity curve reveals the quantized nature of the
layer growth at the surface. (After Ocko et /.(163))
Not much progress has yet been made in giving a quantitative molecular
statistical description of these remarkable surface phenomena.

3
Continuum theory of the nematic state

3.1 The Ericksen-Leslie theory

In this chapter we shall discuss the continuum theory of nematic liquid


crystals and some of its applications. Many of the most important physical
phenomena exhibited by the nematic phase, such as its unusual flow
properties or its response to electric and magneticfields,can be studied by
regarding the liquid crystal as a continuous medium. The foundations of
the continuum model were laid in the late 1920s by Oseen(1) and Z6cher(2)
who developed a static theory which proved to be quite successful. The
subject lay dormant for nearly thirty years afterwards until Frank(3)
reexamined Oseen's treatment and presented it as a theory of curvature
elasticity. Dynamical theories were put forward by Anzelius(4) and Oseen,(1)
but the formulation of general conservation laws and constitutive
equations describing the mechanical behaviour of the nematic state is due
to Ericksen(5>6) and Leslie(7). Other continuum theories have been proposed,(8) but it turns out that the Ericksen-Leslie approach is the one that
is most widely used in discussing the nematic state.
The nematic liquid crystal differs from a normal liquid in that it is
composed of rod-like molecules with the long axes of neighbouring
molecules aligned approximately parallel to one another. To allow for this
anisotropic structure, we introduce a vector n to represent the direction of
preferred orientation of the molecules in the neighbourhood of any point.
This vector is called the director. Its orientation can change continuously
and in a systematic manner from point to point in the medium (except at
singularities). Thus external forces and fields acting on the liquid crystal
can result in a translational motion of the fluid as also in an orientational
motion of the director.

85

86

3. Continuum theory of the nematic state


3.1.1 Conservation laws and the entropy inequality

We begin by writing down the conservation or balance laws (Ericksen(6)).


We shall employ the cartesian tensor notation, repeated tensor indices
being subject to the usual summation convention. The comma denotes
partial differentiation with respect to spatial coordinates and the superposed dot a material time derivative. For example,
Tti = dT/dXt,

vitj = dvJdXj

and
T=dT/dt.
We shall consider the medium to be incompressible (vt i = 0, where vt is
the linear velocity) and at constant temperature (t = T t = 0). We shall
assume further that the director is of constant magnitude. This implies that
the external forces and fields responsible for elastic deformation, viscous
flow, etc., are very much weaker than the molecular interactions giving rise
to the spontaneous alignment of the neighbouring molecules. It is indeed
a valid assumption in all the static and dynamic phenomena discussed in
this chapter. We may therefore conveniently choose n to be a dimensionless
unit vector (ntnt = 1 ) .
Let the material volume be Vbounded by a surface A. The conservation
laws take the following form:
Conservation of mass

/?dF=0,

(3.1.1)

where p is the density.


Conservation of linear momentum

f pv( d V = f f( d V+ f tn dA,,

(3.1.2)

ai

Jv
Jv
JA
where ft is the body force per unit volume and tn the stress tensor.
Conservation of energy
d

fi

dt)v

**

2 X

f
Jv

l %

f
JA

(3.1.3)
where px is a material constant having the dimensions of moment of inertia
per unit volume (ML'1), U the internal energy per unit volume, G{ the

3.1 The Ericksen-Leslie theory

87

external director body force (which has the dimensions of torque per unit
volume since nt has been chosen to be dimensionless), ti = tn Vj the surface
force per unit area acting across the plane whose unit normal is vp and
st = nn Vj the director surface force (which has the dimensions of torque
per unit area). We assume here that there are no heat sources or sinks.
Conservation of angular momentum

jA
=

{^mxJjc + e^n^dV+X
JV

{e^

(3.1.4)

JA

or in vector notation,

dtjy

[p(rxy)+Pl(nxn)]dV

= I [(rxf) + (nxG)]dF+ | [(rxt) + (nxs)]<U.


JV

JA

Finally, we have Oseerts equation:

f
Jv

Pl nt d V

= f (G( +gt) d V+ f w,f d^,


Jv

(3.1.5)

JA

where gt is the intrinsic director body force, which has the dimensions of
torque per unit volume and whose existence is independent of Gt.
Converting surface integrals into volume integrals and simplifying,
(3-1.1)(3.1.5) lead to the following differential equations:
p = 0,

(3.1.6)

PVi=fi + hi,P
U = tn dtj + nn NtjgiNt,

(3.1-7)
(3.1.8)

Pifi^Gi + gt + njij,

(3.1.9)

where
hi - nkj "i, k + gj ni = hi - Xjci "j, ic + gi nP

(3.1.10)

Nt may be interpreted as the angular velocity of the director relative to that

88

3. Continuum theory of the nematic state

of the fluid. It should be emphasized that the stress tensor tn is asymmetric.


When ni = 0, (3.1.6)(3.1.10) reduce to the familiar equations of hydrodynamics for a normal fluid.
In conjunction with the balance equations we make use of the inequality

i I sdv^o,
where S is the entropy per unit volume. Defining the Helmholtz free energy
function per unit volume
F=U- TS,
we obtain for a system in isothermal equilibrium
hi dij + nn Na-gi Nt-F^0.

(3.1.11)

3.1.2 Constitutive equations


In order to develop the theory, it is necessary to set up constitutive
equations for the quantities F, tjt, nn and gi (Leslie(7)). We assume that these
quantities are single-valued functions of
n0

niJ9

nt and vu.

(3.1.12)

We now invoke the fundamental principle of classical physics that material


properties are indifferent to the frame of reference or the observer.(9) Hence
the constitutive equations should be invariant under proper orthogonal
transformations. It is seen that ni and vtJ do not transform as tensors. The
parameters (3.1.12) must therefore be replaced by
ni9

i p

Ni9 a n d dtj.

Thus F may be expanded as


dF
dnt

dF d
dnt dt

l 3

'

dF dF d
87Vt- x ddt- dt ij

But
nt = Ni + Wyii^

Nt-wjtnj9

d
(nt,) = Ntj - wki nk j - (dkj + wkj) nt k

and

dF d ,

dF

dF

dF

dF

3.1 The Ericksen-Leslie theory


Therefore
.
r

dF Ar
~ A

dF

*^ ^

dF
i

V ^ ^

dF .

^ Pi AT

dF .

dtj
i ^ 7\rl
dd " 3v
tj

dF
i> k

^
dn
tk

J*

dF

Hence (3.1.11) becomes

>

(3

-U3)

In view of the constitutive assumptions, it is clear that wji9 Nip Nt and dtj
can be varied arbitrarily and independently of all other quantities and
hence their coefficients must vanish, i.e.,

or
F=F(ni,niJ);
dF

dF

dn^

' 8^>Jfc

Wyhn, *.-

hw*. y -

dF\

' dnk J

(3.1.14)

dF

,-h,

dn

dF

fr-

^M

Yn

(3.1.15)
nji-^L

= 0;

(3.1.16)

and (3.1.13) reduces to


F)F

?\F\

,^0.

(3.1.17)

Let us write the stress and the intrinsic director body force as

where the superscript 0 denotes the isothermal static deformation value


and the prime the hydrodynamic part. Equations (3.1.14) and (3.1.16)
prove that nn does not depend on dtj or Nt so that nn = nQn. Substituting in
(3.1.17)

90

3. Continuum theory of the nematic state

Since dtj and Nt can be chosen arbitrarily and independently of the static
parts fn and gf,

tfg,

(3.1.20)

t'pdv-g'tN^O.

(3.1.21)

Further, using (3.1.10) and (3.1.15)


';<+*;< = '+& / *r

(3-1.22)

The incompressibility condition implies that the stress is indeterminate


to an arbitrary pressure. Similarly there is a certain degree of indeterminacy
in g and nn for if we replace
rfbyy^-jffj/i^

+ s?

and
ni} by p^ + n^
(3.1.8) continues to be satisfied because
n{nt = niNi = n.N^ + n.^N, = 0.
Thus (3.1.19), (3.1.20) and (3.1.16) become
(3.1.23)
n

k,j

t j

nn=P}nt + ^ - ,

(3.1.24)
(3.1.25)

where p, y and fit are arbitrary constants, while the hydrodynamic


components fulfil the inequality (3.1.21).
For an isothermal static deformation, (3.1.9) becomes an equation of
equilibrium:
(3.1.26)
Substituting for g* and nn from (3.1.24) and (3.1.25),
dF\ dF

- + ^ + ^ = 0.

(3.1.27)

3.1 The Ericksen-Leslie theory

91

3.1.3 Coefficients of viscosity


We next consider the nature of t'n and g[, the hydrodynamic components of
the stress tensor and the intrinsic director body force. We assume that they
are linear functions of ni9 Nt and dtj and omit higher order terms :(5>7)
l
XT \ ri
A2
jikiy k~
' jik

/o

AT,
where A and B are functions of ni (at constant T and p). They can therefore
be expanded as
nt + a4 Sik n, + a5 /if ^ /i^,

Sn 3

km

lk = 73 8^ nk + y4 J4Jfc /i^ + yb Sjk nt + y6, /i^ wfc.

Substituting in (3.1.28) and remembering that N^

= du = 0,

t'n = (o + a e 4m nk nm) 5n + (ax + a14rffcOTfc /i m ), ^


+ a13 ^ + a15 ^ /i4 ^ + a16 dki n, nk + a3 nt N, + a4 ^ ^

(3.1.29)

where
a15 = a9 + a10,

a16 = a7 + a n ,

and
^ = (y0 + 7e dkj nk n,) n{ + y9 dik nk + yx Nt

(3.1.30)

where
But in view of (3.1.22)
y9 = a 1 6 -a 1 5 ,

y1 = a 4 - a 3 .

(3.1.31)

In addition, t'n and g[ must satisfy the entropy inequality (3.1.21).


Substituting (3.1.29) and (3.1.30) in (3.1.21) we get
[(a0 + a6 dkm nk nj S{j] dtj + [ax nt n^ dtj - (y0 + ye dkj nk nd) nt
d

km

92

3. Continuum theory of the nematic state

i.e.,
ax nt nj dtj + (quadratic in dip Nt) ^ 0.
As di:j can be chosen arbitrarily, ax = 0. Also the coefficient of Sjf(= p
say) in tn and that of nt ( = y say) in g[ are arbitrary since du = 0 and
ntNt = 0. Putting:
a

Ml =

i4

M* = 13

//2 = a 4

/z5 = a 1 6

we obtain
t'a = Mi nk nm dkm nt H, + // 2 w, A^, + // 3 /i4 ^ + // 4rfi4+ ju5 n, nk dki + n% nt nk dkp
(3.1.32)

g't-^Nt + ^dfl,

(3.1.33)

where, making use of (3.1.31),


^=^2-^3,

/l2 = // 5 -// 6 .

(3.1.34)

We have omitted the terms of the form/7^, in (3.1.32) and 7^ in (3.1.33) as


they do not contribute to the hydrodynamic effects and can be combined
with the corresponding terms in (3.1.23) and (3.1.24) respectively.
Substituting for t'jt and g{ it is found that the left-hand side of (3.1.21) is
a positive and definite quantity if the following inequalities are satisfied :(7)

Finally, from (3.1.23), (3.1.24), (3.1.32) and (3.1.33),

PSji

On

K i + Mink nm dkm nt it,

k,j

+ ju2 ^ Nt + ju3 nt N, + // 4 dn
+ M5njnkdki

+ /i6ninkdkj9

(3.1.36)

goi+gt

yn.-p^-^

+ X^

+^ n ^ .

(3.1.37)

3.1 The Ericksen-Leslie theory

93

The jus represent the six coefficients of viscosity of a nematic liquid crystal.
However, the number of independent coefficients reduces to five if we
assume Onsager's reciprocal relations.
3.1.4 ParodVs relation
From (3.1.21) we observe that the rate of entropy production per unit
volume
TS=t'ndij-g'iNi
where t'n is given by (3.1.32) and g{ by (3.1.33). Since t'n is an asymmetric
tensor it can be resolved into a symmetric component Ytj and an
antisymmetric component Zip where
Y

ij = Mi dkp nk nv nt n, + // 4 dtj + \{pi2 + // 3 ) (nt N, + Nt ft,)


"* *, + dkj nk nt),

As Zn = Z^ and dtj = djt, it follows that


and therefore

^ 4 = 0,
TS-Yfidv-g'tNt.

Thus the entropy production can be separated into two parts, one due to
the linear motion of the fluid and the other due to the orientational motion
of the director. Now h = H x n, where O is the angular velocity of the
director. It is then easily shown that

where Jn = n^ n^ is the torque exerted by the director. Consequently

Y^Jy may be regarded as 'fluxes' and dip (Qtf w y) as 'forces'. The


relation between the fluxes and forces is
(3.1.38)
where
" K/^5 + Me) ($ir Hs Uj + <>js Ur ni)>

%rs = UM2 - Ms) {"i KS Srj + nr nj

is)-

94

3. Continuum theory of the nematic state

From Onsager's reciprocal relations in irreversible processes(10) it follows


that [D] is symmetric, i.e.,
or
// 6 -// 5 .

(3.1.39)

(11)

This is referred to as Parodfs relation.


The number of independent
viscosity coefficients is therefore reduced from six to five.
It is well known that Truesdell(12) is of the view that Onsager's relations
do not apply to phenomena like heat conduction, viscosity and diffusion
since there is no unambiguous way of selecting the fluxes and forces. It
would appear therefore that there may be some doubts as to the validity of
Parodi's relation. Available data indicate that (3.1.39) is satisfied within
experimental limits,(13) but, in any case, the relation has been tacitly
assumed to be true in most discussions.

3.2 Curvature elasticity: the Oseen-Zocher-Frank equations


We have shown that for an incompressible fluid and an isothermal
deformation (see (3.1.14))

We may therefore expand the free energy per unit volume of a deformed
liquid crystal relative to that in the state of uniform orientation as

We neglect higher order terms in the expansion as we are concerned only


with infinitesimal deformations. Since these deformations relate to changes
in the orientation of the director, nitj may appropriately be called the
curvature strain tensor (Frank (3) ). In order to define the components of this
tensor, let us choose a local right-handed system of cartesian coordinates
with n parallel to z at the origin. The components of strain are then
onx

Twist:
Bend:

9w

aw,
6x'
9z'

dz

3.2 Curvature elasticity: the Oseen-Zocher-Frank equations


We ignore dnjdx, dnjdy and dnjdz. The three types of deformation are
shown in fig. 2.3.12. The curvature strain tensor may therefore be written
as
dx
dny
dx

dy
dz
driy dn^
dy
dz

(say).
0

Now any of the three types of deformation destroys the centre of


symmetry of the liquid crystal. The strain tensor is therefore an axial
second rank tensor which vanishes identically under a centro-symmetric
operation. Since the free energy is a scalar, the components of k tj also form
an axial second rank tensor:
Vk
AC n
k \k
^ij

^21

l_/c 31

k12
K
k

k 13 1
AC
k \

^22

^ 2 3 I >

/c 3 2

A:33J

or using an abbreviated one-index notation


Vk

(3.2.1)

K K

In general kt has nine components, but the presence of symmetry in the


liquid crystal reduces this number.(14) The distribution of molecules around
any point is cylindrically symmetric, so that the choice of the x axis is
arbitrary apart from the requirement that it should be normal to the
director axis z. Therefore
-kt

. 0

K
K
0

0
0
0

Additional symmetry operations reduce the number of moduli even


further:
Enantiomorphic and
Enantiomorphic and
Non-enantiomorphic
Non-enantiomorphic

k + 0,
polar
non-polar
x = 0,
and polar
and non-polar kx = 0,

k2 + 0

k2 ==
| 0
k2 = 0
k2 = 0.

The tensor kijlm (or ktj in the abbreviated notation of (3.2.1)) has 81

95

3. Continuum theory of the nematic state

96

components in general, but as a79 a8 and a9 are zero there remain only 36.
The presence of cylindrical symmetry reduces this number to 18 with only
5 independent constants:
kl2
k22
0

0
0

k15
k12
0

k22

~^12

^24

/Cqq

^24

Lo

k12

k12
0

CO

0
0
0

-k12
0

0
0
0
0
0

^33-

where k15 = kxl k22 k2. In the absence of polarity or enantiomorphy,

*2 ==o.o.

The free energy of deformation may therefore be written in the form

+ k12(a + a5) (a2 + aA) - (k22 + 24) (a ah + a2 aj.


As the free energy must be positive definite, it can be shown
algebraical methods (16) that

(15)

(3.2.2)

by standard

^22

In tensor notation (3.2.2) reads as


F = \klx{s + nit ,)2 + \k22{q + n{ em nk ;.)

+ kM)[nijnji-(niif],

(3.2.3)

where s = kjkxi is the permanent splay and q = kjk22 the permanent


twist. There is no physical polarity along the direction n in any known
nematic or cholesteric substance. The molecules themselves may be polar
but the absence of ferroelectricity confirms that there is equal probability
of their pointing in either direction. We shall therefore set kY = k12 = 0.
Substituting for F i n (3.1.27) we then obtain
~

ijk nj\np

pqr Ur, q), fcJ

= 0.

(3-2.4)

It is seen that &24 plays no role in (3.2.4) and can be omitted as far as
equilibrium situations are concerned (Ericksen (17)). When external body
torques are absent {Gi =0), the solutions of (3.2.4) are

n2 = sin (qz+y/\)

3.3 Summary of equations of the continuum theory

97

where

^ + ? S = 0.
dx* dy*

(3.2.6)

Equation (3.2.5) describes a cholesteric structure with a twist per unit


length of kjk22. In the absence of enantiomorphy, k2 = 0 and the structure
is nematic. The solutions of (3.2.6) describe the configurations around line
singularities in the structure which we shall consider in some detail in a
later section (3.5.1).
In vector notation the free energy density may be written in the more
compact form
F = k2(n V x n) + &n(V n)2 + \k22(n V x n)2 + ^ 33 (n x V x n)2, (3.2.7)
with k2 = 0 in the nematic case.

3.3 Summary of equations of the continuum theory


In the following sections of this chapter we shall apply the continuum
theory to study the behaviour of the nematic phase in various physical
situations. For convenience we set out below the most important equations
of the theory which we shall be referring to constantly:
pvt=ft + tjU9

(3.3.1)

px h'i Gi + gi + nn p

(3.3.2)

where p is the density of the fluid (assumed to be incompressible and at


constant temperature), px a material constant having the dimensions of
moment of inertia per unit volume, nt a dimensionless unit vector called the
director, vt the linear velocity, f the body force per unit volume, tn the
stress tensor, Gt the external director body force, gt the intrinsic director
body force and nn the director surface stress.
The stress tensor tn may be separated into a static (or elastic) part and a
hydrodynamic (or viscous) part:
tjt = qt + t'ji9

(3.3.3)

where
$i = -P*ii-^t.i>

(3-3.4)

n, nj + JU2 ^ Nt + jus nt N, + //4 dn


^^^^d^n^n^d^

(3.3.5)

98

3. Continuum theory of the nematic state

F is the free energy per unit volume given by


,-K2)ninjnhinhp
3 3 (nxVxn)

(3.3.6)
,

(3.3.7)

N^rit-w^

(3.3.8)

2dtj = vu + v^

(3.3.9)

IWt^Vij-v^

(3.3.10)

p is an arbitrary (indeterminate) constant, // x ... ju6 are the coefficients of


viscosity (generally referred to as Leslie coefficients), and k119 k22 and k33
the elastic constants (or Frank constants). Similarly the intrinsic director
body force gt may be written in two parts
gi = gi+g't,

where

(3.3.11)

gi=yni-PjniJ-dF/dnt9

(3.3.12)

g'^KNt + ^dw

(3.3.13)

y and Pj are arbitrary (indeterminate) constants, and


(3.3.14)

Also, according to the Onsager-Parodi relation


(3.3.15)

JU2+JU3 = JU6-JU5.

The director surface stress


Uji

= ^jni + dF/dnij.

(3.3.16)

For static deformations,


= 0.

(3.3.17)

3.4 Distortions due to magnetic and electric fields: static theory


3.4.1 The Freedericksz effect
The simplest method of measuring the three elastic constants of a nematic
liquid crystal is by studying the deformations due to an external magnetic
field (Freedericksz and Tsvetkov,(18) Z6cher (2)). The geometry has to be so
chosen that the orienting effect of the field conflicts with the orientations
imposed by the surfaces with which the liquid crystal is in contact. To
develop a static theory of such deformations we apply the equation of

3.4 Distortions due to magnetic and electric

fields

99

equilibrium (3.3.17) where Gt is the external director body force due to the
magnetic field H. If/y and/ are the principal diamagnetic susceptibilities
per unit volume along and perpendicular to the director axis respectively,
G,=/a//,,#

(3.4.1)

where / a = / , , - / .
Let us consider first the case of a nematic film in which the initial
undisturbed orientation of the director is throughout parallel to the glass
plates. The magnetic field H is now applied perpendicular to the director
and to the plates (fig. 3.4.1 (a)). For this geometry, n = (cos #,0, sin 6),
H = (0,0,//) and G = (O,O,/ a // 2 sin0). The free energy of elastic
deformation (3.3.6) reduces in this case to

By straightforward substitution in (3.3.17) and simplification, we obtain


the equilibrium condition

As is to be expected, the deformation involves the splay and bend moduli,


klx and 33 respectively, and not the twist modulus k22. Because of the
orienting influence of the glass surfaces, 0 = 0 at z = 0 and d, where dis the
thickness of the film. Therefore 9 attains a maximum value #max at z = d/2
and from symmetry considerations 0(z) = 6{dz). Since d8/dz = 0 at
z = d/2, we get

Jo

Transforming to a new variable X given by sin X = sin 9/sin 0

r [ M l - sin2 #max sin2 X) +fr33sin2 flmax sin2 Xf


l-sin 2 0 max sin 2 A

cu.

Taking the limit #max = 0 gives

'~AtV

(3A2)

In other words, deformation occurs only above a certain critical field Hc.
This is referred to as the Freedericksz effect. The threshold condition can be
used for a direct determination of the splay modulus klv

100

3. Continuum theory of the nematic state


z

H<H

H<Hn

ZZZ H>HC

1 1 ) 1 ) 1 }
/ / / /
H<HC

>Hc/
/

(c)

/ / /

I I I I I I II

Fig. 3.4.1. The experimental Freedericksz geometries for the determination of the
(a) splay, (b) twist and (c) bend elastic constants of a nematic liquid crystal.
For H > Hc, the deformation at any arbitrary point can be computed
from the expressions*19'20*

J 0

Isin 2 ^_-sin 2 0

= I arc sin
where q =

..., (3A3)

-,

(3.4.4)

(k^-k^/k^.

Two other important geometries are illustrated in fig. 3.4.1. For


n = (cos 0, sin 6,0), H = (0, H, 0) (fig. 3.4.1 (b)),
(3.4.5)

3.4 Distortions due to magnetic and electric

fields

101

and for n = (sin 0,0, cos 0), H = (//, 0,0) (fig. 3.4.1 (c)),

/XJ-

(3A6)

(21)

De Gennes
introduced a parameter which he called the magnetic
coherence length to define the thickness of the transition layer near the
boundary. Consider, for example, a nematic liquid crystal occupying the
half space z > 0. Let the wall, the xy plane at z = 0, impose an orientation
along x and let the magnetic field be applied along y (analogous to the
geometry of fig. 3.4.1 (&)). If cp ( = \n 6) is the angle made by the director
with the field, the equilibrium condition is easily shown to be

where = (k2Jx$H~x. Integrating, subject to the boundary condition


that when z -> oo, cp -> 0 and dcp/dz -> 0,

is the magnetic coherence length. It is usually of the order of a micron for


a field of 104 G and increases with diminishing field. If the sample thickness
d is very much greater than , most of the material will be aligned in the
field direction.
The experiment for the determination of klx or A;33 consists of measuring
the variation of birefringence for light incident normal to the film. With
linearly polarized light incident and a suitable analyser (e.g., a combination
of a A/4 plate and a linear polarizer), the transmitted intensity shows a
sudden change when the field attains the threshold value. A measurement
of Hc in the geometries (a) and (c) of fig. 3.4.1 therefore gives the elastic
constant kxl or ksz directly. As the field is gradually increased further, the
intensity exhibits oscillations because of the change in phase retardation
(fig. 3.4.2). The observed variation in the retardation is found to be in
conformity with that expected from (3.4.3) and (3.4.4).(20) In principle,
measurements beyond Hc using the geometry (a) of fig. 3.4.1 should yield
both klx and A:33.
However, the threshold for a twist deformation cannot be detected
optically when viewed along the twist axis. This is because of the large
birefringence (Sn) of the medium for this direction of propagation (the case
/? <^ y in 4.1.1). Thus with the experimental geometry of fig. 3.4.1 (ft) in
which the director is anchored parallel to the walls at either end and light
is incident normal to the film, the state of polarization of the emergent
beam is indistinguishable from that of the beam emerging from the

102

3. Continuum theory of the nematic state


123.2 C

121.3 Cu

119.6 C
117.5 C

~liAAAA/\

IWIAA/W
JlMAAA/\

-jmmw
,,-jmm/v

115.5 C

106 C
101.3 C

1
Magnetic field (kG)

Fig. 3.4.2. Raw recorder traces of interference oscillations due to the change in the
sample birefringence with deformation for hexyloxyazoxybenzene at various
temperatures. Polarizer and crossed analyser are inclined at 45 to the principal
axes of the specimen. The sudden onset of oscillations occurs at the threshold field.
The increase in the threshold for the successive traces illustrates the rapid
temperature variation of the elastic constant. Sample thickness 45 /zm. 7^ =
128.5 C. (After Gruler, Scheffer and Meier.(20))
untwisted nematic. For this reason Freedericksz and Tsvetkov(18) used a
total internal reflexion technique by letting the light beam fall at an
appropriate angle on the specimen contained between a convex lens and a
prism. A simple and more direct method has been proposed.(22) If the
ellipsoid of refractive index is viewed obliquely, say at 5 or 10 to the
director (fig. 3.4.3), the effective Sn is reduced to a low value and the
deformed medium can be shown to be optically equivalent to a rotator and

3.4 Distortions due to magnetic and electric fields

103

If !:
1
f
- i!
,:i I

IS

Ill'll1

!>: I

'iliii!1

(a)

07/

Fig. 3.4.3. (a) The usual experimental configuration for the optical observation of
the Freedericksz effect. Light is incident normal to the film. However, for reasons
discussed in the text, this arrangement is unsuitable for observing a twist
deformation, (b)l Oblique' configuration which enables the optical detection of a
twist deformation.(22) The magnetic field is perpendicular to the plane of the paper
in both cases.

a retarder (the case /? ~ y in 4.1.1). Hence a twist deformation produces a


change in the state of polarization of the emergent beam which can be
detected by the usual optical methods. In fig. 3.4.4 we present k22 for two
compounds determined by this technique.
When the medium is twisted, the principal axes of the equivalent
retarder will evidently be rotated with respect to those of the undistorted
one. The angle of rotation can be measured by observing the concoscopic
interference figures in a convergent beam. Here again it is essential that the
rays make a sufficiently large angle with the twist axis to reduce the
effective Sn. This method has been used by Cladis(23) to determine k22.
It should be emphasized that if the magnetic field is not strictly
perpendicular to the initial undisturbed orientation of the director, the
distortion does not set in abruptly (fig. 3.4.5) and the experimental

104

3. Continuum theory of the nematic state

-30

-20

-10

r-r NI (c)
Fig. 3.4.4. Twist elastic constant (k22) versus temperature for PAA and PAP.
Squares, circles and triangles represent independent measurements on different
samples. (Karat. (22))

determination of the elastic constants becomes somewhat unreliable.


When the field is applied exactly at right angles, there is equal probability
of the director turning through an angle cp or cp with respect to H.
Consequently a number of disclination walls dividing the domains having
different preferred orientations are formed in the specimen (see 3.5.2).
This serves as a useful criterion for checking the alignment of the field.
Orientation at the glass surfaces
Errors may also be introduced by weak anchoring at the glass surfaces. The
usual procedure for obtaining a planar or homogeneous structure is to rub
the glass (which may be coated with a very thin layer of polymer) with dry
lens tissue or cloth along a fixed direction. (24) Berreman has shown that
geometrical factors play an important role in producing such alignment,
for rubbing tends to corrugate the surface.(24) Assuming that both ends of
the molecule have equal affinity for the surface material so that they lie flat
against the surface, which may not always be true, it is obvious that more
elastic energy is required for the molecules to lie across the rubbed

3.4 Distortions due to magnetic and electric fields

105

0.7
0.6
^

0.5
0.4

0.3
0.2
0.1
0
i

Magnetic field (kG)

Fig. 3.4.5. Theoretical curves illustrating the relaxation of the Freedericksz


threshold as the magnetic field is tilted away from the normal to the initial
orientation of the director. The calculations have been made for PAA in the
twist geometry for a sample of thickness 12.7 jum. (a) Field normal to the director,
<p = 90, (b) q> = 89, (c) (p = 85 and (d) tp = 80. 0max is the maximum deformation
in the midplane of the sample. (Kini and Ranganath, unpublished.)

(a)

(b)

Fig. 3.4.6. The orienting effect of grooves. Extra elastic energy is required for the
nematic director to lie across the grooves on the solid surface as in (a) rather than
to lie parallel to them as in (b).

direction than for them to lie parallel to it (fig. 3.4.6). A simple calculation
shows that this energy difference for the material near the surface is quite
appreciable and almost impossible to overcome by means of a magnetic
field. The anchoring is therefore firm. Deposition of silicon monoxide and
certain other materials on the glass surface at oblique incidence has been
shown to have essentially the same effect as rubbing.(25)

106

3. Continuum theory of the nematic state

If the surface is equally rough in both directions (as can happen if it is


etched and cleaned thoroughly) there would be a tendency for the
molecules to stand upright. Perpendicular or homeotropic alignment may
also be achieved by coating the surface with surfactants in which case the
detailed intermolecular forces probably play a significant part. There is
evidence that homeotropic alignment is not always rigid. The consequences
of weak anchoring on the Freedericksz transition have been discussed by
Rapini and Papoular.(25)
Electric fields
Measurements can also be made using an electric field,(20) or even an
optical field from a laser.(26) There is complete analogy between electric and
magnetic fields as far as the threshold is concerned (except when the
dielectric anisotropy is very large(191)). Above the threshold the analogy
fails in general because the distortion gives rise to a non-uniformity of the
electric field in the medium, a problem which, for all practical purposes,
does not arise in the magnetic field case. An important precaution to be
taken in electric field measurements is that the sample has to be pure to
avoid conduction-induced instabilities (see 3.10).
Capacitance measurements
The deformation at the threshold field in geometries (a) and (c) of fig. 3.4.1
may be detected conveniently by measuring the change in the capacitance/ 2 ^ A slight disadvantage with this technique is that relatively
large areas have to be uniformly oriented which is not easily achieved in
practice. Non-uniform areas and edge effects tend to destroy the sharpness
of the threshold. For optical observations, on the other hand, a well
oriented area of less than 1 mm2 is adequate.

3.4.2 The twisted nematic device


Another configuration of much practical interest is the twisted nematic
film.(2829) The liquid crystal is sandwiched between two glass plates with
the director aligned homogeneously (parallel to the walls). A twist is now
imposed on the liquid crystal by turning one of the plates in its own plane
about an axis normal to the film. A magnetic field above a critical strength
applied along the twist axis results in a deformation as shown schematically
in fig. 3.4.7.
When H = 0, we have n = {cos MX)], sin[#?(z)], 0}, where (p(O) = (po
and cp{d) = #?0, d being the film thickness. When H = (0,0, H), n =

107

3.4 Distortions due to magnetic and electric fields

H < Hc

Hc

Fig. 3.4.7. The twisted nematic cell.


{cos [9(z)] cos [<pO)L C O S [#( Z )1 si n [<P(z)], sin [#(z)]} where # is the angle made
by n with the xy plane. From (3.3.17) we get

and
(3.4.8)
where
f(9) = kxl co
g(9) = (k22 cos2 9 + 33 sin2 9) cos2 9.
After integration, (3.4.7) and (3.4.8) yield
(3.4.9)
and
(3.4.10)
where A and B are constants. Equation (3.4.9) may be rewritten as
(3.4.11)

108

3. Continuum theory of the nematic state

Because of strong anchoring at the walls, 9 = 0 at z = 0 and d, while # max,


the maximum value of 9, occurs at the midplane z = d/2. Since d9/dz = 0
at z = rf/2,
[B2/g(9maK)]+XaH*sm*9max.

A =

Substituting in (3.4.11)

or

Aw)
i//, say.
Similarly from (3.4.8)
7r

Therefore
=2

(3.4.12)
JJ 00

and

(3.4.13)
Transforming to a new variable given by sin X = sin ^/sin 0 max, (3.4.12)
and (3.4.13) become

,=2 r

r__j

Jo L ^ -

and

Jo
where
M{)

T 1

(si
= [fc33 - 2k22 - (kS3 - k22) (sin 2 W + sin 2 emj]/g(6)

g(0m!lx).

3.4 Distortions due to magnetic and electric fields

109

6 h

Fig. 3.4.8. Computed relative capacitance change AC/C 0 and optical transmission
T between parallel polarizers (both parallel to the director at one of the boundaries)
of a twisted nematic film as functions of H/Hc. The total twist angle is n/2. Film
thicknesses 13 and 54 jum. The threshold for optical transmission increases with the
thickness of the film. (After Van Doorn.(27))
Taking the limit 0 m a x ^ ' ', we have
2<po/d,B^2k22<po/d,M(k33-2k22)/k'

=2n

ll9

22

Jo \x.H*- (4tpl/d*){k^-2k22)

or

g(0)^k22, dcp/dz

and

I
(3.4.14)

which is the critical field for the deformation to occur. (28) Thus a
measurement of Hc in this geometry gives k22, if kxl and kZ3 are known.
However, for reasons discussed in 4.1.1, the deformation cannot be
detected optically by observation in polarized light at normal incidence
until the field is well above the threshold value Hc. A more convenient way
of detecting Hc would appear to be by measuring the change in the
capacitance (27) (fig. 3.4.8).
A noteworthy feature of the twisted nematic (TN) is that the intensity of

110

3. Continuum theory of the nematic state

the transmitted light (through a pair of polarizers) as a function of field


shows a 'bilevel' behaviour (fig. 3.4.8). It can therefore act as an electrically controllable optical shutter, as was first demonstrated by Schadt
and Helfrich.(29>30) This principle is now finding extensive application
for making liquid crystal display devices (LCDs) for watches, pocket
calculators, automobile dashboards, miniature TVs, etc. A thin nematic
film of positive dielectric anisotropy is sandwiched between two glass
plates, the inside surfaces of which are coated with transparent conducting
material. By surface treatment, the director is aligned parallel to both
plates except that a 90 twist is imposed on the liquid crystal (fig. 3.4.7,
top). Linearly polarized light incident normal to the plates, say with the
vibration direction parallel to the director on the entrance side of the film,
will emerge with its vibration direction rotated through 90 (see (4.1.15)).
The emergent beam will be transmitted by a second polarizer set at the
crossed position with respect to the first one. If now an electric field is
applied, the molecules in the bulk of the sample tend to align themselves
normal to the electrodes when the voltage exceeds a threshold value,
which, by analogy with (3.4.14), is given by

for 2q>0 = n/2, where ea is the dielectric anisotropy. For voltages of the
order of 2 Vc the twist is lost over most of the sample (fig. 3.4.7, bottom), the
polarization is no longer rotated by 90 and the transmitted light is
extinguished by the second polarizer. With parallel polarizers, one can get
the opposite effect, namely bright field in the ON state and extinction in the
OFF state. LCDs are often operated in the ' reflexion' mode: this is done
by having a diffuse reflector on the rear side of the cell. Unlike the dynamic
scattering mode (see 3.10.1) in which conductivity plays a crucial role, the
TN LCD is a field effect device. Therefore, it is advantageous to have high
purity (low conductivity) materials. A major advance in the materials
development was the discovery by Gray and others(31) of highly stable
mesogens of strong positive dielectric anisotropy, like 5CB, 7CB and
related compounds. A wide nematic range, from about 10 C to 70 C,
that is necessary for most practical devices, is obtained by making suitable
mixtures of these compounds. With a 90 twist there is equal probability of
the medium acquiring a right-handed or left-handed twist, which results in
the formation of disclination walls. To avoid this, a small quantity of a
cholesteric dopant is added to the nematic mixture in order to favour one
sense of twist throughout the sample. Typically, with such materials, the
threshold voltage is about 1 V, the switching time (for a cell of thickness

3.4 Distortions due to magnetic and electric

fields

111

8 //m) a few tens of milliseconds and the power consumption a //W cm" 2 of
display area. A disadvantage with the TN cell is that the viewing angle is
limited to about 45 from the normal, but the advantages far outweigh
this disadvantage. By patterning the electrodes appropriately, it is possible
to display the required information, whether it is digits or letters or any
other symbols.
For higher information content it is convenient to use the dot matrix
configuration, with electrodes as horizontal rows on one glass plate and as
vertical columns on the other. The intersection of a row and a column is a
picture element (pixel). Thus a matrix display with TV rows and M columns
has TV x M pixels but only N+M connections. When the number of pixels
becomes large one resorts to an electronic addressing technique known as
multiplexing: each row in the matrix is selected sequentially, while
appropriate data waveforms coded with the information are applied to the
columns.(32) Because of the slow response times of LCDs (~50 ms), each
pixel responds only to the rms of the resulting waveforms. As the number
of lines TV in the matrix increases, the fraction (I/TV) of the total time that
the selected pixels see the full select pulse decreases, thereby reducing the
ratio J^ ms (sel)/^ ms (unsel). Alt and Pleshko(33) showed that

For TV = 100, this ratio is just about 1.1. If the electrooptic response of the
display were very steep (like a step function), a small change in the voltage
would produce a large change in the director orientation, and it would be
possible to activate a given pixel without altering the state of the other
pixels, but since the response is rather broad (fig. 3.4.8), TV is restricted to
about 100 to get a reasonable contrast ratio. To overcome this limitation,
in the pocket TV, for example, an active addressing technique is used in
which each pixel is backed by a thin film transitor (TFT) which enables one
to apply any desired voltage to the ON pixels and zero voltage to the OFF
pixels. With each column replaced by three which are coated with different
pigments, one can get full colour display. Further, the broad electrooptic
response of the material is actually helpful for producing grey scales.
An advance that has extended the application of LCDs to full page
computer terminals and other high information content displays without
having to resort to TFTs is the supertwisted nematic device/ 34 ' 35) It makes
use of the fact that the electrooptic response of the nematic gets
progressively steeper as the twist angle ^ is increased, until at a certain

112

3. Continuum theory of the nematic state

1.0

1.5

2.0
Reduced voltage (V)

2.5

3.0

Fig. 3.4.9. Computed values of the tilt angle 0max in the midplane of the sample
versus voltage for various twist angles <fi for a standard TN mixture. Cell
thickness/pitch = (j>/2n in all cases and the director tilt at the surface 6S= 1.
(After Scheffer.(32))
value of ^ it becomes infinitely steep. This is illustrated in fig. 3.4.9, which
gives plots of the computed values of the maximum director tilt angle #max
in the midplane of the sample versus voltage for various values of <j> for a
standard TN mixture. In all cases the cell thickness/pitch ratio d/P =
<f>/2n, and the director tilt at the surface 6S= 1, which is the value
generally obtained with standard nematic materials on rubbed polymer
coatings on glass. As remarked earlier, the steeper the response the higher
the multiplexability, and the infinite slope for <j> = 270 represents the best
condition for rms multiplexing. For higher twist angles the slope is
negative, and in practice the device becomes bistable.(36) The precise value
of </> at which the slope becomes infinite depends on combinations of the
relevant material and device parameters,(37) but for most known nematic
materials it lies in the range 180 < <f> < 360. The high twist is stabilized in
the cell by the addition of the appropriate quantity of a cholesteric dopant.
Modified forms of these displays have been developed - e.g., the 'dye'
display which has a pleochroic dye dissolved in the nematic material and
requires the use of just one polarizer - but we shall not be discussing them
here. Analytical expressions have been derived(37) which simplify the
computational effort involved in optimizing the material and device
parameters, but one has to rely on numerical modelling to give a complete
description of the dynamical characteristics of these devices.(38) Certain
unusual dynamical effects observed in the TN device, e.g., the 'reverse-

3.4 Distortions due to magnetic and electric fields

113

twist' and 'optical bounce' effects, will be explained very briefly in a later
section while discussing the dynamics of the Freedericksz transition (see
3.8).

3.4.3 The Freedericksz effect in highly anisotropic nematics: periodic


distortions
The Freedericksz transition discussed in 3.4.1 may be called a 'homogeneous' transition since the distortion occurring above the threshold is
uniform in the plane of the sample. In low-molecular-weight nematics,
which as a rule have relatively small elastic anisotropy (klx ~ k33;k119
k33 ~ 2k22), it is the homogeneous transition that is generally observed.
Some polymer nematics, however, are known to exhibit high elastic
anisotropy - an example is a racemic mixture of poly-y-benzyl-glutamate
(PBG) which has klx/k22 =11.4 and k33/k22 = 13.0(39) - and in such cases
more complex types of field-induced deformations are possible.(40)
Let us consider the geometry of fig. 3.4.1 (a): the initial unperturbed
orientation of the director is along x and the magnetic field is applied along
z. Lonberg and Meyer,(41) who investigated PBG in this geometry, found
that above a well defined threshold there appears a periodic distortion with
the wavevector along y (fig. 3.4.10). It emerged from their theoretical
analysis that when klx/k22 is greater than about 3.3, the periodic distortion,
which involves both splay and twist, has a lower threshold than the usual
homogeneous splay distortion, and moreover, close to the threshold the
deformation angles for splay and twist are out of phase with each other
along the direction of periodicity.
The theory of periodic distortions has been discussed thoroughly. (415)
Because of thermal fluctuations, the director orientation is perturbed and
one may therefore write n = (1, ^, #), where 9 and (/> are, respectively, the
splay and twist angles, which are assumed to be small. Bearing in mind the
experimental result, one may assume that 9 and <f> are functions of y and z.
Retaining terms to the first order in 9, (/> and their derivatives, one obtains
the following equations:

where rj = 2z/d, d being the sample thickness, and the subscripts denote
partial derivatives, e.g.,
etc.

(3.4.16)

114

3. Continuum theory of the nematic state

Fig. 3.4.10. Periodic distortion in PBG. Initial unperturbed director is parallel to


the stripes and the magneticfieldperpendicular to the plane of the sample. Sample
thickness ~ 37 //m and distance between two dark bands 32.5 //m. (Lonberg and
Meyer.(41))
The boundary conditions are 6 = <j> = 0 at n = 1. Seeking solutions of the
form
(3.4.15) and (3.4.16) lead to the compatibility condition (41)
1

^ tanh q 2 tan qx = 0,

(3.4.17)

where

Qy is the dimensionless wavevector of the periodicity. For given k119 k22, d


and / a , (3.4.17) is satisfied by a certain Hz(Qy) for each value of Qy. In the
limit Qy->0, Hz(Qy)-*Hz(0) which is the splay Freedericksz threshold.
Calculations show that when k1:L/k22 is large enough, Hz{Qy) decreases

115

3.4 Distortions due to magnetic and electric fields

0.8

RH

Fig. 3.4.11. Threshold plots of RH (the ratio of the periodic distortion threshold
and the normal (homogeneous) Freedericksz threshold) and of Qyc (the dimensionless wavevector of periodicity at threshold) versus k1/k22. Periodic
distortion is possible only when kxjk22 ^ 3.3; for kxjk22 ^ 3.3 the normal
Freedericksz deformation is favourable. (After reference 42.)
from Hz(0) as Qy increases, reaching a minimum Hz(Qyc) at Qy = Qyc. On
increasing Qy beyond Qyc, Hz(Qy) increases again. Hz(Qyc) is regarded as
the threshold for periodic distortion, and Qyc (and kyc = nd/Qyc) the
wavevector (and wavelength) of periodicity at threshold.
Plots of the ratio RH = Hz(Qyc)/Hz(0) and Qyc as functions of r =
fcn/fc22 are shown in fig. 3.4.11. It is seen that as r decreases RH increases
and Qye decreases. When r tends to a lower limiting value rc ~ 3.3, RH^\,
Qyc-+0, showing that for r < rc, the periodic distortion is no longer
favourable. The transition to the periodically distorted state can be treated
as a second order phase transition with Qy as the order parameter. <43"5)

3.4.4 The constant kls


(46)

Nehring and Saupe put forward the view that linear terms of the second
derivatives of n in the nematic free energy expansion can, in principle,
make contributions of the same order as the quadratic terms of the first

116

3. Continuum theory of the nematic state

derivative, and proposed the following more general expression instead of


(3.3.7):
F = ^ ( V - n ) 2 + y n - V x n ) 2 + ^ ( n x y x n ) 2 + 2yv-(V-n)n]

(3.4.18)

where k'xl = klx-2klz


and k'33 = kS3 + 2kls.
As far as simple splay and bend deformations are concerned, kn and kss
have, in effect, been rescaled and therefore do not yield an estimate of the
absolute value of k13. The term A:13[V (V n) n] satisfies the Euler-Lagrange
equation identically, i.e., it does not contribute to the bulk equilibrium
configuration of the director. Thus kls cannot be measured by the usual
techniques. Its influence on the director configuration can be detected only
if the anchoring at the surfaces is weak. (4748) However, Oldano and
Barbero(49) have pointed out that there is a mathematical difficulty in
estimating its value quantitatively. Assuming that 9, the director tilt, and
d8/dz are independent variables at the boundaries, they have shown that
the variational problem does not have a solution corresponding to the k13
term. This leads to a discontinuous variation of 9 at the boundaries, in
contradiction to the basic principles of the continuum theory. Inclusion of
higher order elasticity and surface terms may perhaps restore a continuous
variation of #?(50'51) but little is known at present about these additional
constants. On the other hand, Hinov(52) has argued that a continuous
solution is possible if 9, dO/dz and their variations are treated as dependent
functions.
A determination of kls is of some interest for it has been suggested that
this constant may play a role in certain special situations, as e.g., in the
measurement of the flexoelectric coefficient of a nematic(53) (see 3.13).
Attempts have been made to estimate k13 experimentally using samples
that are weakly anchored at one or both boundaries.(54"6> As an illustration
of the principle involved, we describe one such experiment(56) employing a
' hybrid' aligned cell with weak homeotropic anchoring on one glass plate
and strong homogeneous anchoring on the other. In such a cell the director
makes a small angle 9 with respect to the normal at the weakly anchored
surface. A magnetic field H is then applied perpendicular to the plates to
change the director profile. The tilt angle 9 is measured by optical methods
as a function of //, and the value of k13 extracted from the surface torque
balance equations (assuming the Euler-Lagrange solution to be valid
right up to the surface of the sample). The value was found to be
( 9 4 ) x 10~7 dyn for /?-cyano-//-heptyl-phenyl-cyclohexane at 25 C. In
view of the difficulties referred to earlier, this can only be taken to be an
approximate effective value which may have contributions from other

3.5 Disclinations

117

terms influencing the director configuration at the surface. The possible


existence of a spontaneously splayed state in a free standing film when kls
is sufficiently large and the film thickness sufficiently small has also been
investigated/ 57 ' 58)
3.5 Disclinations
5.5.7 Schlieren textures
As remarked in chapter 1, the nematic state is named for the threads that
can be seen within the fluid under a microscope (fig. 1.1.6(0)). In thin films
sandwiched between glass plates these threads can be seen end on. A
typical example of the texture in a plane film of thickness about 10 jum
between crossed polarizers - the structures a noyaux or schlieren textures
- is given in fig. 1.1.6(b). The black brushes originating from the points are
due to 'line singularities' perpendicular to the layer. In analogy with
dislocations in crystals, Frank<3) proposed the term 'disinclinations',
which has since been modified to disclinations in current usage.
The brushes are regions where the director (or the local optical axis) is
either parallel or perpendicular to the plane of polarization of the incident
light. The polarization is unchanged by the material in these regions and is
therefore extinguished by the crossed analyser. Some points have four
black brushes while others have only two. The positions of the points
remain unchanged on rotating the crossed polarizers but the brushes
themselves rotate continuously showing that the orientation of the director
changes continuously about the disclinations. The sense of rotation may be
either the same as that of the polarizers (positive disclinations), or opposite
(negative disclinations). The rate of rotation is about equal to that of the
polarizers when the disclination has four brushes and is twice as fast when
it has only two.
The strength of a disclination is defined as s = |(number of brushes).
Only disclinations of strengths s =+%, , + 1 and 1 are generally
observed. Neighbouring disclinations connected by brushes are of opposite
signs and the sum of the strengths of all disclinations in a sample tends to
be zero. At temperatures close to 7^I9 disclinations of opposite signs are
seen to attract each other and coalesce. They may then disappear altogether
(^ + 5*2 = 0) or form a new singularity (s1 + s2 = s').
The significance of these textures was understood by Lehmann(59) and
Friedel,(60) but a mathematical description of the actual configuration
around disclinations was given by Oseen(1) and Frank (3) . The subject has
since been treated in greater detail by Nehring and Saupe,(61) and reviewed
comprehensively in a number of articles.(62~5)

118

3. Continuum theory of the nematic state

Fig. 3.5.1. Director orientation (indicated by arrows) along a polar line making an
angle a. Incident light that is linearly polarized at angle ^ or (j>n/2 will be
extinguished by a crossed analyser and will give rise to a dark brush.
Consider a planar structure in which the director is confined to the xy
plane (the z axis being normal to the film). Taking the components of the
director to be nx = cos(/>, ny = sin (/>, nz = 0, and making the simplifying assumption that the medium is elastically isotropic, i.e., k1 = k22 = 33 = k,
(3.3.7) and (3.3.17) reduce to
(3.5.1)
(3.5.2)

= 0

respectively. We seek simple solutions that are independent of r =


(x2+y2)K The solutions of (3.5.2) are <j> = 0, which is of no interest, and
(3.5.3)

where a = tan" 1 (y/x) and c is a constant. This equation describes the


director configuration around the disclination; the singular line is along
the z axis and the director orientation changes by 2ns on going round the
line. If the orientational order is apolar, it is clear that a rotation of mn
(where m is an integer) in the director orientation </> should correspond to
a rotation of 2n in the polar angle a (fig. 3.5.1). On the other hand, for a
polar medium a change of 2mn in (f> should correspond to a change of 2n
in a. More generally
s = + | , + 1, + . . . ,
s = l, 2 , 3 , . . . ,

with 0 < c < n

(apolar)

with 0 < c < 2n (polar)

s is called the strength of the disclination.


In terms of the Volterra process<66) one can visualize the topological
features of these disclinations in the following way. Cut the material by a
plane that is parallel to the director. The limit of this cut is a line L called

119

5.5 Disclinations

* = *

5=1, C= 0

S=l,C=7T/2

s=\,c=n/4

5=3/2

5=2

Fig. 3.5.2. Molecular orientation in the neighbourhood of a disclination. (After


Frank.(3))
the disclination line. Rotate one face of the cut with respect to the other by
a relative angle 2ns about an axis perpendicular to the director. Remove
material from the overlapping regions or add material to fill in the voids,
and allow the system to relax. If the axis of rotation is parallel to the line
L, as is true for the case under consideration, the disclinations may be
referred to as wedge disclinations.(67) This distinguishes them from twist
disclinations, which will be considered in 3.5.4.
We shall now show that s = ^number of brushes). If light is incident
normal to the film and linearly polarized at an angle (p with respect to the

120

3. Continuum theory of the nematic state

x axis, it is seen from fig. 3.5.1 that the polarization will be unchanged at
all points on the polar line a and hence will not be transmitted by the
analyser. This will result in a black brush at an angle a. A similar situation
will arise when <j> changes by n/2. The angle between two successive dark
brushes is therefore Aa = A<p/s = n/2s. Thus the number of dark brushes
per singularity is 2n/\Aa\ = 4|s|. Also if the polarizers are turned through
angle co the brushes rotate by an angle co/s. The rate of rotation of the
brushes of the two-brush disclination (s = ) is therefore twice as fast as
that of the four-brush variety (s = 1). If the polarizers are kept fixed and
the microscope stage is rotated by co the brushes turn by co{s \)/s.
Observations in polarized light therefore enable one to determine s both in
sign and magnitude. The existence of |.s| = \ in nematic liquid crystals
establishes the absence of polarity in this phase.
If the structure is one in which the director is not normal but inclined
with respect to the z axis (as in the case of smectic C, see 5.8), half integral
values of s are not possible even if the molecular order is apolar.
The molecular orientation in the neighbourhood of a disclination as
given by (3.5.3) is shown in fig. 3.5.2 for a few values of s. For s + 1, a
change Ac in the constant c merely causes a rotation of the figure by
Ac/(1 s), while for s = 1, the pattern itself is changed.
When the strengths of two neighbouring disclinations are equal and
opposite, the brushes connecting them are circular. By superposition of
solutions of the type (3.5.3)
<f> = </>i + 4>2 = ^i <h + s2 a 2 + c o n s t .
Putting sx = s2 = s,
(j> = sfi + const.,

where /? = ax a 2. Thus curves of constant (/> will be arcs of circles passing


through the two disclinations (fig. 3.5.3). For s = \, </> changes by n/2 on
going from one side of the chord to the other (since ft = n/? or
$' = ns <t>) while for s = 1 it is unchanged. An example of such circular
brushes is shown in fig. 3.5.4.
3.5.2 Interaction between disclinations
The energy of deformation of an isolated disclination in a circular layer of
radius R and of unit thickness is(61>68)
W=

\FAxAy.

(3.5.4)

3.5 Disclinations

121

Fig. 3.5.3. Curves of equal alignment around a pair of singularities of equal and
opposite strengths. The orientations marked on the circles refer to the case 5 = 1 ,
The disclination is supposed to have a core whose energy is not known. To
allow for this, we postulate a cut-off radius rc around the disclination and
integrate for distances greater than rc to obtain
W= Wc +

nks2\n(R/rc),

(3.5.5)

where Wc is the energy of the central region. As R^co,


W-^cc
logarithmically, i.e., an isolated disclination in an infinitely extended layer
has infinite energy, but such a situation does not arise in practice because
of the presence of disclination pairs of opposite signs.
The energy of a single defect being proportional to s2 according to the
planar model that we have just considered, defects of strength \s\ > \
should be unstable and should dissociate into |.s| = \ defects. However, as
is evident from fig. 1.1.6(6), stable defects of strength \s\ = 1 (with four
brushes) occur very frequently. The reason for this will be discussed in
3.5.3.
The interaction between disclinations may be calculated by superposing
solutions of the form (3.5.3),
t tan"

+ const.

(3.5.6)

Proceeding as before, we obtain for a pair of disclinations separated by a


distance r12(69)
W = nk{s1 + s2f In (R/rc) - 2nksx s2 In (r 12 /2r c ).

(3.5.7)

The assumption here is that rc <^ r12 <^ R. If s1 = s 2, E becomes

122

3. Continuum theory of the nematic state

{a)

Fig. 3.5.4. Brushes connecting a pair of disclinations of equal and opposite


strengths, s = 1 and 1, in nematic MBBA. Crossed polarizers rotated clockwise
by 22.5 in each successive photograph. In (d) the directions of extinction are
parallel to the edges of the picture. (Nehring and Saupe.(61))
independent of R. (This is also true if there are many defects in the layer
and YJ si O.(7O)) The interaction energy is given by the second term on the
right-hand side of (3.5.7). The force between two singularities is therefore
2nks1s2/r12. Accordingly disclinations of like signs repel and those of
opposite signs attract, the force being inversely proportional to the
distance.
The properties of disclinations in nematics bear some striking similarities
with screw disclinations in crystals (7172) and vortex filaments in superfluids/ 73 ' 74) but at the same time there are important differences that
cannot be overlooked while drawing detailed analogies. (64)

3.5 Disclinations

123

(a)

(b)

SO/mi

Fig. 3.5.5. (a) s = 1 disclinations and (b) s = \ disclination in a nematic film; (left)
between crossed linear polarizers; (right) between crossed circular polarizers.
(Meyer.(76))

5.5.5 Non-singular structures (s = 1): escape in the third dimension


Cladis and Kleman (75) and Meyer(76) showed that the singularity at the
origin of the |^| = 1 defect as given by the planar model can be avoided by
a non-singular continuous structure of lower energy. The director
orientation now 'escapes' in the z direction. Optical observations confirm
that this does indeed happen close to the origin of \s\ = 1 defects (fig. 3.5.5).
Structures of this type are also formed in thin capillaries (fig. 3.5.6).
Let us consider the nematic in a capillary of radius r0 with the director
homeotropically aligned at the wall (i.e., with the director normal to the

Continuum theory of the nematic state

124

I
\

\
(a)

Fig. 3.5.6. (a) Director 'escape' at the centre of a disclination of strength s = 1 in


a thin capillary: the wall alignment is homeotropic and changes by 90 from wall
to axis. (Williams, Pieranski and Cladis(77).) (b) Projection of the structure on a
plane normal to the capillary axis. Nails signify that the director is tilted with
respect to the plane of the paper. Solutions with positive and negative tilts are
equally probable.
surface). With the planar solution this would lead to the structure s = + 1,
c = 0 (see fig. 3.5.2). However, if we allow for the possibility of a director
tilt towards the capillary axis (z axis), we may assume in the region
0 < r < r0, nx = sin <j> sin 0, ny = cos $ sin 0 and nz = cos 0. The energy of
deformation per unit length is then

W = \k\ [(V0)2 + sin2 0(V</>)2 + 2 sin 0cos 0{V<p x V0}] dr.

(3.5.8)

The equations of equilibrium (3.3.17) become


= 0.

(3.5.9)

We look for a solution with (j> = ^(a) and 9 = 6{r). Therefore, (3.5.9)
reduce to
(3.5.10)
- l - f r ? \ - sin 0 cos 9(V</>)2 = 0,

8V = 0.

(3.5.11)

The solution of (3.5.11) is


while 0 = 0(r) can be determined from
1 8 / 80>

= 0.

5.5 Disclinations

^
-

125

S*

T
'
_^ _\^

** **
(

^ ^
^_

^
-

\ \ I//
(a)

(b)

Fig. 3.5.7. Escaped configurations of (a) s = I, c = n/2 and (b) s = 1 disclinations. Nails signify that the director is tilted with respect to the plane of the paper.
Assuming the boundary condition
0
at r = 0,
= n/2 at r = rQ,
we get

(3.5.12)

This is an escape involving bend and splay (fig. 3.5.6).


In (3.5.8), the first two terms in the square brackets represent volume
integrations, which together give 27^1^1, while the third term is a surface
integration whose value is nks. Thus the total energy
_ (3nk for s = + l,

\nk

for s = 1.

Interestingly, the energy is independent of r0. On the other hand, the planar
structure gives

W=nk\n(rjrc)+Wc
for s = 1, where rc is the core radius and Ec the core energy. As rc is
expected to be of molecular dimensions, the planar solution has much
higher energy than the continuous structure if r0 is large enough for optical
observations. On the other hand if the capillary radius r0 is extremely small,
or the elastic constant very large (as can happen in the vicinity of a
nematic-smectic A transition) the planar solution may be more favourable
energetically.
The escaped configurations for s = 1, c = n/2 (involving bend and
twist), and for s = 1 (involving twist, bend and splay) are shown in fig.
3.5.7. Structures with the nails pointing in the reverse direction (i.e., with
nz being replaced by n z) are equally probable.

3. Continuum theory of the nematic state

126

z>0

z>0
z=0
-\//'---

z<0

z>0

= 1

z<0

z>0

\ ^--*
\ -v - ^

z=0

\ w \\\
\\\ \\ \

z<0

z<0

z>0

z>0

\ .-//
\

1 1
1 1

z=0
z<0

'

1 1

z<0

(a)

(ft)

Fig. 3.5.8. Twist disclinations: the director patterns for (<z) s = J, c = 0, n/4 and
7r/2; and (Z?) s = 1, c = 0, rc/4 and n/2. In each case, the disclination line is shown
as a full line in the middle of the z = 0 plane. The patterns for (s, c) and ( s, c)
are mirror images of each other.

3.5.4 Twist disclinations


Equation (3.5.2) admits of another type of singular solution (Friedel and
de Gennes (78) ): the director is parallel to the xy plane as before, but <j> is now
a function of x and z. The solution is
(f> = s t2Ln~\z/x) + c

where s is the strength of the disclination and c a constant. The axis of


rotation (z axis) is now at right angles to the singular line (y axis). Thus this
is referred to as the twist disclination. The director fields for a few values of
s and c are illustrated in fig. 3.5.8. Negative strengths are not shown in the
figure, because the structures of (s, c) and ( s,c) are mirror images of

5.5 Disclinations

127

each other. It is seen that all three types of elastic distortions, splay, twist
and bend, are now present whereas planar solutions for the wedge
disclination involve only splay and bend.
In the elastically isotropic medium, the expressions for the energies and
interactions derived for wedge disclinations are exactly applicable to the
present case. Similarly, the non-singular solution (3.5.11) for s = 1 is valid
in this case, except that the energy for the escaped configuration turns out
to be 2nk\s\ per unit length.(76) The structure of the escaped configuration
is, of course, rather more complicated than that for the wedge disclination.
From the nature of the director patterns it is clear that dark brushes of
the schlieren type will not be seen under the polarizing microscope for light
propagating normal to the film (see 4.1.1). Twist disclinations may
therefore be expected to be less conspicuous than wedge disclinations, and
few observations have been reported of their existence in ordinary
nematics. They do, however, reveal themselves under favourable circumstances in twisted nematics, often as loops separating regions of
different twist.(79)
The Volterra process for creating a loop, i.e., a closed disclination line,
in a nematic is as follows. Let be the surface enclosed by the loop L. Call
the two sides of the surface S + and S~. Rotate the molecules in contact with
I + by an angle sn and those in contact with Z~ by sn about an axis
normal to the unperturbed orientation of the director, where s = |, 1,
etc., is the strength of the disclination line. At finite distances from S the
director will adjust itself and the resulting configuration will be continuous
everywhere except on I .
We now consider a twist disclination loop in a twisted nematic. The
nematic is supposed to have a planar structure with the director parallel to
the xy plane and an imposed twist of q per unit length about the z axis, and
the disclination loop of radius R is supposed to be in the xy plane. The
director distortions are planar, nx = cos^, ny = sin^, nz = 0. On going
once round the disclination line at any point on the loop, the director
orientation <f> changes by 2ns, the sign of which may be either the same as
that of q or opposite.
The net energy of such a structure is(78)
W = n2ks[2sR In (R/rc) - qR2].
When s and q are of the same sign, W has a maximum value when

128

3. Continuum theory of the nematic state

Fig. 3.5.9. Shrinking of twist disclination loops: (a) thin thread |s| = |, (b) thick
thread \s\ = 1 with an escaped (coreless) structure. Each thread was photographed
twice several seconds apart. (Nehring.(80))
The energy decreases for higher and lower values of R. Thus large loops
with R> Ro may be expected to occur. Smaller loops shrink (fig. 3.5.9) and
disappear. When s and q are of opposite signs, loops may not be expected
to occur at all.
A stability analysis(81) has shown that twist disclinations are less
favourable than wedge disclinations in elastically anisotropic media. This
may explain why the former are so rarely seen in ordinary nematics.

3.5 Disclinations

129

(a)

Fig. 3.5.10. Singular points in droplets: (a) spherically symmetric radial (hedgehog)
configuration with the director normal to the surface; (b) bipolar structure with the
director tangential to the surface; (c) singular points in a capillary.
3.5.5 Singular points
Point singularities of strength s = 1 occur in droplets and can also be
seen in thin capillaries (fig. 3.5.10). To obtain their solutions (68) let us set
nx cos (/> sin 9, ny = sin <j> sin 9, nz = cos 9, and use spherical coordinates
x = p sin S cos a, y = /?sin^sina, z = pcosS. If we assume that <j> = ^(a)
and 9 = 0(8)9 we find that

while 9 can be obtained from the differential equation


c)29
oo

:
do

= 0.

Selecting solutions with 9 -> S for p -> 0,


tan

D-M9I

For \s\ = 1,9 = S. On the other hand, if 9 -> n S as p -> 0, then c becomes
c + n.
Fig. 3.5.11 gives the director configurations for some typical cases. The
sections through the (x, y), and the (x,z) or (y,z) planes are identical with
the patterns for the +1 and 1 wedge disclinations in two dimensions.
Any pattern on the left-hand side of fig. 3.5.11 may be combined with any
one on the right to give a possible point singularity.

130

3. Continuum theory of the nematic state


y

s = 1,

c = n

Fig. 3.5.11. Director field around singular points. Any pattern on the left may be
combined with any one on the right to give the field around a possible point
singularity. (Saupe.(68))
The total energy for a spherically symmetric radial configuration is
E = %nkR.
where R is the radius of the sphere.(82) This configuration (sometimes
referred to as the ' hedgehog' point defect) is realized in droplets with the
director normal to the surface (fig. 3.5.10(a)).
If the boundary condition is tangential, point defects are formed at the
two poles (fig. 3.5.10(b)). The director components in cylindrical polars
may be taken as nr = sin 6, na = 0, nz = cos 0, and assuming
0 = tan" 1

rz

R2-r

it turns out that for this bipolar structure(82)

E 5nkR.
Elastic anisotropy modifies the idealized configurations shown in fig.
3.5.10.(83>84) More complex structures with an oblique orientation of the
director at the surface have also been reported/ 85 ' 86)
Point singularities of equal and opposite strengths attract one another
and are annihilated(68'87) (see fig. 3.5.12). As the total energy of elastic
deformation around a point defect increases linearly with the radius of the

3.5 Disclinations

131

Fig. 3.5.12. A sequence of photographs demonstrating the attraction and


annihilation of singular points of opposite strengths. (Saupe.(68))

132

3. Continuum theory of the nematic state

volume enclosed, it has been suggested that the interaction energy of two
defects grows linearly with separation, analogous to quarks interacting
through a gluon field.(65)
3.5.6 Interaction between disclinations and surfaces
Interaction with a plane surface
From the superposition principle (3.5.6) we know that the director pattern
around a pair of like disclinations located at x = d and d is given by

X l

Hence, at all points (0, y) on the midplane z = 0, the director orientation


<p = sn + 0o = 0W. We may therefore conclude that a wall located at the
midplane with the director firmly anchored on it at an angle 0W can be
replaced by an image of the defect. Consequently the wall repels the defect
with a force
/ = nks2/d.
Interaction with an air bubble: formation of a dipole pair
It has been observed that an air bubble in a nematic sample interacts with
a point defect.(87) A (1) point is attracted to the bubble and settles at
some equilibrium distance from it to form a stable dipole pair (fig. 3.5.13).
We shall consider the analogous, though simpler, problem of the
interaction between a line defect and a cylindrical cavity, the line of
singularity being parallel to the axis of the cylinder(88) (fig. 3.5.14). We shall
assume uniform boundary conditions at the surface of the cylinder, i.e.,
that the director is everywhere inclined at the same angle with respect to the
radius vector. Let the disclination of strength s be at A at a distance D from
the centre of the cylinder of radius R (i.e., r = D and a = 0). Now, let us
place a disclination of strength (ls) at the centre of the cylinder and
another of strength s at A' at a distance D' such that DD' = R2. The net
orientation at any point P(x, y) on the cylinder is then
y ,| + ,stan(

-DJ

\x-D

which implies a uniform alignment of the director on the surface of the


cylinder. Thus in the presence of a disclination the cylindrical cavity can be
replaced mathematically by two disclinations, one of strength (1 s) at its
centre and another of strength s at the conjugate point.

3.5 Disclinations

133

(a)

(b)

Fig. 3.5.13. Stable dipole pairs formed between 1 point disclinations and air
bubbles in a nematic between (a) crossed linear polarizers and (b) crossed circular
polarizers. (Meyer.(87))

134

3. Continuum theory of the nematic state


y

<x,y)

Fig. 3.5.14. A line singularity of strength s is located at A at a distance D from the


centre 0 of a cylindrical cavity of radius R.

The net force on the disclination at A is


D

s2
DD

r0 being a unit vector directed away from the centre towards the singularity.
At large distances from the cylinder (D $> R)

Consequently, at far off points the cavity behaves as a + 1 disclination


located at its centre. A negative disclination (s < 0) at a large distance away
will therefore be attracted by the cylinder.
In the neighbourhood of the cylinder (D R),

and the disclination is repelled by the cavity.


At a certain intermediate distance given by

the force / = 0. This represents a position of equilibrium and the


disclination and the cavity form a dipole pair, similar to the ones observed
by Meyer (fig. 3.5.13).

5.5 Disclinations

(a)

135

(b)

(c)

Fig. 3.5.15. Helfrich walls: (a) a twist wall parallel to thefield,(b) a bend-splay wall
parallel to the field, and (c) a splay-bend wall perpendicular to the field.
On the other hand, a positive disclination will be repelled by the cavity
at all distances.

3.5.7 Defects in the presence of an external field


Helfrich walls
When a nematic of positive diamagnetic anisotropy (ja > 0) is placed in a
magnetic field, the director aligns itself parallel or, equivalently, antiparallel with respect to the field direction. A region of parallel alignment
and one of antiparallel alignment can be separated by a wall, inside which
the director turns through an angle n. There are three possibilities. (89)
(a) A twist wall: The field is along the z axis and the director is confined to
the yz plane and twists about x (fig. 3.5.15(#)). The wall, which is
parallel to the field direction, is analogous to the Bloch wall in a
ferromagnet.
(b) A bend-splay wall: The field is along z and the director is confined to
the zx plane (fig. 3.5.15 (b)). The transition from + n to n takes place
mainly through a bend deformation, though some splay is also present.
The wall is parallel to H and may be compared with the Neel wall in
ferromagnetic systems.
(c) A splay-bend wall: In this case the transition from + n to n is
predominantly through splay but there is some bend as well (fig.
3.5.15(c)). It is an 'inversion' wall perpendicular to H.
The thickness of the Helfrich wall is of the order of 2, where is the
magnetic coherence length defined as H'^k/xJ* (see 3.4.1), and the wall
energy per unit length

3. Continuum theory of the nematic state

136

H<HC

H>He

H> Hc

(a)

(b)

(c)

Fig. 3.5.16. Brochard-Leger walls: (a) initial unperturbed orientation of the


director when thefieldHis less than the threshold value Hc; (b) H > Hc: twist wall
parallel to thefield,full line for y > 0 and dashed line for y < 0; (c) H > He: splay
wall perpendicular to the field.
Brochard-Leger walls
Such walls are associated with the Freedericksz deformation. (90) With the
homeotropic geometry of Fig. 3.4.1 (c), the possible distortions for H > Hc
are illustrated in fig. 3.5.16. Since the director tilt has a degeneracy in sign
with respect to H, there can arise twist walls parallel to the field (fig.
3.5.16(&)) or splay walls perpendicular to the field (fig. 3.5.16(c)). Similarly
with the homogeneous geometry, there can arise bend walls.
In this case the wall thickness t ~ /a, where
a2 = 31 1

d2

t diverges as H-+Hc. The wall energy per unit length

W---1
Umbilics
Consider a nematic film of negative dielectric anisotropy (ea < 0) aligned
homeotropically between glass plates. If an electric field is applied along
the director axis (z axis) a distortion will set in when the field exceeds the
critical Freedericksz value given by

Ec =

(n/d)(k3S/Ej.

137

5.5 Disclinations

I I I I I lllll
11111 T T T T T
lllll

TT

TTT

iilll T T T T T
lllll l l l l l

T 1

[1
\

Fig. 3.5.17. The structure of umbilics of strength 5 = 1 . Nails signify that the
director is inclined with respect to the plane of the paper.
(The experiment can also be performed with a magnetic field and a
negative diamagnetic anisotropy material like a discotic nematic, see 6.5.)
Two possible types of distortions are depicted in fig. 3.5.17. In the
distorted state we have a n component which is degenerate in the xy plane.
Therefore, there can be defects in the n field, and because of the symmetry
in the xy plane, only defects of integral strength can occur. Such defects
have been observed*87>68) (fig. 3.5.18) and are called umbilics.m) They are
somewhat similar to the s = + 1 of the schlieren texture but differ in detail.
Over a distance

from the centre of the defect, the director gradually tilts towards the z axis,
the tilt angle becoming exactly zero at r = 0. In this sense, umbilics have
a collapsed core, but the structure is not exactly the same as that of the
s = 1 defects discussed in 3.5.3.
Planar and linear solitons
When a magnetic field is applied normal to a half-integral disclination line
in a nematic, there results a domain wall terminating in a singular line. The
wall thickness is of the order of . Figs. 3.5.19(#) and (b) illustrate the
director patterns for s = \ and \. Such walls have been referred to as
planar solitons.m) On the other hand, a magnetic field acting on a point
defect with a radial configuration will give rise to a cylindrical domain

3. Continuum theory of the nematic state

138

Fig. 3.5.18. Umbilics induced by an electric field in a nematic film of negative


dielectric anisotropy: crossed polarizers. (Saupe. (68))

(a)

Fig. 3.5.19. Planar soliton produced by a field acting on (a) s = \ and (b) s = \
disclination lines.

3.5 Disclinations

139

ccccc
ccccc
Fig. 3.5.20. Linear soliton produced by afieldacting on a point defect of strength
^= 1.
ending in a singular point (fig. 3.5.20). This has been called a linear soliton.
The occurrence of such solitons in the phases of superfluid 3 He,
antiferromagnetics, etc. has been discussed by Mineev(93).
Other interesting cases have been investigated by Sunil Kumar and
Ranganath. (94)
3.5.8 Some consequences of elastic anisotropy
Real nematics are, of course, elastically anisotropic. In certain situations,
as for example at temperatures close to the nematic-smectic transition, the
anisotropy becomes very large and certainly cannot be ignored. We shall
now investigate some of the consequences of elastic anisotropy on the
properties of disclinations.
Solutions for isolated disclinations
As before, we shall begin by considering a planar sample in which the
director is confined to the xy plane. In such a case, a wedge disclination
involves only splay and bend distortions and we need to take into account
only the splay-bend anisotropy (fcn ==
j fc33).
The free energy density (3.3.7) may be written as
F=^[l+cos2(^-a)]/r2,
where (px = d</>/doc, a = tan~ x (y/x), k = (k11 + k33) and e =

(3.5.13)
(k11-k33)/

(&n + 33), and the equation of equilibrium is


-^)sin2(^-a)],
2

(3.5.14)

where ^ aa = d <p/da . For any general value of e, (3.5.14) can be solved


numerically and energies can be evaluated. (95) However, for small e one can
employ a perturbation technique to obtain analytical solutions. (61) For
s #= 1, <f> may be expressed as
e2y/(2)

+....

140

3. Continuum theory of the nematic state

(a)

(b)

Fig. 3.5.21. The structure of s = \ in the extreme limits of (a)


e = (fcn-fcssV^u + A^) = 1 and (6) e = - 1 .
Substituting in (3.5.14) and integrating one obtains to a first order in e
(3.5.15)
where c and c' are constants. The energy of the singularity (for s 4= 1) is
(3.5.16)
which may be compared with (3.5.5) for the isotropic case. It may be noted
that s and s have different energies because of elastic anisotropy.
For s = 1, one obtains the elastically isotropic solution with c = 0 or n/2
to be a solution of (3.5.14) depending upon whether e is negative or
positive, and for s = 2 one gets the isotropic solution irrespective of e.
In the limit of e = 1, analytical solutions can be obtained for s = \
and -l : ( 96 ' 97 >
(3.5.17)
with C = | for s = \ and C = 2 for s = 1. The corresponding energy is
given by
(3.5.18)
For s = |, e = 1, one gets a solution with </> = 0 for a between n/2 and
and ^ = a {n/2) for a between n/2 and 3TT/2. In this solution,
sketched in fig. 3.5.21 (a), splay is avoided as it requires infinitely greater

TT/2,

141

3.5 Disclinations

0.2 -

1.0

Fig. 3.5.22. Dependence of the energy on e = (k1k zz)/(kxl + A;33) for 5 = \ and |
disclinations. Squares are from the Nehring-Saupe approximation and circles from
the perturbation calculation in the neighbourhood of s = 1. (After reference 97.)
energy than bend when e = 1. The solution for s = |, e = 1 is also shown
in fig. 3.5.21(6).
For a general value of e, the corrections can be evaluated by numerical
methods. Fig. 3.5.22 gives the energies of s = + | and s = \ defects in
units of k\n(R/rc). The results of the Nehring-Saupe formula are plotted
as squares, and those calculated from a perturbation expansion in the
neighbourhood of s = 1 are marked as circles.
Interaction between disclinations
The radial force of interaction between disclination will, of course, be
modified because of anisotropy. (97) In addition there will now be an
angular component of the force.(98) The physical basis for the angular force
can be understood by referring to fig. 3.5.23, which shows the director
patterns for two pairs of unlike defects, ( + f, |) and ( + 1 , - 1 ) , each in
two different situations. It is seen that there are significant differences in the

3. Continuum theory of the nematic state

142

(-L+!)

(-1,+

c= 0

Fig. 3.5.23. The director patterns for (, ) and (1, 1) defect pairs in two
situations, c = 0 and c = n/2. The double-headed arrow at the centre indicates the
director orientation far away from the defect pairs. (After reference 98.)

patterns depending on whether the director at large distances is parallel or


perpendicular to the line joining the defects. For the ( + |, |) pair, the
central region is predominantly bend for one configuration and predominantly splay for the other. In the case of the ( + 1, 1) pair, a structure
that is mainly splay becomes one that is mainly bend and, moreover, the
+ 1 defect itself changes from a radial to a circular pattern. Thus depending
on the sign of the elastic anisotropy, one or the other configuration is
energetically favoured. In other words, given a boundary condition, an
angular force comes into play. This force can be computed to a first order
in s. It is given by
where
=

\[(O2x-Ol

with
= s tan" 1

x-d

6X = and
ox

tan"1|

(3.5.19)

0 y = .
dy

The positive sign in (3.5.19) is taken for like singularities and the negative
sign for unlike singularities.

3.5 Disclinations

143

Fig. 3.5.24. Possible helical configurations of disclination pairs in twisted nematics


or cholesterics of large pitch (see figs. 4.2.2 and 4.2.3).
The above arguments do not hold for Yjisii = + 1 o r a n v si = + * This is
because when e =N 0, the structure of s = + 1 can be either radial or circular
but cannot have the intermediate logarithmic spiral pattern, as was, in fact,
first pointed out by Frank. (3) Thus c has to be 0 or n/2 depending on the
sign of .
In summary, the presence of elastic anisotropy favours one particular
value of the constant c for a disclination pair. This may have a bearing on
the structure of disclination pairs in long-pitched cholesterics in which each
layer may be regarded as nematic-like. Since each layer is allowed only one
value of c, and the layers themselves are rotated about the twist axis,
disclination pairs may be expected to adopt helical configurations as
illustrated schematically in fig. 3.5.24. As we shall see in 4.2.1, this
conclusion seems to be in agreement with experimental observations.

3.5.9 The core structure


The nature of the core still remains an interesting unsolved problem. We
have seen in 3.1.1 that director distortions have stresses associated with
them as given by (3.3.4). In the case of a single disclination the stress is a
tension which can be expressed as

in the one-constant approximation. Ericksen(99) postulated that since the


tension increases on approaching the centre of the defect, below a certain
critical radius rc it should be large enough to transform the material from
the nematic to the isotropic phase. Thus the core should consist of a

144

3. Continuum theory of the nematic state

cylinder of isotropic material. The nematic-isotropic transition being of


first order there will be a physical interface between the two regions.
Ericksen showed that the surface-director couple vanishes on this
interface. By the same argument, it is reasonable to expect a spherical
isotropic droplet at the centre of the hedgehog (all radial) singular point
(fig. 3.5.10(a)) and isotropic regions near the surface in the bipolar
structure (fig. 3.5.10(b)).
However, an important parameter that has been ignored in this
approach is the surface tension at the interface. The interfacial tension T
can be taken into account in an elementary way as is generally done for
crystal screw dislocations.(100) The total energy of the disclination in the
one-constant approximation, including the energy at the core surface, is

E=2nrT+nks2

In (R/r)

the minimization of which gives the radius of the core


rc = ks2/2T.
Typically, rc ~ 10~6 cm. Of course, the complete analysis has to include the
surface anchoring energy, latent heat, etc.
3.6 Flow properties
3.6.1 Miesowicz's experiment
We shall now discuss the application of the Ericksen-Leslie theory to some
practical problems in viscometry.(101) Probably the first precise determination of the anisotropic viscosity of a nematic liquid crystal was by
Miesowicz.(102) He oriented the sample by applying a strong magnetic field
and measured the viscosity coefficients in the following three geometries
using an oscillating plate viscometer:
(i) n parallel to the flow;
(ii) n parallel to the velocity gradient;
(iii) n perpendicular to the flow and to the velocity gradient.
In the presence of a strong field, the magnetic coherence length is quite
small and one may without sensible error neglect boundary effects and
director gradients. The observational data for the three geometries can
then be readily interpreted on the basis of the equations given in 3.3.
The apparent viscosity for any geometry
rj =

shear stress
velocity gradient

Lt
2dtj'

3.6 Flow properties

145

Table 3.6.1. Miesowicz's viscosity coefficients (measurements in 10~2 P)

/?-Azoxyanisole
122 C
/7-Azoxyphenetole
144.4 C

Molecules
parallel
to flow
direction

Molecules
parallel to
velocity
gradient

Molecules
perpendicular
to flow direction
and to velocity
gradient

2.4 0.05

9.2 0.4

3.40.3

1.3 0.05

8.30.4

2.50.3

If we take the flow to be along x and the velocity gradient along y,

and
xy

yx

2 x, y

Since director gradients are neglected, the elastic part of the stress tensor
t% = 0. Thus, for n = (1,0,0) we have from (3.3.5)

or
Similarly

= XMiesowicz's results for two compounds are presented in table 3.6.1.


3.6.2 Tsvetkov's experiment
In this experiment, a tube containing a nematic liquid crystal is suspended
in a uniform magnetic field acting in a horizontal plane and is spun at a
constant angular velocity Q about a vertical axis. If the axis of rotation is
along z and the magnetic field along y, the components of the bulk velocity
of the fluid are
vx Cly, vy = Qx

and

vz = 0.

Since the director lies in the xy plane n = (cos cp, sin (p, 0). If the diameter of

146

3. Continuum theory of the nematic state

the tube is large enough, wall effects and director gradients may be
neglected. Therefore (3.3.2) may be written as

where g\ is given by (3.3.13) and G{ by (3.4.1). Therefore


(3.6.1)
where Xx = JU2JUZ. Below a critical angular velocity Q c , (3.6.1) has the
simple solution
sin2<p = Q/ft c ,
(3.6.2)
where
(3.6.3)
Thus when Q < Q c the director makes a constant angle cp with the field.
This has been accurately verified by Leslie, Luckhurst and Smith(103) from
a study of the electron spin resonance spectrum of a paramagnetic probe
dissolved in a nematic liquid crystal which is spun in a magnetic field. The
spacing between the hyperfine lines was found to be in quantitative accord
with (3.6.2).
When Q > Q c (3.6.1) does not have a steady state solution. Assuming
director inertia to be negligible
O.

For Q > Q c ,

^j)

( ^ ^ | Q

) ^ - g ,

(3.6.4)

(3.6.5)

where

assuming the initial condition that cp = 0 when t = O.(103) Hence the


director rotates with a mean angular velocity
co = (Q 2 -Q c 2 )i

(3.6.6)

An alternative method of performing the experiment is to have a


stationary sample in a rotating magnetic field. In point of fact Tsvetkov<104)
(and later Gasparoux and Prost(105)) used this method and measured the
torque exerted by the fluid on the cylinder as a function of Q. With
increasing Q, the torque at first increases linearly, reaches a maximum and
then starts to decrease. From (3.3.5), the stress tensor

3.6 Flow properties

147

The total torque on a cylinder of length L and radius R is therefore

= VX1 Q c sin 2(p

from (3.6.4), where V = nR2L is the volume of the cylinder. When Q < Q,c,
T = -Vkxa.

(3.6.7)

The torque increases linearly with the angular velocity and offers a direct
method of determining Ar When Q = Qc,
T = \VX*H\

(3.6.8)

When Q > Qc (3.6.5) yields the relation


sin 2(p =

2 tan w
l+tan>
2{QC/Q + (1 1 + {Qc/Q + (1 - Q2/Q2)^ tan [(Q2 - Q2)i/ -10]}2'

The mean value of the torque is therefore

When Q ^> Qc,


f=-^Q2.

(3.6.10)

Above the critical angular velocity, the torque decreases with increasing
Q. The predictions are generally in agreement with observations(105) (fig.
3.6.1). However, the shape of the experimental curve at higher angular
velocities appears to be rather sensitive to the nature of the solid surface in
contact with the liquid crystal, showing that a complete theory has to take
into account boundary effects and the production and migration of
disclination walls at the surface.(106)
Another method of determining A1 is by studying the damped torsional
oscillations of the nematic suspended in a static magnetic field.(107) For
i 1, where D is the torsion constant of the wire and V the

148

3. Continuum theory of the nematic state

1.0

0.5

Fig. 3.6.1. Variation of the torque as a function of the angular velocity of the
rotating magnetic field. Open circles: theoretical values; filled squares: experimental values for PAA, T= 112C, / / = 2 9 0 0 G (Tsvetkov(104)); crosses: experimental values for MBBA, T=24C, 77 = 2230 G; triangles: experimental
values for the same compound, T = 22 C, H = 2230 G, but using a solid teflon
cylinder immersed in the fluid to measure the torque. (After Gasparoux and
Prost.(105))

volume of the sample, the decay in the angular amplitude of the oscillations
can be expressed as

where K = Xx V/D, neglecting wall effects, backflow, etc. The method is


particularly useful for materials with large kv

3.6.3 Poiseuille flow


We shall next consider the rigorous theory of Poiseuille flow, i.e., the
steady laminar flow, caused by a pressure gradient, of an incompressible
fluid through a tube of circular cross-section.(108) We shall suppose that the
tube is of infinite length so that end effects can be ignored. Let us choose
a cylindrical polar coordinate system rcpz with z along the axis of the tube.
In the steady state the only component of the velocity gradient is vZt r =
dv/dr. It is natural to expect that the director is everywhere in the rz plane

3.6 Flow properties

149

making an angle 6(r) with z. Thus we seek for the components of the
director and velocity fields the solutions
nr = sin 0(r),
vr = 0,

n9 = 0,
v(p = 0,

nz = cos 9{r).
vz = v{r).

If a magnetic field with components (Hr9 0, Hz) is applied, the external body
force
Gz = x^Hr sin 6 + Hz cos 6) Hz,
where / a is the diamagnetic anisotropy per unit volume. Substituting in
(3.3.1) and (3.3.2) we have in the steady state
2/d0

k
HzsinO) = 0,

ar b\

(3.6.11)

where
(3.6.13)
(3.6.14)

a = -dp/dz,
and ft is a constant. Equations (3.6.11) and (3.6.12) are also applicable to
flow through the annular space between two coaxial cylinders. For flow
through a capillary we put ft = 0 to avoid the singularity in (3.6.12) at
r = 0. The two equations can then be used to obtain the orientation and
velocity profiles. At high flow rates, the contributions of the elastic terms
tend to become small and
Xx + X2 cos 2 0 ^ 0
in the absence of a magnetic field. The director orientation then approaches
an asymptotic value given by
V ^

(3.6.15)

or, assuming Parodi's relation (3.3.15),


(3.6.16)
In ordinary nematics //2 and //3 are both negative and the flow alignment
angle 0O is usually small. This equilibrium orientation of the director is

150

3. Continuum theory of the nematic state

> 4.0
i

(c)

3.0

a,
2.0
0.001

1.0
0.1
AQInR (cm2 s1)
Fig. 3.6.2. Apparent viscosity rj for Poiseuille flow of PAA at 122 C (homeotropic
wall orientation) plotted against the ratio of the flow rate to the radius of the tube.
Open circles are experimental data of Fischer and Fredrickson,(110) (a) values
obtained from table I, (b) values obtained from table II with //1 = 0, (c) values
obtained from table II with ju = 0.038. (After Tseng, Silver and Finlayson.(111))
0.01

Table I
^ = 0.043 (g cm- 1 s-1)
H2 = -0.069
jus = -0.002
ju, = 0.068
AiB = 0.047
/z6 = -0.023
^ - ^ = ^ = -0.067
H - ^ = ; = 0.0705

Table II
ju1 = 0or -0.038 (g cm"1 s"1)
H% = -0.068
//3 = 0.000
/i4 = 0.068
Ai5 = 0.048
//6 = - 0 . 0 2 0
/ o -#i, = A, = - 0 . 0 6 8

attained in practice in relatively thick samples at high flow rates so that the
aligning effect of the walls has negligible influence.
The amount of fluid flowing per second

Jo

v(r)rdr

and the apparent viscosity


n =
By scaling the radius and the time as r' = hr and f = kt respectively, with
k = h2, it is easily shown(109) that in the absence of a magnetic field Q/R is
a unique function of aR3. Consequently n plotted versus Q/R should be a
universal curve for all tube radii and flow rates. This has been confirmed

3.6 Flow properties

0.001

151

0.01

AQInR (cm2 s ')


Fig. 3.6.3. Apparent viscosity rj for Poiseuille flow of PAA versus AQ/nR computed
for a tube of radius R = 55.5 jum. The values of/ a // 2 and 0(R), the orientation at
the wall, are respectively (a) 0, - T T / 2 , (6) 1.23, -n/2,

(c) 0, - T T / 4 , (</) 24.2, - T T / 2 ,

(e) oo, - 7 i / 2 or 0,0. (After reference 112.)

experimentally(110) (fig. 3.6.2). The apparent viscosity increases slightly at


lower pressure gradients.
Equations (3.6.11) and (3.6.12) with H = 0 have been solved numerically
by Tseng, Silver and Finlayson(111) assuming the boundary conditions
v(R) = 0, 0(R) = - 7z/2
and
6(0) = 0.
The computed apparent viscosity versus shear rate is shown in fig. 3.6.2.
The agreement with the experimental data can be seen to be quite good.
Calculations have also been made of the effect of an axial magnetic
field.(112) The apparent viscosity decreases appreciably in the presence of
the field but an experimental study of this effect has not yet been reported.
Curves for rj, the orientation and velocity profiles as functions of shear rate
and magnetic field are presented in figs. 3.6.3, 3.6.4 and 3.6.5.
It is seen from fig. 3.6.3 that the apparent viscosity is very sensitive to

152

3. Continuum theory of the nematic state


0

-0.2
-0.4
-0.6
-0.8

(a)

-1.0
-1.2
-1.4
-a/2

0
-0.2

-0.4
-0.6
-0.8
-1.0
-1.2
-1.4
-ir/2
0.4

0.5

0.6

0.7
r/R

0.8

0.9

1.0

Fig. 3.6.4. Orientation profile for Poiseuille flow for different shear rates. Wall
alignment homeotropic. The values of z a // 2 are 24.2 for (a) and 1.23 for (b). The
values of AQ/nR in (a) are (i) 0.003045, (ii) 0.03354, (iii) 0.1345, and in (b) they are
(i) 0.001245, (ii) 0.004167, (iii) 0.032438, and (iv) 0.1372. (After reference 112.)
9(R), the director orientation at the boundary. Thus imperfect alignment
or weak anchoring can be a serious source of experimental error in the
determination of rjam.
3.6.4 Shear flow
Consider now the steady laminar flow of a nematic fluid between two
parallel plates. If the flow is along x and the velocity gradient along y the
components of the velocity and the director are
v

x =

nx = cos 0(y),

ny = sin 0(y\

nz = 0,

3.6 Flow properties

153

l.O
0.9
0.8
0.7

g 0.5
0.4
0.3.
0.2
0.1

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

y/R

Fig. 3.6.5. Velocity profiles for Poiseuille flow; the curves (a) and (b) are for
different values oiAQ/nR but for the samefield,while for curves (b) and (c) 4 Q/nR
is nearly the same but the fields are different. The values of / a H2 and 4Q/nR
are respectively (a) 1.23, 0.001245, (b) 1.23, 0.004167, (c) 24.2, 0.003045 and
(d) xaH2 = oo or 4Q/nR = oo (truly parabolic). (After reference 112.)
where 6 is the angle made by the director with x. Proceeding as before, the
differential equations for the velocity and the director orientation in the
presence of a magnetic field (Hx9 Hy, 0) turn out to be(7)
-n+Tn\*7.) +(ay +

^1 + A2cos2g

W)

+ 2xa(Hx cos 6 + Hy sin 0) (Hy cos 0-Hx

sin 0) = 0,

(3.6.17)
(3.6.18)

where f[6) and g(9) are defined by (3.6.13) and (3.6.14) respectively and
a = dp/dx is the pressure gradient and c the constant shear stress applied
to the fluid. If both plates are stationary a ==
| 0, and taking y = 0 half-way
between the plates, c = 0. On the other hand, if there is no pressure

154

3. Continuum theory of the nematic state

gradient, and the flow is caused by one of the plates moving at a uniform
velocity in its own plane, a = 0 and c + 0. Equations (3.6.17) and (3.6.18)
can be solved to yield the apparent viscosity, velocity and orientation
profiles under different boundary conditions.(113)
If the plate separation is large enough, boundary effects and elastic terms
can justifiably be neglected in (3.6.17). For large shear rates and zero
magnetic field the director orientation approaches the value 0O, defined by
(3.6.15). But if a magnetic field of moderate strength is applied, the
orientation profile is modified slightly. Gahwiller(114) has studied this
behaviour by measuring the change in birefringence. He used capillaries
(5 cm long) of rectangular cross-section (4 mm x 0.3 mm) and measured
the rate offlowdue to a pressure gradient. If H is along the flow direction
and shear rates are large, we obtain from (3.6.17) and (3.6.18)
(/I1 + /l 2 cos2fl)^ = z a // 2 sin2fl.

(3.6.19)

Gahwiller assumed that the velocity profile may be approximated by the


usual parabolic dependence
v(y) = vo[l-(4y2/d%

(3.6.20)

where v0 is the velocity half-way between the plates and d the plate
separation (though, as emphasized by O'Neill(115), this assumption may
not be strictly valid in practical situations). From (3.6.19) and (3.6.20)
tan0*8^oj;//a//2</2.

(3.6.21)

The phase difference between the two perpendicularly polarized components when light is incident normal to the plates is then
fd/2

\
-d/2

where

[n(0)-no]dy,

1 /n2 = (sin2 6/nl) + (cos2 #/2),

where n0 and ne are the ordinary and extraordinary refractive indices. If the
magnetic field is applied along the velocity gradient,
2 2

(3.6.22)

and S can similarly be calculated. In the absence of a magnetic field,

Thus the three measurements So, <5y and d yield

JU3//I2, JJLJXB.

an(

3.6 Flow properties

155

1.5

1.0

0.7

IT

s 0.4

0.2

0.1

20

30

40
Temperature (C)

50

60

Fig. 3.6.6. The viscosity coefficients rj19 rj2 and rjs of MBBA as functions of
temperature. The temperature scale is linear in T~x. (After Gahwiller.(114))
At high magnetic fields the experiment reduces in effect to Miesowicz's
method except that Gahwiller extended it to arbitrary orientations of the
magnetic field. If 9 is the angle between the director and the flow direction
and cp that between the projection of the director on the yz plane and the
velocity gradient, then, neglecting secondary flow (see 3.6.5), one may
write approximately
rj(9, (p) = jux sin2 9 cos2 9 cos2 cp 1// 2 sin2 9 cos2 cp
+ IJUS COS2 6 +1// 4 4-1//5 sin2 6 cos2 cp +1// 6 cos2 6
= rjx cos2 6 + (rj2 + jux cos2 9) sin2 8 cos2 cp
By choosing 9 and (p appropriately, one can determine rj19 rj2, rjs and juv
Using these two sets of data, Gahwiller was able to determine all five
independent viscosity coefficients as well as / a . Some of his results are
presented in figs. 3.6.6 and 3.6.7.

3. Continuum theory of the nematic state

156

0.1

0.06 -

0.03

80
100
Temperature (C)

60

120

Fig. 3.6.7. The viscosity coefficients rjx, rj2 and //3 of /?-n-hexyloxybenzylidene-//aminobenzonitrile (HBAB) as functions of temperature. The temperature scale is
linear in T~\ (After Gahwiller. (114))
+3 -

80
100
60
Temperature (C)
Fig. 3.6.8. The ratio // 3 /// 2 versus temperature for MBBA (crosses), HBAB (open
circles) and 1:1:1 molar mixture of HBAB, /?-n-butoxybenzylidene-//-aminobenzonitrile and ^-n-octanoyloxybenzylidene-^'-aminobenzonitrile (filled circles).
(After Gahwiller.(114))
40

3.6 Flow properties

157

Gahwiller discovered that nematics that undergo a transformation to


the smectic phase at lower temperatures exhibit an unusual type of
instability as the temperature approaches the transition point. The limiting
value of the flow alignment angle 0O defined as
(see (3.6.16)) decreases rapidly and becomes zero at a certain temperature,
below which the steady laminar flow breaks up into many irregular
domains. He interpreted this as a reversal in the sign of ju3 below the critical
temperature (fig. 3.6.8). Under these circumstances there is no equilibrium
value of 90, and in the absence of an orienting effect due to the walls or a
strong external field the flow becomes unstable/ 116 ' 117) The effect is
associated with the formation of cybotactic (smectic-like) clusters in the
nematic phase in the vicinity of the transition to the smectic phase.
Consequently, such materials have two distinct flow regimes which have
been the subject of several detailed investigations. (155)
When I^/AJ > 1, (3.6.15) is evidently not valid. Steady state solutions
now exist in which the director is rotated by many turns on going from one
plate to the other.(118) However, such a cycloidal configuration may be
expected to be unstable.
3.6.5 Transverse pressure and secondary flow
We now examine in greater detail the case of oblique orientation of the
director in the Miesowicz experiment. Let the flow be along y (in a plane
Poiseuille geometry) and the velocity gradient along z, but the director is
oriented by a strong magnetic field in the xy plane at an angle <f> with respect
to x. (The velocity is assumed to be small enough not to give rise to any
distortion in the director orientation.) Under such circumstances theory
predicts that there should develop a transverse pressure gradient and
secondary flow along x.{119)
Let the sample thickness (measured along z) be d, and its lateral width
(along x) be /. Strictly speaking, the finite lateral width / of the sample and
the appropriate boundary conditions have to be taken into account in
treating this problem exactly, but since / is small we shall make the
simplifying assumption that there is no secondary flow vx but that a
pressure gradient p x builds up. We take n = (cos <j>, sin <j>, 0) throughout the
sample, and thus ignore director gradients.
In the steady state, the x component of the momentum transport
equation (3.3.1) yields
giVy,z=P,*z
(3.6.23)

3. Continuum theory of the nematic state

158

90

Fig. 3.6.9. (a) Ratio of transverse to longitudinal pressure gradient versus director
orientation <j> in plane Poiseuille flow of nematic MBBA. Sample thickness
d = 200 jum. Length of cell L = 4 cm and lateral width / = 4 cm. (b) Deflection
y/ of flow lines with respect to the y axis versus <f> in wide cells (L/l ~ ^ ) . Full line
represents theoretical variation. (From Pieranski and Guyon.(120))

and the y component yields


(3.6.24)

where
l

Oil ~ Vz) S m ^ C 0 S ^

Imposing the boundary conditions vy{\d) = 0, we have from (3.6.24)


vy=pjx*-d*)/2g2
and from (3.6.23) the transverse pressure gradient

The transverse pressure difference was directly demonstrated by Pieranski


and Guyon(120) by measuring the liquid level difference in tubes connected
to holes facing each other across the width / of the cell. Its dependence on
(j) predicted by (3.6.25) was verified (fig. 3.6.9(a)). It vanishes for <p = 0 and
n/2. Further, it was confirmed that it changes sign when the flow is
reversed or when the field is rotated so that </> becomes <f>.
For wide cells (1 |> L, the length of the cell measured along y) it was
found by observing the motion of dust particles that the flow lines are not
along y (in other words, the assumption that vx = 0 is no longer valid). In

3.7 Reflexion of shear waves

159

Transmitted wave (damped)


Nematic

Fig. 3.7.1. Experimental arrangement for studying the reflexion of ultrasonic shear
waves at a solid-nematic interface.
the central region of the cell the flow lines were deflected by an angle y/(<f>)
with respect to y, which again showed the expected dependence on cp (fig.
3.6.9(6)).
Though this experiment deals with a particularly simple situation, the
basic principle that it illustrates is relevant to the discussions that follow on
hydrodynamic instability (3.11).

3.7 Reflexion of shear waves


The viscosity coefficients may also be determined by studying the reflexion
of ultrasonic shear waves at a solid-nematic interface. The technique was
developed by Martinoty and Candau. (121) A thin film of a nematic liquid
crystal is taken on the surface of a fused quartz rod with obliquely cut ends
(fig. 3.7.1). A quartz crystal bonded to one of the ends generates a
transverse wave. At the solid-nematic interface there is a transmitted wave,
which is rapidly attenuated, and a reflected wave which is received at the
other end by a second quartz crystal. The reflexion coefficient, obtained by
measuring the amplitudes of reflexion with and without the nematic
sample, directly yields the effective coefficient of viscosity.
To explain the principle of the method we shall consider the simpler case
of normal incidence. Let a shear wave be incident along z with its vibration
direction along x, and let the nematic director be anchored firmly at the
interface (z = 0) along y. The only non-vanishing component of the
velocity of the fluid is vx = v, and the velocity gradient is along z. From
(3.3.5) the stress across the interface is (neglecting director gradients)
(3.7.1)

160

3. Continuum theory of the nematic state

and from (3.3.1)


(3.7.2)

Now the velocities associated with the incident, reflected and transmitted
waves may be written as
vt = Ai exp ( F s z) exp (icot),
vr = Ar exp (F s z) exp (icot),
vt = At exp ( - Tn z) exp (icot),
respectively, where F may be complex and the subscripts s and n stand for
solid and nematic. Therefore, in the nematic medium
tzx = -ZnAt

exp (-Tnz)

exp (icot),

where Z n = |// 4 F n is the mechanical impedance, and


pvt = X(8 2iV8z 2).
We thus have
or
and

F n = (1 -\-i)(pco/ju^

(3.7.3)

Z n = |(1 + i) (pcoju^y.

(3.7.4)

The complex reflexion coefficient is given by


y _ z

A
r

_ 111 _ ^ s

Ax

^n

Z s + Z n'

where Z s is the mechanical impedance of the solid which may be assumed


to be a real quantity. If
r= |r|exp(-i0),
n

^s

l-Hexp(ifl)
l + |r|exp(i0)
i

. .12

, OII

/3-

(3.7.5)

Since according to (3.7.4) the real and imaginary parts of Z n are of equal
magnitude, it follows that

-^t

(3.7.6)

3.8 Dynamics of the Freedericksz effect

161

Thus a measurement of \r\ at once gives Z n , which, in turn, yields //4.


Similarly, for the director oriented along x at the interface

and for the director along z at the interface

The theory can be generalized to the case of oblique incidence.(122)


Martinoty and Candau found that the viscosity coefficients determined by
the ultrasonic technique compare fairly well with those derived from
capillary flow.
3.8 Dynamics of the Freedericksz effect
3.8.1 Twist deformation
We shall now extend the theory of the Freedericksz effect to study the
dynamical behaviour when the magnetic field is switched on or off
suddenly.(123) The analysis is particularly simple for a twist deformation
(fig. 3.4.1 (b)) because the torsion exerted on the director does not result in
a translational motion of the centres of gravity of the molecules. Neglecting
director inertia in (3.3.2) we obtain the following equation of motion for
this geometry:
22

d28
8 ? *a

.
Sm

C0S

W
dt ~

'

where kx is the twist viscosity defined by (3.3.14). If 8 is small,

where
The most general solution satisfying the boundary conditions 8 = 0 at
z

Neglecting higher harmonics and remembering that 8 has a maximum


value #max at z = 0, we take

8 = 8m^(t) cos (nz/d).

162

3. Continuum theory of the nematic state

Equation (3.8.1) then gives

or
- 1] exp {- (2
Thus 0max(O attains the value 0(oo) with a time constant T(H) given by
H2).

(3.8.2)

If the field is now reduced to a value less than i/ c , the decay rate is still given
by the same expression only with a negative sign.
If the field is switched off from H > Hc to zero,

and the decay rate

T-\0) = (kJX1)(n*/d*\
The twist viscosity can be determined from a measurement of T. (123)
Typically, T(0) for a film of 25 jum is about 10"1 s. This gives an idea of
the order of magnitude of the relaxation time for most nematic liquid
crystal devices.
3.8.2 Homeotropic to planar transition: backflow and kickback effects
The other two geometries used in the Freedericksz experiment are more
interesting as they result in a new effect, namely, hydrodynamic flow
induced by orientational deformation. This is the inverse of the more
familiar property of flow alignment that has been discussed at length in
previous sections.
Let us consider the homeotropic to planar transition (fig. 3.4.1 (c)). For
this geometry, n = (sin 0,0, cos 0), 0 = 0(z), v = vx(z), and vz(z) = 0. Setting
klx = fc33 = k, cos 0 1 and sin 0 0, we get from (3.3.2)
^

^
dz

^
dt

= 0
dz

where

and r = (l2-

(3.8.3)

3.8 Dynamics of the Freedericksz effect

163

Neglecting inertial effects, (3.3.1) reduces in the present case to


where, using (3.3.3),
,dvT

d6

hx = K/^ + ^ - ^ - ^ + ^ Q p
neglecting squares and higher powers of 9. Therefore

where a = f(//4 + //5 fii2) a n d b = ju2.


The boundary conditions for n and v are 9 = 0 and vx = 0 at z = d/2.
The solutions are of the form
9 = 0o[cos qz-cos (qd/2)] exp (t/x\
vx = i;0[sin qz-(2z/d)

sin (qd/2)] exp (t/z).

(3.8.5)
(3.8.6)

Substituting in (3.8.3) and (3.8.4) we obtain the following relations,

i=

(3-8-7)

and
/ H\2

(HJ

Ay/2 (iff/A) tan y/

le

p-tan, '

,.

o o

(18 8)

where A = k'b/a and y/ = qd/2. From numerical calculations, Pieranski,


Brochard and Guyon(123) have shown that the relaxation rate can be
expressed in the form
T~\H) = (xJ^)(H2-Hl\

(3.8.9)

where the apparent viscosity A* is now strongly dependent on H.


The translational velocity (3.8.6) has two components in the simplest
case: one linear in z and the other oscillatory, the wavelength of the latter
diminishing with increasing value of the final field H. The transient velocity
profile is illustrated schematically infig.3.8.1. The effect of this backflow is
to relax the constraints, i.e., to reduce the apparent viscosity.
In the planar to homeotropic transition (fig. 3.4.1 (a)) backflow effects
are not usually so pronounced near the threshold. In this geometry, the
torque exerted by the director on an elementary volume of the fluid is

164

3. Continuum theory of the nematic state

Fig. 3.8.1. Velocity profile in the homeotropic to planar transition (seefig.3.4.1 (c)).
The velocity has two components, one linear in z and the other oscillatory, the
wavelength A of the latter diminishing with increasing value of H.

Fig. 3.8.2. Torques acting on an elementary volume of the fluid when the molecules
are rotating with angular velocity D: (a) homeotropic to planar transition,
F a = |(2 1 A 2)Q; (b) planar to homeotropic transition, Tb = ^(<X1 + X2)Q. As a
rule, |FJ > | r j .
and

where Q is the angular velocity (fig. 3.8.2). In many nematic liquids, Ax and
X2 are of opposite signs and of comparable magnitude (see, e.g., legend of

3.8 Dynamics of the Freedericksz effect

165

30

25

20

r(0)
r(h)

15
10

0 a
10

15

20

Fig. 3.8.3. Theoretical and experimental normalized relaxation rates as functions of


h2 = H2/H\. Open circles and triangles are respectively the experimental and
theoretical values for the homeotropic to planar transition. Closed circles are the
experimental values for the planar to homeotropic transition. The line represents
the variation expected from (3.8.9). The departure from this line for the
homeotropic to planar transition is a consequence of backflow. Material: MBBA.
(After Pieranski et al.a23))

fig. 3.6.2) so that F is small. In other words, A is small and the solutions of
(3.8.7) and (3.8.8) reduce to y/ = n/2 and

as for the twist geometry.


The marked difference in the relaxation rates for the two geometries at
higher fields confirms the existence of backflow as predicted by the Leslie
equations (fig. 3.8.3). Backflow has also been studied by direct observation
of the motion of disclination walls separating two regions of opposite tilt
in a film which is subject to a magnetic field.(124)
It is instructive to consider the geometry of fig. 3.4.1 (a) and to examine
the relaxation when the field is switched off from H 5> Hc to zero. Clark
and Leslie(125) have analysed this problem theoretically and have presented
the equations in a form that reduces considerably the computational effort
required for making detailed predictions in any practical situation. Using

166

3. Continuum theory of the nematic state

Fig. 3.8.4. (a) Velocity profiles at (i) 1.09 s, (ii) 1.744 s, (iii) 2.397 s, (iv) 2.943 s, (v)
4.469 s, (vi) 4.905 s and (vii) 8.61 s after the magnetic field is switched off from
H ^> Hc to zero in the Freedericksz experiment of fig. 3.4.1 (a). The values are
computed for a sample of MBBA of thickness 200 jum using the equations of Clark
and Leslie.(125) (b) Director orientation profiles at (i) 0 s, (ii) 0.109 s, (iii) 0.654 s, (iv)
3.161 s, (v) 6.648 s after switch off. (U. D. Kini, unpublished.)

their equations the velocity and orientation profiles at different instants of


time after the field is switched off have been calculated and are shown in fig.
3.8.4(a) and (b). The velocity is zero at the boundaries (z = d/2) and at
the middle of the sample (z = 0), and the profile is antisymmetrical about
the midplane. At intermediate regions (z ~ d/4) the fluid first moves one
way and then with passage of time reverses direction before coming to rest.
The director orientation in the middle of the sample initially tips over to an
angle greater than n/2 and then gradually relaxes to 6 = 0. These have
been termed as backflow and kickback effects respectively.
The physical explanation of these effects will be clear from fig. 3.8.5. (126)
The initial director orientation profile is shown on the left (fig. 3.8.5(a)).

3.9 Light scattering


d/2

167

-d/2
(a)

(b)

Fig. 3.8.5. Interpretation of backflow and kickback effects (see text(126)).


The elastic torque will be greatest around z = d/4 where the curvature is
greatest, and this is balanced by the magnetic torque. When the field is
switched off the unbalanced elastic torque causes a clockwise rotation of
the director in this region (as indicated in fig. 3.8.5(b)). Because of the
coupling between director rotation and hydrodynamic motion, fluid flow is
induced as shown by the long arrows. The fluid motion, in turn, results in
a counterclockwise torque on the director in the middle region (z ~ 0).
This overcomes the elastic torque, which is weak in this region, and gives
rise to a counterclockwise rotation of the director. The director tilts over to
an angle greater than n/2; when it relaxes back to 6 = 0, it produces a small
amount of fluid motion in the reverse direction. Finally at large values of t
the system settles down to the undistorted equilibrium configuration.
This transient effect manifests itself in a direct way in the behaviour of a
twisted nematic cell (see 3.4.2). When the external field (assumed to be
sufficiently strong) is switched off, the light transmission shows an ' optical
bounce effect', i.e., it does not decrease monotonically but rises again to a
peak before decaying to its 'off' value. Calculations have confirmed that
the peak in transmission corresponds approximately to a perpendicular
alignment of the director in the central portion of the cell. This is caused by
fluid motion, which also gives rise to a reverse-twist. (124127)
3.9 Light scattering
3.9.1 Orientational fluctuations
One of the most striking features of the nematic liquid crystal is its
turbidity. From systematic observations of the Rayleigh scattering from
oriented samples, Chatelain (128) showed that the scattered intensity is
strongly depolarized and exhibits a marked angular variation. An early
model put forward to explain this phenomenon assumed the medium to be
composed of swarms, about 1 jum in diameter, of aligned molecules, the

168

3. Continuum theory of the nematic state

orientations of the different swarms being uncorrelated. However, it is now


well established(129"31) that the light scattering can be interpreted rigorously
in terms of the small amplitude orientational fluctuations as described by
the continuum theory. The intensity of this scattering turns out to be very
much larger than that arising from the density fluctuations in the fluid, so
much so that the latter contribution can be neglected altogether.
Let co0, k0 and i be respectively the angular frequency, wavevector and
unit polarization vector of the incident beam and co^ kx and f the
corresponding quantities for the scattered beam. The scattering process is
associated with an angular frequency change
co = co0 co1
and a wavevector change
q = ko-k1.
We define the differential scattering cross-section per unit volume of the
scatterer, per unit solid angle (Q), per unit angular frequency change as
to

dfidcu
where X = 2nc/co0 is the vacuum wavelength and

is the mean square fluctuation of the dielectric constant at a given point r


and time t.
For the uniaxial nematic medium, the dielectric tensor at any point r can
be written as (see 2.3.1)
eif = s + e&(ntnf-lX

(3.9.2)

where e = (e||+2e 1 )/3 is the mean dielectric constant (at optical frequencies), a = || e_L is the dielectric anisotropy assumed to be <^,
n{: = n i and nf = nf. An electric vector polarized along i induces a
displacement Df = eifEt along f. The director at r
n(r) = n0 + dn,
where, to a first approximation, <5n n0 = 0, since we assume the fluctuations
to be of small amplitude. Thus, neglecting density fluctuations (i.e.,
assuming s and a to be constant) we have from (3.9.2)

3.9 Light scattering

169

(c)

Fig. 3.9.1. (a) The two uncoupled modes Snx and Sn2; (b) components of the
deformation in the SnY mode, bend and splay; (c) components of the deformation
in the 3n2 mode, bend and twist.
where i0 = n0 i and / 0 = n0 f. Therefore
.(r,/)>.

(3.9.3)

The fluctuations can be analysed into Fourier components. For a given


Fourier component of wavevector q, we may conveniently resolve Sn into
two components 5nx and Sn2, the former in the q, n0 plane and the latter
perpendicular to it (fig. 3.9.1). Let us therefore introduce two unit vectors

ex = e2 x n0,
where qL is the component of q perpendicular to n0. Defining ia = e a i and
/a = e a f ( a = l , 2 ) ,

170

3. Continuum theory of the nematic state

Hence
<Se2} = el [ (ijo +/aio)\6nJL09 0) SnJLr, t))].

(3.9.4)

The scattering cross-section is then


dV/dQdco = n*lr*el (i a / 0 + /0/o)24(q,<*>),

(3.9.5)

where
Too

aq,w

j ^

na

q,

and
<Sa(q, /) = (snJLr, t) exp (iq r) dr.

(3.9.7)

Here (iaf0 + i0fa)2 is a purely geometric factor while / a is a correlation


function which describes the power spectrum of the fluctuations.
For a = 1, the director vibrates in the n o ,q plane and the mode is a
superposition of bend and splay. For a = 2, the vibration is normal to the
n0, q plane and the mode is a superposition of bend and twist. The two
modes are shown schematically in fig. 3.9.1.

3.9.2 Intensity and angular dependence of the scattering


It is of interest to consider first the intensity of the scattered light integrated
over time (or frequency). The differential scattering cross-section is then
= 7l2A~*el YJ 0'a/o + *o/a)2<<^a(q)>

(3.9.8)

where

Writing <5n(r, t) = <5nexp [i(q r cot)] and substituting in (3.3.6) we obtain


the free energy of elastic deformation of the system

where

3.9 Light scattering

171

Optic axis Q

Fig. 3.9.2. A typical experimental configuration used by Chatelain in his lightscattering studies: k and i are the wavevector and unit polarization vector of the
incident light, k' and f the corresponding quantities for the scattered light, the
suffixes e and o denote the extraordinary and ordinary polarizations with respect
to the optic axis of the nematic medium and qx is the wavevector change on
scattering.
From the equipartition theorem (which is certainly valid in the present
problem)

where kB is the Boltzmann constant. To get an idea of the order of


magnitude of the scattering cross-section let us suppose that k1 k22
kS3 = k\ then
dcr ^ (nea\2 ^B T

dQ~VlV ~W'

(3.9.10)

On the other hand, the cross-section due to density fluctuation is given by


the well known formula

(3.9.11)
where /? is the isothermal compressibility. Taking p(ds/dp) ~ ea,

da'/da ~ 0kq2.
Typically P ~ 10"11 cm2 dyn"1, k ~ 10"6 dyn, q ~ 104 cm"1, so that

da'/da ~ 10"8.

3. Continuum theory of the nematic state

172

150 r-

100

I
50

1
0

^
50

100

150

200

cot 2 (<p/2)

Fig. 3.9.3. Angular dependence of the intensity of scattering for PAA. Circles give
the experimental values of Chatelain and the line represents the theoretical
variation. (After de Gennes.(129))

Thus the director fluctuations make the predominant contribution to the


light scattering, as was first pointed out by de Gennes.(129)
A comparison of the polarization factors in (3.9.5) and (3.9.11) at once
explains why the light scattering from a nematic liquid crystal is strongly
depolarized. The angular dependence of the light scattering is also
accounted for in a straightforward manner. Let us, for example, consider
one of the geometries used by Chatelain (fig. 3.9.2). The incident and
scattered beams are both normal to z; the incident beam is linearly
polarized in the plane of scattering while the scattered beam is polarized
along z, the optic axis of the medium. If the angle of scattering is <p9

: k0 sin (p/2), qz = 0, ix = cos (<p/2),


i , = / 1 = / 2 = 0,
fz=L

Therefore
dcr

dQ
(3.9.12)

3.9 Light scattering

173

Fig. 3.9.3 compares this relation with the experimental data of Chatelain
and as can be seen the agreement is good. In principle a measurement of the
intensity of the scattering and its angular variation offers a method of
determining the elastic constants.
3.9.3 Eigenmodes and the frequency spectrum of the scattered light
For a given mode of vibration of angular frequency co and wavevector q,
we can write
<5n(r, t) = Sn exp [i(q r cot)],
Substituting in the basic equations of motion (3.3.1) and (3.3.2) we obtain
icopvk = &k-qk(qj^/q2\

(3.9.13)
0.

(3.9.14)

We ignore inertial effects as well as terms quadratic in Sn; J%. = i


where t'ik is the viscous part of the stress tensor defined by (3.3.5), and F
is the elastic energy density. As the fluid is supposed to be incompressible
v
k, k = 0 o r Qk vk = 0. lfv19 v2, vz are the components of v in the frame e19 e2, z,
we have
dzz = kz vz,

dxl = iq v1 = - \qz vz,

where q2 = q\ + q\. Substituting in (3.9.13) and (3.9.14)


iQ(q) vz + [Llx co + ^(q)] Sn, = 0,1
iCa(q)i;2 + [UlG> +fca(q)](J/i2= 0 j
Sn, = 0,1
v2\pco-iP2(q)]-icoQ2(q)Sn2 = 0j
where

K) q\+(K - K

' '

'

174

3. Continuum theory of the nematic state


\+vm q\ ql)/q\

Is =

For compatibility of (3.9.15) and (3.9.16) we have the vanishing of the


determinant, and therefore
[pa - iPa(q)] [xcol, + kJLq)] - C a (q) Qa(q) = 0
or

) = 0,
6

(3.9.17)

which has two roots. Typically, k ~ 10~ dyn, p ~ 1 g cm" ,77 ~ // ^ O.li5,
2

C(q)

Consequently /?A:a(q) is negligible compared to ^ ^(q) and Ca(q)Qa(q), and


therefore the two roots of (3.9.17) are

(3919)

The subscripts s and f denote 'slow' and 'fast', for


cos - ikq2/rj,
and

cot -

ir/q2/p,

cojcot ~ pk/rj2 ^ 1.

We observe that both modes are purely dissipative (non-propagating); cos


involves the elastic constants while cof does not. The slow mode therefore
represents the relaxation of the orientational motion of the director, while
the fast mode may be looked upon as the diffusion of a vorticity but one in
which there is no torque on the molecules. The light scattering, being
dependent primarily on the orientational fluctuations, is controlled entirely
by the slow mode, as we shall proceed to show.
We have resolved the director fluctuations into dnx and Sn2 which from
the symmetry of the problem can be seen to be uncoupled. If Gx and G2 are

3.9 Light scattering

175

the forces responsible for these tilts in the director, then in a first order
theory dn1=%1G19 ^ 2 = / 2 G 2 , where / 1 ? / 2 are susceptibilities; more
generally
We have seen that due to the thermal agitation, there are spontaneous
fluctuations in n whose mean square value is defined by
7a(q, co) = (SnJt - q, 0) Sna(q, co)}.
According to the fluctuation-dissipation theorem,(132) the relation between
h

and

X, is

/a(q, co) = ^ J l m (xJLq, co)),

(3.9.20)

where Im stands for the imaginary part.


We can derive an expression for 7a(q, co) from the equation of motion for
the director in the presence of an external field:

Accordingly (3.9.15) becomes


iC^v^dnJico^

+ k^q)] = G19

iC2(q)v2-\-Sn2[icol1-\-k2(q)] = G2.
Along with (3.9.16), this can be simplified to obtain

AaVM

' '

[pco-iPM][ttiO)+K(<i)]-c(q)QMv

or
7a(q, co) = y^

YT~T\

2\

>

(3.9.21)

where
um = -ico s o c

a n d uta = -icof(X.

(3.9.22)

Thus 7a is a superposition of two Lorentzians. However, as wfa > wsa, we


may ignore the second term in the square brackets of (3.9.21) and rewrite
the power spectrum as

3. Continuum theory of the nematic state

176

200

150

100

50

10

20

30

40

Fig. 3.9.4. Angular dependence of the width of the Lorentzian spectral density for
mode 2 in PAA at 125 C. Open circles denote experimental values in the [ke,k^]
configuration and open squares the values in the [ko, k^] case. The curve is obtained
by a least squares fit with the theory. (After the Orsay Liquid Crystals Group.(131))
The light scattering is therefore determined entirely by the slow mode. The
integrated intensity

in agreement with (3.9.10). Also from (3.9.18), (3.9.19) and (3.9.22)


(3.9.24)

3.10 Electrohydrodynamics

177
(3.9.25)

The integrated intensity gives the elastic constants ku while the half width
yields uS(X. It is therefore possible to measure the viscosity coefficients from
an analysis of the scattered light using appropriate geometries.
As an example, we present in fig. 3.9.4 a convenient geometry for
isolating mode 2. The director is aligned parallel to the walls of the glass
plates. The incident beam is polarized parallel to the director and the
scattered beam perpendicular to it. If ne and no are the extraordinary and
ordinary indices of the liquid crystal, and cp the scattering angle,
ke = Innjk,

k'o =

Innjk,

qz = k'o sin q>9 q = ke- k'o cos q>.


For small angles of scattering,

and from (3.9.25)

which enables Xx to be determined. By going to higher angles, it is possible


to obtain juJ2fil and rjv/2fi\. A typical curve for the angular dependence of
the width of the Lorentzian spectral density is shown in fig. 3.9.4. The
experiments are rather difficult because of the large amount of stray
radiation, especially in the forward direction, arising from defects in the
alignment of the specimen. However, by very careful techniques using a
laser light beat spectrometer, and employing various geometries LegerQuercy(131) has been able to determine four viscosity coefficients of PAA,
//2-//5. The values of//1? rj2 and rj3 calculated from these coefficients are in
reasonably good agreement with those determined by Miesowicz (see table
3.6.1).
3.10 Electrohydrodynamics
3.10.1 The experimental situation
From dielectric studies in the radio frequency region,(133) it has long been
known that PAA is negatively anisotropic, i.e., a = el{ e < 0. However,
in a number of early investigations on the effect of an external DC electric
field it was noticed that the PAA molecules align themselves parallel to the
field, rather than perpendicular to it as would be expected of a material of
negative dielectric anisotropy. This observation gave rise to some contro-

178

3. Continuum theory of the nematic state


Transparent conductive coating

Glass
LIQUID

CR VYSTAL

^> *

Spacers

Glass

Light beam

Fig. 3.10.1. Experimental arrangement for observing Williams domains.


versy in the 1930s but it has since been confirmed by the systematic
experiments of Carr, (134) who proved that the anomalous alignment is due
to the anisotropy of the electrical conductivity of the liquid crystal. His
studies showed that there is a critical frequency of the applied field below
which the alignment of PAA is anomalous, and that this frequency
increases with the conductivity of the material, ranging from 2 to 100 kHz
in the samples examined by him.
Macroscopic motion of the fluid induced by electric fields was observed many years ago by Freedericksz and Zolina, (18) Tsvetkov and
Mikhailov,(135) Bjornstahl(136) and Naggiar.(137) Tsvetkov also noted that
the flow decreases with increasing frequency of the applied field, and
probably recognized the fact that the phenomenon may be connected in
some way with the electrical conductivity. More recently, Williams (138)
discovered that a thin layer of a nematic material of negative dielectric
anisotropy between conducting glass plates forms regular striations when
a DC voltage of sufficient magnitude is applied. At higher voltages, the
regular pattern gives way to turbulence accompanied by intense scattering
of light, which has come to be known as dynamic scattering and has found
practical applications in display devices. (139) Similar observations have
been reported by other authors. (140)
The experimental arrangement for observing the Williams domains is
shown in fig. 3.10.1. The nematic film of negative dielectric anisotropy
(e.g., PAA or MBBA) is aligned with the director parallel to the glass
Fig. 3.10.2. Electrohydrodynamic alignment patterns in nematic liquid crystals, (a)
Williams domains in a 38 jum thick sample of /?-azoxyanisole. 7.8 V, 100 Hz.
(Penz.(141)) (b) Chevron pattern of oscillating domains in MBBA. Sample thickness
~ 100 //m. Distance between bright lines ~ 5 jum. 260 V, 120 Hz. (Orsay Liquid
Crystals Group.(142))

3.10 Electrohydrodynamics

Fig. 3.10.2. For legend see facing page.

179

180

3. Continuum theory of the nematic state

Fig. 3.10.3. (a) Flow and (b) orientation patterns of Williams domains. The
periodic orientation pattern and the consequent refractive index variation has a
focussing action for light polarized in the plane of the paper. This gives rise to the
bright domain lines as indicated by the stars above and below the sample. (After
Penz.(143))
surfaces which are coated with a transparent conducting material. When a
DC or low frequency AC field is applied between the transparent
electrodes, there appears above a threshold voltage a regular set of parallel
striations perpendicular to the initial unperturbed orientation of the
director (fig. 3A0.2(a); see, however, 3.13.2). Dust particles are seen to
undergo periodic motion in the field of view proving that the domains are
due to hydrodynamic motion. The distortion of the director orientation
caused by this motion results in a focussing action for light polarized
parallel to the director(143) (fig. 3.10.3). This is responsible for the
appearance of a set of bright lines with a spacing approximately equal to
the film thickness when the microscope is focussed at the top surface. The
lines are shifted by about half the spacing when the focal plane is moved
down to the bottom surface. The pattern disappears when the light is
polarized perpendicular to the director.
The threshold voltage is usually a few volts and is practically independent
of the sample thickness. It is, however, strongly dependent on the
frequency(142) (fig. 3.10.4). There is a cut-off frequency coc above which the
domains do not appear, the value of coc increasing with the conductivity of

181

3.10 Electrohydrodynamics

400 -

2
"o

100 -

100 (Hz)
400
Frequency (Hz)

600

Fig. 3.10.4. Threshold voltage of the AC instabilities versus frequency for MBBA.
Sample thickness 100 jum. Region I: conducting regime (stationary Williams
domains); region II: dielectric regime ('chevrons')- Full line is the theoretical
curve. The cut-off frequency / c = 89 Hz. (After the Orsay Liquid Crystals
Group.(142))
the sample. Below coc, i.e., in the so-called conduction regime, the regular
Williams pattern becomes unstable at about twice the threshold voltage
and the medium goes over to the dynamic scattering mode. Above coc, in
the dielectric regime, another type of domain pattern is observed. Parallel
striations, again perpendicular to the initial orientation of the director but
with a much shorter spacing (a few microns), are formed in the midplane
of the sample. When the field is increased very slightly above the threshold,
the striations bend and move to form a chevron pattern (fig. 3.10.2(b)). In
this regime, the threshold is determined by a critical field strength rather
than a critical voltage. Both the threshold field strength and the spatial

3. Continuum theory of the nematic state

182
150

Oscillating domains

100
B

vi

-a

Region of stability

50

_#
50

100

200

150

Frequency (Hz)

Fig. 3.10.5. Threshold voltage versus frequency for MBBA. Sample thickness
50/urn. Open circles: sinusoidal excitation; triangles: square wave excitation.
(After the Orsay Liquid Crystals Group. (142))

Approaching - optic axis 11 field


25
No domains
20

"o

o " o

Optic axis_L field


Domains
/
10
--_

No domains

/C7H15

H 15 C 7

0
1

10

100
1000
Frequency (Hz)

10000

Fig. 3.10.6. Frequency dependence of the threshold voltage in /?,//-di-n-heptoxyazobenzene, a nematic of positive dielectric anisotropy. (After Gruler and
Meier.(144))

3.10 Electrohydrodynamics

183

periodicity of the pattern are frequency dependent - the former increasing


with frequency (as OP) and the latter diminishing with it. The relaxation
time of the oscillating chevron pattern is a few milliseconds while that of
the stationary Williams pattern is typically about 0.1 s for a thickness of
25 jum. The oscillating domain regime is therefore sometimes called the fast
turn-off mode. The chevron pattern also gives way to turbulence at about
twice the threshold field. An applied magnetic field parallel to the initial
orientation of the director increases the threshold voltage in the conduction
regime, but has no effect in the dielectric regime except to increase the
spacing between the striations. The threshold curve has a pronounced
sigmoid shape with square wave excitation (fig. 3.10.5) indicating that at
high electric fields there is a quenching of the conductive instability even
when co < coc.

DC and very low frequency AC voltages produce electrohydrodynamic


instabilities in the isotropic phase also (T> Tm), the threshold being
comparable to that in the nematic phase. It has been suggested that this is
due to charge injection at the electrodes. A frequency of about 10 Hz is
usually enough to suppress this effect showing that charge injection is not
the primary mechanism for the AC field instabilities in the nematic phase.
If the initial alignment of the director is homeotropic, domains as well as
turbulence can be produced. The threshold voltage for the domains is
somewhat higher than in the case of parallel alignment, but the patterns
persist even when the voltage is reduced to a lower value.
We have so far discussed only materials of negative dielectric anisotropy.
Electrohydrodynamic distortions are observed even in weakly positive
materials,(144) but only when the initial orientation of the director is
perpendicular to the applied field. Striations appear above a threshold
voltage but vanish at still higher voltages and there is no dynamic
scattering. The frequency dependence of the threshold voltage is shown in
fig. 3.10.6.

3.10.2 Helfrich's theory


The basic mechanism for the electric-field-induced instabilities is now quite
well understood. The current carriers in the nematic phase are ions whose
mobility is greater along the preferred axis of the molecules than
perpendicular to it. The ratio of the conductivities o^oL is usually about |.
Because of this anisotropy, space charge can be formed by ion segregation
in the liquid crystal itself, as was first pointed out by Carr.(134) The manner
in which the space charge can build up due to a bend fluctuation is shown

184

3. Continuum theory of the nematic state

Fig. 3.10.7. Charge segregation in an appliedfieldEz caused by a bend fluctuation


in a nematic of positive conductivity anisotropy. The resulting transversefieldis Ex.
schematically in fig. 3.10.7. The applied field acts on the charges to give rise
to material flow in alternating directions which, in turn, exerts a torque on
the molecules. This is reinforced by the dielectric torque due to the
transverse field created by the space charge distribution. Under appropriate
conditions, these torques may offset the normal elastic and dielectric
torques and the system may become unstable. The resulting cellular flow
pattern and director orientations are sketched in fig. 3.10.3. Even a
conductivity of the order 10"9 Q" 1 cm"1 is enough to produce this type of
fluid motion. Indeed unless very special precautions are taken, the impurity
conductivity is usually greater than this value.
With a DC field, there may be injection of charge carriers at the
solid-liquid interface but its role in the electrohydrodynamics of the
nematic phase is not yet fully understood. However, as remarked earlier, a
frequency of about 10 Hz is enough to suppress charge injection. We shall
therefore neglect it in the present discussion.
We shall now outline the theory of electrohydrodynamic instabilities
proposed by Helfrich(145) and extended by Dubois-Violette, de Gennes and
Parodi(146) and Smith et fl/.(147) We consider a nematic film of thickness d
lying in the xy plane and subjected to an electric field Ez along z. Let the
initial unperturbed orientation of the director be along JC, and let there also
be a stabilizing magnetic field along the same direction. We consider a bend

3.10 Electrohydrodynamics

185

fluctuation in which the director is in the xz plane and makes an angle cp


with x. We ignore wall effects and assume that the deflexion cp is a function
of x only.
Due to the anisotropy of conductivity, space charges will develop as
indicated in fig. 3.10.7 till the transverse electric field stops the transverse
current. The local transverse field in the steady state is easily seen to be
al{ cos2 cp + <7 sin2 cp
where cra = o^ oL > 0, a^ and aL being the principal electrical conductivities along and perpendicular to the local director axis. Also, as soon
as the electric field is applied there will be a transverse field EEX due to the
dielectric anisotropy. Since the transverse displacement is zero,
Eex = [ea cos (p sin ^/(e|( cos2 cp + eL sin2 cp)] Ez.
The space charge per unit area produced on any plane normal to the x axis
is
On cos 2 (p + e sin 2 (p) (Ez -

EEX)/4n,

the positive and negative signs standing for the two sides of the plane under
consideration. The applied field Ez acting on these charges causes a flow
along the z axis; the resulting shear stress is evidently
4 = (e^cos2 (p + eLsm2 cp)(Ex- EEX) EJ An.

(3.10.1)

From the Ericksen-Leslie theory, we know that the viscous torque is


rvisc = n x g ' ,

(3.10.2)

where g' is given by (3.3.13). This is the frictional torque exerted by the
molecules on the hydrodynamic motion. Clearly in the present geometry
(3.10.2) reduces to

^^t)

(3.10.3)

where Ax = H2H3 and X2 = fii~/j,6. Also, in the present geometry, the


viscous stress tensor t'n given by (3.3.5) can be simplified to

where, making use of Parodi's relation (3.3.15),

186

3. Continuum theory of the nematic state

Setting Ty = r elastfy + r dlelfI/ + r m a g f y -r v l 8 C f y = -Aq>9 where A is a force


constant, the condition for instability may be written as(148)
A = -Ty/<p = 0.

(3.10.5)

The elastic, dielectric and magnetic torques can be evaluated from the
functional derivative
where
JV

and Fis given by (3.3.6). We then have


r

elast, y = ~ (^33 C O s 2 <P + Kl

sir

(fc n k33) sin cp cos q>(d(p/dx)2,


r

diei, y = - (47r)- 1 K cos <p sin (p(E2z - El)


- ea(sin2 q> - cos2 ^) ^ Ez],

F mag

= 0fa cos ^ sin cp) H2.

Since we are interested in the threshold conditions, we retain only first


order terms in (p, i.e.,

\\

\\

Also assuming a spatially periodic fluctuation of the form


cp = (p0coskxx,
we have

JEZ)

3.10.3 DC excitation
In the DC case, we may set <p = 0 and the analysis becomes simple. Using
(3.10.1), (3.10.3) and (3.10.4),

3.10 Electrohydrodynamics

187

where rj^ = rj'[2X1/(X1 X2)]2. Applying (3.10.5), we obtain the threshold


field
(3.10.7)
where
El = -^(xaH2

+ k3Sk2x)

(3.10.8)

and C is a dimensionless quantity, called the Helfrich parameter, given by


(3.10.9)
Typically 2 is a small number; e.g., for MBBA it is about 3. For DC (and
low frequency AC), kx n/d. (A numerical solution (149) of the twodimensional problem with appropriate boundary conditions has confirmed
that at the threshold kx is indeed n/d.) Thus when H = 0, we have a voltage
threshold independent of film thickness given by
\%

(3.10.10)

where
888

"

'

We have treated the distortion as a pure bend but this is not exactly true.
Since the thickness of the sample is of the same order as the periodicity of
the distortion, the orientation of the director will vary in the z direction
also. There will therefore be a non-negligible splay component. To allow
for this, it has been suggested(147) that the bend elastic constant k33 should
be replaced by
where the wavevector in the z direction kz ~ n/d. With this correction, the
DC threshold given by (3.10.10) is in good agreement with the experimental
value. For MBBA it is about 8 V.
3.10.4 Square wave excitation
In the AC case, the time dependence of cp cannot be neglected. Using
(3.10.3), (3.10.5) and (3.10.6)
F
l

visc,y

A
H7l

\/p
\\
b

L\

188

3. Continuum theory of the nematic state

In addition we have to take into consideration the conservation of charge.


The charge balance equation reads
q + dJx/dx = 0,

(3.10.12)

where q is the excess charge per unit volume and Jx the electric current given
by
Jx = <7|| Ex + o^Ez(p

retaining only the first order terms in cp and Ex. We suppose that diffusion
currents make a negligible contribution. From the relation

where Dx is the x component of the displacement, we obtain

where y/ = d<p/Qx is the local curvature. Therefore


+ oaE^

= 0,

(3.10.14)

where T is the dielectric relaxation time given by


and
11

\ fin

Neglecting the inertial term, (3.3.1) may be written as


',<,,+/< = 0,

(3.10.15)

where/, is the body force per unit volume, which in this case is equal to qEz.
We are interested only in the z component of this equation. Since the
director orientation q> is assumed to be a function of x only, txz vanishes,
and (3.10.15) reduces to

or, from (3.10.4),

At the threshold, we have the condition


o.

(3.10.17)

3.10 Electrohydrodynamics

189

From (3.10.3) and (3.10.16)


ar v l s c , y

Krj, I

ix1

From (3.10.6) and (3.10.13),

- E2
Therefore, from (3.10.17) we obtain the following equation for the
curvature
(3.10.18)
y + yy/ + ?-Ez = 0,
where

(3.10.19)
T is the decay time for curvature, and rj is an effective viscosity coefficient
given by

Equations (3.10.14) and (3.10.18) are two coupled equations which cannot
be solved analytically for any general value oiEz(i) since y itself depends on
Ez and t. However for a square wave

J + ,
A )

\-E,

0<t<n/co
n/(D<t<2n/oj.

In this case y remains constant in any half-period and the solutions are
simpler and enable a physical interpretation of many of the observed
phenomena. If the solutions are taken in the form (147)
y/ = Cw exp (At/z\

q = Cq exp (At/x),

the general solutions become


/T)9

where

(3.10.20)

190

3. Continuum theory of the nematic state

These solutions are valid as long as E is constant. We observe that as


q=

-<?Jiy/Ex/(\+A1),

a change of sign of E implies that q or y/ changes sign but not both, i.e., if
(q, y/) is a solution for a half-period, it will be (q, y/) or ( q, y/) for the
other half-period. Subject to these conditions we obtain two sets of
solutions given by (3.10.22) and (3.10.23) below:

-expH;

(3.10.22)

1exp
Eo^x

exp

-l-Ao
sx

-exp

A2t

2vx

exp

-\-Ax

where v = co/2n is the frequency of the voltage. Here q changes sign with
E, but y/ does not, which is the situation in the conduction regime.

+ expl ^
\

2 - ^ 1

2VT

(3.10.23)

sx Ax
1+exp
2vx
1

+ exp

^T. n

A2-Ax
2vx

exp

In this case y/ changes sign with E, but q does not. This corresponds to the
dielectric regime in which the director oscillates.
In the above equations, s is a real number which determines whether the
system is stable or not, and which takes the value 0 at the threshold. Setting
s = 0 and defining

3JO Electrohydrodynamics

191

\
\
\
\
\
\

j1

(a)
t

I
^

Fig. 3.10.8. Time dependence of the charge q and the curvature y/ over one period
of the square wave excitation, (a) Conduction regime (coz <4 1, T> z). The charges
oscillate but the domains are stationary, (b) Dielectric regime (coz ^> 1, T < r). The
charges are stationary and the domains oscillate. (After Smith et /.(147))
it can be easily shown that in the conduction regime
<3l0 24)

and in the dielectric regime


+i

= U-l

sinh

TTA

'2COT'

(3.10.25)

These equations show clearly that the problem falls naturally into two
distinct parts. For T > z (3.10.25) has no solution and consequently there
is no dielectric regime, while for T < z (3.10.24) has no solution and there
is no conduction regime. For T = T, neither equation has a solution (except
when co = 0).
The theoretical variation of q and y/ over a full period of the exciting
wave is illustrated in fig. 3.10.8 for two values of co, one at low frequency
in the conduction regime (coz < 1, T > z) and the other at high frequency

192

3. Continuum theory of the nematic state

in the dielectric regime (coz > \,T 4, z). In the former case one obtains
essentially stationary domains and oscillating charges and in the latter case
vice versa. The transition from the conduction to the dielectric regime
occurs at a critical frequency coc such that coc z ~ 1.
Certain other interesting conclusions can be drawn from the above
equations. For example, even at frequencies less than coc, there can be
quenching of the conductive instability at high fields. This is because
T = (AEl + AQ)" 1 decreases with increasing Ez9 and when it becomes equal
to T there can be restabilization. At higher fields, T < z and the system
goes over to the dielectric regime. This gives a physical insight into the
origin of the sigmoid shape of the experimental threshold curve. Indeed,
theoretically the threshold field as a function of frequency for conductive
instability may be expected to form a closed loop.
Another result that follows from the theory is that the conduction
regime can be suppressed altogether by using very thin samples. We have
seen that at low frequencies, kx ~ n/d. Now a decrease in sample thickness
increases Ao, where from (3.10.19)

This, in turn, reduces the curvature relaxation T and when T < z the
conductivity instability is eliminated and only a dielectric instability is
possible. A similar quenching can be achieved by applying a stabilizing
magnetic field.
Greubel and Wolff(150) and Vistin(151) observed that for thin and pure
samples the spatial periodicity of the pattern in DC excitation decreases
with increasing voltage above the threshold. Smith et <z/.(147) have suggested
that this variable grating mode(150) is due to non-linear (saturation) effects
at higher voltages in samples that are so thin and pure that the conductivity
instability has been quenched.
3.10.5 Sinusoidal excitation
The behaviour is qualitatively similar when the exciting field is sinusoidal.
Putting Ez = EM cos cot, the coupled equations for the charge and the
curvature become(146)
q + --\-aIly/EMcoscot

= 0,
coscot = 0,

(3.10.26)
(3.10.27)

3.10 Electrohydrodynamics

193

where

We have seen in the previous section that in the conduction regime, q


changes sign with the field but y/ does not. Let us make the simplifying
assumption that y/ is sensibly constant over a period. Since co ~ 1/T and
T > T, we may replace cos2 cot by its average value f. Also expanding q{t) as

we have

Therefore from (3.10.27),

Integration of (3.10.26) yields


#

i^ M 2

(cos ^ +C0T s m

0J

so that

Using (3.10.17) the threshold field is given by

= (E\t)} = \El = p^JT{-

(3.10.29)

Thus the threshold field increases with co. Though the analysis is not
strictly valid when coT ~ 1, calculations show that (3.10.29) holds good
quite well over the range 0 to coc where
COC = (C2-\)*/T,

(3.10.30)

which represents a critical 'cut-off' frequency beyond which the system


goes over to the dielectric regime (fig. 3.10.4). To discuss the effect of the
magnetic field, we write (3.10.8) as
F2

194

3. Continuum theory of the nematic state

where is the magnetic coherence length (see 3.4.1). For low magnetic
fields, > d, we get a voltage threshold independent of the thickness of the
sample:
__v_w_fv

(3.10.31)

where
47TII

j
V

33 '

For higher magnetic fields,

HV>

O.10.32)

where // c = (n/d)(kzz/xd*- T n e threshold voltage therefore increases with


H in agreement with observations. For very high magnetic fields, H ^> Hc,
theory predicts a field threshold independent of thickness but proportional
to if.
When COT P \,q may be taken as constant over a period. Expanding y/(t)
as a Fourier series
00

y/(f) = ^ ( ^ cos co/ + \i/"n sin co/)


n=0

and replacing y/(t) cos cot by its average value \y/[,

Equation (3.10.27) can now be integrated but solutions can only be


obtained by numerical techniques.(146) Calculations show that in the
dielectric regime (a) the threshold field Eth is independent of the wavevector
kx, which results in a,fieldthreshold (as distinct from a voltage threshold as
in the conduction regime), (b) E*h varies linearly as co, (c) for a given //, k\
varies linearly as co, and (d) for a given co, %a H2 + k3S k\ is a constant. These
predictions have been confirmed experimentally.
Other aspects of the problem, e.g., the dependence of the instabilities on
the magnitude and sign of the dielectric anisotropy and on the conductivity,
the behaviour in the vicinity of the threshold on either side, the effect of
external stabilizing fields, etc., have been discussed extensively in a number

3.11 Hydrodynamic instabilities

195

of articles. (1525) Some results on the influence of flexoelectricity will be


described in 3.13.2.
3.11 Hydrodynamic instabilities
3.11.1 Homogeneous instability in shear flow
The anisotropic properties of nematics give rise to certain novel instability
mechanisms that are not encountered in the classical problem of hydrodynamic instability in ordinary liquids. Theoretical work on electrohydrodynamic instability stimulated systematic studies on two other
types of convective processes, viz, thermal and hydrodynamic instabilities,
and it was soon established that the basic mechanisms involved in all three
cases are closely similar.(156~9) In this section we shall examine the problem
of hydrodynamic instabilities in nematics.
We shall discuss first what is called homogeneous instability (HI), which
is the hydrodynamic analogue of the Freedericksz transition/158'159)
Consider simple shear flow between two plane parallel plates (caused by
moving the plates parallel to each other at a constant relative velocity v).
Let the director be initially along x, the flow along y, and the velocity
gradient along z (fig. 3.11.1). At low values of vy the sample retains its
planar configuration, but above a critical velocity vc - or a critical shear
rate Sc = vc/d, where d is the gap width - the director tilts towards the yz
plane, Sc varying inversely as the square of the sample thickness. In the
presence of a strong stabilizing magnetic field Hx along x, Sc increases
as//;.
The physical interpretation of this distortion is as follows. (We shall
assume throughout this discussion that the material is an ordinary nematic
like MBBA with jus < 0.) Let us suppose that there is a fluctuation 0 > 0 in
the xy plane that rotates the director from its initial orientation (1,0,0).
The flow now exerts a viscous torque, which is given by Fz = /z2 SO (fig.
3.11.1). Since ju2 < 0, Tz > 0 and gives rise to a small twist <j> > 0. Now, a
deflection ^ results in a viscous torque Ty = //3 S(p, and since //3 < 0, the
sign of Ty is such as to increase 6 further. Thus the viscous torques have a
destabilizing effect, and above a critical shear rate they overcome the
elastic and magnetic torques and the system becomes unstable.
There is, in fact, another torque that comes into effect. As we have seen
in 3.6.5, a deflection of <j> in the director orientation gives rise to a
secondary flow vx. The secondary velocity gradient vXt z causes a torque Ty
that makes an additional destabilizing contribution (except close to the
boundaries).

196

3. Continuum theory of the nematic state

rv<o

(a)

Fig. 3.11.1. The mechanism for homogeneous instability: the flow is along y and
the velocity gradient along z. (a) An angular fluctuation 6 = nz > 0 results in a
viscous torque Yz > 0 such that a small twist (/> = ny > 0 is produced, (b) a
deformation <j> > 0 results in a torque Yy < 0 such that the initial deformation 0 is
enhanced.
The rigorous theory of HI threshold has been developed by Manneville
and Dubois-Violette(160) and by Leslie(161), but for the present purpose
it is enough to discuss the approximate treatment given by Pieranski
and Guyon(158) neglecting secondary flow. Taking n = [l,^(z),0(z)] 5
v = (0, Sz, 0), and balancing the viscous torque against the elastic and
magnetic torques, one obtains the coupled equations
(3.11.1)
(3.11.2)
Putting 6 = 61 cos (nz/d) and <j> = <t>x cos (nz/d), we get from the condition
for the compatibility of (3.11.1) and (3.11.2) the threshold shear rate for
H=0:
5 =

(3.11.3)

and in the presence of the magnetic field


(3.11.4)
Hcl and H are the Freedericksz threshold fields for the splay and twist
geometries. It is seen that Scccd~2, and for large Hx, ScccH2x; these
predictions are in agreement with observations. For a sample thickness d
of 200 //m, the zero field critical threshold velocity v c= 11.5 jum s"1 for
MBBA.(159)

3.11 Hydrodynamic instabilities

197

100
80
60
40
20

2
Frequency (Hz)

Fig. 3.11.2. Experimental roll instability threshold curves for MBBA as functions
of the applied voltage and frequency of shear for different values of the effective
plate velocity veU; d = 240 //m, Hx = 3200 G. (From Pieranski and Guyon.(159))
It is possible to define a dimensionless quantity called the Ericksen
number
Er = (Sd*/4) (fi2 jus / * u k22)i
which from (3.11.3) is seen to be n2/4 at the threshold.(162) The more
rigorous calculation, taking into account secondary flow, yields a critical
value of Er = 2.309 for MBBA at the threshold.(160) It follows from the
exact solutions that ET is not a universal number but varies from one
substance to another.

3.11.2 Roll instability in shear flow


We have seen that under steady (DC) shear, the HI threshold increases
with increasing strength of the stabilizing magnetic field Hx. However, if
the field becomes large enough, the instability does not appear as a
homogeneous distortion but as a regular series of bright and dark lines
parallel to y, the direction of primary flow.(159) This is referred to as the roll
instability (RI). The spacing between the lines is of the order of the sample
thickness and the pattern is reminiscent of the Williams domains (see fig.
3.10.2 (a)). The lines arise from a cellular motion of the fluid in the xz plane
superposed on the primary flow along y. A laser beam produces a
diffraction pattern, and as in the case of the Williams domains, the

3. Continuum theory of the nematic state

198
i
100

^ ^ ^

t?eff = 0.148 c m s " 1

v'3
Y

>

| Stable

TR
.'

0.6

v3

Y
0.8

Frequency (Hz)

Fig. 3.11.3. Roll instability threshold curve for vett = 0.148 cm s 1 from fig. 3.11.2
showing the regimes Y,Z and TR (see text). (From Pieranski and Guyon.(159))
diffraction pattern vanishes when the light is polarized perpendicular to the
long axes of the molecules.
Another way of producing this pattern is to apply an AC shear. In this
case the RI appears even in the absence of a magnetic field (except, of
course, at very low frequencies, in which case it reduces, in effect, to DC
shear and HI is regained). Two distinct regimes can be identified by optical
observations. These are referred to as the Fand Z regimes. In the Y regime
the y component of the director ny changes sign at each half-period of the
AC shear but the z component nz does not, and vice versa in the Z regime.
Nematic MBBA was used in the experiment with a stabilizing magnetic
field Hx as well as a high frequency stabilizing electric field Ez (the latter is
stabilizing because a < 0 for MBBA). Sinusoidal shears of up to about
2 Hz were applied. The threshold curves for different values of the effective
velocity vett of the upper plate ( = total plate displacement per half-period)
as functions of the frequency of shear and of the applied voltage Vz are
shown in fig. 3.11.2. To the left of each curve is the domain of existence of
the RI. At larger frequencies rolls do not develop.

3.11 Hydrodynamic instabilities

199

Fig. 3.11.4. Afluctuation^ = ny that is spatially periodic in x results in a secondary


velocity vz that is also spatially periodic. This can have a destabilizing effect and
generate convective rolls.
To understand these curves let us examine the curve for a given vm (fig.
3.11.3). Let us suppose that the frequency of shear is kept constant at
0.8 Hz and the voltage Vz is increased. At V = 0, the instability is of the Y
type; at Vx the instability disappears; at V2 > Vx the instability reappears
but is now of the Z type; and finally at V3 > V2 it disappears again. The
spatial periodicity of the rolls remains of the order of the sample thickness,
though it is somewhat larger in the branch just above the cusp. At a
frequency slightly less than that at the cusp, 0.6 Hz say, the instability is of
the Y type at V = 0. At much larger voltages it is of the Z type, but for
V = V's the rolls disappear. The changeover from Y to Z takes place in a
transition regime, TR, that is defined by extrapolating the two branches to
lower frequencies. In TR the rolls do not extend over large regions but
seem to correspond to an interchange between the Y and Z types along a
given roll. The laser diffraction pattern now has satellite spots in the y
direction and also shows that the spacing between the rolls has doubled.
The principal mechanism for the onset of RI is shown schematically in
fig. 3.11.4. A spatially periodic <j>fluctuation(or ny fluctuation) of the form
cos qx x results in a secondary velocity vz that is also periodic in x. This
effect has been referred to as hydrodynamic focussing. Under appropriate
conditions, vz can have a destabilizing effect and generate convective rolls.

200

3. Continuum theory of the nematic state

c = 45.70

o)Ty/2n

Fig. 3.11.5. Calculated roll instability curves fitted to the results of fig. 3.11.3 by
adjusting the values at zero frequency, at the cusp and at a point P to define the
frequency scale. (From Pieranski and Guyon.(159))

For a given <p there is the torque ju3 S</> due to the primary velocity (as in HI)
as well as a torque due to the secondary velocity gradient vZt x and together
the total viscous torque Yy brings about a distortion 6. In turn, a 6
distortion gives rise to a torque Tz that increases <f>. (We have already seen
in the case of HI that Tz = ju2 S6, but this value is now modified slightly
by a torque due to vy z.)
The complete calculation of the threshold for RI involves a solution
of the form exp \{qx x + qz z) with boundary conditions for 6, </>, vx9 vy,
and ^.<159'160'163> We present only the salient results of the theory.
Assuming fcn = k22 = fc33 = k, and ignoring material derivatives in (3.3.1),

3.11 Hydrodynamic

instabilities

201

the behaviour under A C shear c a n be expressed in terms of two coupled


equations with ny( = <p) a n d nz( = 0) as the variables;
y

09
0,

(3.11.5)
(3.11.6)

where the relaxation rates are

yy and yz are effective viscosity coefficients, A and B are functions of the


viscosity coefficients and of the wave vectors qx and qz, and q2 = q2x + q\.
The value of xz is influenced by the external electric field whereas xy is not;
xz > xy gives the Y regime and xz < xy the Z regime.
It is seen at once that (3.11.5) and (3.11.6) bear a close similarity to
(3.10.14) and (3.10.18) for electrohydrodynamic instability; ny, nz and S
correspond to the charge, curvature, and electric field, and the Y and Z
regimes correspond to the conduction and dielectric regimes, respectively.
Let the shear rate S = S0coscot (co = 2n/T). At steady state ny and nz
are periodic functions of time. Suppose that xz > ry, xz > T, and that nz is
constant in time. Then from (3.11.6)
(3.11.7)
;o

Using (3.11.7) in (3.11.5) and integrating


Yly

n*j

AS0 fl,[(cos cot)/ty + co sin cot] x\


-

\J.Y

n n f i

Y.O)

Compatibility between (3.11.7) and (3.11.8) leads to


ABS2cxzxy/2(\+co2x2y)=l

(3.11.9)

which is the threshold condition for the Y regime. On the other hand, with
xy > TZ, Xy> T one gets in a similar fashion
ABS2cxyxz/2(l+co2x2)

= 1

(3.11.10)

as the threshold condition for the Z regime. Equations (3.11.9) and


(3.11.10) can be cast in a parametric form by defining x = coxy, y = (xy/xz)\
c = ABS2cx2y/2.Then
2

= c/(l+x2),

xpy2,

y>\

(3.11.11)

describes the Y regime, while


x 2 + y * - c y 2 = 0, x p l , y < \

(3.11.12)

202

3. Continuum theory of the nematic state

describes the Z regime. The theoretical diagram for threshold is shown in


fig. 3.11.5. These expressions do not, of course, describe the complex
behaviour in the transition region between the two regimes.
3.12 Thermal instability: stationary convection
Let us first recall briefly the classical Benard-Rayleigh problem of thermal
convection in an isotropic liquid.(164) When a horizontal layer of isotropic
liquid bounded between two plane parallel plates spaced d apart is heated
from below, a steady convective flow is observed when the temperature
difference between the plates exceeds a critical value ATc. The flow has a
stationary cellular character with a spatial periodicity of about Id. The
mechanism for the onset of convection may be looked upon as follows. A
fluctuation T' in temperature creates warmer and cooler regions, and due
to buoyancy effects the former tends to move upwards and the latter
downwards. When AT < ATC, the fluctuation dies out in time because of
viscous effects and heat loss due to conductivity. At the threshold the
energy loss is balanced exactly and beyond it instability develops.
Assuming a one-dimensional model in which T' and the velocity vz (normal
to the layer) vary as exp (iqyy) with qy n/d, the threshold is given by the
dimensionless Rayleigh number
Rc = ATcd*pag/Kri = 7i\

(3.12.1)

where p is the density, a the coefficient of thermal expansion, g the


acceleration due to gravity, K = K/pC the thermal diffusivity, K the
thermal conductivity, C the specific heat and n the viscosity coefficient.
Rigorous calculations show that for rigid conducting boundaries 7r4 of
(3.12.1) should be replaced by 1704. For a liquid having material
parameters comparable with the average values of MBBA (K ~ 10~3,
rj ~ 1 and a - 10"3 cgs) the modified form of (3.12.1) gives ATC 2 C for
d = 1 cm and ATcx2x
103 C for d = 1 mm.
The situation is altered profoundly in the case of a nematic because of its
anisotropic transport properties. Dubois-Violette(156) was the first to give
an approximate theoretical treatment of thermal convection in a planar
(homogeneously aligned) nematic and to show by consideration of torques
that such a system will be unstable against cellular flow when the film is
heated from below if Kn > 0, or when it is heated from above if Ka < 0,
where K^ = K^ K is the anisotropy of thermal conductivity (which is
positive for all known nematics(165)). Dubois-Violette also showed that the
critical temperature gradient fte( = ATc/d) should be much less than that

3.12 Thermal instability: stationary convection

203

(a)

T+AT

(b)
Fig. 3.12.1. (a) Thermal instability in a nematic heated from below. Initial
orientation of the director is horizontal (along y). A bend fluctuation causes 'heat
focussing' because of the anisotropy of thermal conductivity (K^ > K) and gives
rise to warmer ( + ) and colder ( ) regions. The warmer regions move up and the
colder regions down due to buoyancy effects and this, in turn, results in a velocity
vy (the fluid being assumed to be incompressible). The transverse velocity gradient
vz y induces a major destabilizing viscous torque, while the vertical gradient vy z
induces only a very weak stabilizing torque. The resulting torques, shown by the
curved arrows, are destabilizing. Here the long straight arrows represent the
translational velocities vz and vy9 the short straight arrows the heat fluxes, (b) The
same geometry as in (a) but with the top plate at a higher temperature. The system
is stable.

for an isotropic liquid. Experimental observations of convective rolls in a


homogeneously aligned film of MBBA heated from below were reported
by Guyon and Pieranski. (157166) As expected, the threshold value of the
thermal gradient fic was about 10~3 times that for an isotropic liquid of
comparable average physical properties. Optical observations confirmed
that the rolls are essentially as depicted in fig. 3.10.2 (a) for the Williams
domains. When the bottom plate was heated to a temperature greater than
the nematic-isotropic transition point the convection disappeared, showing that it is the anisotropy that is responsible for the low threshold.
Figures 3.12.1 and 3.12.2 illustrate the destabilizing mechanisms
involved. Because of the anisotropy of thermal conductivity, a thermal
fluctuation of the director along y creates warmer ( + ) and cooler ()

204

3. Continuum theory of the nematic state

\
T+AT
(a)

T+AT

Fig. 3.12.2. (a) The initial orientation of the director is along z. A splay fluctuation
gives rise to warmer and colder regions. In this case, vyz causes a dominant
stabilizing viscous torque while vz y causes only a very weak destabilizing torque.
The system is therefore stable against stationary convection, (b) The same geometry
as in (a) but with the top plate at a higher temperature. The system is unstable (see
legend of fig. 3.12.1
regions. This is referred to as heat focussing. Because of the buoyancy
effect, the warmer regions move up and the cooler regions move down
creating a velocity fluctuation vz9 which in turn gives rise to a torque Tx
that stabilizes or destabilizes the orientation depending on the sign of
Ka. We seek solutions of the form n = [0, l,0(y,i)]9 \ = [0,0,vz(y,t)],
T=-pz+T'(y,t),
and using (3.1.6)(3.1.9) obtain for a variation
Qxp(iqyy) of the fluctuation(152)
exp (iqyy) [vz0 + (vjzv)
f

- ocgT'o] - (jujp) y/ = 0,
= 0,
= 0,

(3.12.2)
(3.12.3)
(3.12.4)

where y/ = d6/dy is the director curvature and 019 vz0, T'o are amplitudes of
fluctuations. In (3.12.4) T = ^x/kzzq^ is the relaxation time for the
director in the absence of any coupling. Ignoring vz0 in (3.12.2) and

3.13 Flexoelectricity

205

eliminating vz0 in (3.12.3) and (3.12.4) one finds the condition for threshold
to be
where r a = (K^q\XJii^f1. In an isotropic liquid Ka = 0,r a = oo, and we
recover (3.12.1). For a nematic film of thickness d ~ 1 mm, r ~ 103 s,
i a ^ 1 s, and ATe reduces to 10~3 times the isotropic value. It is the large
ratio T /r a that leads to a very low threshold. A magnetic field Hx
(stabilizing) or Hz (destabilizing) can be used to decrease or increase T^, or,
equivalently, increase or decrease A Tc; the critical temperature difference
varies linearly as H\ (or # z 2 ). (166)
Pieranski et al.ae7) then showed that when a homeotropically aligned
film of MBBA is heated from above, the orientation becomes unstable
above a critical value of AT of the same order as that for planar
orientation. They also verified that the film is stable when heated from
below. A model of the destabilizing mechanism is no longer purely onedimensional involving only vz (fig. 3.12.2). A thermal fluctuation causes a
fluctuation vz as before, but because of the incompressibility of the fluid
this, in turn, causes a velocity fluctuation vy that contributes the major
destabilizing torque Fx~ju2vyz
for Ka > 0. Again with a stabilizing
magnetic field Hz there is a linear relationship between A Tc and H\. A field
Hy favours rolls with axes normal to y, but in the field-free case the rolls
degenerate into a square pattern that may be regarded as a linear
superposition of crossed convection rolls. When AT is increased well
beyond A Tc a complex hexagonal structure is found with a nematicisotropic interface if the temperature of the upper plate is large enough.
The studies outlined here are the most important ones that established
the fundamental of principles of thermal instability in nematics. A number
of theoretical and experimental investigations on these and other geometries have since been reported. (155) A particularly interesting study is
that of Lekkerkerker(168) who predicted that a homeotropic nematic heated
from below (which, it will be recalled, is stable against stationary
convection) should become unstable with respect to oscillatory convection.
The phenomenon was demonstrated experimentally by Guyon et al.a69t 170)

3.13 Flexoelectricity
3.13.1 Determination of the flexoelectric coefficients
If the molecule possesses shape polarity in addition to a permanent electric
dipole moment then the possibility exists that a splay or bend deformation

206

3. Continuum theory of the nematic state

(c)
id)
Fig. 3.13.1. Meyer's model of curvature electricity. The nematic medium composed
of polar molecules is non-polar in the undeformed state {{a) and (c)) but polar
under splay (b) or bend (d). (After Meyer.(171))

(b)

Fig. 3.13.2. Interpretation of the origin of flexoelectricity in an assembly of


quadrupoles: (a) in the undeformed state the symmetry is such that there is no bulk
polarization, (b) a splay deformation causes the positive charges to approach the
upper plane and to be pushed away from the lower one. This dissymmetry gives rise
to a dipole moment.
will polarize the material, and conversely that an electric field will induce
a deformation (fig. 3.13.1). This phenomenon, which is somewhat
analogous to the piezoelectric effect in solids and is therefore termed as the

3.13 Flexoelectricity

207

Fig. 3.13.3. A hybrid aligned cell for the determination of the anisotropy of the
flexoelectric coefficients. In this geometry, the director has a splay-bend distortion
which gives rise to a flexoelectric polarization Px. On applying an electric field Ey,
the director is twisted by an angle $ oc (e1 e3) which can be measured optically.
(Dozov, Martinot-Lagarde and Durand. (174) )

flexoelectric effect, was first proposed by Meyer.(171) Subsequently Prost


and Marcerou(172) showed that electric quadrupole moments can also
contribute to this effect (fig. 3.13.2). Since, as a rule, all molecules
have non-zero quadrupole moments, it follows that flexoelectricity is a
universal property of nematics. The observation of flexoelectricity in a
nematic composed of symmetric, non-polar molecules(173) has confirmed
the existence of a quadrupolar contribution, which turns out to be
comparable in magnitude to that from molecular dipoles.
In a first order theory, the polarization P should be proportional to the
distortion:
)] + e8[n-Vn],
(3.13.1)
where ex and e3 are the flexoelectric coefficients corresponding to splay and
bend respectively. In an external electric field E, the flexoelectric effect
leads to a free energy density F = P E .
A simple method of measuring (e1 e3) is as follows.(174) The nematic is
taken in a ' hybrid' aligned cell with the director oriented homogeneously
on one surface, and homeotropically on the other. The anchoring at the
boundaries is assumed to be firm, and in the absence of an external electric
field the director is confined to the xz plane (fig. 3.13.3). The director field
in such a cell has a splay-bend distortion which gives rise to a flexoelectric
polarization along x. If, now, a DC electric field E is applied along y, the
nematic acquires a twist about z due to the action of E on P. (The influence
of dielectric anisotropy can be neglected as long as the field strength is

3. Continuum theory of the nematic state

208

<
-30
-50

Electric field (V mm"1)

Fig. 3.13.4. ^(0) versus is for two twin domains in the sample. The linearity of the
slope confirms the validity of (3.13.2). (Dozov, Martinot-Lagarde and Durand. (174) )

small, since the dielectric free energy is proportional to E2.) The twist angle
is maximum close to the bottom plate (fig. 3.13.3) where the curvature is
maximum and is given by
Ed

(3.13.2)

where k is the elastic constant (in the one-constant approximation) and d


the sample thickness. The angle ^(0) can be determined by measuring the
rotation of the plane of polarization of linearly polarized light incident
along z (see 4.1.1). (Strictly speaking, the adiabatic approximation is not
valid close to the bottom plate of fig. 3.13.3, for in this region the director
orientation is homeotropic and the effective birefringence for light
propagation along z is small. However, the authors have shown that for
thick samples the resulting error is negligible.) A plot of ^(0) versus E is a
straight line (fig. 3.13.4), the slope of which yields (e1 es). The value was
found to be 1 x 10"4 dyn* for MBBA.
The sum of the flexoelectric coefficients (ex + e3) was first measured by
Prost and Pershan.(175) A homeotropically aligned sample is taken between
two glass plates and a periodic electrostatic potential is applied by means
of interdigitated electrodes coated on one of the plates, as shown in fig.
3.13.5. The flexoelectric effect being linearly proportional to the applied
voltage, the resulting distortion has a periodicity 2d, where d is the spacing
between neighbouring electrodes. On the other hand, the distortion due to

3.13 Flexoelectricity

209

(a)

(b)

Fig. 3.13.5. A periodic electrostatic potential applied by means of interdigitated


electrodes coated on one of the plates gives rise to (a) SLflexoelectricdistortion
having a periodicity Id, where d is the spacing between the electrodes and (b) a
dielectric distortion having a periodicity d. (Prost and Pershan.(175))

dielectric alignment, which is proportional to the square of the applied


voltage, has a periodicity d. By optical diffraction it was possible to
distinguish between the two types of distortion. Further, by using optical
heterodyne detection techniques it was shown that the diffracted intensity
due to flexoelectric distortion increases linearly with voltage, as expected.
By evaluating the distortion of the uniformly aligned nematic under the
action of the field gradient, the magnitude of (e1 + es) was estimated to be
2.5 x 10~4 dyn* for MBBA.
A method of determining (e1 + e2) both in magnitude and sign was
devised by Dozov et al.(176) The sample, homeotropically aligned with firm
anchoring at the boundaries, was placed in a quadrupolar field distribution
obtained with two pairs of electrodes. The electric field gradient causes a
tilt of the director at the centre of the sample and hence a tilt in the
conoscopic pattern seen between crossed polarizers. A reversal of the field
gradient reverses the tilt, and therefore the sign of (e1 + e^) could be
determined unambiguously. For MBBA (e1 + e3) was found to be
1 x 10~4 dyn*. Methods requiring weak anchoring boundary conditions
have also been used. (177178) It may be noted, however, that there are some
inherent difficulties, e.g., the large screening effect due to the ions, which
introduce considerable uncertainty in all these determinations.

210

3. Continuum theory of the nematic state

d = 35 /^m
/ c = 120 Hz

Normal
rolls

50
Frequency (Hz)

90

Fig. 3.13.6. The formation of oblique convective rolls in electrohydrodynamic


instability. Below the point M, which may be called a 'triple point', the transition
takes place directly to oblique rolls ( / = 10 Hz). Beyond M, normal rolls are
obtained first, followed by undulatory rolls which then change continuously into
oblique rolls ( / = 60 Hz). (Ribotta et al.a81))
3.13.2 Influence offlexoelectricity on electrohydrodynamic instability
In our discussion of electrohydrodynamic instability (3.10) we have
throughout assumed that the convection rolls (or Williams domains) are
normal to the initial orientation of the director n. However, in many
experiments oblique rolls, whose wavevector makes an acute angle a with
n0, have been observed at low frequencies. (179~81) Detailed studies by
Ribotta, Joets and Lei(181) have shown that these oblique rolls appear at the
threshold up to a critical frequency/ 0 in the conduction regime (fig. 3.13.6).
F o r / > / 0 normal rolls (a = 0) are obtained at the threshold, but on
increasing the field strength these normal rolls first become undulatory and
then oblique. Solutions corresponding to oblique rolls can be obtained by
extending the Orsay model to three dimensions and imposing boundary
conditions at the two surfaces, but the predicted values of a and f0 are

3.13 Flexoelectricity

211

about an order of magnitude smaller than the experimental values.(182)


Later investigations*183"7* have shown that flexoelectricity plays an important role in the formation of oblique rolls.
Flexoelectricity enters the problem in two ways. Firstly, the flexoelectric
polarization arising from the periodic distortion at the onset of instability
gives rise to additional space charges, and secondly the applied electric field
and the induced field gradients lead to flexoelectric bulk torques. The
theory now involves three linear coupled equations.(184) It turns out that
two regimes of instability - the conduction and dielectric regimes - are
possible as in the Orsay model (3.10.4). For frequencies greater than a
certain value f0 (which is much lower than the cut-off frequency fc
separating the conduction and dielectric regimes) one regains the normal
rolls at threshold. Calculations*185"7* including boundary conditions have
confirmed that the model is able to explain the occurrence of oblique rolls
and its dependence on frequency. The critical frequency/ 0 is proportional
to the average conductivity of the sample, which probably accounts for the
fact that while some observers have reported oblique rolls in MBBA,(181)
others have not.<154)
3.13.3 Order electricity
Meyer's idea of flexoelectricity has been generalized to include a contribution due to the gradient of the orientational order parameter/ 188 ' 189)
The polarization in this case arises not from the curvature distortion of the
director but from the spatial variation of the degree of orientational order
of the molecules. In a simple first order theory, one may take P oc Vs,
where s is the order parameter as defined in 2.3.1. This effect has been
termed as 'order electricity'.
Order electricity may be expected to manifest itself at the nematicisotropic (or air) interface where, as discussed in 2.7, the order
parameter changes rapidly across the transition zone from one phase to the
other. Let us make the simple assumption that at the N - I interface the
gradient of the order parameter ~s/, where is a coherence length. If Pz
is the component of the polarization normal to the interface created by the
order parameter gradient, and the director at the interface is tilted at an
angle 9 with respect to z, the dielectric energy due to order electricity will
be proportional to (189)

212

3. Continuum theory of the nematic state

where e is the dielectric constant (taken to be isotropic). Other factors may


come into play, but neglecting their contributions, it is seen that the energy
is minimum when the tilt assumes the 'magic' angle 0 ~ 53 such that
cos2 0 = \. Interestingly, this is in very good agreement with the observed
value of 52 for some cyanobiphenyl compounds. (190)

4
Cholesteric liquid crystals

4.1 Optical properties


The unique optical properties of the cholesteric phase were recognized by
both Reinitzer and Lehmann at the time of their early investigations which
culminated in the discovery of the liquid crystalline state. When white light
is incident on a 'planar' sample (whose optic axis is perpendicular to the
glass surfaces) selective reflexion takes place, the wavelengths of the
reflected maxima varying with angle of incidence in accordance with
Bragg's law. At normal incidence, the reflected light is strongly circularly
polarized; one circular component is almost totally reflected over a
spectral range of some 100 A, while the other passes through practically
unchanged. Moreover, contrary to usual experience, the reflected wave has
the same sense of circular polarization as that of the incident wave.
Along its optic axis, the medium possesses a very high rotatory power,
usually of the order of several thousands of degrees per millimetre. In the
neighbourhood of the region of reflexion, the rotatory dispersion is
anomalous and the sign of the rotation opposite on opposite sides of the
reflected band. The behaviour is rather similar to that of an optically active
molecule in the vicinity of an absorption. Following the theoretical work
of Mauguin, (1) Oseen(2) and de Vries(3) these remarkable properties can
now be explained quite rigorously in terms of the spiral structure
represented schematically in fig. 1.1.4.
4.1.1 Propagation along the optic axis for wavelengths <^ pitch
Basic theory
We shall first consider the propagation of light along the optic axis for
wavelengths much smaller than the pitch so that reflexion and interference
213

214

4. Cholesteric liquid crystals

effects may be neglected. The problem was investigated by Mauguin (1) with
a view to explaining the optical rotation produced by twisting a nematic
about an axis perpendicular to the preferred direction of the molecules. He
used the Poincare sphere(4) and ' rolling cone' method, but we shall adopt
an identically equivalent formalism, viz, the Jones calculus/ 5 ' 6)
The basic principle underlying the Jones method is that any elliptic
vibration can be represented by the column vector

where Ax and A2 are the resolved components of the electric displacement


vector D along x and y, and are, in general, complex quantities. The
intensity is |y41|2 + |^42|2, while the complex ratio A1/A2 describes its
polarization state. The azimuth A (i.e., the angle which the major axis of
the ellipse makes with x) and the ellipticity co are related as follows: if
tana = |^41|/|^42| and A is the phase difference between Ax and A2,
tan 2 A = cos A tan 2a,
sin 2co = sin A sin 2a.
The effect of an optical system - comprised, say, of birefringent, absorbing
and dichroic plates - is to change A1 and A2, so that
D = JD,
where J is a 2 x 2 matrix with complex elements. In what follows, we shall
adopt the convention of describing the optical system as viewed by an
observer looking at the source of light.
Our aim is to evaluate the matrix J for the cholesteric liquid crystal when
light is incident along the helical axis.(7) As demonstrated by Jones, (8) a
problem of this type can be solved by regarding the medium as being
composed of a large number of infinitesimally thin sections, each section
representing an optical element, in this case a linearly birefringent or
retardation plate. For the purposes of this calculation, therefore, we treat
the liquid crystal (fig. 1.1.4) as a pile of very thin birefringent (quasinematic) layers with the principal axes of the successive layers turned
through a small angle /?.
Let the principal axes of the first layer be inclined at an angle /? with
respect to x,y. If light is incident normal to the layers, i.e., along z, the

4.1 Optical properties

215

Jones retardation matrix for the first layer referred to its own principal axes
is given by
0
G[exp(-iy)
[
0
exp(iy)
where y is half the phase difference between the waves linearly polarized
along the principal axes after passing through a single layer of thickness /?,
i.e., y = ndnp/k, Sn = na nb being the layer birefringence and I the
vacuum wavelength. The retardation matrix with respect to x9y is then
Jx = SGS"1,

where

and S"1 is the inverse of S so that SS"1 = S-1S = E, the unit matrix.
If Do is the complex column vector with respect to x, y representing the
incident light, the emergent light after passing through the first layer is
where, as we are interested at present only in the state of polarization
of the emergent beam, we neglect the phase factor exp( ir/), where
rj = n(na + nb)p/A. (Throughout, we follow the convention of representing
the phase factor at any point +z by exp( \2nnz/X).) Let Dx be now
incident on a second birefringent layer whose principal axes are inclined
at 2/? with respect to x,y. The Jones matrix for this layer is
and the emergent vector is
D2 = S2GS 2DX = S2GS 2SGS
= Sa(GS-1)aD0 = J 2 D 0 ,

where J2 = S2(GS~1)2 is the appropriate Jones matrix for this system of two
layers. In general, if we have a pile of m layers, where the principal axis of
the sth layer is inclined at sfi with respect to x, y (s = 1,2,..., m), the Jones
matrix for the pile is evidently

J m = Sm(GS~T = [l *] say.

(4.1.1)

It can be shown from the theory of matrices<7'9) that


E,
where
cos 6 = cos p cosy.

(4.1.2)

216

4. Cholesteric liquid crystals

Now, as stated earlier, the layer thickness is assumed to be very small, say
a few A, while the pitch P is taken to be at least a few wavelengths of light,
so that both ft ( = 2np/P) and y are small quantities. Therefore
(4.1.3)
From (4.1.1) and (4.1.2)
a = cos rap cos

_ tan/? .
_.
_ . sinra# .
,
rat/H -sin rap sinrat/ i sinycos(ra + l)p,
tan 6
sin 6
(4.1.4)

b=

tan/?
_.
_ .
_
_ .sinra# .
. ,
-cosmpsmmu smmpcosmu I- sinysin(ra + l)p,
tan 6
sin 6
(4.1.5)

c = b* and

d=a*,

where a* and &* are respectively the complex conjugates of a and b.


It is a standard result in optics (6) that such a system can, in general, be
replaced by a rotator and a retarder. If p is the rotation produced by the
system, 2(p the phase retardation and y/ the azimuth of the principal axes of
the retarder,
m

fcos y/
[sm^

siny/ifcos/?
c o s ^ J Lsin/?

sin/?l
cos/? J

ip)
L

0 1 _ r
exp (i#>) J [ sin y/

_ r
cos^J

From (4.1.1) and (4.1.6)

c = -6*

and

rf=fl*.

(4.1.9)

Equating the real and imaginary parts of (4.1.4) and (4.1.7) and of (4.1.5)
and (4.1.8), we obtain after simplification(10)
p = m(J3- 6') radians,
1

(p = cos" (sec rafl'/sec m6)\


y/ = \[{m+ \)pp],
where
m

(4.1.10)
(4.1.11)
(4.1.12)

4.1 Optical properties

217

I/A2 (Mm-2)

Fig. 4.1.1. Rotatory dispersion of a cholesteric liquid crystal for wavelengths <^
pitch. Solution of poly-y-benzyl-L-glutamate (PBLG) in chloroform (18 g/100 g).
(After Robinson.(12)).
In these equations m represents the total number of layers in the system.
Since the layer thickness is taken to be a few A, it turns out in actual
practice that even with the thinnest specimens employed, m is usually a
very large number. We shall therefore assume m to be large throughout this
discussion.
Optical rotatory power
When /? > y, 0' 6. This condition is satisfied when \PSn <^ A, i.e., when
the pitch is not too large. The optical rotation produced by m layers is then

p = m(p-0) = m[fi
and the phase retardation 2(p 0. Thus the system behaves in effect as a
pure rotator. If m is the number of layers per turn of the helix, rnfi = 2n and
mp = P, so that the rotatory power in radians per unit length
p = -n(dn)2P/4l\

(4.1.13)

the negative sign indicating that the sense of the rotation is opposite to that
of the helical twist of the structure. (11) Typically, dn 0.05 for a cholesteric;

218

4. Cholesteric liquid crystals

0.1

0.2

0.3

0.4

0.5

Inverse pitch (jim"

Fig. 4.1.2. Optical rotation per unit length in a twisted nematic film versus inverse
pitch for light of wavelength 0.5 jum. The incident linear polarization is parallel to
the director on the entrance side of the film. Layer birefringence Sn = 0.1. Film
thickness (a) 1.0 /urn, (b) 1.25 jum and (c) 1.5 jum.

taking P = 5 jum and X = 0.5 jum, p ~ 2000 mm" 1. This equation has been
verified experimentally in considerable detail by Robinson (12) in lyotropic
systems and by Cano and Chatelain (13) in thermotropic systems.
Robinson discovered that solutions of some polypeptides in organic
solvents, for example, poly-y-benzyl-L-glutamate (PBLG) in dioxan,
methylene chloride, chloroform etc., spontaneously adopted the cholesteric mesophase above a certain concentration. Under suitable conditions
the solutions exhibited equi- spaced alternate bright and dark lines when
observed through a microscope (see fig. 4.2.7). The lines may be interpreted
as a view of the structure at right angles to the screw axis, so that the
periodicity of the lines is equal to half the pitch. Robinson confirmed this
interpretation by observations between crossed polaroids and also by the
use of a quartz wedge; the retardation plotted against distance in a
direction perpendicular to the lines had an oscillating value, as is indeed to
be expected from the structure. The pitch for any given polypeptide
depended on factors such as concentration, solvent, temperature etc. When
viewed along the screw axis no lines were seen, but a very high optical
rotatory power was present. The rotation in every solution, with a very
wide range of values of P was found to be proportional to \/k2 (fig. 4.1.1).
Robinson substituted the observed values of p and P in (4.1.13), and

4.1 Optical properties

219

calculated the layer birefringence Sn per volume fraction of the polypeptide


in solution. The birefringence was remarkably constant despite the widely
varying values of p and P. He then prepared a solution with equal
quantities of the D and L forms (PBDG and PBLG) which too under
certain conditions adopted the spontaneously birefringent phase, only, in
this case, it was not the twisted cholesteric, but the untwisted nematic
structure. He was therefore able to measure the birefringence directly and
the value agreed well with that derived from (4.1.13).
Similar studies have been carried out by Cano and Chatelain(13) on
mixtures of nematic and cholesteric liquid crystals. The birefringence of the
nematic being very large, Sn of the mixture could be assumed without
sensible error to be equal to that of the nematic itself. The pitch P of the
mixture was measured directly from the Grandjean-Cano steps formed in
a wedge (fig. 4.2.9). The values of Sn and P when inserted in (4.1.13) gave
a rotatory power in quantitative accord with observations.
When P is comparable to or less than y9 the system is no longer a pure
rotator. For large values of the pitch, (4.1.1) may be expressed as(14)
fcos mp
I sin mp

+ sin m6

sin ra/TI

cosra//?

Texp(-im(9)
L
0

\sin 9

[Sl exp( -iy)]


\sin0

When /? and ft/y are extremely small,


m

Tcosra/?

~ [sinm/?

(4.1.14)

y and (4.1.14) reduces to

sin m/TI fexp ( imy)

cosmfi J[

0
1
exp (im6)\

exp (imy)J'

(4.1.15)

Equation (4.1.15) implies that at any point in the medium there are two
linear vibrations polarized along the local principal axes. The polarization
directions of these two vibrations rotate with the principal axes as they
travel along the axis of twist and the phase difference between them is the
same as that in the untwisted medium. This result was first derived by
Mauguin(1) and is sometimes referred to as the adiabatic approximation. It
is this property that is made use of in the twisted nematic device discussed
in 3.4.2.
Fig. 4.1.2 illustrates the theoretical variation of the optical rotation with

220

4. Cholesteric liquid crystals

pitch and sample thickness derived from (4.1.14) and (4.1.15) for very thin
cholesteric or twisted nematic films (the sample thickness being smaller
than the pitch). (1415) Here the rotation is defined as the quantity that is
most conveniently measured, viz, the angle between the incident plane of
polarization, assumed to be parallel to the director axis on the entrance
side of the film, and the major axis of the emergent ellipse. Unlike p given
by (4.1.10), this rotation is not independent of the azimuth of the plane of
polarization of the incident light/ 16 ' 17) but the general conclusions that we
may draw from these curves are still valid. We observe firstly that in
contrast to an ordinary optically active substance the optical rotation per
unit length in the twisted nematic (or cholesteric) is a function of the
thickness. A noteworthy feature is the reversal of the sign of rotation as the
pitch is varied.(17) In the neighbourhood of the Mauguin limit, the rotation
has the same sign as that of the helical structure. With decrease of P the
rotation drops to zero and then reverses sign; thus for lower values of P the
sense of the rotation is opposite to that of the helix, in conformity with
(4.1.13). With increase of sample thickness the peaks in the rotation
increase in height and become sharper. The predicted trends have been
confirmed experimentally.(18)
Absorbing systems: circular dichroism
When linearly dichroic dye molecules are dissolved in a cholesteric liquid
crystal the medium exhibits circular dichroism because of the helical
arrangement of the solute molecules in the structure.(19) The theory
developed above can be extended to take into account the effect of
absorption by treating the layers as both linearly birefringent and linearly
dichroic.(19'20) Assuming that the principal axes of linear birefringence and
linear dichroism are the same, the Jones matrix of any layer with reference
to its principal axes is
G

i *-v WJ

0 l[exp(-A: a /?)
p(iy)J|_
0
exp(iy)J

exp(-iy)
0

0
1
exp(-kbp)\

where ka, kb are the principal absorption coefficients of the layer, and

y=

y-lKK-

4.1 Optical properties

221

Proceeding as before, the Jones matrix for m layers is


j m = Sm(GS"1)ro.

(4.1.16)

If Xx and X2 are the eigenvalues of (GS)" 1, it can be shown that

r = ^f-GS-'-AxA/ 1 .

7 E,

(4.1.17)

where
^ = exp (-<![) exp (i0),
A2 = exp(-)exp(-i/9),
and
cos^ = cos 7 cos )8.
Using (4.1.16) and (4.1.17)

J^exp(-^S-[^GS-^Sin(m-1)'El.
L sin 6
sin 9
J

(4.1.18)

Such a system can be resolved uniquely into a rotator, a retarder, a


circularly dichroic plate and a linearly dichroic plate. The unique matrix
resolution is given by

where
_ [cos ^
[sin y/

- sin y/1
cos y/ J '

_ [cos/?
[sin/7

sin/7
cos/7

[ cosh a/2
i sinh a/2l
[ i sinh a/2 cosh a/2 J'
[
K =

[exp(-/c)

exp (icp) J'


0 1

exp(fc)J'

Here p is the rotation, a the circular dichroism, 2#? the linear phase
retardation, 2K the linear dichroism, / the attenuation coefficient and y/ the

222

4. Cholesteric liquid crystals

azimuth of the retardation plate (or, equivalently, of the linearly dichroic


plate). From (4.1.18) and (4.1.19),
-id = m(p U),
-IK =

with

_x/sec2^
COS"1
r m0/
Vsec

(4.1.20)

X=

ICLL

tan<9

When/ ? is much larger than y


= K

and
Hence
and

a = -myju/p.

Therefore, the linear dichroism of the layers not only results in circular
dichroism but also makes a small contribution to the optical rotation
which is opposite in sign to that due to linear birefringence. However, this
contribution is usually negligibly small. Another consequence of the theory
is that o changes sign whenever /? or ju changes sign. Further, the parameter
o exhibits a marked dependence on pitch and sample thickness. These
predictions are in qualitative agreement with observations.(18"20)
4.1.2 Propagation along the optic axis for wavelengths ~ pitch: analogy
with Darwin's dynamical theory of X-ray diffraction
When the wavelength is comparable to the pitch, the optical properties are
modified profoundly. Before discussing the rigorous electromagnetic
treatment of the problem it is instructive to examine it first from the
standpoint of X-ray diffraction theory/ 7 ' 21) Since the dynamical theory of
X-ray diffraction from perfect crystals and its applications are now quite
thoroughly understood, this approach may be useful in elucidating the
optical behaviour of cholesterics and in looking for new optical analogues
of certain well established X-ray effects. An example of a new phenomenon
is the Borrmann effect in cholesterics.(22)

4.1 Optical properties

223

Kinematical theory of reflexion


The theory discussed in 4.1.1 shows that as long as the pitch is not too
large compared with the wavelength, i.e., when \Pdn <^ 1, the liquid crystal
can be treated to a good approximation as a pure rotator for light
propagating along the helical axis. In other words, right- and left-circular
waves| travel without change of form but at slightly different velocities.
The refractive indices for the two components are respectively

and the rotatory power


(4.1.21)
where Sn = na nb and n = \{na + nb).
We shall now give a simple interpretation of how under certain
conditions reflexion of one of the circularly polarized components takes
place. Let right-circular light given by D o = [*] referred to x,y be incident
along z. We shall suppose that the structure is right-handed, i.e., ft is
positive. To calculate the reflexion coefficient at the boundary between the
(s+ l)th and (s + 2)th layers, we resolve the incident light vector along the
principal axes of the (s+ l)th layer which are inclined at angle (s+ 1)/? with
respect to x, y. The resolved components are

(4.1.22)

where cps+1 = 2nnn(s+ \)p/X, p being the thickness of each layer. At the
boundary, the vibration emerges from a medium of refractive index
na and the rj vibration from a medium of refractive index nh. Qualitatively,
it is obvious that since the principal axes of the (s + 2)th layer are rotated
slightly with respect to those of the (s+ l)th layer, one of the components
of (4.1.22) will on emerging from the (s+ l)th layer meet a 'rarer' medium
while the other will meet a ' denser' medium. One component therefore gets
reflected without any change of phase and the other with a phase change of
n. Thus, in contrast to reflexion from a normal dielectric, the sense of
circular polarization remains the same after reflexion. Applying the
f Right- and left-circular polarizations are denned from the point of view of an observer
looking at the source of light. If the electric vector rotates clockwise with progress of time,
then it is right-circular. Thus, at any instant of time, the tip of the electric vector forms a
right-handed screw in space for right-circular polarization.

224

4. Cholesteric liquid crystals

standard formulae for reflexion at normal incidence from the surface of a


non-aborbing anisotropic crystal, the reflected components ', rj\ referred
to the principal axes of the (s + 2)th layer, are

rf

where q = fidn/2n. We make the approximation here that sin/? /?, since ft
is assumed to very small (~ 10" 2 rad). On reflexion a very slight ellipticity
is introduced in the transmitted beam but we shall ignore it in the present
discussion. Transforming back to x,y the reflected wave on reaching the
surface of the liquid crystal will be

GI--UJ

which represents a right-circular vibration travelling in the negative


direction of z. Clearly the phase difference between this wave and that
reflected at the boundary between the first and second layers is 2(sp(p s).
When h = nnP, we have 2nnnp/A = /? and <ps = s/3 (since mp = P and
mp = 2n, where m is the number of layers per turn of the helix). Hence the
phase factor exp[2i(s/$(p 8)] becomes unity irrespective of the value of s,
and there results a strong interference maximum. For a left-handed
structure, /? is negative and (s^(p s) does not vanish; therefore, the waves
from the different layers will not be in phase and the vibration will be
transmitted practically unchanged.
Using the kinematical approximation, i.e., neglecting multiple reflexions
within the m layers, the reflexion coefficient per turn of the helix is
|g| = m\q\ = ndn/n.

(4.1.23)

Dynamical theory of reflexion


The complete solution of the problem has to take into account the effect of
multiple reflexions. This can be done by setting up difference equations
closely similar to those formulated by Darwin (23) in his dynamical theory of
X-ray diffraction. For the purposes of this theory we shall regard the liquid
crystal as consisting of a set of parallel planes spaced P apart. Each plane
therefore replaces the m layers per turn of the helix. We ascribe a reflexion
coefficient \Q per plane for right-circular light at normal incidence.
Assuming a kinematical approximation for the m layers, Q is given by

4.1 Optical properties

,., | f

225

Sr+1
r+1

Fig. 4.1.3. Notation for the primary (T) and reflected (S) waves in the dynamical
theory.
(4.1.23). We can then write the difference equations in a simple manner
because, as stated earlier, circularly polarized waves travel practically
without change of form, so that the interference of the multiply reflected
waves with one another and with the primary wave can be evaluated
directly.
We shall suppose, as before, that the structure is right-handed and that
right-circular light is incident normally. Let Tr and Sr be the complex
amplitudes of the primary and reflected waves at a point just above the rth
plane, the topmost plane being designated by the serial number zero (fig.
4.1.3). Neglecting absorption, the difference equations may be written as
Sr = - i f i r r + e x p ( - i p ) S r + 1 ,
Tr+1 = exp ( - iq>) Tr - xQ exp ( - 2ip) S r+1 ,

(4.1.24)
(4.1.25)

where cp = 2nnnP/A. The reflexion coefficient is here taken to be the same


on both sides of the plane. Replacing r by r 1 in (4.1.24) and (4.1.25),
substituting and simplifying, we obtain

where

Tr^ + Tr_x=yTr,

(4.1.26)

Sr+1 + Sr_1=ySr,

(4.1.27)

y = exp (icp) + exp ( - i(p) + Q2 exp ( \<p).

(4.1.28)

Suppose that the liquid crystal is a film consisting of v planes. Putting


Sv = 0, we have from (4.1.27)
Sv_2 = ySv_!,

and
-i

v-2.._, ,(v-4)(v-3)

_
(4.1.29)

226

4. Cholesteric liquid crystals

Similarly from (4.1.25), (4.1.26) and (4.1.28)

^v-3 = [ ( / - 1) exp (icp) -y] Tv, etc.

and

To = (fly) exp (i(p) -fUy))

Tv.

(4.1.30)

Since, from (4.1.24)


Sv_, = -iQTv_x = - i e e x p ( i ^ ) Tv9

the ratio of the reflected to the incident amplitudes is

Let us assume a solution in the form


Tr+l = xTr,

(4.1.32)

where x is independent of r. Hence x must satisfy


=y=
We have seen that the reflexion condition is nR P = 20 or q>0 = 2n.
Accordingly, we may write
where
which is a small quantity in the neighbourhood of the reflexion. Therefore
x + (\/x) = exp(i^) + exp(-i^) + g 2 exp(-ie).

(4.1.33)

This suggests that in the neighbourhood of the reflexion we may put


x = exp(-)exp(-i<? 0 ) = exp(-),
where is small and may be complex. From (4.1.33) and (4.1.34)
When
y = exp ( 0 + exp (-) = 2cosh,
the series in (4.1.29) is given by(24)

(4.1.34)

227

4.1 Optical properties

1.0
0.8

0.6

:J

(a)

9S 0.4 -

0.2 -

8 o.o
o
'%

<S

^
1.0 -

ib)

0.8 0.6 _A 1

0.4 0.2 00
0.45

0.49

0.47

0.51

0.53

0.55

Wavelength (jum)

Fig. 4.1.4. Reflexion coefficient 0t at normal incidence versus wavelength for a nonabsorbing cholesteric: (a) semi-infinite medium, (b) film of thickness 25P, where P
is the pitch. Curves are derived from the dynamical theory; circles represent values
computed from the exact theory (4.1.3) assuming that the medium external to the
cholesteric (e.g., glass) has a refractive index 1.5. The parameters used in the
calculations are n = 1.5, Sn = 0.07, X0 = nP = 0.5 jum. (After reference 21.)
Substituting in (4.1.31) and simplifying

So_

-iQexpjie)

(4.1.36)

or

+ cf cothV
For the semi-infinite medium, v = oo in (4.1.36) and
So

ei(Q2-ey

To
When - Q < e < Q, is real and

01= = 1.

(4.1.37)

4. Cholesteric liquid crystals

228

0.75

0.50

0.25

0
0.70

0.68

0.66

0.64

0.62

0.60

0.58

0.56

Wavelength (jum)

Fig. 4.1.5. Reflexion spectrum from a monodomain cholesteric film at normal


incidence. Full curve: experimental spectrum for a mixture of cholesteryl
nonanoate, cholesteryl chloride and cholesteryl acetate in weight ratios 21:15:6 at
24 C (intensity in arbitrary units). Broken curve: spectrum computed from the
exact theory for a film thickness of 21.0 //m and pitch 0.4273 jum. (After Dreher
et

The reflexion is total within this range. The spectral width of total reflexion
is therefore A/I = Qkjn. Using (4.1.23),
A/I = Pdn,
(3)

(4.1.38)

a result first derived by de Vries. Outside this range, the reflexion


decreases rapidly on either side. When X > Ao, e is negative and hence the
negative value of the square root in the denominator of (4.1.37) has to be
taken because @ can never exceed unity; when X < l0 the positive root has
to be taken.
Illustrative curves of M as a function of wavelength are shown in fig.
4.1.4. The semi-infinite medium gives the familiar flat topped curve of the
dynamical theory, while the thin film gives a principal maximum
accompanied by subsidiary fringes. The fringes are somewhat difficult to
observe as even slight inhomogeneities in the specimen and small variations
in its thickness tend to obliterate them, but careful experiments by Dreher,
Meier and Saupe(25) have confirmed that they do occur (fig. 4.1.5).

4.1 Optical properties

229

Primary extinction and anomalous rotatory dispersion


If reflexions are neglected, the optical rotation per thickness P of the liquid
crystal is \{(pn (pL) and the rotatory power is given by (4.1.21). Near the
region of reflexion, the dynamical theory shows that the right-circular
component suffers anomalous phase retardation and, under certain
circumstances, attenuation as it travels through the medium. Left-circular
light, on the other hand, exhibits normal behaviour throughout, and as a
consequence the rotatory dispersion is anomalous around the reflecting
region.
(i) We shall consider first the semi-infinite case. According to (4.1.32),
the amplitude of the right-circular wave as it passes from one plane to the
next is given by
T

xT

where
x = exp(-)exp(-i<p 0 ),

Inside the totally reflecting range, is real and the wave is strongly
attenuated. In X-ray diffraction theory, this phenomenon is referred to as
primary extinction. The extinction length, defined as the distance over
which the amplitude of the incident wave decreases to l/e of its value, is
P/Q at the centre of the reflexion band.
Outside the range of total reflexion, is imaginary and primary
extinction vanishes. Fig. 4.1.6 gives a plot of the wavevectors Kn
( = 2nnn/X) and KL ( = innJX) for right- and left-circular polarizations
respectively, as functions of the wavelength. The real part of Kn shows a
gap within the reflexion band - analogous to the familiar band gap in solid
state physics - while the imaginary part grows rapidly in the same region.
When e2 > Q2, ^ i(e2 Q2)* and may be positive or negative. The
optical rotation per pitch is clearly

and hence the rotatory power in radians per unit length


P=

n(5n)2P
niX X
7TT- +
5-

230

4. Cholesteric liquid crystals


17.0 -

18.0 -

19.0 -

&

20.0 -

21.0?0.45

0.47

0.53

0.49
0.51
Wavelength (jim)

0.55

Fig. 4.1.6. The wavevectors Kn and K^ of the normal waves as functions of


wavelength in a semi-infinite non-absorbing medium. Curves are derived from the
dynamical theory; circles represent values computed from the exact theory, n, Sn
and XQ same as in fig. 4.1.4.(21)

When Q2 > e2, i.e., within the region of total reflexion, p given by
(4.1.39) becomes complex, showing that the medium is now circularly
dichroic. The real part which represents the rotatory power is

n{Sn)2P
U2
'

Pk

(4.1.40)

'

(ii) For a thin film, we have from (4.1.30) and (4.1.35)


sinh(v

*r

cosech v^
(4.1.41)
k 4- coth v

and

- 1.
Consequently the oscillations which appear in the reflexion curve should

4.1 Optical properties

0.45

0.49
0.51
Wavelength (jim)

0.47

231

0.53

0.55

Fig. 4.1.7. Rotatory power versus wavelength for a non-absorbing cholesteric: (a)
semi-infinite medium, (b) film of thickness 25P. Curves are derived from the
dynamical theory; circles represent values computed from the exact theory. , Sn
and Xo same as in fig. 4.1.4.(21)
also be seen in transmission and in circular dichroism. Equation (4.1.41)
may be expressed as

where
tan vy/ =
The optical rotation per thickness P is therefore
and
P

n{8nfP

yj-e

W~ ~^p~'

The theoretical variation of p with X is shown in fig. 4.1.7. As observed

4. Cholesteric liquid crystals

232
-30

-20

0.8
0.6

-10

0.4
0.2
0.0

10

20

0.4

0.5

0.6

0.7
0.8
Wavelength (//m)

0.9

1.0

Fig. 4.1.8. Experimental circular dichroism (open circles) and rotatory dispersion
(closed circles) of cholesteric cinnamate at 177 C. Sample thickness ~ 3 jum.
(After reference 26.)
already in 4.1.1 the rotatory power of a cholesteric liquid crystal, unlike
that of an ordinary optically active substance, is a function of the sample
thickness. Some measurements(26) of the optical rotation right through the
reflexion band using thin films are represented in fig. 4.1.8. The oscillations
in the theoretical curve for p appear to be smeared out, probably owing to
slight imperfections in the sample, but the trends are in agreement with
theory. There is also some evidence of subsidiary maxima in the circular
dichroism which again is to be expected theoretically.
Absorbing systems: the Borrmann effect
The Borrmann effect is the anomalous increase in the transmitted X-ray
intensity when a crystal is set for Bragg reflexion.(27) An analogous optical
effect in absorbing cholesteric media in the vicinity of the reflexion band
has been predicted and confirmed experimentally/ 22 ' 2829) The origin of the
effect can be readily understood by extending the dynamical theory to
include absorption. However, in contrast to the X-ray case, the polarization of the wave field and the linear dichroism play an essential part.
Suppose that the birefringent layers are also linearly dichroic and that
the principal axes of linear birefringence and linear dichroism are the same.
All the equations obtained for the non-absorbing medium hold good in

233

4.1 Optical properties

-1.0
0.45

0.47

0.49

0.53

0.51

0.55

Wavelength (jim)

Fig. 4.1.9. (<z) Reflexion coefficient 0t at normal incidence, (Z?) imaginary parts of Kn
and ATL, (c) rotatory power /? (the real part of />), plotted against the wavelength for
an absorbing semi-infinite medium calculated from the dynamical theory, K = 0.02,
SK = 0.028 and other parameters same as in fig. 4.1.4. (After reference 21.)

this case also except that Q, q>n,


corresponding complex quantities:

etc. have to be replaced by the

Q = nSn/n,
<Pn =

with

2nnnP
X

KR = innJX,
yi

yi

a na

2nnP
X

n(Sn)2P2
4/1 2

2nnP

n(Sn)2P2
4X2

KL = Innjk, etc.,

\rr

1K

a>

yt

b ~

yi

i -is

lK

b>

where Ka,Kb are the principal absorption coefficients. Fig. 4.1.9 gives the
reflexion coefficient 0t and the dependence of the real part of p and the

234

4. Cholesteric liquid crystals

0.00 -

0.45

0.47

0.49

0.51

0.53

0.55

Wavelength (jjm)

Fig. 4.1.10. Transmission coefficients, 5"R and ^ L for right- and left-circular waves
calculated for a film of thickness 25P; (a) non-absorbing and (b) absorbing.
Parameters same as in fig. 4.1.9. The enhanced transmission for the right-circular
component in (b) is the analogue of the Borrmann effect. (After reference 21.)
imaginary parts of Kn and K^ on wavelength. Here Sn, n and P are taken
to be the same as for the non-absorbing case (see fig. 4.1.4) and in addition
it is assumed that K = \{Ka + Kb) = 0.02 and SK = Ka Kb = 0.028. The
interesting result is obtained that on the shorter wavelength side Im (K n) is
less than Im(^ L ), i.e., the right-circular component is less attenuated than
the left component, while on the longer wavelength side the opposite is
true. To observe this effect thin films have to be used.
For an absorbing film of thickness vP,
cosech v

The theoretical dependence of 3Tn and ^ L on X is shown in fig. 4.1.10 for


both the non-absorbing and absorbing cases. The structure being righthanded, the right-circular wave is reflected, and hence in a non-absorbing
film (K = SK = 0) ^"R is always less than ^L. On the other hand, in the

235

4.1 Optical properties

-0.8
0.31

0.33

0.35
0.37
Wavelength (//m)

0.39

0.41

Fig. 4.1.11. Circular dichroism versus wavelength computed for different K and SK.
Sample thickness 25/>, n = 1.5, Sn = 0.07 and X0 = 036jiim. The absorption
coefficients were assumed to be Gaussian curves having a maximum at 0.36 jum and
of halfwidth 0.06 jum. The maximum values of K and SK are, respectively, as
follows: (a) 0.0125, 0.0157; (b) 0.0250, 0.0314; (c) 0.0375, 0.0471; (d) 0.0500,
0.0628; (e) 0.0625, 0.0785. (After reference 28.)

absorbing case 3T^ shows an enhanced value on the short wavelength side
of the reflexion band, which is the analogue of the Borrmann effect.
The phenomenon is shown up even more convincingly in the circular
dichroism curves (fig. 4.1.11). For a left-handed structure (i.e., negative)?),
yL exhibits an anomalous increase and for SK < 0, the peak transmission
occurs on the long wavelength side of the reflexion.

4. Cholesteric liquid crystals

236

-0.4 0.2 -

0.1

0.0
-0.1 -0.2 -0.3
-0.4
0.32

0.34

0.36
Wavelength (//m)

0.38

Fig. 4.1.12. Experimental circular dichroism curves versus wavelength, (a) Pure
cholesteryl nonanoate (CN), (b) CN + 0.98 per cent by weight of PAA. Sample
thickness in both cases 6.5 jum. (After reference 28.)
The first experiments demonstrating the effect were conducted on
cholesteryl nonanoate in which was dissolved small quantities of PAA or
n-/7-methoxybenzylidene-/?-phenylazoaniline.(22'28) The temperature of the
system was adjusted so that the reflexion band overlapped with the
strongly linearly dichroic absorption band of the solute molecule. Under
these circumstances, the circular dichroism exhibits the features predicted
by theory (fig. 4.1.12). Similar measurements were reported subsequently
by Aronishidze et a/.(29) (fig. 4.1.13).

4.1 Optical properties

237

4.1.3 Exact solution of the wave equation for propagation along the optic
axis: Mauguin-Oseen-de Vries model
We next consider the exact solution of the wave equation for propagation
along the optic axis. The complete theory is contained in the papers of
Mauguin,(1) Oseen(2) and de Vries,(3) and has since been presented in
various forms by other authors. (30~3) We shall discuss an elegant treatment
of the theory developed by Kats, (30) and by Nityananda. (31)
We represent the dielectric tensor by a 'spiralling ellipsoid' whose
principal axis Oc is always parallel to z; the other two principal axes Oa
and Ob (with principal values sa and eb) spiral around z with a twist angle
q = 2n/P per unit length. If Oa, Ob are taken to be along x, y at the origin,
the tensor s at any point z may be expressed with respect to x, y as
fcos^z
[sin#z

sin#zir a
cos qz J L 0

[_ asin2#z

01 f cos^z
e j [ singz

singzl
cosgzj

ocsin2qz ,
e acoslqz

(AAA2)

where sa = n2a, sb = n\, e = %ea + eb), a = \(ea-eb) = \{na + nb)(na-nb)


nSn. The wave equation for propagation along z is

e2E
=--jeE.
8?2
c

(4.1.43)

We introduce the variables

E" =

2~\Ex-xEy\

E' is right-circular and E" left-circular for propagation along -fz and vice
versa for propagation along z. Substituting in (4.1.43) we obtain
co2[
e
aexp(2ig.
,22 l ^ ^ /
-:^_\
c [aexp( 2\qz)
e

Ilr7//I'

(4.1.44)

The solution of (4.1.44) is of the form

[A'cxp[i(k-q)z]\'
which is a superposition of two waves of opposite circular polarizations

238

4. Cholesteric liquid crystals

-0.5

\JLJLJLJ

I
500

600
A (nm)

700

Fig. 4.1.13. Circular dichroism in a mixture of 91.5% nematic ROTN 103 (of
Hoffman-La Roche) +7.5 % optically active L-menthol + 1 % dye/?-nitrobenzenebis(benzalazo)-/?'-dimethylaniline. Sample thickness 11.3/mi. Crosses represent
experimental values and the solid curve gives the theoretical variation. (After
Aronishidze et al.{29))

with wavevectors differing by 2q. It is readily verified that when substituted


into (4.1.44) the (k + q) component of (4.1.45) suffers a wavevector shift of
2q and is converted into a(k q) wave, and vice versa, so that together the
two components form a closed set. Equation (4.1.45) therefore represents
a true normal wave which can satisfy (4.1.44) with a proper choice of A'
and A". From (4.1.44) and (4.1.45), we have
(4.1.46)

(k-qf-t

where K = 2n/X and Km = InnIX, X being the wavelength in vacuo. The


condition for (4.1.46) to yield non-vanishing solutions of A' and A" is

[(k + qY- Kl] [(k -qf-Kl\-

o?K* = 0,

whose roots are

kt, k, =

(4.1.47)

4.1 Optical properties

239

Corresponding to the kx and k2 solutions we have respectively


*K2

A'
A"
and

}
A"
A'

(4.1.48)

-qf-Kl')

When a is small, (4.1.47) and (4.1.48) give for the k2 solution A' I A" ~ a,
and for the k solution A" I A' ~ a. In other words, each normal wave is
made up of two oppositely polarized circular components with one of the
components generally dominating. The mixing of these two components
with wavevectors differing by 2q is a consequence of the Bragg reflexion.
Equation (4.1.45) may conveniently be rewritten as

U_ T

U
I

where
Kx = k1+q

exp(i^z)

exp(iK2z)

]'

= q + [K2m + q2-(4Klq*
2

+ a*K^]K
2

K2 = K ~ q = - q + [Kl + q + ( 4 ^ q + a K*)i\\

(4.1.51)
(4.1.52)

The fact that the wavevectors Kx and K2 are different is responsible for the
optical activity of the medium, the optical rotation per unit length being
p = K ^ ^g) rad. However, as emphasized in previous sections the phenomenon is not identical with natural optical activity because the normal
waves are not pure circular waves.
The de Vries equation
If now we make the approximation that (Kx K2)/q <^ 1, or that the
rotation per pitch is small compared with n, which is certainly valid in most
cholesterics,
P=
where

240

4. Cholesteric liquid crystals

When x2 < a2K*, p becomes a complex quantity. The real part gives the
rotatory power and the imaginary part the circular dichroism. Since no
dissipative mechanism is built into the model it follows that the imaginary
part of p is associated with the reflexion of one of the components. The
reflexion band is centred at x = 0, i.e., at Km = q or Xo = nP where Xo is the
wavelength in vacuo. The range of reflexion extends from x = +aK2 to
x = OLK2, i.e., from + q\dn/ri) to -q\dn/ri). Since
3x = S(KJ2 = 2Km(SKm) ~ 2q(SKJ,
the spectral width of total reflexion is given by
AX = PSn

(4.1.54)

as was first shown by de Vries(3) (see (4.1.38)).


When x21> a 2 ^ 4 , which is not valid close to or inside the reflexion band,

n{dnfP

(4.1.55)

This is known as the de Vries equation. The sign of the rotatory power
reverses on crossing the reflexion band (Ao). When X <^ Ao, (4.1.55) reduces
to (4.1.21), and when X > 20, p tends asymptotically to 0. The behaviour on
either side of the reflexion band has been confirmed experimentally(26) (fig.
4.1.8).
Thin films
(16)

Nityananda and Kini have applied the theory to obtain exact solutions
for reflexion and transmission by a plane parallel film bounded on either
side by an isotropic medium. The treatment allows for the contribution due
to reflexion at the cholesteric-isotropic interface. In general, for each
circular polarization at normal incidence the reflected and transmitted
waves consist of both circular polarizations. Four coefficients, two for
reflexion and two for transmission, are required to describe the problem
fully and the solution consists of matching the incoming and reflected
waves on one side of the slab with four waves within the slab (two in the
forward direction and two in the backward direction) and the transmitted
wave on the other side. An extension of the treatment to absorbing media
yields the theory of the Borrmann effect.(22)

4.1 Optical properties

241

Some calculations for the semi-infinite medium and for the thin film are
shown in figs. 4.1.4, 4.1.6 and 4.1.7. In these calculations, the isotropic
medium external to the liquid crystal is assumed to have a refractive index
equal to n = \{na + nb) so that the contribution of the ordinary Fresnel
reflexion coefficient at the cholesteric-isotropic interface is eliminated. It is
clear from these figures that the results of the exact theory differ only
slightly from those of the dynamical model, indicating that the latter is
probably adequate for most practical calculations. However, the simple
formulation of the dynamical model presented in 4.1.2 does have some
inherent limitations: (i) it is valid only for integral values of the pitch, (ii)
it is developed for small e (= 2U{X XQ)/X) and therefore does not give
exact values for wavelengths away from the reflexion band, and (iii) it fails
when the film thickness is very small, or when the extinction length is of the
order of a pitch. These limitations arise primarily because of the
kinematical approximation made for the reflexion from the m layers per
turn of the helix. We shall now show that when multiple reflexions within
the m layers are also included, the simple difference equations become
matrix difference equations and the resulting solutions turn out to be fully
equivalent to the exact electromagnetic treatment. Proofs of this result
have been given by Joly(32) and by Nityananda. (3134) We shall follow the
latter's treatment(34) which is simpler.

4.1.4 Equivalence of the continuum and the dynamical theories


We go back to the model discussed in 4.1.1, viz, a twisted stack of thin
birefringent layers. The principal refractive indices of each are na and nb,
and the angle of twist between the successive layers is /?. Let ra, ta, rb and tb
be the reflexion and transmission coefficients for a single layer for light
linearly polarized along its principal axes at normal incidence. For
polarization along the a axis,
rn = -/ =

(+!)*-(-1)* T J'

where xa exp (ina Kp), K is the wavevector in vacuo and p the layer
thickness.(35) Exactly similar expressions may be written for the other
polarization also.
We define Is and Js as the amplitudes incident on the sih layer in + z and

242

4. Cholesteric liquid crystals


s

s+\

Fig. 4.1.14. Notation for the incident and reflected waves. Is and Js are the
amplitudes of the waves incident on the 5th layer in the positive and negative
directions respectively, and Es and Ss the amplitudes emerging from the 5th layer in
these two directions.
z directions respectively, and similarly Es and Ss as the amplitudes
emerging from the sth layer in the + z and z directions respectively (fig.
4.1.14). Therefore, for the sth layer we have
a _
s ~

f
l l

ja
s

(4.1.56)

The first two equations of (4.1.56) can be combined and written as

The second two lines can also be similarly combined. However, in what
follows we shall write them in the form
(4.1.57)
with the understanding that E, I, S and J* are each column vectors ( 2 x 1
matrices), and that r and t are each 2 x 2 diagonal matrices.
Now, the emergent wave E s in the + z direction from the ^sth layer is
physically the same as I s+1 , the wave incident on the (s+l)th layer, but
(4.1.57) will apply to the (s+ l)th layer only if I s+1 is referred to its own

4.1 Optical properties

243

principal axes which are rotated through /? with respect to those of the sth
layer. We therefore write
IS+1 = SES
(4.1.58)
and similarly
J^S-1^,
(4.1.59)
where
cosy?
sin/?l
sin/? cos/?J
Strictly these equations should include a phase factor allowing for the air
gaps between the layers, but the gap may be taken to be infinitesimally
small compared with the thickness of the layer, which itself tends to zero.
Using (4.1.58), (4.1.59) and (4.1.57)

Because of the periodicity with respect to s, the difference equations can be


solved in the form
Es = Eexp(i^).
This yields

Now the total field in the gap is F = E +/, so that


F - J ^ = tSexp(-i^)(F-j^) + rJ^,
Jf = S" 1 rS(F-./) + S- 1 texp(ip)./.

(4.1.60)
(4.1.61)

From (4.1.60) we get


Using this to eliminate from (4.1.61)
{[l+S- 1 rS-S- 1 texp(i(^)][l+r-tSexp(-i^)]- 1
x [1 - tS exp ( - \(p)\ - S'^S} F = 0. (4.1.62)
As before we effect a change of variables:
F=2-kFa\Fb),
i.e.,

M -3EMS-*

244

4. Cholesteric liquid crystals

The matrices in the difference equations should also be transformed to the


new variables. For example r should be replaced by

i = [i(',. + O K ' a - O i r r Sr/2]


~YWa-rb)

Wa + rb)\

[Sr/2 f \

where r = \{ra + rb) and Sr = ra rb9 and S by


ASA"1 =

i/0

0 "I
exp(iy9)J'

Since r and t are functions of the thickness p of the layer, we expand them
in powers of/?. It is sufficient to retain the first power in each case:
F=l(s-l)iKp,
dr/2 = \xaKp,

dt/2 = (ta - Q/2 = \i<xKp,

where as before e = |(a + eb) and a = |(e o ). Writing fi = qp and q> = kp


(and remembering that since r is of order p, rS ~ r to this order) (4.1.62)
may be expressed as
(4.1.63)
where R simplifies to (4.1.64) below:
-1
K

Uk + qf-eK*
-*K* 1
2
1 -a*
(k-q)*-eK'\-

6 )

Since the first matrix on the right-hand side is non-singular we may


premultiply by its inverse to obtain (4.1.63) in the form

Uk + qf[ -<xK2

(k-q)2-sK2\[F_\

'

which is precisely the same as (4.1.46). The dynamical theory applied in this

245

4.1 Optical properties

manner to a twisted pile of birefringent layers is then exactly valid for any
arbitrary thickness of the sample and for the entire range of wavelengths.
4.1.5 Oblique incidence
The theory of propagation inclined to the optic axis is, of course, very
much more complicated, and analytical solutions have not so far been
found.<36) The first attempt at solving the problem numerically was by
Taupin,(37) but the most complete calculations are those of Berreman and
Scheffer<38) who also carried out a precise experimental study of reflexion
from monodomain samples at oblique incidence. Fig. 4.1.15 presents their
observed reflexion spectra for two polarizations.
Berreman used a 4 x 4 matrix multiplication method. Assuming the
incident and reflected wavevectors to be in the xz plane, z being along the
helical axis of the cholesteric, the dependence on the y coordinate may be
ignored altogether. Writing Ex = Exexp [i(cot kx)] etc., it is easily verified
that Maxwell's equations can be reduced to the matrix form
.kceT

1-

\HV

CO

CO

-e

kcY
CO

CO 1

Ex

iHy

Ey

or
QZ

where, assuming a spiralling ellipsoid model, the components of the


dielectric tensor are given by
exx =
eyy = e occos2qz,

and all other components are zero (see (4.1.42)).

4. Cholesteric liquid crystals

246

([

0.2 -

00

0.2

0.6 _

jV

r
I

First
order

1.0

0.6

ji^

Second
order

Observed spectra

I
1.4

Computed spectra

Fig. 4.1.15. First and second order reflexion spectra of a cholesteric liquid crystal
film (0.45:0.55 mole fraction mixture of 4/-bis(2-methylbutoxy)-azoxybenzene and
4,4'-di-n-hexoxyazoxybenzene) 15 pitch lengths or 11.47 jum thick. Angle of
incidence 45. Polarizer and analyser are parallel to the plane of reflexion for 0tn and
normal to it for 0to measurements. The small oscillations are interference fringes
from the two cholesteric-glass interfaces. (After Berreman and Scheffer.(38))

4.1 Optical properties

247

To a first order of approximation

where 3P is a 4 x 4 propagation matrix and E is the unit 4 x 4 matrix.


Repeated matrix multiplication in very small steps Sz gives the matrix for
the total film, and by taking into account the appropriate boundary
conditions on either side (glass) one can work out the transmitted and
reflected waves. Of course, cyclic and other symmetry properties of 3?(z)
reduce the number of matrix multiplications to a reasonable number in
practical cases. In the actual calculation, Berreman and Scheffer included
a second order term subject to the symmetry property
&>(z,Sz) =

0>-1(z,-dz).

The results of their computations are shown in fig. 4.1.15. The agreement
with the experimental spectra can be seen to be good. There is a difference
in the intensities, which may conceivably be due to thin regions near the
surface with anomalous dielectric properties or due to the neglect of
absorption.
An important fact that emerged from this study is that the observed
features were best reproduced when the local dielectric ellipsoid was taken
to be a prolate spheroid, with the principal axis Oc parallel to z and
sc = s a. Thus the assumption that is generally made that the local
dielectric ellipsoid is uniaxial would appear to be valid to a very good
approximation as far as optical calculations are concerned (see, however,
4.10).
4.1.6 Propagation normal to the optic axis
When light is incident normal to the optic axis polarized diffraction
maxima are seen in transmission.(39) For the electric vector polarized
parallel to z the refractive index is independent of the z coordinate, whereas
for the vector polarized perpendicular to it the refractive index varies
periodically from na to nb. The wavefront having the latter polarization
therefore suffers changes of phase which vary along z with a periodicity
equal to half the pitch. There can also be changes of amplitude as parallel
rays tend to acquire a slight curvature when travelling in a medium in
which the gradient of refractive index is normal to the direction of
propagation. (40) This periodicity in the phase and amplitude gives rise to
polarized diffraction effects. Approximate expressions for the intensity of
the maxima may be derived by applying the Raman-Nath theory of the
diffraction of light by ultrasonic waves.(41)

4. Cholesteric liquid crystals

248
0.3

0.2
0.1

-0.1

-0.2

-0.3

30

34

38

42

46

50

54

Temperature (C)

Fig. 4.1.16. Variation of inverse pitch with temperature in a 1.75:1 weight mixture
of right-handed cholesteryl chloride and left-handed cholesteryl myristate as
determined by laser diffraction. The mixture becomes nematic at 42 C. (Sackmann
et
Sackmann et al.(39) have investigated the temperature variation of the
pitch of a mixture of right-handed cholesteryl chloride and left-handed
cholesteryl myristate by this method. At a certain temperature (7^) there is
an exact compensation of the two opposite helical structures and the
sample becomes nematic. At this temperature only the central spot (zero
order) is observed, while at the other temperatures, polarized diffraction
maxima of higher order make their appearance. The inverse pitch varies
almost linearly with temperature passing through zero at 7^ (fig. 4.1.16).

4.2 Defects
4.2.1 %-disclinations
We now consider defect structures in the cholesteric liquid crystal. Treating
the cholesteric as a spontaneously twisted nematic,
nx = cos 0, ny = sin #,
6 = qz,

nz = 0,

q = 2n/P,

P being the pitch. The free energy density is then expressible as


(4.2.1)

4.2 Defects

249

Fig. 4.2.1. Director patterns for .s = and 1 /-screw disclinations in a cholesteric.


In the one-constant approximation,
(4.2.2)

and V26> = 0.
disclinations
Such disclinations are closely analogous to nematic wedge disclinations
(3.5.1). The singular line is along the z axis (parallel to the twist axis) and
the director pattern is given by
nx = cos 9, ny = sin 6, nz =
0 = s tan" 1 (y/x) + qz

(4.2.3)

s = N/2,
where N is an integer. The presence of the disclination does not alter the
pitch. Fig. 4.2.1 illustrates the director patterns for disclinations of strength
s = 1 and 1. Many of the conclusions arrived at in 3.5.1 are valid in this
case as well, and in the one constant approximation the expressions for the
energies and interactions are the same as for nematic wedge disclinations.
As already indicated briefly in 3.5.8 the effect of elastic anisotropy has
some interesting implications for cholesterics, especially for long-pitched
structures. We have seen that disclination pairs in nematics have angular
forces in the presence of elastic anisotropy. For all practical purposes, the
solutions that were obtained for nematics will hold good for each nematiclike cholesteric layer, except that the layers now twist continuously in the

250

4. Cholesteric liquid crystals

(a)

Fig. 4.2.2. A pair of like /-screw disclinations forming a stable double helix in a
cholesteric: (a) a pair of s = \ disclinations (after Cladis, White and Brinkmann(42)),
(b) a pair of s = 1 disclinations (after Rault(43)); see fig. 3.5.24.
medium. Therefore, for a pair of disclinations, angular forces will lock the
line joining the disclinations at the same orientation with respect to the
local director n (of the defect-free sample) in every layer. Hence while single
disclinations may be straight, pairs of disclinations in a cholesteric may be
expected to have a helical configuration (see fig. 3.5.24). For a pair of like
disclinations, mutual repulsion increases the separation between them, but
in the helical state this will be opposed by the line tension in each
disclination. In the end the two opposing processes should balance to result
in a stable double helix. This is indeed found to be the case experimentally

4.2 Defects

251

&
Fig. 4.2.3. A helical %(s = 1) disclination wound round a straight %(s = +1)
disclination in a cholesteric liquid crystal (Rault(44)); see fig. 3.5.24.
(fig. 4.2.2). On the other hand, if the individual disclinations of the pair
have different line tensions (as is the case for 1 and 1, both of which are
escaped structures) then the disclination with a lower tension should wind
around the one with the higher tension. This again appears to be in
conformity with experimental observations (fig. 4.2.3). The escaped
configurations for 1 and 1 in the cholesteric are rather complex,(45) which
may perhaps account for the fact that the two do not annihilate one
another.

252

4. Cholestenc liquid crystals

-I

h- I- h- I-

h-

Fig. 4.2.4. The director pattern for s = \ /-edge disclination in a cholesteric. Dots
signify that the director is normal to the plane of the diagram, dashes that it is
parallel to and nails that it is tilted.
X~edge disclinations
In this case the singular line is perpendicular to the twist axis. On going
round this line, one gains or loses an integral number of half-pitches. The
director pattern around the /-edge disclination was first worked out by de
Gennes(46) who proposed a nematic twist disclination type of solution:
nx = cos 0,

ny = sin 0,

nz = 0,>

9 = s tan"1 (z/x) + qz,

(4.2.4)

s = N/2.
The cholesteric pitch is altered around the singular line where TV is an
integer. The pattern for s = \ is shown in fig. 4.2.4. Again, the energies and
interactions in the one-constant approximation are the same as for nematic
twist disclinations. A somewhat more elaborate treatment of this model
has been presented by Scheffer(47) and the effect of elastic anisotropy has
been investigated by Caroli and Dubois-Violette.(48)
The Volterra process for creating these disclinations is the same as for
nematic disclinations. For the screw disclination the plane of cut is parallel
to the cholesteric twist axis while for the edge disclination it is perpendicular
to it.
4.2.2 Lattice disclinations
The cholesteric may also be regarded as having a layered structure with a
periodicity of P/2 along the z axis. This lattice can have disclinations just
as in smectic A and the Volterra process for creating them is also essentially
the same (see 5.4.4). If the cut is such that the line L is along the local
molecular axis, the disclinations so created are designated as X+ and A~,
{X standing for ' longitudinal' and the plus (or minus) sign indicating that

4.2 Defects

253

(b)

(a)

\ N

id)
+

Fig. 4.2.5. The configurations for (a) A", (6) A , (c) T" and (d) T+ disclinations in a
cholesteric. Dots, dashes and nails have the same significance as in fig. 4.2.4.

(a)

(d)
Fig. 4.2.6. Examples of pairing of X and z disclinations of opposite signs in a
cholesteric. Edge disclinations composed of (a) X' and A+, (b) T~ and T+, (C) T~ and
A+, (d) X~ and T+ and (e) pincement composed of z+ and T~.

material has to be removed (or added) to arrive at the final configuration)


but if L is perpendicular to the local molecular axis they are designated as
T+ and T~ (fig. 4.2.5) (T standing for 'transverse'). The Is are coreless and
therefore have lower energies than the is which do have cores.

254

4. Cholesteric liquid crystals

As or rs of opposite signs occur in pairs to form dislocations and


pincements (fig. 4.2.6). Such pairing can be observed directly in the
fingerprint textures which are exhibited by cholesterics of large pitch when
the helical axis is parallel to the plates (figs. 4.2.7 and 4.2.8).
4.2.3 Dislocations
Because of the layered structure, defects in the cholesteric can be likened in
many respects to those in smectic A. Both of them exhibit focal conic
textures(52) and both allow for the existence of screw and edge dislocations.
To discuss these similarities we employ a 'coarse-grained' approximation
in which the cholesteric distortions are considered to be small and to vary
slowly over a pitch. In this approximation the free energy of distortion may
be expressed in terms of layer displacement u parallel to the twist axis:

where B = k22 q2 and K = |fc33 (see (4.6.22)). This expression first derived
by de Gennes is exactly of the same form as (5.3.3) for smectic A. We shall
now consider some applications of this model.(53)
Screw dislocations
The deformation around a screw dislocation may be written as
NP
u = -^Uur1(y/x)9

(4.2.5)

where PQ is the pitch and TV is an integer. The singular line is along the twist
axis. A circuit around the singular line results in a displacement of the
layer along the helical axis by an integral number of half-pitches. Equation
(4.2.5) can be recast to give the director orientation, and it is easily verified
that it becomes identical to (4.2.3) for the /-screw disclination and leads to
same results for the energies and interactions.
Edge dislocations
Let the layers be in the xy plane and the line of singularity along y.
Following the theory for smectic A (5.4.2), one may write,
p Y If 00 H/r
1
u(x,z) = - 1 + T-exp(-Aic a |z|+iK*)
where Po is the pitch of the undistorted

cholesteric and

X=

4.2 Defects

255

(a)

(b)

(iv)

Fig. 4.2.7. Disclinations and dislocations in (a) fingerprint textures of cholesterics.


(b) interpretation of (i), (ii) and (iv) of (a). (After Bouligand and Kleman.(49))

256

4. Cholesteric liquid crystals

Fig. 4.2.8. Fingerprint textures of cholesterics showing pincements, (a) after


Robinson, Ward and Beevers,(50) and (b) after Bouligand.(51)

4.2 Defects

257

(P0/4n)(3kS3/2k22)i. From this one gets the layer tilt 6 (with respect to the
z axis) and the layer dilatation S = du/dz as

)\

<4 2 6)

- -

The energy of a single dislocation in an infinite medium is then


h

0.6nkP

where Ec is the energy of the core and C the core radius, which in this model
can probably be assumed to be of the order of Po. Hence the energy turns
out to be finite and does not diverge logarithmically with sample size as
does a nematic-lilce solution of the form (4.2.4).
If there are two like dislocations, one at (0,0) and the other at (x0, z0) the
forces Fx and Fz (along and perpendicular to the layers) acting between
them are

kPl

For a pair of like dislocations Fx is always repulsive, while Fz is repulsive for


x\ < 2A0z0 and attractive for x\ > 2Aozo. Thus, as in the case of smectic A
(see fig. 5.4.7), there can result a clustering of like edge disclinations to form
a 'grain boundary'. Such clustering is often observed in fingerprint
textures (fig. 4.2.7).
When a cholesteric sample is prepared in the form of a thin wedge
between two glass plates inclined at a small angle, and the twist axis is
approximately normal to the plates, regular striations are seen running
across the film when viewed under a polarizing microscope (fig. 4.2.9(a)).
These striations, usually referred to as the Grandjean-Cano pattern, are
due to edge dislocations. If 9 is the wedge angle, the number of edge
dislocations per unit length that are required to minimize the energy is
d'1 = 26/P0 where d is the average distance between dislocations. The
darker lines in the photograph are dislocations having double the Burgers
vector of the weaker ones. They are composed of /l~A+ pairs, which as we
have seen already are devoid of cores.(55) When a magnetic field is applied

258

4. Cholesteric liquid crystals

Fig. 4.2.9. Grandjean-Cano pattern of edge dislocations in a cholesteric wedge


formed between a cylindrical lens and a flat plate (a) without a magnetic field and
(b) with a magnetic field normal to the spiral axis and the dislocation line. The
' double' dislocations - the darker lines - become zig-zag in the presence of the
magnetic field while the 'single' dislocations remain stable. (The Orsay Liquid
Crystal Group.(54))
normal to the helical axis and to the dislocation line, the 'double'
dislocations get distorted (fig. 4.2.9 (b)). This is because a cholesteric
composed of molecules of positive diamagnetic anisotropy prefers to have
its helical axis normal to the field, but the region between the disclination
pair has its axis along the field. Energetically, therefore, it is favourable for
the lines to adopt a zig-zag shape. Qualitative arguments suggest that the
lowering of the energy is greater the larger the separation between the
disclinations. This probably explains why the weaker lines, which are
composed of Xx pairs, remain stable.

4.3 Leslie's theory of thermomechanical coupling

The fact that the properties of the cholesteric liquid crystal are not
invariant with respect to reflexion introduces some additional complexity
into the equations of the continuum theory. The possibility now exists of

4.3 Leslie's theory of thermomechanical coupling

259

a coupling between thermal and mechanical effects which symmetry


considerations rule out automatically from the theory of the nematic state.
Thus translational or rotational motion of the fluid can, in principle, cause
heat transfer, and equally a thermal gradient can create motion. (56)
Consider an incompressible cholesteric fluid with a non-polar director of
constant magnitude. The basic equations developed in 3.1 need to be
modified slightly because of the absence of a plane of symmetry. Allowing
for heat flux and thermal gradients the conservation laws are
/> = 0,

(4.3.1)

U=Q-qi,i + tjidti + njiNtj-giNi,

(4.3.3)

where Q is the heat supply per unit volume, qt the heat flux vector per unit
area per unit time and the other symbols have the usual meanings (see
3.3). If /?^ is the entropy flux vector per unit area per unit time, it is
convenient to introduce a vector cpi such that

The entropy inequality may then be written as


t^

+ nttNv-gtNt-PtTt-F-St-v^

> 0,

(4.3.5)

where F = U TS is the free energy function. Making use of the fact that
Ti9 dtj and nu can be chosen arbitrarily and independently, we obtain the
following equations:

with

<Pt = ^m^(Nk

+ dkpnp),

where a is a constant. The inequality (4.3.5) then becomes

260

4. Cholesteric liquid crystals

Resolving tip gt and pt into their static and hydrodynamic parts,

we have

~~^nk,p

(4.3.8)

Pt = 0 ,

(4.3.9)

t'iidii-g',N<-(p'i+%j!) T{ > 0.

(4.3.10)

fn + gi^ = t'ji + g'jni,

(4.3.11)

and (4.3.6) reduces to

Also

and the entropy generation per unit volume


TS = Q-qitt

+ <KtjJnNk + dkpnp)li + fJtdv-g'iNt.

(4.3.12)

The hydrodynamic components of the stress tensor etc., are


t'n = Mi nk % dkv nt n, + ju2 N< n, + //3 N, nt + //4 dn + ju5 dik nk n5
+ ^edjknkni

+ M7etpqTQnjnp

+ /i8eJpqTqntnp9

(4.3.13)

g\ = k1Ni + A2 ^ rij + /l3 ^ M /i p Tq9

(4.3.14)

ft = ^ r , + ^ 2 /ifc r f c ^ + ^ 3 etpq np NQ + A:4 eipq nv dqr nr,

(4.3.15)
(4.3.16)

where // 15 ...,// 6 are the six viscosity coefficients already defined in the
continuum theory of the nematic state; ju7 and ju8 are two additional
viscosity coefficients coupling thermal and mechanical effects; A^1?...,AT4
are the coefficients of thermal conductivity.
The free energy of elastic deformation per unit volume is given by (3.2.7).
Martin, Parodi and Pershan(57) and Lubensky(58) developed a general

4.3 Leslie's theory of thermomechanical coupling

261

Fig. 4.4.1. Lehmann's diagrams depicting the rotation phenomenon in open


cholesteric droplets heated from below. (After Lehmann.(59))
hydrodynamic theory of layered systems, and showed that symmetry
permits a thermomechanical coupling between the phase of the layers and
the temperature gradient. For a cholesteric, the phase is determined by the
azimuthal angle of the director, and this again leads to the same conclusions
as Leslie's theory.

262

4. Cholesteric liquid crystals


4.4 The Lehmann rotation phenomenon

An example of this type of thermomechanical coupling appears to have


been observed by Lehmann(59) in cholesteric liquid crystals very soon after
their discovery. He found that droplets of the material when heated from
below seemed to be rotating violently, but from optical studies he
concluded that it was not the drops themselves but the structure that was
rotating. Fig. 4.4.1 shows a few of the many sketches that he made
depicting his observations. Leslie's equations(56) offer a simple explanation
of the phenomenon; because of the absence of mirror symmetry, an
applied field, which is a polar vector, can result in a torque, which is an
axial vector.
Let the cholesteric film be bounded between the planes z = 0 and z = /*,
and let there be a temperature gradient along the screw axis z. The
components of the director in a right-handed cartesian coordinate system
are {cos#(z, t), sin#(z, i), 0}. We assume that there are no heat sources
within the liquid crystal, no external body forces and that the velocity
vector is zero. Hence T = T{z)Ji = G{ = dtj = wtj = 0. Thus from (4.3.4)
820

A _ 87*

+.

^ M* *)
_ 60

8/

80\

(4A)

and from (4.3.12)

^)-{x^-x^k>

(4A2)

where

and Fo is the free energy in the absence of elastic deformation. In static


situations, 820/8f2 = dO/dt = 0, and (4.4.1) and (4.4.2) yield

Af -k
dz\

giving

CT

Jo

de

\-)idT-o

22

dz)

6z

4.4 The Lehmann rotation phenomenon

263

and
^ Jo

22 \ Jo

Jo ^22

where ^4, 5 , C, D are constants which can be determined from the


boundary conditions.
If T(0) = To, T(h) = T and the director is assumed to be completely free
at the bounding surfaces (i.e., the torques TU = eijkntnlk at the boundaries
are zero),
^k

+ oc)

= [k22-k 2

+ oc)

=0.

One can then work out a simple solution if the material constants are
assumed to be independent of temperature. Let 9 = cot +/(z) and T = g(z);
(4.4.1) and (4.4.2) then reduce to

^ 2 2 0 - ^ +^
a2?

a?

oz

oz

= 0,

(4.4.3)

Equation (4.4.4) gives


g = (V^s) ^

+ ^ exp (A8 G > Z / ^ ) + 5,

and (4.4.3) then gives


/ = (Kx A/cok22) exp (A3 coz/KJ + Cz + D.
From the boundary conditions, one obtains finally
r=ro+[(r1-ro)z//i],

(4.4.5)

e = eo + ^ + [(*a-a)z/fc22],

(4.4.6)

where 60 is the orientation at z = 0, and


.

(4.4.7)

Thus the director rotates about Oz with an angular velocity co, which
explains Lehmann's observations.
In the absence of a temperature gradient (4.4.6) reduces to
0 = 60 +

[(k2-a)z/k22]

which describes the normal cholesteric structure with a pitch P =


2nk22/(k2 a). The coefficient a has not yet been determined experimentally

264

4. Cholesteric liquid crystals

Fig. 4.4.2. A spherical drop of a long-pitched cholesteric liquid crystal showing the
characteristic/-line (from Robinson(62)). The structure of this defect was explained
by Frank and Pryce(61).
and it is not known whether its contribution is of practical significance.
According to the Oseen-Zocher-Frank elasticity equations a = 0, and in
the absence of any evidence to the contrary it is generally neglected in most
discussions.
The Lehmann rotation phenomenon has never been reproduced since its
discovery. However, the experiment has been successfully repeated in the
author's laboratory (60) using a DC electric field instead of a thermal
gradient. Difficulties arising from anchoring effects at the boundaries were
eliminated by forming cholesteric drops suspended in the isotropic phase.
This was achieved in the following manner. The material was carefully
chosen to avoid other electrical effects on the orientation of the director.
A binary mixture of alkoxyphenyl-rras-alkylcyclohexyl carboxylates
(supplied by Merck) was prepared to give a room temperature nematic
with dielectric anisotropy 1. The mixture did not exhibit any
electrohydrodynamic instabilities up to DC voltages of ~ 8 V. Addition of
~ 5 wt % of cholesteryl chloride resulted in a left-handed cholesteric liquid
crystal with pitch P % 5//m, while ~ 10% of methylbutylbenzoyloxyheptyloxycinnamite yielded a right-handed cholesteric with the same P.

4.4 The Lehmann rotation phenomenon

265

Fig. 4.4.3. Photographs demonstrating the Lehmann rotation effect in cholesteric


drops under the action of a DC electric field. (a)-(c) taken about 30 s apart,
illustrate a clockwise rotation of the structure, and (d)(f) an anticlockwise
rotation when the voltage is reversed. The material used was a binary nematogenic
mixture of alkoxyphenyl-fraws-alkylcyclohexyl carboxylates (supplied by Merck) to
which was added ~ 5 wt% of cholesteryl chloride. (After reference 60).
The cholesteric material was then doped with a small quantity of a nonmesomorphic epoxy compound, Lixon, which lowered the cholestericisotropic transition temperature and gave rise to a broad two-phase
region. Because the glass plates have greater affinity for Lixon than for the
liquid crystal compound, the cholesteric drops were surrounded on all
sides by the isotropic phase. With thick cells, spherical drops with the
characteristic /-line of strength 2 were formed (fig. 4.4.2). In thin cells
(~ 8 jum thick) flattened drops were obtained in which the central portion

266

4. Cholesteric liquid crystals

had an essentially planar structure with the helical axis normal to the flat
region and the /-line was confined to the periphery. Since the anchoring
energy at the cholesteric-isotropic interface may be expected to be
negligible, these flat drops proved to be most suitable for studying the
phenomenon. In principle, any transport current can produce a crosscoupling effect and hence a DC electricfieldwas used for convenience. At
2 V DC, the dark brushes emanating from the /-line became curved, and
then the whole structure started to rotate, apparently without any further
distortion. Fig. 4.4.3 shows the photographs of the rotating drops, which
can be seen to be closely similar to Lehmann's diagrams reproduced in fig.
4.4.1. Systematic observations on a number of drops established the
following results: (a) all the drops rotate in the same direction for a given
sense of the field: the right-handed helix has an anticlockwise rotation
when viewed along the field direction. When the voltage is reversed, the
curvature of the dark brushes and the sense of rotation of the structure are
reversed; (b) the angular velocity increases linearly with applied voltage up
to ~ 3.5 V, beyond which the structure of the drop changes and the
rotational velocity becomes a non-linear function of the applied voltage
(fig. 4.4.4); (c) nematic drops do not rotate under the action of E; (d) when
the handedness of the helix is reversed, the angular velocity also reverses
sign for any given sense of the field E; (e) the angular velocity does not
depend on the radius of the drop, showing that it is a rotation of the
structure rather than a rigid body rotation of the drop as a whole. All these
observations are in conformity with the theory. Though the angular
velocity was approximately the same for all drops, some drops which had
dust particles attached to them rotated with a lower velocity. For the sake
of completeness, even these values have been plotted in fig. 4.4.4. The
extrapolated angular velocity becomes zero for V 1.9 V (fig. 4.4.4). The
last point indicates that the DCfieldis totally screened up to 1.9 V and
that the redox potential of at least one of the components in the mixture is
about 1.9 V.
For a defect-free planar structure, the angular velocity in the presence of
an electric field E is in analogy with (4.4.7),

where vE is the electromechanical coupling coefficient. However, in the


actual experiment a line defect at the periphery of the drop also rotates
with the structure. The effective friction coefficient for the slow motion of
a nematic line singularity has been estimated by Imura and Okano(63) and

4.5 Flow properties

267

0.6

OX

0.5 A

0.4 0.3 0.2 0.1

J^

3
Voltage (V)

Fig. 4.4.4. The rotational velocity against applied voltage. The different symbols
denote measurements on different drops. The angular velocity was noticeably less
for drops which accidentally had dust particles attached to them (see the circles in
thefigure),but only the data for the fastest rotating drops were considered in the
calculations. Between 3.5 and 5 V there was visible disturbance within the drop and
measurements were not possible. At 5 V and above, the drop regained a uniform
texture. (After reference 60.)
by de Gennes.(64) Extending this theory to the case of the /-line of strength
2, it turns out that in the presence of the line defect rotating with the
structure (60)

Using the slope of the linear part of the |o>| against |E| curve (fig. 4.4.4),
and putting Xx = 0.7 P, it was found that |vE| = 0.28 cgs, and vE/q =
2 x 10~5 cgs for the material used in the experiment.
4.5 Flow properties
The flow properties of a cholesteric liquid crystal are surprisingly different
from those of a nematic. Its viscosity increases by about a million times as
the shear rate drops to a very low value(65) (fig. 4.5.1). One of the difficulties
in interpreting this highly non-Newtonian behaviour is the uncertainty in
the wall orientation which cannot be controlled as easily as in the nematic
case. Some careful measurements of the apparent viscosity ^ app in Poiseuille
flow have been made by Candau, Martinoty and Debeauvais (66) of a

268

4. Cholesteric liquid crystals

105

Is

1O4

103

102

Cholesteric

Isotropic

lO"1
I

io- 2

100

110

i l l

120

^
1

130

140

Temperature (C)

Fig. 4.5.1. Apparent viscosity of cholesteryl acetate versus temperature. Capillary


shear rate (s"1): open inverted triangle, 10; cross, 50; open square, 100; filled
inverted triangle, 1000; filled square, 5000. Rotational shear rate (s"1) open
triangle, 104; open circle, high shear rate and normal liquid behaviour. (After
Porter, Barrall and Johnson.(65))
nematic-cholesteric mixture whose pitch could be varied by changing the
composition. Microscopic examination revealed that the helical axes were
oriented radially in the capillary so that the cholesteric layers were rolled
up in the form of coaxial cylinders. The flow direction was therefore
normal to the helical axis at every point. In this geometry there is a slight
dependence of the viscosity on the shear rate and on the pitch, but the
significant fact emerges that //app is approximately of the same order of
magnitude as that for a nematic even at low shear rates (figs. 4.5.2 and
4.5.3).
The application of the Ericksen-Leslie equations to cholesteric flow is
less straightforward than in the case of nematics and no detailed solutions
have so far been possible even for simple geometries. However, the
behaviour in certain limiting situations can be explained qualitatively.

4.5 Flow properties

269

0.30 -

0.28

0.26

0.24

fe)
0.22

0.25

0.50
0.75
Shear rate (103 s"1)

1.00

1.25

Fig. 4.5.2. Apparent viscosity in Poiseuille flow as a function of shear rate. Flow
normal to the helical axis (see text). Pitch (//m) = (a) 1.9, (b) 2.6, (c) 3, (d) 3.9, (e)
6, (/) 9.1 and (g) oo. (After Candau et a/.<66

1.5 V-

1.0

0.5

10

15

20

lAP 2 (10 6 cm- 2 )

Fig. 4.5.3. Threshold of shear rate above which the fluid becomes non-Newtonian
plotted against 1/P2, where P is the undistorted value of the pitch. The arrows
represent the upper and lower limits of the shear rate (see fig. 4.5.2). (After Candau
et

66

270

4. Cholesteric liquid crystals

Fig. 4.5.4. Helfrich's model of'permeation' in a cholesteric liquid crystal. At low


shear ratesflowtakes place along the helical axis without the helical structure itself
moving.
4.5.1 Flow along the helical axis
Helfrich(67) proposed a simple physical mechanism, which he called
permeation, to account for the very high apparent viscosity at low shear
rates. He suggested that flow takes place along the helical axis without the
helical structure itself moving owing to the anchoring effects at the walls
(fig. 4.5.4) and that the velocity profile is flat rather than parabolic. Under
these circumstances, the translational motion of the fluid along the
capillary can be directly related to the rotational motion of the director.
The energy gained by the motion in the pressure gradient should be equal
to that dissipated by the rotational motion. Now the viscous torque
exerted by the director
T = n x g',
where g' is given by (3.3.13). In the absence of velocity gradients this torque
is evidently Xxqv, where v is the linear velocity and q = 2n/P is the
cholesteric twist per unit length. In nematics, Xx < 0; we shall assume this
to be true in the present case also. Thus

where dp/dz is the pressure gradient. The quantity of fluid flowing per
second is
Q = _nR\dp/dz)
where R is the radius of the capillary. Applying Poiseuille's law,
/

/app =

(4.5.1)

Typically, R ~ 500 jum and P = 2n/q - l / / m s o that ?/app - - 10% which


explains the very high viscosity at low pressure gradients.

4.5 Flow properties

271

We shall now show that the essential features of Helfrich's model can be
derived on the basis of the Ericksen-Leslie theory. (68)
Flow between parallel plates
We shall consider flow between two parallel plates, caused by a pressure
gradient. Choosing a right-handed cartesian system such that the plates
occupy x = h, we seek solutions of the form
nx = cos (qz + (p) cos 0, ny = sin (qz + q>) cos #,
nz = sin 0,

vx = 0,

vy = 0,

vz = w,

with 0 = 6(x,z), cp = (p(x,z) and w = w(x). This gives a cholesteric of pitch


P = 2n/q with the helical axis along z for 0 = cp = 0. Strictly speaking, a
general theory should allow for non-zero values of vx and vy but as we shall
see shortly the present approximation is valid except in a negligibly thin
layer very near the boundary.
We consider very low pressure gradients and retain only the first powers
of 0, cp and w. Then
nx = CcpS,

ny S+cpC,

nz = 6,

where C = cos qz and S = sin qz. Neglecting director inertia and product
terms involving w6, w X6 i9 6 XX9, w X6 x etc., (4.3.4) reduces to

ezx(k11-k22s2)-(pxx(k11s+k33sc2)-lp^k22s-d^(k!>3+^2)qsc
-2<p ^k^qC-^wqS+yiC-Sg?)

= 0, (4.5.2)
2S2)q]
= 0,

(4.5.3)

0.

(4.5.4)

In the above equations 9 x = d9/dx, 9 xt = 9 ^/8x9z etc. Similarly, under


the same approximations, (4.3.2) becomes
P,x = -%(*+Hz)+ vAqSCwtX,
py = [M.C'-^ + lfi.iC'-S^qw^

P,z =

(4.5.5)
(4.5.6)

S^ik^-k^qS-cp^ik^S^k^C^q-y^k^q
-^, I (A; 22 + fc33)^C + > ^

+ ^6-/u2)C2].

(4.5.7)

From (4.5.2) and (4.5.3) we get


and from (4.5.7) and (4.5.8)

2 2 + k33)qC-A,

^ - p . , = 0.

Wq

= 0,

(4.5.8)
(4.5.9)

272

4. Cholesteric liquid crystals

We make a 'coarse-grained' approximation and replace |[// 4 + (JU5JU2) C2]


by an average value if and rewrite (4.5.9) as
nw.xx + n*i42-P.z = 0.

(4.5.10)

A solution of (4.5.10) with the boundary conditions w(h) = 0 is(68)


w(x) = T-i2 l
rrr '
(4.5.11)
^
\
cosh Kh)
where K2 = Xx q2/ff. The velocity is symmetric about x = 0. The amount
of fluid flowing per second in the z direction is
Q=

w(x)dx = 2

w(x)dx

Hence the apparent viscosity


/app

3Q

3[1

Taking 2h = 100 jum and P= 1 jum, the velocity attains 0.99 of the
maximum value within a thickness of 0.5 //m of the boundary. Thus in all
practical situations the velocity profile is flat over most of the region
between the plates and
*

a p p

- ^

(4.5.14)

which is the analogue of (4.5.1).


Poiseuille flow
In cylindrical polar coordinates, we seek solutions of the form
nr = cos(qzy/ + <p)cos8, n = sin(qzy/ + (p),
nz = sin 0,
where 9 and (p are functions of r, y/ and z. Considering very small pressure
gradients we obtain to a first order in 6 and (p,

where C = cos(qzy/) and S = sin(qzy/). For the velocity field we


assume
vr = 0,

v = 0,

vz = w(r).

4.5 Flow properties

273

Proceeding as before we obtain


p z = \w rr [// 4 + (ju5 - /z2) C2]
^

(4.5.15)

Again replacing the coefficients of wrr and w r by an average value


Wrr +

IWr

0.

(4.5.16)

A well behaved solution of (4.5.16) is(60)

where A is a constant, / 0 is the modified zero order Bessel function of the


first kind and K2 = Xx q2/fj. Using the boundary condition w(R) = 0,
A =

~ffIQ(KRY

and

where R is the radius of the capillary. The quantity of liquid crystal flowing
per second

KI0{KR)\
where Ir is the modified first order Bessel function of the first kind.
Hence
/aPP

8{ x

[2ii(KR)/KRI0(KR)]}'

Again, in practical situations the velocity profile is flat except very near the
boundary and

which is identical with (4.5.1). Thus Helfrich's idea of permeation along the
helical axis of a cholesteric can be justified in terms of the Ericksen-Leslie
equations.

274

4. Cholesteric liquid crystals


4.5.2 Flow normal to the helical axis

The general theory of shear flow normal to the helical axis has been
discussed by Leslie.(69) An interesting feature that comes out of this
analysis is that a shear in the xz plane can give rise to secondary flow along
y. (See 3.6.5; in principle secondary flow should occur in the Helfrich
configuration also, but only in a negligibly thin layer very near the
boundary where the velocity profile is not flat.) We shall now present a
simplified version of Leslie's theory ignoring thermomechanical coupling.
Consider a cholesteric film between two plane parallel plates, one of
which is moving with constant velocity V in its own plane. The plates
occupy the planes z = h. We examine solutions of the form
nz = cosOcosp,
vx = u(z),

ny = cos 6 sin <p, nz = sin0,


vy = v(z),

vz = 0,

(4.5.17)
(4.5.18)

where 6 = 6{z) and (p = cp{z). Then, tzx = a (constant shear), tzy = 0 and
tzz = p (an arbitrary constant). Using (4.3.7) and (4.3.13), we get
(H1 + H2 cos 2 cp) + H2 rj sin cp cos cp = a,

(4.5.19)

H2 sin cp cos cp + (H1 + H2 sin 2 (p) rj = 0 ,

(4.5.20)

where
2 = dw/dz,
and

2r] = dv/dz,

H2 = (2//x sin2 6 + jus + ju6) cos2 6.

From (4.3.4) we obtain


d20
dz2

XdFJddY
2 dO \dzj

\dF2(d(p\2
2 d6 \dz)

Ay
dz

h A2 cos 20) (<!; cos p + iy sin p) = 0,

(4.5.21)

and
i ^ ^ 2 uifl 1/COS (/~

sin (p-rj cos (p) = 0,


where

(4.5.22)

F = kxl cos2 6 +fc33sin2 6,


F2 = (k22 cos2 9 + &33 sin2 0) cos2 0.

From the symmetry of the problem it is clear that 6 and v should be even

4.5 Flow properties

275

functions of z, while u f Kand cp are odd functions of z. Equations (4.5.19)


and (4.5.20) yield
f = 41 +(// 2 /// 1 )sin>]/(i/ 1 + // 2 ),
rj = -aKHJHJsinvcosvy^

(4.5.23)

+ HJ,

(4.5.24)

which immediately give the velocity profiles


II = 2

P <Jdz, r = - 2

77 dz.

(4.5.25)

It is seen that v 4= 0 even though the shear is confined to the zx plane; in


other words, secondary flow occurs.
Using (4.5.23) and (4.5.24), (4.5.21) and (4.5.22) can be simplified to
U

1 Uij

I KXIS 1

1 VJ.J.

I \J.$

\dz/
d^?
2

dz

= 0 (4.5.26)

and
2

dz2

d# dz dz

2 sin#cos#-;al

dz

= 0,

(4.5.27)

where
Q = ( ^ _|_ ^ 2 cos20)/(/f 1 + /f2)
and
P = (A 2 -.^ s i n ^ c o s ^ / ^ .
Leslie assumed the following boundary conditions:

fd<p\
dz)h

/ d ^ fc2 }
\dz)_h

(4.5.28)

k22

Now
V=4\
Jo

and the apparent viscosity


17app

_ a
~ K/2A
2

(4.5.29)

4. Cholesteric liquid crystals

276

1.2

l.l

1000

100

Fig. 4.5.5. Theoretical variation of the apparent viscosity //app with pitch P = 2n/q
for flow normal to the helical axis of a cholesteric (or twisted nematic) at low shear
rates. Plot ofrjapp(q)/rjapp(0) versus P for twisted PAA. The separation between the
plates = 100 jum. The horizontal dashed line corresponds to f] (co)/rfapp(0). (After
ref. 70.)

3.7

3.3

2.9

T
2.5
'50

51

52
T(C)

53

54

Fig. 4.5.6. Evidence of oscillatory behaviour in the variation of the apparent


viscosity with temperature in a cholesteryl chloride-cholesteryl myristate mixture
in the neighbourhood of the nematic point 7^ (see fig. 4.1.16). (After Bhattacharya,
Hong and Letcher.<72))

4.6 Distortions of the structure by external

fields

277

Detailed analytical and numerical calculations have been presented by


K i n i

(7o,7i)

i s

It is seen that for pitch values of the order of the sample thickness, ?/app
should exhibit oscillatory behaviour with varying pitch because of the term
sin qh/qh in (4.5.30). A representative theoretical curve is presented in Fig.
4.5.5. This prediction has been verified qualitatively. (72) Measurements of
the apparent viscosity on the cholesteryl chloride-cholesteryl myristate
mixture, whose pitch, as we have seen earlier, is sensitive to temperature
(fig. 4.1.16), showed evidence of oscillations as the temperature was varied
(fig. 4.5.6).
Often, in practical cases, qh is so large that sin qh/qh is negligibly small
and ?/app approaches a maximum limiting value which is independent of the
pitch or gap width. This value is several orders of magnitude less than that
for flow along the helical axis and is comparable to that for a nematic.
When a is sufficiently large
q> = 0

and

0 = 00 = i c o s - 1 ^ / ^ ) ,

i.e., the helix is unwound completely except in a layer of thickness of the


order of the q~x at the boundaries, and

which is again independent of the pitch or gap width. This lower limit is
reached when a & k22 q2 or more. All these predictions are in qualitative
agreement with the observations of Candau et #/.(66) (see fig. 4.5.2).
4.6 Distortions of the structure by external fields
4.6.1 Magnetic field normal to the helical axis: the cholesteric-nematic
transition
When a magnetic field is applied at right angles to the helical axis of an
unbounded cholesteric liquid crystal composed of molecules of positive
diamagnetic anisotropy (xa = X\\~X > 0) the structure gets distorted as
illustrated schematically in fig. 4.6.1. As the field strength approaches a
certain critical value Hc the pitch increases logarithmically; for H > Hc the
helix is destroyed completely and the structure becomes nematic. (73) The
dependence of the pitch on the field strength was calculated by de
Gennes(74) and by Meyer.(75)

4. Cholesteric liquid crystals

278

>H

>H

H=0
0<H<Hc
H>HC
Fig. 4.6.1. Schematic representation of the effect of a magnetic fieldapplied normal
to the helical axis of a cholesteric liquid crystal composed of molecules of positive
diamagnetic anisotropy. For H > Hc the cholesteric is transformed into a nematic.

Taking the helical axis to be along z, H = (H, 0,0) and n = (cos cp, sin q>,
0), the free energy of the system is

= ((Fdz = 1 J[(^f-

-/aH2

sin2 J d p + constant,

where q0 = 2n/P0, Po being the pitch of the undistorted structure in the


absence of a magnetic field. The equation of equilibrium is therefore
X H sin

* *

which yields

COS 9 =

z = 2AZK(A),

where
22/

and

72

K(A) =
is the complete elliptic integral of the first kind; A is a constant which can

4.6 Distortions of the structure by external fields

279

1.5

1.0

H/Hc

Fig. 4.6.2. Dependence of the pitch P on the magnetic field strength H in PAA
mixed with a small quantity of cholesteryl acetate. Curve represents the theoretical
variation predicted by de Gennes's equation (4.6.1). (After Meyer.(76))

be determined from the condition that J^ should be a minimum; z = P/2


the half-pitch of the distorted structure. It is assumed that the sample is
sufficiently thick for boundary effects to be neglected. Therefore

where

=2

rnr.

Jo

4
?)2 dcp =
A

and

E(A) =
is the elliptic integral of the second kind. The condition
to the relations

= 0 leads

280

4. Cholesteric liquid crystals


H

V///////////////////////A

W7/7///////////////////.
2*

Fig. 4.6.3. Deformation of a planar structure due to a magnetic field acting along
the helical axis of cholesteric liquid crystal composed of molecules of positive
diamagnetic anisotropy. A similar deformation superposed in an orthogonal
direction results in the square-grid pattern (see fig. 4.6.4). (Helfrich. (79))

Putting z0 = n/q0 = |P 0 , we have


(4.6.1)
and

2E(A)'
When

(4.6.2)

l, E(A) -> 1, K(A) -> oo and H-> Hc, so that h = n/2 or

Hc = Inqo(k22/XJK

(4.6.3)

which is the critical field at which the structure becomes nematic.


The variation of pitch with magnetic field strength predicted by (4.6.1)
has been verified experimentally (76'77) (fig. 4.6.2). It has also been confirmed
that Hc is inversely proportional to Po the pitch of the undistorted
structure.(77) It turns out that with the usually available magnetic field
strengths, the experiment is conveniently performed only with cholesterics
of relatively large pitch. For example, in a typical measurement using
nematic PAA doped with a small amount of cholesteryl acetate Hc was
8.3 kG for PQ = 26 jum.

4.6 Distortions of the structure by external

fields

281

4.6.2 Magnetic field along the helical axis: the square grid pattern
We next examine the effect of a magnetic field acting along the helical axis
of a cholesteric film having a planar texture. If / a > 0 and boundary
constraints are absent, there is a possibility of a 90 rotation of the helical
axis because \{X\\+X) > X- I o n t n e other hand, boundary effects are
such as to maintain the orientation of the helix, an expected type of
deformation is for the director at every point to be tilted towards the field,
i.e., a conical distortion/ 75 ' 78) However, it was pointed out by Helfrich(79)
that yet another type of deformation can set in, viz, a corrugation of the
layers (fig. 4.6.3). This has since been confirmed experimentally using both
magnetic(80) and electric fields/81'82) and takes place at a much lower
threshold. It results in the so-called square grid pattern (fig. 4.6.4), the
theory of which was first proposed by Helfrich(79) and subsequently
elaborated by Hurault. (84) We shall discuss first the magnetic field case.
For the unperturbed cholesteric
n = (cos# 0 z,sin# 0 z,0),
where we take z normal to the film. For small perturbations we may put
nx = cos(q0z + (p) cosq 0 z cpsinq0z,
ny = sin (q0 z + (p) sin q0 z + q> cos q0 z,

(4.6.4)

nz = 9 cos qoz.
Substituting in (4.2.1) the local energy density

M\

?>d(

d<p\

+A: 33 cos 4 ?0 z^J j ,

^ + smqozcosqozJ

(4.6.5)

and the total energy


#- =

\FdV.

Now
-^-=

\AdV

and

^=

\BdV,

(4.6.6)

282

4. Cholesteric liquid crystals

Fig. 4.6.4. The square grid pattern in a cholesteric liquid crystal induced by (a) a
magnetic field and (b) an electric field. (Rondelez.(83))

4.6 Distortions of the structure by external


where

(ffm

fields

283

dQ

A = - (fc u sin 2 q0 z + fc33 cos 2 q0 z) ( ^ + q0

dx

(4A7)

and

8>

Aoz

c)2Q
rvn wS 2 ^ 0 Zr.
oz

(4.6.8)

For a given #, ^ is a minimum when cU^/8^ = Oov A = 0. We consider the


perturbations 6 and ^ to be dependent on z and x (where x is any arbitrary
direction in the plane of the layers) and write them in the form
6 = 00 sin kx x cos kz z,

(4.6.9)

<p = (<p0 + cpx c o s 2q0 z) c o s kx x sin kz z,

where kz = mn/d, d being the film thickness and m an integer. We shall


confine our discussion to m = 1. In practice,
K<kx<<lo

(4.6.10)

and we shall assume this to be true in what follows. The condition


= 0 yields
- ^ i i ) ^ > i = 0, (4.6.11)
Kl^ql <P, - q0 K 0O)+K*S3 - *n) ( ^ ^o - q K e0)+ife+*n)

/c2 ^ = o,
(4.6.12)

where we make a 'coarse-grained' approximation, i.e., take into consideration only the slowly varying parts of A. The minimum energy density
O 00 ~ K <Po) (*11 + ^

-ko K <Pi(2k22+^33 - ^11)] cos K z s i n K x-

( 4 - 6 - 13 )

284

4. Cholesteric liquid crystals

From (4.6.11) and (4.6.12) we get


^i^Z^o

(4.6.14)

and
ffo ^o ~ \i(ic

+L. \ + k ^\(k

+k } -

(k

k ^*1* ~ 1

because of (4.6.10). Therefore

Thus, in the coarse-grained approximation the energy density becomes


\

(4.6.16)

An alternative derivation of (4.6.16) has been given by de Gennes. We


consider the cholesteric to be a quasi-layered structure and write the energy
density in terms of the displacement u(r) of each plane in the following
form:

where B is an elastic coefficient associated with the compression of the


layers. Terms involving (du/dx)2 and (du/dy)2 are not included as they
correspond to a uniform rotation of the layers and do not contribute to the
free energy. Introducing a unit vector h along the helical axis, (4.6.17) may
be expressed as
(

where P is the local value of the pitch. We know that the twist energy of
deformation of the cholesteric structure can be written in terms of the pitch
as \k22(q q0)2 where q = 2n/P and q0 = 2n/P0. Comparing this with the
first term on the right-hand side of (4.6.18), it is clear that
B*k22q*.

(4.6.19)

In order to evaluate K, one may use the following argument. Suppose that

4.6 Distortions of the structure by external

fields

285

the film is rolled up into a cylinder, i.e., the screw axes are oriented radially
about the cylinder axis and the layers form coaxial cylinders. In cylindrical
coordinates, the components of the director are now
nr = 0,

riy = cos 0(r),

and

nz = sin 6{r)

and the local free energy is given by


sin0cos0\ 2

1f

cos 4 0

^-^^-

(4-6.20)

The optimum value of 6{r) compatible with the periodicity 6{r) = 6{r + Po)
corresponds to
d0
1 . .
.
- = # n + - s m 0 c o s 0 .
dr
r
Averaging over cos4 0,
K = \k33,

(4-6.21)

so that

We may take u as our variable and write


U UQ COS kx x cos kz z,

(4.6.23)

where kz = n/d. It is seen at once that (4.6.22) becomes equivalent to


(4.6.16) if we replace 0 by kx u.
In the presence of a magnetic field applied normal to layers the total free
energy
where

()V

(4-6.24)

^ a = ^ ~x being the anisotropy of diamagnetic susceptibility. Applying


the condition dF/du = 0 and using (4.6.22), (4.6.23) and (4.6.24)
= 0.

(4.6.25)

Thus H^oo when kx^co


as also when kx->0. This is because the
perturbation u is made up of two components, bend and twist: kx -> oo

286

4. Cholesteric liquid crystals

excites the bend mode while kx -> 0 excites the twist mode whose amplitude
diverges in the limit. The optimum wavevector corresponds to an
admixture of both modes and is given by
8

2 / 2

tA<LK\

kx = ^ q\ k\,

7 4

^ 2 2

(4.6.26)

or
kxK(Pod)-K

(4.6.27)

and the threshold field


1

H^ = (6k 22 k 33 ) 2 q 0 k z ,

(4.6.28)

77HocOP0 </)-".

(4.6.29)

/a

or
It is interesting to note that this threshold field is lower than that for a
conical distortion or that for cholesteric-nematic (unwinding) transition.
For a conical distortion, the theory is closely similar to that discussed in
3.4.2 and has been treated by Leslie (78); the critical field is given by

HI = (k, J
/a

For the cholesteric-nematic transition, we have from (4.6.3)


G ~

^ 2 2 V/0>

Aa

and in view of (4.6.10), 77H is much less than HF or HG.


The experimentally observed square grid pattern corresponds to two
such distortions, which are orthogonal.

4.6.3 Electric field along the helical axis


Electric field effects are more complicated because of conduction. (84) The
Carr-Helfrich instability, which occurs in nematics of negative dielectric
anisotropy (see 3.10.2), may be expected to take place in this case too,
only the bend and twist distortions are now coupled. Moreover, the fluid
motion along z can occur only by the process of permeation (4.5.1).
We shall consider the DC case first (neglecting, of course, any charge
injection at the boundaries). If al]h and alh are the conductivities along and

4.6 Distortions of the structure by external

fields

287

perpendicular to the helical axis (which, as before, is taken to be parallel to


z), the electric current
Jx = *Ex-(\\*-Ort)E^

(4.6.30)

which should be zero (V-J = 0). Here Eo is the applied field and Ex that
caused by the Carr-Helfrich mechanism. Therefore

OX

(Tl

The charge density is given by

, c 4v
( E) =
An

(4.6.32)

An e.

and the electric force


fc=PeE0.

(4.6.33)

The dielectric torque

U^

An -\E0

from which it follows that the contribution to the vertical force is


ox

<+n alh ox

The net electric force


/elec

/c ' /diel-

From (4.6.22) the elastic restoring force

The threshold field is obtained by setting / e l e c +/ e l a s = 0. The optimum


wave vector kx is given by (4.6.26) and the thresholdfieldby
/72 ^^th

~~

-*-h r3/r ^ "jirp //V 1

_87T 3 <7 | | +a 1

^2^22 ^ 3 3 /

V^0u/

)-\

(4.6.34)

where crn and aL are the conductivities parallel and perpendicular to the

288

4. Cholesteric liquid crystals

preferred molecular direction and el{ and e are similarly defined. It may be
noted here that we have neglected the contribution of the viscous torque
altogether. This is because, as remarked earlier, fluid motion takes place
only by permeation, and, moreover, the distortions are infinitesimal and of
such long wavelength {kx <4 q0) that the effect of shear flow will be very
small indeed. The dependence of the spatial periodicity of the pattern and
the threshold field on d and Po is similar to the magnetic field case, and is
borne out by experiments/ 81 ' 8285)
Hurault(84) has extended the treatment to AC fields. The method is
somewhat analogous to the one discussed in 3.10 for nematic instabilities
and leads, in the conduction regime, to the following threshold for
distortion:
_8n3e]]+s

1+Q) 2 T 2

^(PdY1

(4 6 35)

where

and T is the dielectric relaxation time given by

For negative dielectric anisotropy (j < 0) the conduction regime


occurs when co < coc, where

These results are also in quantitative agreement with observations.(85)


However, the behaviour at higher frequencies does not appear to be fully
understood.
In the case of dielectrically negative molecules, the square grid pattern
changes as the voltage is raised above the threshold and the helical axis
rotates by 90. (The tilted structure is metastable and relaxes to the planar
texture if the field is switched off but only after a long time.) At even higher
voltages turbulence sets in and the system goes over to the dynamic
scattering mode. On switching off, the liquid crystal relaxes to the focal
conic texture and the scattering persists. This has been described as the
storage mode or the memory effect in cholesteric liquid crystals.(86) An

4.7 Anomalous optical rotation

289

audio frequency (~ 10 kHz) pulse then restores it to the planar texture. In


combination with a photoconductor, this effect has been made use of to
construct image storage panels. (87)
4.7 Anomalous optical rotation in the isotropic phase
We have seen in 2.5 that the orientational correlations between the
molecules give rise to certain remarkable pretransitional effects in the
isotropic phase of nematic liquid crystals. Similar correlations exist in
the cholesteric phase as well, except that by virtue of the chiral nature of the
interactions the local order lacks a centre of inversion. Cheng and Meyer (88)
discovered that these correlations give rise to an enhancement of the
optical activity in the isotropic phase. The correlation length increases as
the temperature approaches the isotropic-cholesteric transition point, the
local helical ordering builds up and the optical rotation increases
accordingly. The magnitude of this effect is just barely observable in most
cholesteric materials as the anisotropy of molecular polarizability of these
compounds is usually rather small. For this reason Cheng and Meyer used
a specially synthesized nematogen with an optically active end group: pethoxybenzal-/?'-(/?-inethylbutyl)aniline. This molecule forms a cholesteric
phase and has the advantage of having a high anisotropy. Cheng and
Meyer's experimental results are shown in fig. 4.7.1. The natural optical
activity of the molecule (in the absence of correlations), determined by
measuring the rotatory power of dilute solutions of varying concentration
and extrapolating to 100 per cent concentration, was just about 1 cm" 1 as
compared with a total rotation of nearly 40 cm" 1 close to the transition,
proving that correlations play the predominant role.
The theory of Cheng and Meyer is rather elaborate and will not be
discussed in detail here. We shall merely indicate the major steps in the
calculations. We consider a system of identical molecules and for simplicity
neglect the contribution of the natural optical activity to the total rotation.
If Eo is the externally applied field, the net field F, acting on a molecule at
x, is
F(x) = E o + T P ( x ) G(x - x', k0) d V ,
J V

where P(x) = 7Va(x) F(x) is the polarization (or the dipole moment per
unit volume) at x, a is the polarizability of the molecule at x, G(x x', k0)
a tensor representing the field at x due to a dipole at x', k0 the wavevector
of the incident radiation, v the volume of the Lorentz cavity which is not
supposed to contribute to the effective field and V the total volume.

4. Cholesteric liquid crystals

290

35

40
35
30
25

I 20
o
"S3

60.5

15

60

65

70

75

80

61.0 61.5 62.0 62.5 63.0


Temperature (C)

85

90

95

100

105

63.5

110

115

Temperature (C)

Fig. 4.7.1. Anomalous optical rotation in the isotropic phase of a cholesteric liquid
crystal. Open and closed circles are measurements on two different samples
appropriately normalized. Cholesteric-isotropic transition temperature 60.57 C.
X = 0.6328 /an. (After Cheng and Meyer.(88))

Writing

) + <SP(x).

where Po is the polarization in the absence of correlations and SP


a small correction term,

f dV<<5a(x)-G-<Ja(x')>-F(x'),

J V

neglecting higher powers of SOL. Assuming a Lorentz-Lorenz type of


relationship for the polarization field in the medium, expressing Sen in terms
of the tensor order parameters s (see, for example, 2.3.1) the susceptibility
d3/?G(R, k0) exp (mk0 R)<s(0) s(R)> ,
where n is the refractive index of the isotropic phase, a the mean molecular
polarizability and aa = a,, a the polarizability anisotropy. Expressing
the integral in terms of the Fourier components,
0

J.

+ q) <s*(q) s(q)>

4.7 Anomalous optical rotation

291

Hence the dielectric tensor may be expressed as

where A is a tensor of the form

In order that the medium be non-absorbing we must have

so that for an isotropic medium

8 =

-iA;;
L-iAL

S+A;,

-iA;' 2

iA;; .
e + A^

Such a system will exhibit circular birefringence,

or an optical rotation

The rotation exists because of the correlations (s%asyfiy which are nonvanishing in the case of a cholesteric. The averages may be evaluated on
the basis of de Gennes's model(89) (see 2.5). To allow for the noncentrosymmetric ordering in the cholesteric, we include an additional term
of the form s V x s in the free energy of the isotropic phase
/? c = /5 N + 2 t f 0 L ' ^ ^ - i f e ,
oxy

(4.7.1)

where FIN is the free energy per unit volume in the isotropic phase of a
nematic given by (2.5.15) and (2.5.22), q0 is a pseudo-scalar and L is a
constant. The average can then be worked out. For example

292

4. Cholesteric liquid crystals

BPII

Disordered
-

BPIII

* = = --

BPI

Helicoidal

Inverse pitch

Fig. 4.8.1. Schematic phase diagram, showing the three experimentally observed
blue phases (BP).

where

A = a(T-T*),

l\ = LJA,

and L l 5 L 2 are the constants occurring in (2.5.22). Therefore the optical


rotation increases rapidly as the temperature approaches T*.
4.8 The blue phases

The blue phases occur in cholesteric systems of sufficiently low pitch, less
than about 5000 A. They exist over a narrow temperature range, usually
~ 1 C, between the cholesteric liquid crystal phase and the isotropic liquid
phase (see (1.3.5)). The first observation of a blue phase was described by
Reinitzer(90) himself in his historic letter to Lehmann as follows: 'On
cooling (the liquid phase of cholesteryl benzoate) a violet and blue
phenomenon appears, which then quickly disappears leaving the substance
cloudy but still liquid.' Although Lehmann (91) recognized it as a stable
phase, not until the 1970s was it generally accepted that the blue phases are
thermodynamically distinct phases. The nature of these phases has now
become a subject of considerable interest to condensed matter physicists.
Fig. 4.8.2. (a) Faceted single crystals of BP I in equilibrium with the isotropic phase
(58 wt% mixture of cyano-4-methylbutylbiphenyl (CB15) in nematic ZLI 1840 of
Merck). (From Cladis, Pieranskio and Joanicot(95).) (b) Optical Kossel diagram of
BP I, {110} direction, X = 5290 A (42.5 wt% mixture of CB15 in nematic E9 of
BDH). (From Cladis, Garel and Pieranski.(96))

293

(a)

Fig. 4.8.2. For legend see facing page.

294

4. Cholesteric liquid crystals

Fig. 4.8.3. Unit cells of BP disclination lattices. O2 is simple cubic, O5, O 8 + and
O8 are body-centred cubic. The tubes represent disclination lines whose cores are
supposed to be isotropic (liquid) material. (From Berreman.(98))
Three distinct blue phases have been identified: BP I, BP II and BP III,
occurring in that order with increasing temperature (fig. 4.8.1). All of them
are optically active but isotropic. (They may have colours other than blue,
but are still referred to as blue phases). From observations of optical Bragg
reflexions(92) and other studies, it is found that BP I is a body-centred cubic
lattice (crystallographic space group I4X32 or O8), BP II a simple cubic
lattice (P4232 or O2) and BP III probably amorphous. The suggestion has
been made that BP III, called the blue fog, may be quasi-crystalline.(93)

4.8 The blue phases

295

Fig. 4.8.4. A representative phase diagram calculated from the Landau theory.(110)
The normalized ordinate t is proportional to T T* and the normalized abscissa K
is proportional to 1/P, where P is the cholesteric pitch. The diagram illustrates how
different BPs can occur by changing the chirality. (From Crooker.(104))
Striking confirmation of the cubic structures of BP I and BP II was
obtained by Onusseit and Stegemeyer(94) and others, who succeeded in
growing beautiful single crystals of up to a few hundred microns in size(95)
(fig. 4.8.2(#)). Optical Kossel diagrams, analogous to the Kossel lines
observed in X-ray diffraction from crystals, have confirmed their symmetry (96) ( f i g 4.8<2(6)).
Topologically, it turns out that the helical structure of the cholesteric
cannot be deformed continuously to produce a cubic lattice without
creating defects. Thus BP I and BP II are unique examples in nature of a
regular three-dimensional lattice composed of disclination lines. (97) Possible unit cells of such a disclination network, arrived at by minimizing the
Oseen-Frank free energy, are shown in fig. 4.8.3.(98) The tubes in the
diagram represent disclination lines, whose cores are supposed to consist
of isotropic (liquid) material. Precisely which of these configurations
represents the true situation is a matter for further study.
The occurrence of the BPs can also be described in terms of the Landau
theory. (99103) The free energy expansion contains the usual nematic terms

296

4. Cholesteric liquid crystals

and an additional term of the form s - V x s to allow for the noncentrosymmetric ordering (see (4.7.1)). The order parameter is expressed as
a Fourier series, and by choosing the appropriate (hkl) reciprocal lattice
vectors for the Fourier components and minimizing the free energy, one
can generate phase diagrams. A representative phase diagram illustrating
how the different phases can occur by changing the chirality is given in fig.
4.8.4. Calculations show that there can be stable body-centred cubic and
simple cubic structures, and in the presence of an electric field an hexagonal
phase as well. All these predictions are in broad agreement with the
experimental facts. There has been a surge of activity on various aspects of
the BP problem - phase stability and its dependence on pitch, phase
diagrams, electric field effects, single crystal morphology, growth rate,
mechanical properties, etc.(104)

4.9 Some factors influencing the pitch


In this section, we present a brief survey of experimental studies on the
dependence of the pitch on temperature, composition, etc.
Dependence of pitch on temperature: applications to thermography
In most pure cholesteric materials, the pitch is a decreasing function of the
temperature. An elementary picture of the temperature dependence of the
pitch can be given in analogy with the theory of thermal expansion in
crystals.(105) Assuming anharmonic angular oscillations of the molecules
about the helical axis, the mean angle between successive layers

where A is the coefficient of the cubic anharmonicity term, (106) a>0 the
angular frequency and / the moment of inertia of the molecule. Thus the
pitch P (oc 1/(6}) may be expected to decrease slightly with temperature.
However, in many substances the rate of variation is extremely high. It is
now established that if the cholesteric phase is preceded by a smectic phase
at a lower temperature, the pitch increases very rapidly as the sample is
cooled to the smectic-cholesteric transition point (see fig. 5.5.3).
The strong temperature dependence of the pitch has practical applications in thermography, as was first demonstrated by Fergason. (107108)
The material has to be so chosen that the pitch is of the order of the
wavelength of visible light in the temperature range of interest. This is
achieved by preparing suitable mixtures. Small variations of temperature

4.9 Some factors influencing the pitch

297

are shown up as changes in the colour of the scattered light and can be used
for visual display of surface temperatures, (109) imaging of infrared(110) and
microwave(111) patterns, etc.
Dependence of pitch on pressure
Pollmann and Stegemeyer(112) investigated the effect of pressure on the
pitch of cholesteryl oleyl carbonate (COC) mixed with cholesteryl chloride
and found that the pitch increases very rapidly with pressure, the effect
being more pronounced the greater the concentration of COC. This
appears at first quite surprising, but, in fact, the explanation is straightforward.(113) We have emphasized that the pitch diverges as the temperature
approaches the cholesteric-smectic transition point. Now pure COC
exhibits a smectic A phase below 14 C at atmospheric pressure. This
temperature may, of course, be somewhat lower in the case of the mixture.
However, as the pressure is raised the temperature of transition goes
Updi4,ii5) s o ^ a ^- t k e p^ch at room temperature may be expected to rise
accordingly.
Mixtures: dependence of pitch on composition
We have seen in 4.1.6 that a mixture of right- and left-handed cholesterics
adopts a helical structure whose pitch is sensitive to temperature and
composition. This result was first described by Friedel. (116) For a given
composition, there is an inversion of the rotatory power as the temperature
is varied, indicating a change of handedness of the helix. The inverse pitch
exhibits a linear dependence on temperature, passing through zero at the
nematic point where there is an exact compensation of the right- and lefthanded forms (fig. 4.1.16).
A similar reversal of handedness takes place as the composition is
varied.(117) The inverse pitch shows a nearly linear relationship with
composition around the nematic point, but there are significant departures
when one of the components has a smectic phase at a lower temperature. (118) The anomaly may again be attributed to smectic-like shortrange order.
It is well known that a nematic liquid crystal readily adopts a helical
configuration if a small amount of a cholesteric is added to it. For low
concentrations of the cholesteric, the inverse pitch is a linear function of
the concentration, but at higher concentrations the linear law is not
obeyed.(119) The curve attains a maximum at a certain concentration,
beyond which it decreases. It turns out, for example, that the twist per unit
length of pure cholesteryl propionate is actually less than that of its

298

4. Cholesteric liquid crystals

mixture with a small amount of MBBA. Saeva and Wysocki (120) have
observed that a compensation occurs even in an MBBA + cholesteryl
chloride mixture when the MBBA concentration is about 30 per cent by
weight. They suggest that certain optically inactive molecules may become
chiral in the helical environment of a cholesteric liquid crystal. In this case,
the MBBA in the mixture would appear to have a chirality opposite to that
of the cholesteryl chloride.
A small quantity of a non-mesomorphic optically active compound may
also transform a nematic into a cholesteric. (121) However, the handedness
of the helix does not seem to be directly related to the absolute
configuration of the solute molecule, as has been shown by Saeva. (122)
For example, (S)-s-amyl-/?-aminocinnamate and (S)-2-(octyl)-/?-aminocinnamate, both of which have the same absolute configuration, result
in helices of opposite senses when dissolved in MBBA. Impurities, indeed
even the vapour of an organic liquid coming into contact with a
cholesteric,(108) have a profound influence on the pitch.
4.10 Molecular models
The first attempt to develop a statistical model of the cholesteric phase was
by Goossens(123) who extended the Maier-Saupe theory to take into
account the chiral nature of the intermolecular coupling and showed that
the second order perturbation energy due to the dipole-quadrupolar
interaction must be included to explain the helicity. However, a difficulty
with this and some of the other models that have since been proposed <124)
is that in their present form they do not give a satisfactory explanation of
the fact that in most cholesterics the pitch decreases with rise of
temperature.
The local order in a cholesteric may be expected to be very weakly
biaxial.(125) The director fluctuations in a plane containing the helical axis
are necessarily different from those in an orthogonal plane and result in a
'phase biaxiality'. (126) Further, there will be a contribution due to the
molecular biaxiality as well. It turns out that the phase biaxiality plays a
significant role in determining the temperature dependence of the pitch.
Goossens(127) has developed a general model taking this into account. The
theory now involves four order parameters; the pitch depends on all four
of them and is temperature dependent. However, a comparison of the
theory with experiment is possible only if the order parameters can be
measured.
Interesting evidence in regard to the factors responsible for the helical

4.10 Molecular models

299

molecular arrangement has been reported by Coates and Gray.(128) They


have demonstrated that a hydrogen-nleuterium asymmetry in a molecule is
sufficient to produce a cholesterogen. An example is given below:
O

II

N = C Q C H = N C H = C H C O C CH 2 CH 2 CH 3

X
X = H, nematic; X = D, cholesteric.
The CH and CD bond lengths are the same, 1.085 A,(129) when one
takes into account the anharmonicity of the vibrations. Thus it would seem
that steric effects are not essential for the helical arrangement.

5
Smectic liquid crystals

5.1 Classification of the smectic phases


Present classification of smectic liquid crystals is based largely on the
optical and miscibility studies of Sackmann and Demus.(1) The miscibility
criterion relies on the postulate that two liquid crystalline modifications
which are continuously miscible (without crossing a transition line) in the
isobaric temperature-concentration diagram have the same symmetry and
therefore can be designated by the same symbol. It is not clear whether this
criterion is valid regardless of the differences in the molecular shapes and
dimensions of the two components, but empirically Sackmann and Demus
have found that in no case does a phase of a given symbol mix continuously
with a phase of another symbol. The method is simple and has been used
for the identification of a number of new phases, but, of course, it does not
throw light on the precise nature of the molecular order in these phases.
Systematic X-ray investigations have been carried out during the last
decade at several laboratories, and particularly with the availability of
synchrotron X-ray sources, considerable progress has been made in
elucidating the structures/ 2 ' 3)
The notation of Sackmann and Demus is according to the order of the
discovery of the different phases and bears no relation to the molecular
packing. The broad structural features of these phases are summarized in
table 5.1.1. A more detailed description of these structures may be found
in the excellent reviews by Pershan(2) and by Leadbetter.(3)

300

5.7 Classification of the smectic phases

301

Table 5.1.1. Structural classification of smectic liquid crystals


Smectic A(SA)

Liquid-like layers with the molecules upright on the average


(fig. 1.1.5(a)); negligible in-plane and interlayer positional
correlations. Thus the structure may be described as an
orientationally ordered fluid on which is superimposed a onedimensional density wave. A number of polymorphic types of
smectic A have been discovered (see 5.6).
Two distinct types of smectic B have been identified: {a)
Smectic B(SB)
Crystal B - three-dimensional crystal, hexagonal lattice with
upright molecules. Though the structure has threedimensional long-range positional order, the interlayer
ordering is extremely weak energetically because of the weak
interlayer forces, (b) Hexatic B - stack of interacting 'hexatic'
layers with in-plane short-range positional correlation,
negligible interlayer positional correlation and long-range
three-dimensional six-fold bond-orientational order (see fig.
5.7.1). Here, the term 'bond' signifies the line joining the
centres of mass of the nearest neighbours.
Liquid-like layers, as in SA, but with the molecules inclined
Smectic C(SC)
with respect to the layer normal (fig. 1.1.5 (/?)).
Smectic C*(SC*) Chiral S c with twist axis normal to the layers.
Cubic lattice with about 103 molecules per unit cell. (46) The
Smectic D(D)
detailed molecular arrangement is not known, but is generally
assumed to be of the micelle type. Thus this phase should
probably be labelled as ' D ' rather than 'Smectic D \ At
present only four compounds are known to exhibit this phase,
4'-n-hexadecyloxy- and 4 /-n-octadecyloxy-3 /-nitrobiphenyl-4carboxylic acid and two similar acids with CN replacing NO 2 .
Interestingly, D occurs between S c and SA or between S c and
the isotropic phase. The manner in which such a structural
rearrangement takes place is yet to be resolved.
Three-dimensional crystal, orthorhombic with interlayer
Smectic E(S E)
herringbone arrangement of the molecules. (2)
C-centred monoclinic (a > b) with in-plane short-range
Smectic F(SF)
positional correlation and weak or no interlayer positional
correlation: tilted hexatic.
Smectic F*(SF.) Chiral SF with twist axis normal to the layers.
Three-dimensional crystal, C-centred monoclinic (a > b).
Smectic G(SG)
Three-dimensional crystal, monoclinic (a > b), herringbone
Smectic H(S H)
structure.(2)
Smectic H*(SH*) Chiral SH with twist axis normal to the layers.
C-centred monoclinic (b > a), tilted hexatic with slightly
Smectic I(ST)
greater in-plane correlation than SF.
Chiral Sz with twist axis normal to the layers.
Smectic I*(SIA)
Three-dimensional crystal, C-centred monoclinic (b > a).
Smectic J(ST)
Smectic J*(SJ#) Chiral Sj with twist axis normal to the layers.
Three-dimensional crystal, monoclinic (b > a), herringbone
Smectic K(SK)
structure.(2)
Smectic K*(SK*) Chiral SK with twist axis normal to the layers.

302

5. Smectic liquid crystals

5.2 Extension of the Maier-Saupe theory to smectic A: McMillan's


model
McMillan(7) proposed a simple and elegant description of smectic A by
extending the Maier-Saupe theory to include an additional order parameter for characterizing the one-dimensional translational periodicity of
a layered structure. A similar but somewhat more general treatment, based
on the Kirkwood-Monroe theory of melting,(8) was developed independently by Kobayashi(9), but McMillan's approach lends itself more
easily to numerical calculations and comparison with experiment.
The anisotropic part of the pair potential is conveniently taken in the
form
3cos 2 0 1 2 -l
^ ,

__,.
(5.2.1)

where the exponential term reflects the short-range character of the


interaction, r12 is the distance between the molecular centres and r0 of the
order of the length of the rigid part of the molecule.
If the layer thickness is d, we may write the self-consistent single particle
potential, retaining only the leading term in the Fourier expansion, as
follows:
^(z,cos/9) = - F0[^ + (7acos(27rz/^)]|(3cos26>-l),

(5.2.2)

where
(7rr0A02].

(5.2.3)

As will be seen later (5.3.1), experiments have confirmed that the


density wave in smectic A is, in fact, very well represented by a sinusoidal
function, indicating that higher terms in the Fourier expansion can be
neglected. The form of the potential (5.2.2) ensures that the energy is a
minimum when the molecule is in the smectic layer with its axis along z; s
and o are order parameters which we shall define presently.
The single particle distribution function is then
/i(z, cos 0) = exp [ - Fx(z, cos 0)/kB T]
and self consistency demands that
_ /3cos2fl-l\

(5.2.4)

5.2 Extension of the Maier-Saupe theory to smectic A

303

1.0
s

tic

<
3

'20

opi

0.5

is - 15

on

>

- 10

1 -

0.8

0.9

1.0

Reduced temperature

Fig. 5.2.1. Order parameters s and a, entropy S and specific heat cv versus reduced
temperature kB T/0.2202 Vo predicted by the model for a = 1.1 showing afirstorder
smectic A-isotropic transition. S and cv are expressed in terms of Ro, the gas
constant. (After McMillan.(7))
3 cos2 0 - 1
o = /cos(27iz/d)f :

(5.2.6)

where the angular brackets denote statistical averages over the distribution
fv The parameter s defines the orientational order, exactly as in the
Maier-Saupe theory, while a is a new order parameter which is a measure
of the amplitude of the density wave describing the layered structure. The
last two equations can be solved numerically to obtain the following types
of solutions:
(i) o = s = 0
(isotropic phase)
(ii) a = 0, s =1= 0 (nematic phase)
(iii) a H= 0, s ^F 0 (smectic phase).

5. Smectic liquid crystals

304

0.8

0.9
Reduced temperature

1.0

Fig. 5.2.2. Order parameters s and a, entropy S and specific heat cv versus reduced
temperature for a = 0.85 showing first order smectic A-nematic and nematicisotropic transitions. (After McMillan.(7))

The free energy of the system can be calculated in the usual manner:
F=U-

where

TS,
(5.2.7)

and
-TS

= NV0(s2

T dz T d(cos 0)/^, cos 0)1.


Jo

Jo

J
(5.2.8)

The two parameters characterizing the material are Vo, which determines

5.2 Extension of the Maier-Saupe theory to smectic A

305

1.0

0.5

)I
4

o?

<

- 11 //

.a
s

20

a,
o
- 15
- 10
-

1 -

1.0
0.9
Reduced temperature
Fig. 5.2.3. Order parameters s and cr, entropy S and specific heat cv versus reduced
temperature for a = 0.6 showing a second order smectic A-nematic transition and
a first order nematic-isotropic transition. (After McMillan. (7))

0.8

the nematic-isotropic transition temperature, and a, a dimensionless


interaction strength, which can vary between 0 and 2. Experimentally the
layer thickness d is of the order of the molecular length. Neglecting the
odd-even effect (see 2.3.3) the energy associated with smectic ordering
tends to increase if a (and hence d) is larger. Thus a increases with
increasing chain length of the alkyl tails.
Curves of the order parameters, entropy and specific heat for three
representative values of a are presented in figs. 5.2.1, 5.2.2 and 5.2.3. For
a > 0.98, the smectic A transforms directly into the isotropic phase, while
for a < 0.98 there is a smectic A-nematic (A-N) transition followed by a
nematic-isotropic transition at higher temperature. For a < 0.70 and
87 t n e m
^AN/^NI < 0d e l predicts a second order A-N transition. Hence

306

5. Smectic liquid crystals

l.l

Isotropic

1.0

Nematic

Smectic A

0.8

Transition temperature

0.9 -

0.7

\.

Isotropic

Nematic\.
or
\ ^ ^
cholesteric^^"^^^^-^^
/

Smectic A

Alkyl chain length

1.0
1.2
1.1
0.7 0.8 0.9
Model parameter a
Fig. 5.2.4. Phase diagram for theoretical model parameter a. Inset: typical phase
diagram for homologous series of compounds showing transition temperatures
versus length of the alkyl end-chains. (After McMillan.(7))

0.5

0.6

a = 0.7 (and TA^/TNI = 0.87) corresponds to a tricritical point (10) at which


the line of first order transition goes over to a line of second order
transition.
The phase diagram of transition temperature versus a or alkyl chain
length is shown in fig. 5.2.4. There is broad agreement with the trends in
thermodynamic data, though the theoretical A - N transition entropy
versus TAN/Tm is somewhat higher than the observed values (fig. 5.2.5). To
improve the agreement, McMillan used, in a later paper (11), the modified
pair potential
K12(r12,cos012) = _

(5.2.9)

There are now three model-potential parameters which are fixed by


requiring the theory to fit 7^N, Tm and 5 AN . The results are essentially the

5.2 Extension of the Maier-Saupe theory to smectic A


1.5 r

307

1.0

0.5

0.85

0.90

0.95

1.00

Fig. 5.2.5. Smectic A-nematic (or cholesteric) transition entropy versus ratio of
transition temperatures TAN/Tm. Solid line is theoretical curve taken from fig.
5.2.4; open circles are experimental values of Davis and Porter (MoL Cryst. Liquid
Cryst., 10, 1 (1970)); open triangles are data of Arnold (Z. Physik. Chem. (Leipzig),
239, 283 (1968); ibid., 240, 185 (1969)). (After McMillan.(7))

200
(a)

S 100

Isotropic

Smectic A

I
60

70

80

90

Temperature (C)
Fig. 5.2.6. Measured intensity of X-ray scattering at the Bragg angle versus
temperature for cholesteryl myristate. The dashed line is the calculated diffuse
scattering and fluctuation scattering contribution. The full lines represent the
theoretical curves for the total intensity due to Bragg, diffuse and fluctuation
scattering derived from (a) the simpler model potential and (b) the refined one. The
theoretical intensity has been adjusted to be equal to the experimental value at the
lowest temperature. (After McMillan. (11))

5. Smectic liquid crystals

308
0.8
0.6

(a)
0.4
0.8

Z. 0.6
0.4
0.8
0.6
0.4
0.8

(c)

0.9
Reduced temperature

1.0

Fig. 5.2.7. Experimental values of the orientational order parameter s obtained


from NMR measurements for the 4-n-alkoxy-benzylidene-4/-phenylazoaniline
series, (a) C14; (b) C10 (open circles), C7 (X), C3 (filled circles); (c) C2. The solid
curves give the values predicted by McMillan's model. (After Doane et al.(22))
same as those obtained with the simpler model but there are some
quantitative improvements. A number of other refinements and extensions
have been proposed, (1221) but McMillan's model remains the simplest
which brings out all the qualitative features of the A-N and A-I transitions.
A direct method of studying the translational order (or the amplitude of
the density wave) is by measuring the intensity of the Bragg scattering from
the smectic planes. McMillan's experimental results on cholesteryl myristate(11) are shown in fig. 5.2.6 and as can be seen there is excellent
agreement with the refined model. The X-ray intensities reveal an
appreciable pretransitional smectic-like behaviour in the cholesteric
(nematic) phase. This aspect of the problem will be dealt with in a later
section.
The orientational order parameters in the smectic and nematic phases,
studied by magnetic resonance and other techniques also follow the

5.2 Extension of the Maier-Saupe theory to smectic A


1.5

309

N==C
O-C 8 H 17

O
1.0

0.5
1.2 -

90 C
J
100
110
80
90
Temperature T (C)
Fig. 5.2.8. Temperature dependence of the orientational order parameter s
determined by 14 N quadrupolar splitting measurements for CBOOA. The inset
shows the discontinuity in slope at the smectic A-nematic transition. (After
Cabane and Clark. (23) )
0

50

60

70

predicted type of behaviour as the length of the alkyl end-chain is


increased. In particular, a continuous change of s at 7^N, as expected of a
second order transition, has been found (within experimental limits) for a
number of cases. Figs 5.2.7 and 5.2.8 present the data for n-/?-ethoxybenzylidene-/7-phenyl-azoaniline(22) and 4-cyano-benzylidene-4 /-octyloxyaniline (CBOOA) (23). Thus this simple molecular model correctly leads to
the existence of a tricritical point, but it makes no predictions regarding the
critical exponents. Being a mean field theory, it may be expected to yield
y = 1, V,| = v = 0.5, which, as we shall see later, is not in accord with the
experimental results (5.5).

310

5. Smectic liquid crystals


5.3 Continuum theory of smectic A
5.5.7 The basic equations

The stratified structure of a smectic liquid crystal imposes certain


restrictions on the types of deformation that can take place in it. A
compression of the layers requires considerable energy - very much more
than for a curvature elastic distortion in a nematic - and therefore only
those deformations are easily possible that tend to preserve the interlayer
spacing. Consider the smectic A structure in which each layer is, in effect,
a two-dimensional fluid with the director n normal to its surface. Assuming
the layers to be incompressible, the integral
i-dr

(5.3.1)

represents the number of layers crossed on going from A to B, where d is


the layer thickness.(24) In a dislocation-free sample, this number should be
independent of the path chosen so that
and hence

Vxn = 0

n-Vxn = 0 I
and
\
(5.3.2)
n x V x n = 0.J
In other words, both twist and bend distortions are absent, leaving only
the splay term in the Oseen-Frank free energy expression (3.3.7). It is seen
from fig. 5.3.1, that by merely bending or corrugating the layers a splay
deformation can be readily achieved without affecting the layer thickness.
A more complete description of smectic A needs to take into account the
compressibility of the layers, though, of course, the elastic constant for
compression may be expected to be quite large. The basic ideas of this
model were put forward by de Gennes.(24) We consider an idealized
structure which has negligible positional correlation within each smectic
layer and which is optically uniaxial and non-ferroelectric. For small
displacements u of the layers normal to their planes, the free energy density
in the presence of a magnetic field along z, the layer normal, takes the form

/6M

8w\

(a?+#

.....

(5 3 3)

'-

where the first term is the elastic energy for the compression of the layers

5.3 Continuum theory of smectic A

311

nTTTTTTT
Fig. 5.3.1. Flexibility of smectic A layers: only such deformations as preserve the
interlayer spacing take place readily.

and/ a the anisotropy of diamagnetic susceptibility. When H = 0, there will


be no terms in (du/dx)2 or (du/dy)2 as a uniform rotation about y or x does
not affect the free energy. The last two terms are usually negligible and may
be omitted. Also, the physically reasonable assumption is made that twist
and bend distortions given by (5.3.2) are not allowed despite the fact that
V x n does not strictly vanish when the layers are compressible.
5.3.2 The Peierls-Landau instability
Equation (5.3.3) is analogous to the Peierls-Landau (25) free energy
expression for a two-dimensional crystal, and leads to a logarithmic
divergence of the mean square fluctuation <w2> as H->0. Writing the free
energy in terms of the Fourier components of u
uq = u(r) exp (iq r) dr,

(5.3.4)

and substituting in (5.3.3) we get in the harmonic approximation,


F = \ Yu \uq\2 [Bg2 + n(#i + ~2) #j.L
1

where q\ = q x + ql and = {klx/xJiH'


/

(5.3.5)

is the magnetic coherence length.

From the equipartition theorem

kBT

(5.3.6)

from which the mean square fluctuation


<u2(r)} =

(5.3.7)

where d is the layer spacing, assuming the sample to be infinitely large. As

312

5. Smectic liquid crystals


I n , l III 1 , 1 , 1 , 1 ,

Fig. 5.3.2. A diagram depicting the molecular arrangement in (a) the nematic and
(b) the smectic A phases. In the nematic phase the molecules are randomly
distributed so that any horizontal line intersects the same number of molecules. In
the smectic A phase the number of molecules intersected by a line varies sinusoidally
with the position of the line, being 50 per cent more than the average at the
positions indicated by the arrows and 50 per cent less than the average half way
between. This represents a density wave whose amplitude is 50 per cent of the mean
density, which is far greater than what actually occurs near the A-N transition in
a real system. (After Schaetzing and Litster.(26))

i/-*0, <w2>->oo showing that such a structure cannot be stable. For a


sample of dimension L
kBT

(5.3.8)

in the absence of a magnetic field. The layer fluctuations therefore diverge


logarithmically with sample size.
These results imply that smectic A does not possess true long-range
translational order. Hence the conventional picture of the smectic A
structure with the molecules forming well defined layers (fig. 1.1.5 (a)),
though useful conceptually, is far from accurate. Fig. 5.3.2 gives a more
realistic representation of the actual situation. X-ray experiments confirm

5.3 Continuum theory of smectic A

313

that the density wave is, in fact, very well described by a sinusoidal
function; the higher order diffraction maxima are either absent or, when
they do occur, extremely weak, about 10~3 or 10~4 times weaker than the
intensity of the principal peak. (27)
In the harmonic approximation, the displacement-displacement correlation function may be defined as
G(r) = exp{-[^ 2 <lW-t/(0)| 2 >]}.

(5.3.9)

For the crystalline lattice, this is the familiar Debye-Waller factor. In the
case of smectic A, one gets from (5.3.4)
(5.3.10)
This was evaluated by Caille (28): for d <^ \r\ < L, the correlation functions
for the two principal directions vary as
and

G(r)ocz-

r = ( * + , ) * - 0,1
2

G(r) oc rL \

z~0,

( 5 3 U )

where
n = ~~zi,

(5-3.12)

which is usually a small number ( ~ 0.1-0.4).


The Fourier transform of G(r) yields the intensity of scattering:
lK\qz-q0\-*+i,
Ioc\q\-*^,

q = 0,
qz = 0,

(5.3.13)
(5.3.14)

where
q0 = In Id.
The sharp Bragg peak characteristic of the long-range ordered crystal
lattice is now replaced by strong thermal diffuse scattering with a ' powerlaw' singularity.(29) The excellent high resolution X-ray measurements of
Als-Neilsen et al.m) have confirmed this prediction (fig. 5.3.3).

533

The Helfrich deformation

If a magnetic field is applied parallel to the smectic planes in a


homeotropically aligned sample and / a > 0, one may expect a Helfrich

314

5. Smectic liquid crystals


10

: / \
6
10"1

\\ v
M

\ \

10"2

\
\

A
\\ V
\\
\

\ 1

ry

= 0.38
O

= 0.17

-1

Fig. 5.3.3. X-ray scattering intensity profile from the smectic A phase of 80CB at
two reduced temperatures, ( r A N - T)/TAN = 9 x 10"4 (filled circles) and 10"6 (open
circles). The dashed line is the experimental resolution function as would be seen if
smectic A had true long-range order. The full lines are the best fits of the theoretical
line shape \qz qo\~2+n folded with the resolution function. The values of rj so
obtained agree with those calculated from (5.3.12) using experimentally determined
values of q0, B and k1. (Als-Nielson et al.(m)

type of deformation to set in above a critical field, as in the case of


cholesterics (fig. 4.6.3). Assuming a distortion of the form
u(x, z) = u0 sin kz z cos kx x,
where kz = n/L and L is the sample thickness, and evaluating the average
free energy, one obtains
(5.3.15)

F =where
X=

(5.3.16)

is a characteristic length of the material of the order of the layer thickness.


Proceeding as in 4.6.2, the optimum value of the distortion wavevector is
k\ = kjk = n/kL.

(5.3.17)

5.3 Continuum theory of smectic A

315

In other words, the spatial periodicity of the deformation is proportional


to the geometric mean of the layer thickness and the sample thickness. The
critical field given by
is, however, rather large compared to that for cholesterics. For example,
taking k = 20 A and L = 1 mm, Hc ~ 50 kG, and thus far no experimental
studies of this effect appear to have been carried out.
The same type of distortion can be achieved more easily by mechanical
means, i.e., by increasing the separation between the glass plates (3132) (fig.
5.3.4). In the absence of a magnetic field, and taking u to be independent
of y, (5.3.3) reduces to
(5.3.18)
However, in the present case we note that a bending of the layers alters the
effective layer spacing along z and therefore makes a second order
contribution to the layer dilatation in that direction. Hence (5.3.18) needs
a correction term:

ir = i^r^_I/ r ^yi 2 _ h fc n ^Y

(5.3.19)

The displacement u is now given by


u = sz + u0 sin kz z cos kx x,

(5.3.20)

where s = A/L, A being the plate displacement. The problem is analogous


to the magnetic case except that Bs replaces z a / / 2 . The threshold value of
the strain
sc = 2nk/L
(5.3.21)
and the plate displacement A = Ink ~ 150 A, which is easily realized in
practice. Experiments have been done on CBOOA to verify these
conclusions/ 31 ' 33) The plate separation was increased in a controlled
manner by piezoelectric ceramics. When the dilatation reached a certain
value, there appeared two transient bright spots in the laser diffracted
beam confirming the onset of a spatially periodic distortion above a
threshold strain. A transient periodic pattern was also visible under a
microscope. It was verified that kx cc L'1 in accordance with (5.3.17) (fig.
5.3.5). The measurements yielded k = 22 3 A for CBOOA at 78 C. Also,
assuming kxx ~ 10~6 dyn, B was estimated to be 2 x 107 cgs.

316

5. Smectic liquid crystals


AA

V/////////////////7/A

Y////X
Y/////. 777/77//A
277/*,

Fig. 5.3.4. The Helfrich deformation in a smectic A film subjected to a mechanical


dilatation A.

50

100

150

Fig. 5.3.5. Dependence of the wavevector of the Helfrich deformation in the


smectic A phase of CBOOA on the sample thickness d. The slope yields a value of
X = 2 2 3 A. (After Durand. (33) )

5.3 Continuum theory of smectic A

317

5.3.4 Fluctuations and Rayleigh scattering


If a smectic A sample is homeotropically aligned between two glass plates
having a separation L, the boundary conditions require that qz = mn/L,
where m is an integer. We shall confine the discussion to m = 1. When
H = 0 we have from (5.3.6)
(\uQ\2y = ^B T/iBql + k^q^).

(5.3.22)

We know that the intensity of light scattering is proportional to the mean


square fluctuation of the director (see 3.9):

k T

(5.3.23)

The elastic energy is minimized when qL = qc, which represents the


optimum wavevector. When qz = 0,
<|<N2> = KT/kuq\

(5.3.24)

which is a large quantity as in a nematic liquid crystal, since it involves only


the splay coefficient. On the other hand, when qz and q are comparable,
<53 25)

is quite small; consequently, in certain geometries, e.g., when a homeotropically aligned sample is held against an extended source of light and
viewed normally, the medium will not appear very turbid.
Strictly speaking, one should take into account the contributions of k 22
and k33, since the layers are assumed to be compressible. Following the
procedure outlined in 3.9, the fluctuations may be decomposed into two
modes, and choosing the wavevector q in the xz plane, one gets the general
expressions(34)
,,* , , s =

xl?

k q i + k^

kBT[l+(B/D)(qJqLn

+ iB/DHD

+k q l +k ^ D i q J q r

where D is the elastic constant associated with the fluctuations of the


director away from the layer normal.

318

5. Smectic liquid crystals

Fig. 5.3.6. Distortion caused by an irregularity on the glass surface. The bending of
the layers can be relaxed only by compression, which requires considerable energy,
so that the surface distortion extends to appreciable depths inside the specimen.
In a perfect sample of smectic A, the most important contribution to
light scattering comes from the undulation mode with the wavevector
parallel to the layers (qz = 0). However, unless very special precautions are
taken, it is difficult to observe this scattering, for even small irregularities
on the surfaces of the glass plates cause static distortions of the layers
which extend to appreciable depths inside the specimen (3532) (fig. 5.3.6).
The reason for the high penetration length of a surface distortion is easily
understood. A corrugation of the layer which involves a nematic-like splay
elastic constant and requires very little energy can be relaxed only by
compression, which needs considerable energy and therefore occurs over a
large distance. If a static undulation of this type is written as
u = u0 exp ( - z / 0 cos qx,
where C is the 'attenuation' length, substitution in the free energy
expression (5.3.18) gives

or

C = (V)
(whereas in nematics, ~ q x ). These long-range static undulations give
rise to an intense scattering of light which completely swamps that due to
the thermal fluctuations. Indeed in the first experiments*36'37) it was the
static effects that were observed as was confirmed by a temporal analysis of

5.3 Continuum theory of smectic A


105

319

r-

104

103

77 C

102

79 C

10

0.1

74 C

I
10"

io

101

102

Shear rate (s" )


Fig. 5.3.7. Apparent viscosity versus shear rate for cholesteryl myristate at different
temperatures: full line, smectic A; dashed line, cholesteric; chain line, isotropic.
(After Sakamoto, Porter and Johnson.(41))
the scattered light, using a laser beat spectrometer and correlator.
However, in subsequent studies(38) the scattering due to thermal fluctuations were detected by choosing optical glass plates of very high quality.
The temporal analysis showed an exponentially decaying correlation
function characteristic of a dynamic undulation.
These modes are highly damped, the relaxation time T being of the order
of IO"3 s. To discuss the damping quantitatively we have to consider the
hydrodynamics of smectic A. The most general formulation of the theory
is due to Martin, Parodi and Pershan(39) but we shall present the relevant
equations in a simplified form using de Gennes's notation.

320

5. Smectic liquid crystals


5.3.5 Damping rate of the undulation mode

We write the energy density in a more general form(24) to include volume


dilatation 6:

For static isothermal deformations, F may be minimized with respect to 6


to give
Aodz
so that

B = B0-(2C20/A0)

(5.3.29)

and (5.3.28) reduces to the simpler form (5.3.3) assumed previously.


If vt is the velocity of the particle, the equation of motion is
/>*i = -P.i + gi + t'Jij>

(5.3.30)

where gt is the force on the layers and t'n the viscous stress tensor which, in
contrast to the nematic case, is assumed to be symmetrical and dependent
on the velocity gradients only. In the problem under consideration we have
(5.3.31)
and
g = -SF/Su.

(5.3.32)

The force g normal to the layers will be associated with permeation effects.
The idea of permeation was put forward originally by Helfrich(40) to
explain the very high viscosity coefficients of cholesteric and smectic liquid
crystals at low shear rates (see figs. 4.5.1 and 5.3.7). In cholesterics,
permeation falls conceptually within the framework of the Ericksen-Leslie
theory(42) (see 4.5.1), but in the case of smectics, it invokes an entirely new
mechanism reminiscent of the drift of charge carriers in the hopping model
for electrical conduction (fig. 5.3.8).
The rate of entropy production may be written as (see 3.1.4)
TS=t'ndij+g(u-vz),

(5.3.33)

where dtj = \{v{ j + Vj t), and (u vz) describes the permeation. Treating t'n

5.5 Continuum theory of smectic A

321

Fig. 5.3.8. Helfrich's model of permeation in smectic liquid crystals. The flow takes
place normal to the layers in a manner similar to the drift of charge carriers in the
hopping model for electrical conduction.
10 --

A
\
<\
\
L

5 T

10

15
q2 (108 cm- 2 )

1
20

25

Fig. 5.3.9. Damping rate of the undulation mode in the smectic A phase of CBOOA
determined by laser beat spectroscopy for two different sample thicknesses, 200
and 800 /zm. Solid lines represent the theoretical curves calculated from (5.3.40)
and (5.3.41). (After Ribotta, Salin and Durand.(38))
and g as fluxes and dtj and (u vz) as forces, and expanding t'n one obtains
the following relations:
^ = Mi sn

Sjz dzi + Siz dzj) + // 5 Sjz Si

+ M2

(5.3.34)
u-vz

= v p g,

(5.3.35)

where //15 ...,// 5 are five viscosity coefficients and vp the permeation
coefficient.

322

5. Smectic liquid crystals

As far as the highly damped undulation modes are concerned, the


volume dilatation can justifiably be neglected and the isothermal approximation is probably satisfactory. The equation of motion then reduces to
pvz = -rjq2vz + g,

(5.3.36)

where

g = -kliq*u

= (u-vz)/vv

(5.3.37)

and
* = SG<s + / O .

(5.3.38)

Neglecting the inertial term in (5.3.36) and eliminating vz,

For small q, the last term may be neglected (except very near the boundary,
but we shall ignore the boundary layer). This results in a purely damped
mode whose relaxation rate is
1

L=
T

n2

h3_.

rj

(5.3.39)

More generally, taking into account the thickness L of the sample one may
write

where qc is defined in (5.3.23). The relaxation rate should have a minimum


value
^
TC

t]lL

2 ^
rj

where

1 = qc = (n/lL)i

(5.3.42)

Experiments with specimens of different thicknesses have confirmed this


prediction(38) (fig. 5.3.9). At very low q, the boundary effects quench the
fluctuations and T"1 increases sharply. At high q9 T"1 becomes a linear
function of q2 from the slope of which klx/rj may be determined. This value
is comparable to that for a nematic ( ~ 2 x 10~6 cgs) as is to be expected.
The relaxation time TC decreases linearly with decreasing sample thickness
as predicted by (5.3.41), and the wavevector qc shows the expected
dependence on L. The measurements yield a value of I which is in
reasonable agreement with that obtained from (5.3.21).

53 Continuum theory of smectic A

323

5.3.6 Ultrasonic propagation and Brillouin scattering


For an arbitrary direction of the wavevector there are two acoustic
modes.(24) Neglecting viscous effects and permeation, putting y = du/dz,
and using the conservation law
divv + 0 = 0,

(5.3.43)

we obtain the following equations:

y and 6 being the independent variables. This leads to the secular equation
co{\pco2 - (Bo + Co) q% (po? - Ao q\) -(Au + Co) q\{paf + Co q\)} = 0.
(5.3.45)
We are interested here in propagating modes and ignore the case co = 0.
Solving for the velocities c = co/q, we get
c\ + c\ = p-\AQ cos2 <p + (Ao + Bo + 2C0) sin2 p],
2

c c =/7- (^0^0-C )sin

cos ^,

(5.3.46)
(5.3.47)

where cp is the angle between q and its projection on the layer.


Ultrasonic velocity measurements have been reported on oriented
smectic A samples of 4,4/-azoxydibenzoate. The first studies by Lord(43) for
two values of (0 and 90) established the anisotropy of the velocity of
propagation. Subsequently, Miyano and Ketterson(44) investigated the
angular dependence of the sound velocity and from a least squares fit with
(5.3.45) were able to determine the elastic coefficients Ao, Bo and Co. By
analogy with the elasticity theory of solids(45), we may write

^33

Their estimates show that Ao is much larger than Bo or Co (fig. 5.3.10). This
enables us to give a simple physical interpretation of the two branches

5. Smectic liquid crystals

324
1.6

1.5

1.4
110

115

120

Temperature (C)

0.3 r

0.2
X

0.1

8- -8-

no

120
115
Temperature (C)

Fig. 5.3.10. Temperature dependence of the elastic constants of the smectic A phase
of diethyl 4,4/-azoxydibenzoate determined by ultrasonic velocity measurements:
open circles, 2 MHz; filled circles, 5 MHz; triangles, 12 MHz; crosses, 20 MHz.
(After Miyano and Ketterson.(44))
described by (5.3.46) and (5.3.47): the density and layer oscillations are, in
effect, uncoupled. Hence one of the branches corresponds to the normal
longitudinal wave whose velocity can be seen from (5.3.46) to be
Cl

(AJp)\

(5.3.48)

which is practically independent of the direction of propagation. The other


branch corresponds to changes in the layer spacing, without appreciable
density changes and may be compared with the phonon branch in
superfluids known as second sound.m) The velocity of this mode
c 2 (B0/p)* sin q> cos <p

(5.3.49)

is strongly orientation dependent. It becomes zero for propagation along

5.3 Continuum theory of smectic A

325

the layers as well as perpendicular to them. When the wavevector is


parallel to the layers (qz = 0), it becomes the highly damped undulation
mode which we have already examined in detail. When q is normal to the
layers permeation effects set in and the wave is again strongly damped.(39)
Neglecting density changes for this mode, it is clear from (5.3.43) that for
q along z,vz = 0. Thus from (5.3.35)
or using (5.3.24)

u = vpg

(5.3.50)

u = vpBd2u/dz2.

(5.3.51)

Therefore, when the wavevector of second sound is normal to the layers, it


becomes a purely dissipative mode with a relaxation rate
\/T = vvBq\

(5.3.52)

Furthermore, second sound is a critical mode, i.e., its velocity goes to zero
as r - TAN.
A direct confirmation of the existence of these two branches has been
found by Liao, Clark and Pershan(47) from their Brillouin scattering
experiments on a monodomain sample of ^-methyl butyl /?((/?-methoxybenzylidene)amino) cinnamate. This compound shows the nematic,
smectic A and smectic B phases. Choosing both the incident and the
scattered light to be polarized either as ordinary or extraordinary waves,
they observed two peaks corresponding to the two modes, the angular
dependence of which is in excellent agreement with the theory (fig. 5.3.11).
5.3.7 Breakdown of conventional hydrodynamics
Theoretical developments*48'49) in the early 1980s showed that the nonlinear
interaction of thermally excited layer undulations, which as we have seen
have large amplitudes because of the Peierls-Landau instability, leads to
interesting new effects in the hydrodynamics of smectic A at small
wavevector and frequency. We present below a very brief outline of the
physical arguments involved.(50)
In writing the expression for the free energy density (5.3.18) we neglected
terms involving (du/dx)2 and (du/dy)2 since a uniform rotation of the layers
does not cost energy (assuming, of course, that there is no externally
applied magnetic field). Consequently in the harmonic approximation,
<|wJ2> takes the form (5.3.22). However, while discussing the undulation
instability in smectic A (5.3.3), we observed that a bending of the layers
alters the effective layer spacing along z and therefore makes a second

5. Smectic liquid crystals

326

2.0 H

1.5

1.0

(a)

0.5

1.0

2.0

2.0

1.5

1.0
(b)
0.5

1.0

2.0

Sound velocity (105 cm s"1)

Fig. 5.3.11. Dependence of the sound velocities on the polar angle q> from Brillouin
scattering experiments on ^-methyl butyl /?((/>-methoxy-benzylidene)amino) cinnamate. (a) Smectic A (T = 60.7 C), (b) smectic B (T = 48.1 C). The dashed lines
are calculated from theory. The presence of a third component in (b) indicates that
the shear modulus does not vanish in smectic B at these very high frequencies.
Circles, triangles and squares represent measurements at different scattering
angles. (After Liao, Clark and Pershan.(47))
order contribution to the layer dilatation in that direction. The corrected
form of the free energy density is given by (5.3.19), which we write here as
follows:

(say),

(5.3.53)

where the subscripts h and a stand for the harmonic and anharmonic parts

5.4 Defects in smectic A

327

of F, and V and A are the gradient operator and Laplacian respectively


in the xy plane. Using a renormalization procedure, Grinstein and
Pelcovits(48) showed that the effect of Fa at non-zero temperature is to
change (5.3.22) into a form where the elastic constants are replaced by q
dependent functions with

for asymptotically small q.


This surprising result prompted Mazenko, Ramaswamy and Toner(49) to
examine the anharmonic fluctuation effects in the hydrodynamics of
smectics. We have already shown that the undulation modes are purely
dissipative with a relaxation rate given by (5.3.39). To calculate the effect
of these slow, thermally excited modes on the viscosities, we recall that a
distortion u results in a force normal to the layers given by (5.3.32). This is
the divergence of a stress, which, from (5.3.53), contains the non-linear
term 82(V u)2. Thus, there is a non-linear contribution (V uf to the stress.
Now the viscosity at frequency co is the Fourier transform of a stress
autocorrelation function,(51) so that Ar/(co), the contribution of the
undulations to the viscosity, can be evaluated. It was shown by Mazenko
et al.m) that Arj(co) ~ \/co. In other words, the damping of first and second
sounds in smectics, which should go as rj{co) co2, will now vary linearly as co
at low frequencies.
A similar calculation for discotics(52) yields An(co) ~ co~*. The original
work of Mazenko et al.m) argued that one of the shear viscosities should
also diverge, but Milner and Martin(53) showed that this was not the case.
This remarkable l/co divergence of the viscosities of a smectic at low
frequencies is now confirmed by several independent experiments(54) using
ultrasonic attenuation and second sound resonance.
5.4 Defects in smectic A
5.4.1 Focal conic textures
We have noted that the smectic A layers are flexible and easily distorted,
but tend to preserve the interlayer spacing (fig. 5.3.1). Moreover, as the
layers can slide over one another, the structure adjusts itself readily to
surface conditions. For example, when there is a centre of attachment at
the glass surface the molecules adopt a radiating or a fan-like arrangement
and the layers form a family of equi-spaced surfaces normal to the
molecular directions. Under a polarizing microscope such distortions give

5. Smectic liquid crystals

328

(a)

(b)

Circle L(.
Hyperbola Lh

to

Ellipse Le

Fig. 5.4.1. (a) Smectic layers in concentric cylinders to form a myelin sheath with
a singular line L along the axis; (b) the cylinders are closed to form tori: there are
two singular lines, a circle Lc and a straight line Ls; (c) the general case when the
smectic layers form Dupin cyclides: the circle becomes an ellipse L e and the straight
line a hyperbola Lh.

rise to beautiful optical patterns known as focal conic textures. These were
studied in considerable detail by Friedel, (55) to whom we owe the
explanation of their origin. We shall now examine the structures of these
focal conic domains. (56)
(a) The simplest case is shown in fig. 5.4.1 (a). The layers are wrapped
round in concentric cylinders to form a myelin sheath and there is a
singular line L along the axis, (b) When the myelin sheath is bent and closed
to form tori, the singular line L becomes a circle Lc (fig. 5.4.1 (b)). In
addition, the structure now has another singular line Ls - a straight line
through the centre of the circle and perpendicular to it. Such a structure is

5.4 Defects in smectic A

{a)

329

(b)

Fig. 5.4.2. A rare example of a pair of singular lines, one almost straight and the
other almost circular in a toric domain in the (a) smectic A and (b) smectic C
phases. Additional disclination lines develop near the centre in the smectic C phase
for reasons discussed in 5.8.3. (From A. Perez, M. Brunet and O. Parodi, J. de
Physique Lettres, 39, 353 (1978)).

observed only very rarely (fig. 5.4.2). (c) In the most general case, the
smectic layers lie on Dupin cyclides and not on tori. The circle Lc is then
transformed to an ellipse Le and the straight line into a hyperbola Lh (fig.
5.4.1 (c)). Le and Lh are conic sections in perpendicular planes, one going
through the focus of the other (figs. 5.4.3 and 5.4.4). It is these structures
that are most commonly seen under the polarizing microscope (fig.
1.1.7 (a)).
Based on the fact that the layers are bent into Dupin cyclides, one can
derive expressions for the principal curvatures ox and cr2, and thus work out
the energy,

330

5. Smectic liquid crystals

(a)

Fig. 5.4.3. (a) Geometry of a pair of focal conies. (After Friedel.(55)). (b) A section
of (a) in the plane of the hyperbola, showing parts of Dupin cyclides. (After
Bragg.(56))
For the general focal conic structure, Kleman (57) has shown that

where e is the eccentricity of the ellipse, p its perimeter, a its semi-major


axis, and rc the core radius. This leads to the interesting result that E
decreases with increasing e. In other words, the circle (e = 0) and straightline combination has higher energy than the ellipse and hyperbola
combination. Hence the possibility of Dupin cyclides degenerating into
tori is not energetically feasible. This is in accord with observations, for the
most commonly observed texture in smectic A is the one in which the focal
lines are ellipses and hyperbolas.

5.4 Defects in smectic A

331

(a)

Fig. 5.4.4. (a) Arrangement of a set of cones within a pyramid. The hyperbolas
belonging to the ellipses meet at the vertex of the pyramid, (b) The base of the
pyramid and the axial directions radiating from the foci at the base of the
hyperbolas. (After Bragg.(56))

In principle, the focal lines can both be parabolas. Such parabolic focal
conies have been observed by Rosenblatt et al.{58) but they are not
common. Kleman(57) has shown that in this case

4/rJ'
where/is the focal length of the parabola and R the sample radius. The
energy increases with R and diverges as f-+ 0. Thus, in general, such
structures are not energetically favourable. However, it turns out that
when R < 3/, they can have smaller energies and Kleman argues that this
may explain the observations of Rosenblatt et al
A remarkable property of the smectic liquid crystal is that an open

332

5. Smectic liquid crystals

Fig. 5.4.5. Optical interferometric patterns showing terraces on the surface of a


homeotropically aligned smectic liquid crystal formed by a solution of potassium
oleate in aqueous methyl alcohol, k = 0.5893 jum. (Reference 59.)

5.4 Defects in smectic A

333

droplet, homeotropically aligned on a perfectly clean glass plate or on a


fresh cleaved mica sheet, forms terraces (fig. 5.4.5). This is known as the
stepped drop, goutte a gradins, and was discovered by Grandjean.(60) From
surface energy considerations it can be argued that the occurrence of steps
is a consequence of the layered structure.(61) Under sufficient magnification
the terraces are seen to be fringed by a chain of focal conic patterns
indicating that the layers are crumpled at the edges.
When the isotropic melt is cooled, the smectic phase often makes its
appearance in the form of needle-shaped particles showing evidence of
focal conic structure. Examples of such needles, termed bdtonnets, can be
seen in fig. 2.1.15 (b).
5.4.2 Edge dislocations
Fig. 5.4.6 shows a simple edge dislocation in smectic A with the dislocation
line parallel to the smectic planes. In fig. 5.4.6(a) the extra half-plane is to
the right of the observer looking down the dislocation line L, and in fig.
5.4.6(b) it is to the left. Consider now the contour integral
-cpn-dr = TV

dj

along a closed path about L in the anticlockwise direction. Clearly N = + \


in fig. 5.4.6 (a) and 1 in fig. 5.4.6 (b). Accordingly, we may refer to the two
cases as positive and negative edge dislocations of unit strength respectively. Topologically these defects are very similar to their counterparts
in crystals, but they differ significantly from the point of view of energetics,
as we shall see presently.
From the free energy expression (5.3.18), we obtain the equation of
equilibrium

where
A. = (k11/B)* and u = u(z, x).
We now proceed to work out the singular solutions of (5.4.1). For an edge
dislocation with the extra half-plane on the side x > 0, the boundary
conditions are
\0
for i < 0
m
"(0, x) = i
[b/2 for x > 0 .

334

5. Smectic liquid crystals


n

(a)

(b)

Fig. 5.4.6. (a) Positive and (b) negative edge dislocations of unit strength.
Also, u(z,x) = u( z,x). We look for a solution of the form<62)
b

b f , x fexp (i<7.x)l ,
( 5 A 2 )

We then have

(5.4.3)

The tilt in the layer normal 6 = du/dx, and the dilatation in the layer
spacing S = Ad/d = Xffl/dx; 6 and 5 can be derived explicitly from the
form of u:
|=

b__ex

J
x

i__x'

(5.4.4)

Xb

(5.4.5)

From (5.4.5) it is seen that there is a compression of the layers for x > 0 and
an extension for x < 0.
The stresses associated with the dislocation cannot be derived uniquely.
However, one acceptable set of values can be obtained from the equation
of motion. (63) They are
Bb

-x
17-3 e x P
2

XBb

-x

/15Z? z I

(5.4.6)

-:

72

2 AX\z exp
x
2X\z exp

- AX\z\
- 4A|zj

(5.4.7)
(5.4.8)

Apart from the exponentially decaying term, a33 varies as l/n, GX1 and a13

as \/n.

In contrast, all these stress components vary as 1 /r in the case of crystal

5.4 Defects in smectic A

335

edge dislocations. The displacement u around the crystal dislocation is


given by

where v is Poisson's ratio. Again the form of this expression can be seen to
be quite different from (5.4.2) for u in smectic A.
The energy of a single edge dislocation
The total energy is given by

E= I \Fdxdz.

(5.4.10)

Using (5.3.18) and (5.4.2) and integrating,

where rc is the core radius, which is expected to be of the order of the


interlayer spacing d, and Ec the core energy. The total energy of an isolated
edge dislocation is therefore independent of the sample size for smectic A,
whereas it increases logarithmically with sample size for a crystal edge
dislocation.
Interaction between parallel edge dislocations
Since (5.4.1) is linear in w, we may invoke the superposition principle to
evaluate the interaction between dislocations.* 63'64) We superpose the
displacements due to one dislocation line at (0,0) and another at (x o ,z o ).
The total displacement is
for z < 0,

u = - 2 ^ l + 4^ \{exp[-(iqxo + Aq\)]-e}[exp(iqx + kq2z)]j?-,


for 0 < z < z0,
u = 2 <*i.-+ 4^ K e x P ( ~ V ^ ) + exp[-i#x o +
for z > z0,

(5.4.12)
Aq2(z0-z)]}[exp(iqx)]jt,
(5.4.13)

u = - 3 > + I {1 - s exp ( - iqx0 + A#2z0)}[exp (iqx - lq*zj\ -^-.


(5.4.14)

336

5. Smectic liquid crystals

Here e is 1 for like dislocations and 1 for unlike ones. The interaction
energy

Ex= J \Fdxdz-E0,

(5.4.15)

where Eo is the self energy of two isolated dislocations. Therefore


|[exp(-Vkol)]cos(^ 0 )d^.

(5.4.16)

This enables us to calculate dislocation interactions but it must be borne in


mind that the theory is not strictly valid for distances of the order of the
core radius.
When the two edge dislocations are in the same horizontal plane, i.e.,
z0 = 0, the interaction is negligible. When z0 > rl/4n2l, i.e., the vertical
separation between the dislocations is much greater than the core size, Ex
can be simplified to

^ ( ^ f f ^ )

(5.4.17)

From this we get two forces Fx and Fz along and normal to the smectic
planes:

Since k.JX2 = b,
and

Fx = -ebaS3

(5.4.20)

Fz = ebals.

(5.4.21)

Here cr33 and <713 are the stresses at the origin, where the first dislocation is
situated, due to the presence of the other dislocation at (x o,zo). This is
known as the Peach-Koehler force(65) on a dislocation arising from the
stress field of the other. This can be generalized to mean that under
application of a stress oip a dislocation experiences a force Fi whose exact
relationship is given by the above expressions.
It is seen from (5.4.18) that Fx is always repulsive for a pair of like
dislocations and always attractive for an unlike pair. On the other hand,
Fz can be attractive or repulsive depending on the relative positions of the
two dislocations. For example, Fz is repulsive for like dislocations when
xl < 2Xz and attractive when x\ > 2Az0. This provides a mechanism for the

5.4 Defects in smectic A

337

Fig. 5.4.7. Clustering of like edge dislocations in smectic A to form a domain wall
or 'grain boundary'. (After Pershan.(63))

clustering of dislocations to form a domain wall or a grain boundary (63)


(fig. 5.4.7).
Another interesting implication of the theory is that under an applied
stress <T33 = Bdu/dz, which can be generated by stretching or compressing
the smectic lattice along the layer normal, the dislocation experiences a
force Fx parallel to the layers. The force changes sign with the sign of the
edge dislocation. This effect is similar to the magnus force acting on
vortices in superfluids. In the case of smectic A, a constant stress induces
a constant flow of dislocations from the boundaries of the container. The
dislocations move to release the stress. As a result, if the dislocations are
not pinned the applied stress will always be expelled by the movement of
dislocations. This has been termed as ' dielasticity' by analogy to
diamagnetism. Any measurement of the intrinsic elastic constants of
smectic A should therefore be carried out at frequencies (or other
experimental conditions) such that the dislocations are immobile.
The above phenomenon has a bearing on the behaviour of a dislocation
loop under an applied stress. For a given dislocation loop, the extra halfplane may be either (i) inside the loop or (ii) outside it. Consequently,
subjecting the system to a tensile stress (normal to the smectic layers)
results in an expansion of the loop in situation (i) and a shrinking of it in
situation (ii). Loops can also occur spontaneously to relax an applied
stress.
Pershan(63) has investigated the interaction of edge dislocations with
different kinds of boundaries in various geometries. Williams and
Kleman(66) have observed, by phase contrast microscopy, walls of
dislocations (probably of large Burgers vector) which are strongly attracted
to the boundaries (the glass substrate or the smectic-solid interface).

5. Smectic liquid crystals

338

(a)

vllllIMIIIII
^11111111111

dmmmii

1 III 1
11111
r 1HIM

1,111,1.1,1,1,1.1.1

II

(c)

(d)

Fig. 5.4.8. Disclinations in smectic A: (a) la(n); (b) Qb(n); (c) Qo( - n); (</) Qb( - n).

5.4.3 Screw dislocations


These are exactly analogous to screw dislocations in crystals. There is a
spiral arrangement of the smectic layers around the dislocation line L
which is normal to the layers. The associated deformation is given by

u =

tan"1 r-

2n
\x
where u is the layer displacement along z, and b the Burgers vector is equal
to an integral multiple of the layer spacing d. For a positive screw
dislocation one gains a step of height b on going once round the line L in
the anticlockwise direction, while for a negative dislocation one loses a
step.
A noteworthy aspect of the above solution is that it does not involve
either lattice dilatation du/dz or layer undulation V-n. Therefore, within
the approximations of the linear theory considered here, screw dislocations
in smectic A have no self energy (apart from the core), nor do they interact
amongst themselves. In this respect they are entirely different from screw
dislocations in crystals.

5.4.4 Disclinations
The geometrical process for creating a disclination in smectic A is as
follows.(67) Cut the material by a semi-infinite plane that runs parallel to the
layers, the limit of the cut being the disclination line L. Rotate the two faces

5.4 Defects in smectic A

rr
IT

1,

^\\\\ II

11II

^X//lll

II

1111

i i II
I I I II 1 IN

n1

1111 1 II MI

11 111 i i n i
Y l III I I I I
II i

sttt

III -UrrU
41 El III! I I I II
(b)

()

-rrfiSl) 111 '

339

i IIII

II ,1,1,1
II
i 1111
HIM II
III
111

{d)

(c)

Fig. 5.4.9. Edge dislocations in smectic A composed of disclination pairs: (a)


a(-n); (b) na(n) + nb(-n); (c) Qb(n) + 4 (-n); (d) Q.b(n)+ ila(-n).

(a)

(b)

Fig. 5.4.10. 'Pincements' in smectic A composed of disclination pairs:


(a) n a (-7r) + Q ; (b) Qb(-7r)
of the cut about L through a relative angle + Nn, N being an integer. Fill
in the voids (or remove overlapping material) to get positive (or negative)
disclinations, and allow the system to relax. The structures so obtained
depend on whether the cut is made at the extreme end of the molecule or
through its middle. The resulting disclinations, designated as Q a and Q&,
are illustrated in fig. 5.4.8. It is obvious that Q a and Qb are not the same
- their core energies are different, and they also have different configurations at large distances. The creation of Q disclinations also gives rise
to disclinations of strength s = | in the smectic director n.
The energy of an isolated disclination being large, disclinations of
opposite signs may be expected to occur in pairs to form edge dislocations
and what Bouligand has called 'pincements' (see figs. 4.2.6 and 4.2.8).
Some possibilities are shown in figs. 5.4.9 and 5.4.10.

340

5. Smectic liquid crystals


5.5 The smectic A-nematic transition
5.5.7 Phenomenological theory of the smectic A-nematic transition

We now proceed to consider a Landau type of phenomenological


description of the A-N transition. This approach to the problem was due
to de Gennes (6869) and to McMillan,(11) both of whom recognized the
analogy with phase transitions in superfluids. We shall follow de Gennes's
treatment.
We start with the density wave in the smectic phase
p(z) = ^ [ 1 + 2 % ! cos (? s z-?>)],

(5.5.1)

where p0 is the mean density, \y/\ the amplitude and qs = 2n/d the
wavevector of the density wave, d the interlayer spacing and (p a phase
factor which gives the position of the layers. Thus the smectic order can be
fully specified by the complex parameter
y/ = Hexp(i<p).

(5.5.2)

Near the transition, the free energy density may be expanded in powers of
y/ as follows
....
(5.5.3)
From symmetry considerations it is clear that only even powers of y/ may
be included. The coefficient /? is always positive. At a certain temperature
7"*, which is the second order transition point, a = 0. Accordingly, as
explained in 2.5.1, we set in the mean field approximation,
a = a(T-T*).
To allow for the coupling between \y/\ and the orientational order
parameter s, the free energy is expressed as(69)
F= \\*+P\y\* + [(dsy/2x\-C\w\>ds,

(5.5.4)

where Ss is the change in the nematic order brought about by layering of


amplitude y/, x a response function which depends on the degree of
saturation of nematic order, and C a positive constant. Minimization with
respect to Ss leads to
2
F
\
where
For a wide nematic range x(TAN) is small, /?' > 0, and the transition is of
second order. For a small nematic range x(TAN) is large, /?' < 0, and the
transition is of first order; in this case one must add a positive sixth order

5.5 The smectic A-nematic transition

341

term in the free energy to ensure stability. The tricritical point occurs when
/?' = 0, i.e., x(TAN) = 2/3/C2. This is in agreement with the prediction of the
microscopic theory, according to which the tricritical point occurs at
7; N /r NI = 0.87 (see 5.2). /(7^ N ) may be altered by varying the length of
the end-chain, or by preparing mixtures, or by the application of
pressure.(70)
5.5.2 Pretransition effects in the nematic phase
In 5.2, we referred briefly to McMillan's X-ray evidence(11) for the growth
of smectic-like short-range order in the nematic (or cholesteric) phase near
the A-N transition (see fig. 5.2.6). This pretransition effect manifests itself
more strikingly in the temperature dependence of the elastic properties of
the nematic phase, as was first demonstrated by Gruler.(71) The twist and
bend distortions, which are normally disallowed in the smectic phase,
become increasingly difficult as the smectic clusters build up and these two
elastic constants rise much more rapidly than expected from the simple s2
law discussed in 2.3.5. To investigate these properties, we expand the free
energy in powers of y/ and its gradients. We shall suppose for the present
that the A-N transition is of second order. For a fixed orientation of the
director,

In the smectic phase (T < T*), the amplitude of the density wave may be
taken to be constant, so that only q> varies. The gradient terms of F
therefore become

Comparing this with (5.3.3) it is at once clear that cp is related to the layer
displacement u:\q>\2 = q2\u\2 and B =\y/\2 ql/Mv.

The terms d<p/dx and

dcp/dy represent the tilt of the layers with respect to the director. If the
director orientation is not fixed, it is the relative tilt between the layers and
the director that should be considered and therefore (5.5.6) takes the
generalized form

^(^)\
2Mv\dz)

2M T '

(5.5.7)

where VT is the gradient operator in the plane of the layers. This equation

342

5. Smectic liquid crystals

is reminiscent of the Landau-Ginsburg expression(72) for the free energy of


superconductors; n corresponds to the vector potential A, V x A being the
local magnetic field.
The analogy may be extended further. By including the Frank elastic
free energy terms in (5.5.6), we may define (as already shown in 5.3.3) a
characteristic length X = (k/B)*. Making use of the condition dF/dy/0 = 0
and ignoring the difference between MT and Mv,

where k is an appropriate elastic constant. For a twist or bend, both of


which involve V x n, X may be interpreted as the depth to which the
distortion penetrates into the smectic material. Thus X is equivalent to the
penetration depth of the magnetic field in superconductors.
Above 7"*, we may ignore the term involving the fourth power in y/ and
from the equipartition theorem obtain in the usual manner
,

(5.5.9)

where the half-widths or the associated coherence lengths may be related to


Mv, MT and a as follows :
2=l/2MFa,
$1 = l/2MT(X.

(5.5.10)
(5.5.11)

Since a ^ 0 as T^ T*, <|^(#)|2>, ^y and L diverge near the transition. The


variation of {\y/(q)\2} reveals itself in the intensity of the Bragg scattering
(see fig. 5.2.6).
Critical divergence of the elastic constants
To discuss the critical behaviour of the twist and bend elastic constants in
the nematic phase, we observe that the Frank free energy expression should
include the contribution due to smectic short-range order:
F=\k\Vnf

+ Fs{yy\

(5.5.12)

where k is the usual nematic elastic constant in the absence of smectic-like


order, and Fs(y/) when averaged over all y/ is of the form
1V1

<f(Vn)2,

(5.5.13)

where we have replaced Sn by Vn and ignored the difference between Mv

5.5 The smectic A-nematic transition

343

and MT. For a correlated region of volume 3, it can be shown, using


(5.5.9), that in the mean field approximation
<M 2 >oc/; B 7yacf.

(5.5.14)

Thus from (5.5.12), (5.5.13) and (5.5.14), the effective elastic constant for
twist or bend will be k + Sk, where
Skoc^t

(5.5.15)

Since the coherence lengths diverge rapidly near the A-N transition, the
elastic constants for twist and bend should also show critical behaviour. In
the mean field approximation
oc ( r - r * ) ~ 1 / 2 .

(5.5.16)

However, invoking the analogy with superfluids, de Gennes predicted


oc (T- T*)~2/\

(5.5.17)

Using scaling arguments, Jahnig and Brochard(73) have shown that in


the anisotropic case
<J* M ocft/,,
(5.5.18)
<**,

(5.5.19)

and that, below T* (i.e., in the smectic phase),


Bazit/Zl,

(5.5.20)

Dccl/r

(5.5.21)

The theory continues to be valid even if the A-N transition is weakly first
order, except that T* represents a hypothetical second order transition
temperature slightly below 7^N.
The first X-ray scattering experiments of McMillan(74) on /?-n-octyloxybenzylidene-/?-toluidine, which exhibits a first order A-N transition,
agreed with the mean field theory. On the other hand, CBOOA showed
appreciable anisotropy in the temperature variation of the longitudinal
and transverse coherence lengths. Later studies on a number of materials
have confirmed that anisotropy is a general feature exhibited by all of
them.
The increases in k22 and fc33 in the nematic phase near the A-N transition
over and above that given by the usual s2 law are now well established. Fig.
5.5.1 presents the data of Cheung, Meyer and Gruler(75) for CBOOA; &33

5. Smectic liquid crystals

344
70 r-

60
50
40

Ax
30

20

10
,, 1

0.1

1.0

r-r A

10

30

Fig. 5.5.1. The temperature dependence of the splay and bend elastic constants, kxl
(crosses) and k33 (circles) respectively, in the nematic phase of CBOOA prior to the
smectic A-nematic transition. The values are plotted as &../A/ where A/ is the
anisotropy of the diamagnetic susceptibility of the nematic. The full line shows the
order parameter s, normalized to fit fcn/A/ at higher temperatures. The splay
constant klx deviates only very slightly from ordinary nematic behaviour while the
bend constant kZ2> exhibitsa critical increase near 7^N due(75)to pretransition
fluctuations. (After Cheung, Meyer and Gruler. )

diverges rapidly while fcn exhibits normal behaviour. Similarly, it has been
verified that the layer compressibility constant B shows a critical variation
in the vicinity of the transition.
Critical divergence of the viscosity coefficients
Certain nematic viscosity coefficients also exhibit critical behaviour. The
origin of this effect may be explained physically as follows. Take, for
example, the frictional torque associated with the twist viscosity coefficient
Xx defined by (3.3.14). The formation of smectic clusters results in an extra
torque due to the flow of the liquid normal to the smectic planes. This extra
torque increases as the clusters become larger and longer lived, and in
consequence there is a net enhancement of the effective Xx. To estimate this
enhancement/ 73 ' 76) consider a slowly rotating magnetic field having an
angular velocity Q. We have seen in 3.6.2 that the director follows the field
at a constant inclination, the torque being given by 2XQ. Now, the
layered structure of the smectic-like regions makes an additional con-

5.5 The smectic A-nematic transition

345

tribution to the torque, say F s . If the angle between the layer normal and
the director is 0, then
6= QV
(5.5.22)
where TV is the relaxation time of \y/\. The torque Ts may be derived from
(5.5.13) with 0 = <jVn:
M

< w 2 >

(5 5 23)

- -

Using the value of \i//\2 from (5.5.14), the excess viscosity


8X1 oc rji.

(5.5.24)

The temperature dependence of SX1 depends on both z and . The


theory has been worked out in detail by McMillan(76) using the mean field
approximation and by Brochard(73) assuming dynamical scaling laws. The
critical exponents for the divergence of the visosity as predicted by the two
theories are different:
S^ ~ (T- r*)- 5 0
SXX ~ (T- r*)--

33

(mean

field)

(dynamical scaling).

(5.5.25)
(5.5.26)

Another difference between these two approaches lies in the behaviour of


the director relaxation time r. For example, for a twist deformation we
know that (3.8.1)
T~1 =
-k22q2/Al.
It is seen that in the mean field approach T is independent of temperature
since near T* both k22 and lx diverge similarly. On the other hand, with
helium-like exponents
T-1 oc (T- r*)- 3 3 An anomalous increase of Xx has been observed but the principal
experimental difficulty in determining the critical exponent accurately is
that the normal nematic viscosity (in the absence of correlations) is itself
strongly temperature dependent and the anomalous part forms only a
relatively small contribution. However, careful measurements(77) have
indicated a mean field behaviour.
Comparison between theory and experiment
Halperin, Lubensky and Ma(78) have argued that when director fluctuations are taken into account, the A-N transition should be at least

346

5. Smectic liquid crystals

Reduced temperature [t = (T/Tc) -1]


Fig. 5.5.2. Temperature dependence of the susceptibility (x) and the correlation
lengths ^ and for smectic A short-range order in the nematic phase of
butoxybenzylidene-octylaniline. Open circles were deduced from X-ray scattering,
and filled circles from light scattering. The smectic densityowavevector q0 remains
essentially constant, but f(| increases from about 100 A to 1.5 jum over the
temperature range shown in the diagram. Here Tc represents the A-N transition
point. (After Litster and Birgeneau.(88))

weakly first order. The shift in the transition point due to this effect,
relative to the second order transition point T*, was estimated to be of
the order of 10 mK. Whether or not the A-N transition can, in principle,
be of second order has been the subject of some discussion. (7981) From
high resolution experimental studies it appears that the second order
nature of the transition has been more or less established in at least a
few cases(82) (though Anisimov et al.m) have claimed to have observed
the 'Halperin-Lubensky-Ma effect' in three different systems). A
number of theories have been proposed, which have been reviewed in
authoritative articles,(84~6) but a precise description of the A-N transition
still remains elusive. We shall give a very brief summary of the current
situation.

347

5.5 The smectic A-nematic transition


Table 5.5.1. The A-N transition: summary of experimental data(S9
a

0.07
0.05
-0.007
0
0.15
0.15
0.2
0.22
0.31
0.45
0.50
-0.007
0.5

0.70
0.69
0.78
0.83
0.70
0.70
0.71
0.71
0.67
0.61
0.57
0.669
0.5

0.65
0.61
0.65
0.68
0.62
0.57
0.58
0.57
0.51
0.51
0.39
0.669
0.5

1.22
1.22
1.46
1.53
]1.30
1.31
1.32
1.31
11.26
]1.10
11.10
] .316
11.0

Compound
T8
T7
40.7
8S5
CBOOA
40.8
80CB
9S5
8CB
10S5
9CB
XY model
Tricritical

0.66
0.71
0.926
0.936
0.94
0.958
0.963
0.967
0.977
0.983
0.994

a + 2v + v|| Reference
1.93 + 0.15
2.06 + 0.15
2.07 + 0.1
2.19 + 0.16
2.09 + 0.14
1.99 + 0.18
2.07 + 0.17
2.01 0.18
2.00 + 0.13
2.08 + 0.18
1.85 + 0.18
2.0
2.0

92
92
89,93
94,95
96,97
98
96,99
95, 100
101,82
95, 100
100, 82
102, 103

Tn = 4-alkoxybenzoyloxy-4/-cyanostilbene n = 7,8
40.m = butyloxybenzylidene alkylaniline m = 7,8
nS5 = 4-n-pentylphenylthiol-4/-alkyloxybenzoate n = 8,9,10
Reduced temperature: t =
(T-TAN)/TAN.
Specific heat at constant pressure: Cv = A\t\~
Correlation length,
longitudinal (parallel to n): ^ = ^\t\'vw
transverse (perpendicular to n):

(5.5.27)

= \

Susceptibility: x = Xo\*\~

Anisotropic hyperscaling relation,


vy+2v_L + a = 2.

(5.5.28)

Measurements of the intensity and the width of the X-ray scattering


yield the susceptibility (x) and the correlation lengths () respectively.
Light scattering, which gives the divergence of the elastic constants, can
also be used to measure . The values of determined by the two methods
are in excellent agreement (fig. 5.5.2). Table 5.5.1 presents the relevant
exponents derived from the best available measurements for nine compounds. It is seen that the data for the different materials are by no means
'universal'.
Present theories of the A-N transition indicate two possibilities. Firstly,
that it belongs to the 'inverted' XY class(103) which leads to the critical
exponents(80)
a = -0.007,

= 0.67

and

y =1.32.

348

5. Smectic liquid crystals

Secondly, that it belongs to an anisotropic class with(104)


v(| = 2v,
(though some doubts have been raised as to the validity of this
relation(105)). Another proposal for the anisotropic class is that (106)
v,, = ivxy = 0.8,

v = lvxy = 0.53.

We note from Table 5.5.1 that v is anisotropic for all the materials, v|( v
being about 0.13. In no case is v(| = 2v. Also, all the materials satisfy the
anisotropic hyperscaling relation (5.5.28) to within experimental limits.
For two compounds, 407 and 8S5, a agrees with the predicted value of
-0.007, but the vs do not. Brisbin et al.{95) have suggested that the A-N
transition for some of the compounds, e.g., 10S5 and 9CB, is near a
tricritical point as the values of TAN/TNl are high for these compounds. This
may well be true, for y, a and (v)( + 2v )/3 for both 10S5 and 9CB are close
to the tricritical values of 1.0, 0.5 and 0.5 respectively. But this still leaves
the question of anisotropy unexplained.
From (5.5.20) and (5.5.21) it is seen that B oc t*9 where ^ = 2 v - v,, and
D oc fw. If V|| = 2v1? B should be finite at the transition temperature.
However, experimentally/ 107'108) it appears that B at the transition is
almost vanishingly small within experimental limits. Few measurements
are available on D to draw any definite conclusions. In any case, as pointed
out earlier, the exponents are neither universal nor do they agree with the
predictions of any of the theoretical models. Vithana et al.ao9) have
suggested that the widely differing values of the exponents for the different
compounds may be a consequence of the fact that one is measuring
effective values associated with crossover effects between the XY class and
a tricritical point. A further complication is that the experiments of
Evans-Lutterodt et al.(92) appear to indicate that the occurrence of different
forms of smectic A (see 5.6) may have a bearing on the nature of the A-N
transition.
In summary, therefore, it would be fair to conclude that no theory
predicts all the exponents correctly for any of the systems. Thus the A-N
transition, which is probably the most extensively studied transition in
liquid crystals, still remains a major unsolved problem in condensed matter
physics.
The smectic A-cholesteric transition
Lubensky(110) has shown that, in principle, the smectic A-cholesteric
transition should always be of first order, in analogy with the behaviour of
a superconductor in an external magnetic field. The relative shift of the

5.5 The smectic A-nematic transition

349

0.40

0.35

0.30

0.25

72

76

80
84
Temperature (C)

Fig. 5.5.3. Pitch versus temperature in cholesteryl nonanoate prior to the smectic
A-cholesteric transition (74 C); the crosses are values obtained from observations
of the Grandjean-Cano walls and the circles from the wavelengths of maximum
reflexion. (After Kassubek and Meier.(111))
transition temperature AT/T* due to this effect is estimated to be ~ 10~3,
but no experimental study of the effect appears to have been reported.
Near the transition the free energy of the cholesteric may be written as
F =\k\2q2 k\2qQq + FJ<y/)

(5.5.29)

omitting the ' background' term, where q0 is the equilibrium value of the
twist per unit length in the absence of smectic-like short range order and q
the actual value. Therefore
(5.5.30)
Minimizing with respect to q, we get

/Coo I* Ofcn

(5.5.31)

which decreases rapidly as the temperature drops to the smectic A-cholesteric transition point, or, in other words, the pitch P = 2n/q increases.
Fig. 5.5.3 presents the temperature dependence of the pitch of cholesteryl
nonanoate, (111) a compound which shows the smectic A phase below 74 C.

350

5. Smectic liquid crystals

Pindak, Huang and Ho (112) have reported that dk22 oc r 0 6 7 for this
compound. The high temperature sensitivity of the pitch of such materials
has applications in thermography, as we have already seen in 4.9.

5.6 Smectic A polymorphism


5.6.1 Smectic A phases of strongly polar molecules
We have so far regarded the molecules to be symmetric and non-polar, and
therefore assumed the smectic A layer spacing (d) to be approximately
equal to the molecular length (/). However, if the molecules have a strong
longitudinal dipole moment there will be near-neighbour antiparallel
correlations, as discussed in 2.6.2, and this can result in subtle changes in
the structure.
The first evidence that more than one form of smectic A exists came from
an observation of Sigaud et a/.(113) of a phase transition that was detected
by calorimetry, but could not be observed optically. X-ray studies revealed
that this was a transition between two forms of the A phase: the higher
temperature phase was characterized by a pair of reflexions corresponding
to a layer spacing d ~ /, and the lower temperature one by two pairs of
reflexions corresponding to d ~ I and 2/ respectively. The former type of
smectic A is called the monolayer (Ax) phase and the latter the bilayer (A2)
phase. A third type which has a layer spacing intermediate between / and
2/, has been identified and is called the partially bilayer (Ad) phase. The
structures of these phases are represented schematically in fig. 5.6.1.
Prost,(114) and later Prost and Barois,(115) developed a phenomenological
theory of the A phases in which the free energy density is expressed in terms
of two coupled order parameters, p(r) describing the mass density wave,
and ^(r), the dipolar density wave arising from the antiparallel associations
of neighbouring molecules. The total free energy is written as
(5.6.1)
where
F2 = \{A2p2 + \B2p* + C2[(V2 + k22)p}\
F12 = A12pj> + D12<l>*p + D21p^ + \B12p^\

(5.6.3)
(5.6.4)

Ax = a-^TTj), A2 = a2(TT2), and T19 T2 are the mean field transition


temperatures for the two types of ordering.
The elastic terms C1[(V2 + ^ ) ^ ] 2 and C2[(V2 + k22)p]2 describe spatial
modulations of <f> and p with wavevectors k1(= 2n/l') and k2(= 2n/l)

5.6 Smectic A polymorphism

351
A^

in,
1
1

Ad

1
I
1

ri

i, - .
i, , , ,
ii

!!

!
I

I I P IIII

ii

ji

II

K vi

II

if
'!

>;

\i

II

I
i

ii

" "h i
.

l> '
i;
ii

11

i
i

Fig. 5.6.1. The different forms of smectic A composed of polar molecules:


monolayer A1? bilayer A2, partially bilayer A d and A phases, the last being similar
to A2 but with a transverse modulation of the structure.

respectively. The importance of the different coupling terms in (5.6.4)


depends on the relative values of/' and /. When /' 2/, it is readily shown
that A12p(r)(/>(r) and D21p*(r)(p(r) can be neglected. It then follows that
p ==
| 0, <f> = 0 corresponds to the monolayer Ax phase in which the dipoles
can point up or down with equal probability within each layer (fig. 5.6.1);
p =t= 0, </> =N 0 with k2 = 2k1 represents A2 which has an antiferroelectric type
of ordering; and <j> ^> p yields A d . The elastic terms favour modulations
of (f> and p with wavevectors kx and k2 which are incommensurate, while
the coupling terms favour a lock-in of kx and k2 such that 2kx = k2, which
as we have already seen leads to the A 2 phase. The system gains energy
through the coupling term at the expense of the elastic energy, and vice
versa, and the competition between these two causes ' frustration' in the
system. Different ways of satisfying these opposing tendencies result in the
A phase which has a transverse modulation of the structure (fig. 5.6.1) or
in a smectic A phase with two collinear incommensurate periodicities. Two
types of incommensurate phases are envisaged, the weakly coupled one in
which the two interpenetrating density waves are almost independent of
each other (fig. 5.6.2(a)) and the strongly coupled one in which the phases

5. Smectic liquid crystals

352

Soliton region

D3I

II'7T-1I
Locked phase
region

JIT"

nx.

Wi Wn'fi
Soliton region (a)

Fig. 5.6.2. Schematic representation of the molecular arrangement in (a) the weakly
coupled incommensurate smectic A, and (b) the strongly coupled incommensurate
smectic A. (After Prost and Barois. (115))

of the two waves are modulated (fig. 5.6.2(b)) exactly in accordance with
the theory proposed originally by Frank and Van der Merwe. (116) In the
latter case, the structure consists of large regions of A 2 separated
periodically by defect walls or phase solitons. Experimentally, however,
the occurrence of incommensurate smectic A phases has yet to be
established conclusively. (117~119)
Barois, Prost and Lubensky(120) have used the phenomenological model,
within the framework of the mean field theory, to construct phase diagrams
involving the polymorphic forms of the A phase. They have predicted three
kinds of critical points, the A d -A 2 critical point, the A x -A 2 tricritical point
and the A 1 -N-A d bicritical point, the salient features of which are
summarized below.
Since Ad and A2 have the same symmetry, it is to be expected that a first
order A d -A 2 phase boundary should terminate at a critical point. This has
been confirmed(121) in a binary liquid crystal system. Plots of wavevectors
against temperature for different concentrations (X) in the vicinity of the
critical point are shown in fig. 5.6.3. It is seen that the first order A d -A 2
transition manifests itself as a jump in the wavevector, accompanied by a
two-phase coexistence region. On increasing X the magnitude of the jump
as well as the width of the two-phase region decrease, until finally only a

5.6 Smectic A polymorphism

119

120

121

353

124

125

126

127

122

123

124

125

Temperature (C)
Fig. 5.6.3. Plots of the magnitudes of wavevectors q0 and q'Q (corresponding to A2
and Ad periodicities respectively) against temperature (7") for mixtures of 4-nundecyloxyphenyl-4/-(4//-cyanobenzyloxy)benzoate (11OPCBOB) and 4-n-nonyloxybiphenyl-4'-cyanobenzoate near the A d -A 2 critical point (CP). The mole
fractions of 11OPCBOB in the mixtures are (a) 0.550, (b) 0.571, (c) 0.597, (d) 0.619,
(e) 0.642, (/) 0.715, (g) 0.80 and (h) 1.0. The first order A d -A 2 transition is identified
for plots (a)-(d) by a two-phase region consisting of both Ad and A2 periodicities
(shown as closed circles) and also by a jump in the wavevector. As CP is
approached the two-phase region shrinks with an accompanying decrease in the
difference between q0 and q'o. At CP (plot (e)) only a vertical inflexion is seen.
Beyond CP, Ad evolves continuously into the A2 phase (plots (/)-(//)). The
behaviour is similar to that seen in the vicinity of the gas-liquid critical point.

(After reference 121.)

vertical inflexion point is seen. Beyond this critical point, Ad evolves


continuously into the A2 phase. The analogy with the density curves near
a gas-liquid critical point is obvious. However, Park et al.(122) have pointed
out that when the coupling of the order parameter with the elastic degrees
of freedom and with the fluctuations in the layer spacing is taken into
account, the A d -A 2 critical point may belong to a new universality class
with the upper marginal dimensionality dc = 6.
The transition between the Ax and A2 phases can be of first or second
order because of the exact doubling of the layer periodicity (see Landau
and Lifshitz,(25) p. 468). Thus the A x -A 2 phase boundary - say in a binary

5. Smectic liquid crystals

354

120

I
5

10
15
Concentration (molar % TBBA)

Fig. 5.6.4. Binary phase diagram for mixtures of 4-n-hexylphenyl-4/-cyanobenzoyloxybenzoate and terephthal-to-butylaniline (TBBA) showing the A^-Ag
tricritical point (TCP). The solid line represents the first order phase boundary
and the dashed lines represent the second order phase boundaries. (After Chan
et

tf/.(123))

system in the temperature (^-concentration (X) plane-can exhibit a


tricritical point. This has, in fact, been observed by the high resolution
X-ray experiments of Chan et al.(123) The phase diagram is shown in fig.
5.6.4. It was found that the correlation length exponents (associated with
the divergence of the A2-like fluctuations in the Ax phase) are isotropic, i.e.,
v|t = v = 0.74. This value is reasonably consistent with the Fisherrenormalized Ising exponent, indicating that the A x -A 2 transition probably
belongs to the Ising universality class.
The mean field theory predicts a bicritical point when the second order
N-A 1 and N-A d phase boundaries meet a first order A d -A 1 boundary.
However, when the effect of fluctuations is taken into account, the
existence of such a point becomes doubtful. (124) On the experimental side,
phase diagrams with an A 1 -N R -A d point have been reported, (125) (where
N R is the re-entrant nematic, see 5.6.2) the topology of these diagrams
resembling that of the magnetic bicritical point. (126) But high-resolution
experiments(127) carried out subsequently in the immediate vicinity of an

5.6 Smectic A polymorphism

52

53

54
Concentration, X (mol %)

355

56

Fig. 5.6.5. High resolution temperature-concentration (T-X) diagram for binary


mixtures of 4-n-octyloxy- and 4-n-decyloxy-phenyl-4 / -nitrobenzoyloxybenzoate in
the vicinity of the A 1 - N R - A d point. The solid lines denote first order phase
boundaries. The critical end point (CEP) for the A d - N R b o u n d a r y and the
approximate location of the tricritical point (TCP) for the A X - N R boundary are
indicated in the diagram. (After reference 127.)

A 1 -N R -A d point show that the bicritical point has, in fact, split into a
tricritical point (for the A X -N R boundary) and a critical end point (for
the A d -N R boundary, fig. 5.6.5). This new result, though predicted
theoretically for magnetic systems,(128) has not been envisaged in any of
the theories of frustrated smectics.
5,6.2 The phenomenon of re-entrance
When a compound exhibits both nematic and smectic phases, then, as a
rule, the nematic occurs at a higher temperature. Exceptions to the rule
were discovered by Cladis(129) in certain strongly polar materials. The first
observations were on binary mixtures of two cyano compounds: over a
range of composition the sequence of transitions on cooling was as follows :
iso -> N -> SA -> N R -> crystal,
where N stands for the usual nematic and N R for a second nematic, called

5. Smectic liquid crystals

356

1.4

1.0

0.6

Isotropic

Smectic A
0.2

50

70

90

110

Temperature (C)
Fig. 5.6.6. Experimental pressure-temperature diagram for 80CB showing the reentrant nematic phase. (After Cladis et al.a30))

C 9 H 19 O

OOC

OOC

Fig. 5.6.7. The structural formula of 4-nonyloxyphenyl-4 /-nitrobenzoyloxybenzoate, a compound that shows multiple re-entrant behaviour.

the re-entrant nematic, which appears at a lower temperature. The SA phase


occurring between two nematics was identified to be the partially bilayer
A d . Later, a similar effect was reported by Cladis et al.am in a pure
compound at elevated pressures (fig. 5.6.6).
These observations stimulated a great deal of interest, and much more

5.6 Smectic A polymorphism

(a)

357

(b)

Fig. 5.6.8. Schematic representation of (a) a dimer unit consisting of two


antiparallel molecules, (b) the mechanism of destabilization of the smectic A phase.
(After Cladis.(135))
complex examples of re-entrance have since been found. (125131) Probably
the most spectacular case is the compound (132) whose molecular structure
is shown in fig. 5.6.7, which exhibits three nematic, four smectic A and two
smectic C phases in the following sequence :<133)
iso -> N -> Ad -> N R -> Ad -> N R -> Ax -> C -> A2 -> C2 -> crystal,
where C and C2 are two different forms of S c (fig. 5.8.3). It turns out that
re-entrant behaviour and the closely related phenomenon of smectic A
polymorphism are both extremely sensitive to the molecular structure. For
example, the reversal of the sense of the longitudinal component of the
dipole moment of just one of the bridging groups of the molecule, relative
to that of the end group, may suppress completely the occurrence of the
A-A transition or re-entrance or both. (134)
From the molecular point of view, only an approximate qualitative
explanation of the phenomenon has been possible. Fig. 5.6.8 illustrates the
elements of a simple model proposed originally by Cladis. (135) The basic
idea underlying this model is that because of antiparallel correlations, the
molecules form dimers, which are assumed to be somewhat bulgy in the
middle (fig. 5.6.8 (a)). Once the smectic phase is formed, the bulgy parts are
lined up in a plane, but the alkyl chains cannot fill the rest of the space.
With increasing dimer formation (i.e., with decreasing temperature) and
also possibly with the stiffening of the end-chains, the packing becomes so
unfavourable that the A d phase is destabilized and the nematic re-enters
(fig. 5.6.8 (b)).
A number of other more elaborate models,(136~9) which take into account
attractive forces and hard core repulsions, (136) dipole-induced dipole
interactions (which, interestingly, can favour a parallel configuration at
small intermolecular separations (139)), etc., have been proposed sub-

358

5. Smectic liquid crystals

sequently, but as remarked earlier, only a qualitative description of the


effect has emerged until now.(140) The frustrated spin-gas model of Indekeu
and Berker(137) is noteworthy in that it is able to account for multiple reentrance of the kind shown by the compound of fig. 5.6.7 and the sensitive
dependence of re-entrant polymorphism on molecular chain length and
pressure.(133)
In principle re-entrant phases can occur in non-polar systems as
we U U36,i4i) Dowdi^D has discussed this possibility by treating the
molecules as hard rigid cores with semiflexible tails and interacting via
segmental hard repulsions. In the SA phase, the molecules are segregated
into layers. Calculations show that as the tails become more rigid at lower
temperatures the lamellar packing may become disadvantageous and the
SA phase can be destabilized. Experimental evidence of re-entrant phases in
non-polar systems has been presented by the Halle group.(142)
Another example of a special type of molecular interaction concerns the
'induced' smectic phase, i.e., the appearance of a smectic phase in a binary
mixture even though neither component shows this phase in the pure state.
This effect occurs most commonly in mixtures with one component having
a strongly polar terminal group and the other a non-polar terminal
group. (14356) Obviously, dipole-induced dipole interaction plays a part in
phase induction. There is also evidence of charge transfer complex
formation, the polar molecule acting as the acceptor and the other as the
donor. (143151157) However, phase induction has been observed in other
types of mixtures as well. (15860) For example, mixtures of two cyano
compounds have been found to give rise to an induced S c phase. Thus no
generalizations are possible as yet and the precise nature of the molecular
interactions and correlations responsible for promoting phase induction is
not clear.
5.6.3 Smectic A* or the twist grain boundary phase
A significant result that emerged from the phenomenological theory of the
A-N transition discussed in 5.5.2 was that the director n in smectic A
plays the same role as the magnetic vector potential in a superconductor.
Recognizing this analogy, de Gennes(68) predicted that a twist or bend
distortion should depress the A-N transition point, 7^N, relative to that of
the curvature-free sample. The nature of the phase diagram depends on
whether the material is of type I or type II, i.e., on whether the
Landau-Ginsburg parameter, i / , is less or greater than I / A / 2 , where X is
the penetration depth defined by (5.5.8) and the coherence length. If the

5.6 Smectic A polymorphism

359
-COOC*H(CH 3 )C 6 H 13

C 13 H 27 O

81.6 C
88.3 C
94.1 C
*>C*
* A*

33.7 C

s4 <

s,

43.7 C

Fig. 5.6.9. The structural formula of (R)- and (S)-l-methylheptyl 4'-(((4"-nalkoxyphenyl)propioloyl)oxy)biphenyl-4-carboxylates, which show the smectic A*
phase.

Fig. 5.6.10. The structure of smectic A* or the twist grain boundary phase.
A-N transition is of second order, or nearly so, the material may be
expected to be of type I.(69) Measurements(161) on CBOOA, which has a
nearly second order A-N transition, have confirmed that the variation of
7^N with the twist per unit length does indeed reproduce the theoretically
predicted trend for A/f < \/y/2.
For a type II material, de Gennes envisaged the possibility of a regular
network of dislocations being formed in smectic A under an imposed twist
or bend distortion. A model for such a network, similar to the Abrikosov
flux lattice phase(162) in a type II superconductor subjected to an external
magnetic field, was discussed by Renn and Lubensky.(163) Soon afterwards,
a phase consisting of a regular array of dislocations was discovered by
Goodby et al.am who designated it as smectic A*.
Smectic A* is composed of optically active molecules. The molecular
structure of one of the compounds studied by Goodby et #/.(164) and the
phase transitions occurring in it are given in fig. 5.6.9. The structure of
smectic A* is depicted schematically in fig. 5.6.10. Unlike smectic C* whose
helical axis is normal to the layers (see fig. 5.10.1) smectic A* has its helical
axis parallel to them. Hence the structure may be looked upon as a series
of smectic A blocks or grains separated by twist grain boundaries. The
director (which in this case is the layer normal) is rotated by a constant
angle on going from one grain to the next. From observations of the

360

5. Smectic liquid crystals

reflexion of circularly polarized light incident along the helical axis,(165) it


is found that the pitch is of the order of 1 jum and from X-ray
measurements(165) that the grain size is about 180 A. Renn and Lubensky
have called this phase the 'twist grain boundary' (TGB) phase. The
discovery of this phase has raised a number of fundamental questions of
considerable interest to condensed matter physicists.
5.7 The hexatic phase
(166)

Kosterlitz and Thouless


proposed that the unbinding of dislocation
pairs can lead to a continuous melting transition from a two-dimensional
solid to a two-dimensional liquid. Halperin and Nelson(167) then worked
out the details of a defect-mediated melting theory in two dimensions and
predicted the existence of an intermediate phase between the twodimensional solid and liquid. This phase, referred to as the hexatic phase,
has short-range positional order, but long-range 'bond-orientational'
order (bond, in this case, implying the line joining the molecular centres of
nearest neighbours). Its structure is illustrated in fig. 5.7.1. Based on this
idea, Birgeneau and Litster<168) suggested the possibility of a threedimensional liquid crystal phase consisting of a stack of two-dimensional
hexatic layers which interact to produce three-dimensional bond-orientational order. Experimental evidence of this type of phase was presented
by Leadbetter, Frost and Mazid,(169) but its existence was established
conclusively only later by the high resolution X-ray studies of Pindak
et al.(170) on free standing films. This phase, called the hexatic B phase,
occurs between SA (which is a fluid phase with negligible in-plane and
interlayer positional correlations) and a three-dimensional long-range
translationally ordered 'liquid crystal' (e.g., crystal B or E). It has threedimensional sixfold bond-orientational order, short-range in-plane positional correlation and no interlayer positional correlation.
Fig. 5.7.2 presents the experimental results of Pindak et al.(170) on free
standing films. The /-scan profile in the SA phase is a diffuse ring
independent of the angle /- As hexatic ordering develops, the ring exhibits
a sixfold modulation, which eventually breaks up into six diffuse spots.
The degree of hexatic ordering can be defined quantitatively(171) by
fitting the /-scans to a Fourier series of the form

Six) = I0\c0+ t

*ncos6,2(90-/)l + /bg,

(5.7.1)

where Co is a constant and Ihg the background intensity. The coefficients

5.7 The hexatic phase

361

(a)

Orientational long-range order, positional short-range order

0'

(b)

Orientational short-range order, positional short-range order

Fig. 5.7.1. (a) The structure of the hexatic phase of a two-dimensional lattice. The
orientation of the lattice vectors a and b is preserved over a long range, but the
molecular positions are correlated only over a short distance fp. (b) The
orientational order as well as the positional order are short range. (After Litster
and Birgeneau.(88))
C6w measure the degree of 6-fold ordering. An application of (5.7.1) to
study the bond-orientational order across the Sj-S c transition(172) will be
described in 5.8.1.
The hexatic B-S A transition is expected to belong to the XY universality
class,(171) but high resolution calorimetric investigations(173) appear to
indicate that this may not be the case. Tilted analogues of hexatic B,
namely SF and SI5 have also been identified.(174)

362

5. Smectic liquid crystals


(a)

4 -

7=68.1 C

I o
15
(c)

10

r=63.9C
ir/6

Rotation angle, x
Fig. 5.7.2. A 60 segment of a/-scan about an axis normal to the smectic layers of
the intensity of X-ray scattering from n-hexyl-4'-n-pentyloxybiphenyl-4-carboxylate exhibiting the smectic A-hexatic B phase transition, (a) The scattering in
the A phase is a diffuse ring independent of/. For temperatures below the smectic
A-hexatic B transition the ring develops a sixfold modulation (Z>), which
eventually breaks up into six diffuse spots (c). (From Pindak et al.ai0))

5.8 Smectic C
5.8.1 Description of the structure
In S c the molecules are disordered within the layers, as in SA, but inclined
with respect to the layer normal. The structure has biaxial symmetry of the
monoclinic class, with a plane of symmetry as indicated in fig. 5.8.1. The
origin of the biaxiality may be explained as follows. The tilt angle 6 is
directly coupled with the layer thickness whereas the azimuthal angle <p is
not. Therefore, at any given temperature, the amplitudes of the 6
oscillations of the director are small compared with those of the (/>

5.8 Smectic C

363

^ Symmetry plane
' Molecular axis

'///////////////////A

V///////////////////,
V////////////////A

Fig. 5.8.1. The structure of smectic C.


oscillations, with the result that the uniaxial symmetry about the mean
molecular direction disappears, and there exists only a plane of symmetry.
Because of this, and also because of possible anisotropic polarization field
effects, S c is optically biaxial. The optic axial angle is generally quite small
(-10).
S c is often followed by SA at a higher temperature, and in such a case the
tilt angle decreases to zero gradually with rise of temperature. If the C-A
transition is of second order, the tilt 6 decreases smoothly to zero, whereas
if it is of first order 6 drops abruptly from a finite value to zero at the
transition point.
If S c undergoes a transition directly to the nematic phase, 6 is generally
found to be temperature independent and usually about 45. According to
the Landau rules, the C-N transition can be continuous, but when
fluctuations(175) are taken into account it is predicted to be of first order.(176)
Experimentally, only first order C-N transitions have been observed.
Some compounds exhibit transitions from S c to the isotropic phase.
Interestingly, a slight increase of 6 with increasing temperature has been
reported for two such compounds.(177)
As in the SA phase, Sc lacks true long-range translational order because
of the Peierls-Landau instability. Theoretically, the existence of molecular
tilt implies that there must be a certain degree of bond-orientational order
in the S c phase. (171178) This has been verified experimentally by a high
resolution synchrotron X-ray study of the transition from S c to Sx in a
monodomain freely suspended film.(172) It will be recalled that Sx is a tilted
hexatic phase. The bond-orientational order parameter C6 (as defined in
(5.7.1), but appropriately modified to allow for the fact that the molecules
are tilted) is plotted as a function of temperature infig.5.8.2. The weak
bond-orientational order in S c evolves continuously into ST showing

364

5. Smectic liquid crystals

"2
o

75

77
Temperature (C)

79

81

Fig. 5.8.2. The hexatic order parameter C6 as a function of temperature in the S c


and Sj phases of racemic 4-(2-methylbutyl) phenyl 4 / -(octyloxy)-(l,r)-biphenyl-4carboxylate. The weak bond-orientational order in S c evolves continuously into Sx
showing that thermodynamically the two phases are not distinct in this compound.
(After Brock et a/.(172))

that thermodynamically the two phases are not distinct. Tilt angle
measurements (179) (using X-ray methods) have confirmed that there is
indeed a continuous evolution of S c into SI5 as the temperature is lowered.
Generally speaking, compounds exhibiting the S c phase have transverse
components of permanent electric dipole moments. A number of molecular
statistical models (including hard rod theories for systems composed of
oblique cylinders) have been developed. (1808) Goossens(189) has proposed a
model composed of ellipsoidal molecules with attractive interactions
arising from anisotropic dispersion forces as well as permanent quadrupole
moments. His calculations show that the interaction between the permanent quadrupole moments can produce a tilting of the molecules, but a
detailed comparison of the predictions with experimental data has yet to be
made.
When the molecule has a strong longitudinal dipole moment, several
modifications of the S c structure have been identified. They are shown in
fig. 5.8.3, and as can be seen they are closely analogous to the SA phases of
polar molecules (fig. 5.6.1). However, the theoretical situation for S c is not
as clear as for SA. Another modification of S c is the chiral S c or the Sc*
phase which will be discussed in 5.10.

5.8 Smectic C

365

Fig. 5.8.3. The different forms of smectic C composed of polar molecules:


monolayer C1? bilayer C2, partially bilayer Cd and C phases, the last being similar
to C2 but with a transverse modulation of the structure. These are analogous to A1?
A2, Ad and A phases depicted in fig. 5.6.1.

5.8.2 Continuum theory of smectic C


We define a unit vector c to represent the preferred orientation of the
projection of the molecules on the basal (xy) plane. This vector is referred
to as the c-director. It is clear that c is somewhat analogous to the nematic
director n in a homogeneously aligned sample. For example, as already
pointed out, the orientational fluctuations of c (in the xy plane) can be
large, and the S c phase appears quite turbid in certain geometries. Also S c
exhibits schlieren textures similar to, though not exactly like, those seen in
a nematic. Again, in principle, a Freedericksz transition should be
observable in S c .
To construct a continuum theory of S c , we have to take into account
firstly, the orientational fluctuations of the director about the layer normal
(z axis), and secondly, as in smectic A, the distortions of the layers
themselves. Expressions for the former were given by Saupe, (190191) but the
complete theory including the latter contributions and the coupling
between the two was derived by the Orsay group.(192) We chose a cartesian

366

5. Smectic liquid crystals

(arbitrary units)

Fig. 5.8.4. Plot of qxl* versus qzl^ from light scattering measurements in the
smectic C phase of di(4-n-decyloxybenzal)-2-chloro-l-4-phenylene diamine. The
wavevector q relative to the layers is indicated in the inset at the top left-hand
corner. The experimental points lie on an ellipse as expected theoretically. However,
the minor axis of the ellipse does not coincide exactly with the molecular axis,
which is assumed to be inclined at 45 to the layer normal. (After Galerne et a/.(193))
coordinate system such that the projection of the mean molecular direction
on the basal (xy) plane is along x. If the layer displacement along z is
represented by w, we observe that
du/dy = lx,

where Q x and Qy represent rotations about the x and y axes respectively.


Therefore
(5.8.1)

5.8 Smectic C

367

Making use of (5.8.1), and assuming the layers to be incompressible, the


free energy of elastic distortion may be written in the form

Here the A terms describe curvature distortions of the smectic planes, the
B terms the distortions of the director when the smectic planes are
unperturbed, and the C terms the coupling between these two types of
distortions. All the coefficients are approximately of the same order of
magnitude as the nematic elastic constants. A term of the type \B(<du/dz)2
may also be included to allow for the compression of the layers, but we
shall neglect it in the present discussion.
The fluctuations of the director are evidently related to fluctuations in Q.
From (5.8.2) and the equipartition theorem, we obtain for a general
wavevector q

Light scattering studies using monodomain samples of smectic C have


been reported by Galerne et al.(193) They have confirmed that the intensity
of scattering arising from the director fluctuations in the vertical
(scattering) plane (the ke, k'e configuration, e standing for extraordinary) is
extremely weak, whereas it is quite large due to fluctuations normal to the
scattering plane (the ke,k'o or ko,k'e configuration). Now, for the ke,k'o
configuration, we may write
(BlCos2 6 + B2sm20 + 2Blssin0 cos 6) q2I=B(6)q2I=

const,

(5.8.4)

where 6 = tan" 1 (qz/qx) and / the intensity. Thus a plot of qx I* versus qz I*


should give an ellipse. This has been verified to be true (fig. 5.8.4). The
damping time of these modes also shows a similar angular dependence.
From an analysis of the data Galerne et al have estimated that the
viscosity coefficients are an order of magnitude larger than for a typical
nematic.
Again, as in SA, the undulation mode (qz = 0) may be expected to make
an important contribution to the scattering.
The hydrodynamical properties of S c have been discussed by Martin,

5. Smectic liquid crystals

0)
(ii)
Fig. 5.8.5. Disclinations in the odirector field of smectic C. (a) 5 = 1 wedge
disclination with a radial configuration: (i) sink, (ii) source and (iii) meridian
section of (i). (b) s = 1 wedge disclination with a circular configuration: (i) vortex,
(ii) antivortex and (iii) meridian section of (i); the nails signifying that the director
is tilted with respect to the plane of the paper, (c) (i) s = 1 wedge disclination and
(ii) s = 1 twist disclination.

Parodi and Pershan (MPP), (39) and by others.(194) Leslie, Stewart and
Nakagawa (195) have formulated a general non-linear theory, which in the
static case becomes identical with that proposed by the Orsay group. The
full implications of the dynamical aspects of this theory and its comparison
with the MPP theory have yet to be worked out.

5.8 Smectic C

369

5.8.3 Defects in smectic C


Disclinations in the c-director field
As mentioned earlier, schlieren textures, very similar to those of a nematic,
are seen in planar samples of S c . However, since c and c correspond to
opposite tilts, they are not equivalent; consequently, as explained in 3.5.1,
only defects of integral strength (i.e., four brush disclinations) can occur in
S c . Moreover, as the c-director is confined to the xy plane, the usual escape
mechanism for nematic disclinations of unit strength is prohibited in this
case (see, however, 5.8.4).
The polarity of the c-director introduces some additional complications.
For example, the s = 1 disclination with a radial configuration can now
have two independent solutions, the source and the sink. Also, s = 1 no
longer has four-fold symmetry (fig. 5.8.5). The structures are closely
similar to defects in ferromagnetic planar spin systems.
The c-director field can also have twist disclinations, the structures of
which are like those shown in fig. 3.5.8, except that the director is polar. (196)
The director pattern for s = 1, c = 0 is shown in fig. 5.8.5.
S c can have lattice disclinations as well, but they are perfect and
energetically favoured only along the twofold axis (which is normal to the
plane containing the layer normal and the c-director). (197) Focal conic
textures are also seen, though they are always accompanied by additional
singular lines of disclination, which arise because of the molecular tilt (fig.
5.8.6).
Single edge dislocations have been observed by simple optical microscopic techniques(199) at temperatures near the C-A transition. Fig. 5.8.7
shows a photograph of an array of parallel dislocations in a sample
prepared in the form of a wedge. The reason for the appearance of these
lines is the following. We know that around an edge dislocation in SA there
is a stress field, the components of which are given by (5.4.6)-(5.4.8). On
one side of the dislocation line (the side having an extra half-plane) there
is compression, while on the other side there is dilatation. This can give rise
to a phase transition, as discussed in 5.8.4. The sample being at a
temperature close to the C-A transition point, one may expect that in the
region of compression the material has already undergone transition to S c ,
while in the region of dilatation it still remains S A. As the phase change is
accompanied by a molecular tilt it immediately reveals itself optically.

5. Smectic liquid crystals

370

(a)

(b)

Fig. 5.8.6. Disposition of the molecules and layers in elliptical domains in (a)
smectic A and {b) smectic C. The mismatch of the molecular orientations in smectic
C gives rise to additional singular lines. (Bourdon, Sommeria and Kleman. (198))

Fig. 5.8.7. Edge dislocations in smectic C at a temperature very close to the smectic
C-smectic A transition. The sample, prepared in the form of a wedge, consists of
two domains of opposite molecular tilts. (Lagerwall and Stebler. (199))

5.8.4 The smectic C-smectic A transition


The order parameter for the C-A transition has two components and may
be written as(69)
).
(5.8.5)
This at once brings out the analogy with the superfluid transition in
helium (the XY model).(103) Therefore, for T < TCA one expects the order
parameter
0=\\~\t\'9
/f = 0.35,
where t = (T- TCA)/TCA, and a = -0.007, y = 1.32 and v = 0.67, a, y and
v having the same meanings as in (5.5.27).
Though evidence of helium-like critical behaviour has been reported, (200)
high resolution experiments indicate that the transition is of the mean field
type. Safinya et al.(201) have argued that the bare correlation length of the

5.8 Smectic C

371

tilt fluctuations is so large that the true critical region, as defined by the
Ginsburg criterion, should be very small and only a mean field behaviour
should be observable. Huang and Viner(202) and Birgeneau et al.(20S) have
demonstrated that this is, in fact, the case, and that the transition is well
described by a mean field expression but with an unusually large sixth
order term.
We write the free energy as
F= Fo + atO2 + b6* + c6% + . . . .

(5.8.6)

Minimizing with respect to 6

0 = 0,

where

T>TCA
3A1/2

1--J

-11/2

- l j , T<TCA,

(5.8.7)

R = (b/3cr\
t0 = b*/ac.
Substituting (5.8.7) in (5.8.6) and making use of the standard relation
we get
fC
C,=

T>T

CA

"
l.C 0 H-^ii(i m 7 J

(5.8.8)

i < i CA

where
A = ^3/2 /[2(3c) 1/2 r3/A2],
x = |^,

r>rCA.

(5.8.9)

The parameter /0 represents the full width at half maximum of the excess
specific heat curve. It also represents the cross-over temperature from
mean field to tricritical behaviour, for
9 - t1/2, t < t0 (mean field)
6 ~ J1/4,

t > t0 (tricritical).

Fig. 5.8.8 presents the experimental data for the tilt angle, specific heat,
and the susceptibility (the last being measured from the magnetoclinic
effect, the magnetic analogue of the electroclinic effect, see 5.10.1) along
with the curves fitted to (5.8.7), (5.8.8) and (5.8.9). It is seen that the
agreement is very satisfactory.

5. Smectic liquid crystals

372

1.0

0.6

-0.2

0.2

1.0

T-TC{K)
Fig. 5.8.8. (a) Tilt angle 6 versus temperature (open circles and filled circles
representing two different determinations) in butyloxybenzylidene heptylaniline
(40.7). The solid line is a fit to (5.8.7) with t0 = 1.3 x 10"3. The triangles are the
reciprocal of the susceptibility / in arbitrary units measured for two different
samples, (b) Heat capacity near the C-A transition in 40.7. The dashed curve is the
background scaled from 40.8 and the solid line is the fit to (5.8.8). (After Birgeneau
et a/.(203))

C-A transition induced by layer compression


Ribotta, Meyer and Durand (204) observed that a compression applied
normal to the layers of SA induces a transition to S c when the stress exceeds
a threshold value. The effect is particularly easy to observe very near T CA.
If s is the strain normal to the layers, the energy density may be written as
F = \B(s2 + 2ocs92) + a62 + b0\

(5.8.10)

where B is the elastic constant for compression. It is assumed that the


reduced layer thickness due to the tilting of the molecules is dilaO 2),
where d is the normal layer thickness and a a constant which depends on

5.8 Smectic C

373

the molecular axial ratio. For a second order C-A transition b > 0, and in
the high temperature phase a > 0, and 6 = 0. Clearly for a large enough
compression (i.e., large negative s), there is a finite tilt angle
0 = [-(a + asB)/2b]*

(5.8.11)

when s is greater than a threshold value


sth = -a/a/l.

(5.8.12)

C-A transition near the core of a disclination


Let us suppose that the C-A transition is continuous. Near TCA, i.e., for
small 0, one may write the free energy density as(69)

(5.8.13)
where a = oc0(T TCA), k119 k22 and A:33 are the elastic constants for in-plane
distortions, nx = 0cos(f>9 ny = 0sin(/> and nz&\. Assuming the oneconstant approximation and that the distortions are such that there is no
z dependence of 0 and <fi, the equations of equilibrium become
\k[V20-0(V(f>)2]-(x0-2p03

=0

(5.8.14)

and
= 0.

(5.8.15)

Equation (5.8.15) allows solutions of the type


cj) = Nco,

where co = tan -1 (y/x) and N is an integer. These describe the disclinations


of integral strength in the odirector (nx,ny) field. However, it is seen from
(5.8.14) that 0 depends on r, the distance from the centre of the singularity.
The variation of 0 with r (for \N\ = 1) is given by(205)
/ 2 ) = o
2

'

(5 8 i6)

- -

where / = 0/0o, = r/0, r = (x +y )*, 0O is the tilt angle at r-> oo, and
o = { k/2(xf is the coherence length. Equation (5.8.16) is of the same
form as the Ginsburg-Pitaevskii equation for a quantized vortex filament
in a superfluid or a magnetic fluxoid in a type II superconductor. The tilt
angle drops to zero at the centre of the singularity, most of the variation
taking place over a distance of the order of <J0 from the centre (fig. 5.8.9).

374

5. Smectic liquid crystals

This result appears to account for the experimental observation (206) that
there is an appreciable region near the centre of the + 1 singularity over
which 6 is very nearly zero, i.e., the material is smectic A-like.
5.9 The nematic-smectic A-smectic C multicritical point
The nematic-smectic A-smectic C (NAC) point is the point of intersection
of the N-A, A-C and N - C phase boundaries in a thermodynamic plane say, e.g. in the temperature (T) against pressure (P) or concentration (X)
diagram. It is termed a multicritical point because close to it all three phase
transitions are continuous, and at the point itself the three phases are
indistinguishable. (207) It has been observed in the T-X diagram of binary
liquid crystal mixtures(208) as well as in the P-T diagram of a singlecomponent liquid crystal system.(209) High resolution studies have been
made in both T-X(210) and P-T(209) planes, and it is now well established
that the topology of the phase diagram in the vicinity of this point exhibits
universal behaviour (fig. 5.9.1).
Another type of multicritical point, namely the re-entrant nematicsmectic A-smectic C (RNCA) point has also been observed. In the first
studies,(211~13) it was noted that the singularities in the phase boundaries of
the kind seen near the NAC phase were conspicuously absent near the
point. However, it was demonstrated subsequently (214) that this was
because the RNAC point was far away from the N-I transition, so much
so that the Brazovskii fluctuations had a negligible influence near the
multicritical point and hence the critical region was unobservably small. By
investigating different binary systems and 'tuning' the experimentally
observable critical region it was shown that singularities do appear in the
immediate vicinity of the RNCA point as well. Analysis of the topology of
these phase diagrams confirmed that the RNCA point also exhibits the
same universal behaviour.
Earlier, Hornreich, Luban and Shtrikman (215) had proposed a new type
of higher order critical point, which they called the Lifshitz point, by
considering a Landau-Ginsburg free energy expansion in terms of a scalar
order parameter M for a magnetic system
F = ^ 2 M 2 + 4 M 4 + 6 M 6 + ... + c1(VM)2 + c2(V2M)2.

(5.9.1)

The phase diagram involves the ferromagnetic, paramagnetic and helicoidal phases, the last being the modulated form of the ordered phase
described by the gradient terms. The Lifshitz point occurs when the
coefficients a2 and cx vanish, and is characterized by the fact that the

5.9 The nematic-smectic A-smectic C multicritical point

375

Fig. 5.8.9. Variation of the superfluid order parameter with distance from the
centre of a quantized vortex filament as given by the Ginsburg-Pitaevskii equation
(5.8.16). The same curve describes the variation of the tilt angle in the core region
of the smectic C disclination.(206)
wavevector of the modulated phase increases from zero as the system
moves away from this point. Soon afterwards Chen and Lubensky(216)
predicted that the point at which the nematic, smectic A and smectic C
phases meet and become identical is a possible realization of a Lifshitz
point in a liquid crystal system. A noteworthy feature of their model is that
there is a single order parameter defined as
.8exp(ikT)/>(k)d8fc,

(5.9.2)
-/.
where /?(k) is the Fourier transform of the centre of mass density p(r). The
domain of integration D takes into account the fact that the X-ray
scattering in the S c phase will be a ring defined by (qp qcos <fi, qsin (/>). The
general form of the free energy is written as
87T

C2
+ D[V2y/-n

(n

3
+ Fel d r,

(5.9.3)

where Fel is the Frank elastic energy for distortions of the nematic director.
The coefficients C , D and D are positive and a = ao(T 7^ A )/7^ A . The
type of fluctuations in the nematic phase is thus determined by the
coefficient C , i.e.,
if C > 0, SA-like fluctuations
if CL < 0, Sc-like fluctuations.

5. Smectic liquid crystals

376
D

4 3 -

<>

1-

0 /DvM

MX

-1
-2

X-

X*A*

(4 data sets)
l

(a)

_2

*\
1

x\

-1

151.4

150.6

149.8

(b)

275

295
305
Pressure (bar)

315

Fig. 5.9.1. For legend see facing page.

325

5.9 The nematic-smectic A-smectic C multicritical point

311

Hence the CL = 0 point is a fluctuation cross-over point. The locus of such


points (a > 0, C = 0) in a phase diagram, say the T-X diagram, gives the
Lifshitz line, and the point where both a and CL are zero is the Lifshitz
point.
The X-ray scattering intensity in the nematic phase can be obtained by
applying the equipartition theorem to (5.9.3):
k T
For C > 0 this gives a peak at k = q]{ with a Lorentzian profile and
corresponds to SA-like fluctuations, and
q, = (q/2/),,)*.
For C < 0, (5.9.4) can be rewritten as
k T
where the transverse wavevector

and

q\ = \C\/2D9
a = a'0(T-T

Hence, when C is negative, I(k) has a maximum on the two rings


k = (qpqcos(/>,qsm(/>) as expected for Sc-like fluctuations in the
nematic phase.
At the Lifshitz point 7^N = 7^N and C = 0, and (5.9.4) reduces to

Fig. 5.9.1. Topology of phase diagrams in the vicinity of the nematic-smectic


A-smectic C (NAC) multicritical point, (a) The temperature-concentration (T-X)
data for four binary liquid crystal systems: (1) 4-n-pentylphenylthiol-4 /-nonylbenzoate and 7S5, (2) 4-propionylphenyl-trans-(4-n-pentyl)cyclohexane carboxylate (XC) and 4-n-octyloxyphenyl-4 /-n-octyloxybenzoate, (3) XC and 7S5 and
(4) 80CB and 2/?-n-heptyloxybenzylidene aminofluorenone (Brisbin et al.(210)); (b)
the pressure-temperature (PT) data for a single component system: 4-nheptacylphenyl-4/-(4//-cyanobenzoyloxybenzoate). Analysis of the phase boundaries yields identical exponents for both the T-X and PT diagrams showing that
the NAC point exhibits universal behaviour. (After reference 209.)

378

5. Smectic liquid crystals

The transverse mass density fluctuations then become purely quartic. Thus
the model makes the important prediction that there will be a change from
Lorentzian to quartic in the scattering profile near the Lifshitz point.
All these predictions are generally borne out by high resolution X(208 22 2)
' - and light scattering studies. (223"5)
ray( 2i7-i9) calorimetric
Near the A-N transition, the present theory leads to the same
conclusions as de Gennes's model discussed in 5.5.2; the divergent
contributions to k22 and k33 in the nematic phase are given by (5.5.18) and
(5.5.19). Near the C-N transition, all three Frank constants diverge:
T

(5.9.7)
k T
At the Lifshitz point itself, there is a drastic change and Sk33 oc , while dklx
and Sk22 oc In .
Grinstein and Toner(226) applied the renormalization group technique
and presented a dislocation loop model for the NAC point. The most
striking conclusion of this theory is that a biaxial nematic phase N b should
intervene between the N and S c phases. Thus four phases, N, N b , SA and
S c , meet at a point giving rise to a tetracritical point topology. A
fluctuation corrected mean field theory developed by Lubensky(227)
supports this prediction, but points out that it may be difficult to detect the
presence of the N b phase through X-ray and calorimetric measurements.
All experimental studies but one have so far failed to detect the N b phase.
The one study is by Wen, Garland and Wand(228) whose high resolution
specific heat measurements have revealed anomalous variations near the
C-N transition very close to the NAC point, which they suggest may be
related to biaxial fluctuations. An attempt has been made to explain the
universal topology of the phase diagram in the vicinity of this multicritical
point.(258)
5.10 Ferroelectric liquid crystals
5.70.7 The properties of smectic C*
An interesting modification of smectic C is the Sc* phase, which has a twist
axis normal to the layers (fig. 5.10.1 (a)). The possibility of such a structure
being formed by the addition of optically active molecules to the ordinary

5.10 Ferroelectric liquid crystals

\\L

urn ii II

11/111/111]

WWWY
\

379

Helical
pitch, p

1111111111
miiiiui
inn IIui

mmmi

IIIIIIIIII
im/yini

iiii/wmr

mil! mi'

{a)

(b)

Fig. 5.10.1. (a) Helicoidal structure of the ferroelectric smectic C* phase, (b) a
' poled' sample with the helix unwound by an electric field applied normal to the
helical axis.
smectic C was envisaged by Saupe,(190) but the first identification of this
chiral phase in pure optically active compounds was by Helfrich and
Ok (229) J^Q optical properties of this structure are similar to those of a
cholesteric, though there are some obvious differences. In the cholesteric,
the local dielectric properties can be represented by an ellipsoid of
revolution with ea 4= sb = sc, the principal axis Oc being parallel to the
helical axis Oz. In Sc* the local dielectric ellipsoid is triaxial, ea =t= eb 4= sc,
with Oa making an angle 6 with Oz. As far as light propagation along the
twist axis is concerned, Sc* is identical to the cholesteric, but at oblique
incidence some additional features may be expected/ 230 ' 231)
Meyer et al.(232) demonstrated that the Sc* phase has ferroelectric
properties. The first studies were carried out on DOBAMBC (pdecyloxybenzylidene-/7/-amino-2-methylbutylcinnamate) which shows the
following transitions
crystal -

isotropic

Both Sc* and Sj* (the chiral forms of S c and Sx) were shown to be
ferroelectric.

380

5. Smectic liquid crystals

In SA the molecules are upright, and since there is no head-to-tail


ordering (the director being apolar) there is no polarization normal to the
layers. Moreover, even if the molecules themselves are chiral, there is equal
probability of their assuming any orientation about their long axes. Hence
the transverse component of the dipole moment is averaged out and there
is no net polarization parallel to the layers.
In the Sc* phase, on the other hand, the molecules are tilted, and their
rotation about their long axes is biased. The symmetry plane of the
ordinary S c structure (fig. 5.8.1) is now absent because the molecules are
chiral. The only symmetry element that remains is a twofold axis parallel
to the layers and normal to the long molecular direction. This allows the
existence of a permanent dipole moment parallel to this axis. (Of course,
these arguments apply to the Sx* phase as well.)
Thus in Sc* each layer is spontaneously polarized. Since the structure has
a twist about the layer normal, the tilt and the polarization direction rotate
from one layer to the next (fig. 5.10.1 (a)). This implies that there is a
constant bend around the helical axis, which gives rise to a flexoelectric
contribution to the polarization.
When an electric field E is applied normal to the helical axis, the helix
gets distorted in a manner somewhat analogous to that depicted in fig.
4.6.1 for the cholesteric case. Above a critical field given by

where p is the pitch of the undistorted helix and k the 'twist' elastic
constant (which may be expected to vary rapidly with the tilt angle 6), the
helix is completely unwound and the sample is poled (fig. 5.10.1 ()). The
molecules are then aligned in a plane perpendicular to E with a tilt 6 with
respect to the layer normal. When the field is reversed, the polarization P
reverses direction, and because of the coupling between polarization and
tilt, the molecular orientation switches from 6 to 6. As we shall see later,
this effect finds important applications in electro-optic display devices.
The value of the spontaneous polarization in these materials is quite
small, usually between 10 and 1000 nC cm" 2, i.e, about one or two orders
of magnitude less than that for a solid ferroelectric like KH 2 PO 4 .
The coupling between P and 6 manifests itself even above the C*-A
transition: an electric field induces a tilt in the A phase as well. This is called
the electro clinic effect, and was first demonstrated by Garoff and Meyer.(233)
Induced tilt angles as high as 10 have been observed in high polarization
materials. Due to its submicrosecond response and its linear dependence

5.10 Ferroelectric liquid crystals


60

381

15

-r(K)
Fig. 5.10.2. Variation of the polarization P with relative temperature in the smectic
C* phase of DOBAMBC. The curve represents thefitobtained with the generalized
Landau theory. (After Dumrongrattana and Huang.(236))

15

Fig. 5.10.3. Variation of the tilt angle 6 and the ratio P/0 with relative temperature
in the smectic C* phase of DOBAMBC. The curves represent thefitsobtained with
the generalized Landau theory. (After Dumrongrattana and Huang.(236))
on the applied field, the electroclinic effect has proved to be useful in
making optical modulator devices.
The C*-A transition may be of first or second order/ 234 ' 235) The
polarization, tilt and pitch are, of course, temperature dependent and drop
to zero at the transition. The experimental curves for these three parameters
for DOBAMBC. which exhibits a second order C*-A transition, are

5. Smectic liquid crystals

382
10

Is
f 5

15
10
TC.A-T(K)
Fig. 5.10.4. Variation of the pitch with relative temperature in the smectic C* phase
of DOBAMBC. (Ostrovskii et al.(237)). The curve represents the fit obtained with
the generalized Landau theory. (After Dumrongrattana and Huang. (236))

20

T(C)
Fig. 5.10.5. The temperature variation of the field-induced tilt (or the electroclinic
effect) in the smectic A phase of 4-(3-methyl-2-chlorobutanoyloxy)-4 /-heptyloxy
biphenyl. (After Bahr and Heppke. (238))

shown in figs. 5.10.2, 5.10.3 and 5.10.4 and the variation of the electroclinic
effect with temperature in a high polarization material is shown in fig.
5.10.5. The static dielectric constant (for measuring field parallel to the
layers) increases slightly with temperature initially, but drops rapidly close
to 7^*A (fig. 5.10.6). From the dielectric relaxation one can study the two
important director modes, viz, the symmetry-recovering Goldstone mode
and the soft mode, in the vicinity of the C*-A transition. The Goldstone
mode is associated with the fluctuations of the azimuthal angle (<f>) of the
director, and the soft mode with the tilt (9) fluctuations. The former owes
its origin to the helical structure, and hence its frequency and strength

5.10 Ferroelectric liquid crystals

300
0.3 kHz
240 -

180

~"

0.4 kHz

120

0.5 kHz

60 10 kHz
0

1
-

i
2

i
1

i ,
1

r-r c . A (K)
Fig. 5.10.6. The dielectric constant for measuring field parallel to the layers as a
function of temperature in the smectic C* phase of [S]-4'-(2-chloro-4-methylpentanoyloxy)phenylftYtfw-4//-n-decyloxycinnamate.The variation of the dielectric constant with frequency is a consequence of the Goldstone mode relaxation.
(After reference 239.)
vanish in the SA phase (fig. 5.10.7(<z)). As for the soft mode, its frequency
decreases and strength increases on approaching the transition point from
either side (fig. 5.10.7 (b)). The coefficients of viscosity, y^ and y0, associated
with these modes are also strongly temperature dependent (fig. 5.10.8).
By virtue of their symmetry, ferroelectric smectics are piezoelectric.
Polarization can be induced by mechanical shear (2413) and, conversely, an
electric field can produce shear flow. They also possess pyroelectric
properties/ 244 ' 245)
An antiferroelectric smectic phase has been identified recently. (246) The
structure is expected to have an interlayer herringbone arrangement.
5.10,2 The smectic C*-smectic A transition
As the difference in the transition temperatures between the chiral and
racemic forms of the same compound is found to be only of the order of
1 C, it is clear that the transition is driven by intermolecular forces
responsible for the tilted smectic phase and not by the dipole-dipole

5. Smectic liquid crystals

384
640

o o o 0 O Oo
oo
#

|320

OOQQ

#
#

0.5

DC

160 -

(a) 0

1.0

1
%
1

-6

10

-4
-2
T- TC*A (K)

0.16

200

0.12
w

<

< 0.08 -

0.04

150

o
100

kHz)

480

o
o

iI

50

o ooc, 6>

b)

-0.5

0.5

1.0

T--T

15

A(K)

Fig. 5.10.7. The variation of the frequency (open circles) and dielectric strength
(filled circles) of (a) the Goldstone mode and (b) the soft mode in the vicinity of the
C*A transition. Material same as in Fig. 5.10.6. (After reference 239.)

10

fe, 1.0

0.1

90

80

70

T(C)
Fig. 5.10.8. The temperature variation of the Goldstone mode and soft mode
viscosity coefficients, y^ and y0, in the smectic C* phase of DOBAMBC. (After
Pozhidayev et a/.(240))

5.10 Ferroelectric liquid crystals

385

(ferroelectric) coupling. Liquid crystalline ferroelectrics may therefore be


classified as 'improper' ferroelectrics: the primary order parameter is the
tilt (0) and not the polarization (P). Meyer (247) proposed a simple
phenomenological model that accounts for some of the observed phenomena.
We shall suppose that the transition is of second order, and consider
first the spatially homogeneous case (i.e., the helix unwound completely,
q = 27r/pitch = 0), ignoring the effect of the spontaneous helical torsion
that appears below the transition. The free energy may then be written
as (247)

^eErj0P,

(5.10.2)

oil

where A = a(T TCA). The first two terms are the leading terms of the usual
Landau expansion, the next three terms describe the electrostatic free
energy and the last term expresses the coupling between P and 6.
Minimizing with respect to P and 0,

and hence,

P = X(E+r,6),
e = t,XE/A,

(5.10.3)
(5.10.4)

^!iy

(5.10.5)

We may draw the following conclusions: (a) from (5.10.5), we see that
the PB coupling produces a divergent component in the dielectric
constant, (b) from (5.10.3), that polarization can be induced by shear stress
in the absence of an electric field; this is the piezoelectric effect discussed
earlier, and (c) from (5.10.4), that the field-induced tilt, or the electroclinic
effect in the SA phase should diverge as the temperature approaches 7^*A.
However, this treatment is inadequate in many respects. For example,
fig. 5.10.5 shows that the dielectric constant does not exhibit a divergence
at all on approaching TC*A. In order to give a better description of the
properties of the Sc* phase, Pikin and Indenbom (248) proposed the following
free energy expression:
F = \a92 + \b6* + \kq292 - A02q + ^- P2-juP9q - CP6.

(5.10.6)

Here a = ao(T To), To being the transition temperature for the corresponding racemate, b > 0 and is temperature independent, k the elastic
constant, s the high temperature dielectric constant (i.e., in the absence of
ferroelectricity), ju and C are the constants describing respectively the

386

5. Smectic liquid crystals

flexoelectric and piezoelectric types of coupling between P and 0; A is


called the Lifshitz invariant parameter characteristic of chiral molecules
and produces the helicoidal structure. Minimizing with respect to q, P
andfl,
A + suC
q = -^r= constant,
2

k-sju

(5.10.7)

The main predictions of the model are: (i) the transition temperature of the
chiral compound is higher than that of the racemate, (ii) the tilt angle 0 and
hence P (which is proportional to 6) exhibit a power law variation with
temperature with the mean field exponent of 0.5, and (iii) the pitch and the
ratio P/0 are independent of temperature.
While this theory is an improvement over the previous one, its
predictions are still not quite in agreement with observations. The exponent
governing the 6 variation is generally found to be smaller than the mean
field value. Further, the theory is unable to account for the non-monotonic
temperature variation of the pitch and the anomalous behaviour of P/0.
More elaborate models have been proposed(249236) involving additional
phenomenological coefficients, including a sixth order term in 0 and a term
involving P292. A 'generalized' expansion of the form
F = \aO2 + \b6* + \c6 - AqO2 + \Kzq202
+ ^P2-HqP6-CP6-\ClP262

+ \riP*-dq0*

(5.10.8)

has been shown to be necessary to give a satisfactory explanation of the


various properties of the ferroelectric phase(236'250) (see figs. 5.10.2. 5.10.3
and 5.10.4).
5.10.3 Applications of ferroelectric liquid crystals
Smectic C* is proving to be of great practical importance as a material for
fast electrooptical switching.(251) The principle of operation of this switch
can be readily understood on the basis of the following simplified model.
The ferroelectric liquid crystal is taken in a thin cell, usually about 2-3 jum
thick. As long as the cell thickness is less than the pitch of the helix, the
strong surface forces align the molecules parallel to the substrate and the

5.10 Ferroelectric liquid crystals


\\\\\\\\\\\\\\\\\\\\\\\

wwwwwwwwwww
\\\\\\\\\\\\\\\\\\\\\

387

11 ii mi mini mi

wwwwwwwwwwww

Fig. 5.10.9. The SSFLC cell: The bookshelf geometry of a thin film of smectic C*
sandwiched between two glass plates (a) field ' up' (normal to the plane of the
diagram) and (b) field 'down'.
helix is unwound. The molecules are then arranged in the so-called
'bookshelf geometry, all inclined one way at an angle of about 20 with
respect to the layer normal as illustrated in fig. 5.10.9 (a). The spontaneous
polarization of every layer will therefore point in the same direction (say
upwards). If now a reverse electrical pulse is applied across the film, the
polarization will switch over to the opposite direction (downwards) and
the molecular tilt will change by 26 ~ 40 (fig. 5.10.9 (b)). By using a pair
of polarizers and adjusting the cell thickness so that, in effect, the liquid
crystal film becomes a half-wave plate, high optical contrast can be
achieved. The remarkable feature of the device - which is known as the
SSFLC (surface stabilized ferroelectric liquid crystal) device - is that the
switch on and switch off times are just a few tens of microseconds, i.e.,
about 1000 times faster than the twisted nematic device. For a given cell
thickness, the material parameters that determine the switching speed are
the polarization and viscosity. The system is bistable and has memory as
well.
However, in practice, it is found that the molecules adopt a chevron type
of arrangement(252) rather than the bookshelf geometry and this, of course,
affects the contrast ratio slightly. The mechanism of reorientation from
one stable state to the other on reversing the voltage is a complex one
involving the formation of defects.(252) Though the SSFLC cell is more
difficult to fabricate than the twisted nematic or supertwisted nematic cell,
its very fast switching time makes it extremely useful for large area, high
information content displays.(253~6)
Another phenomenon that has potential applications (257) is the fieldinduced tilt or the electroclinic effect. Unlike the SSFLC device, this effect
does not possess bistability but it has a faster (submicrosecond) response.
By using the same bookshelf geometry and a suitable polarizer and
retarder arrangement, the electroclinic effect can be used for modulating a
light signal with a transmitted intensity linearly proportional to the applied
voltage or as a tunable colour filter.

6
Discotic liquid crystals

6.1 Description of the liquid crystalline structures


Since the early investigations of Lehmann(1) and others(2) elucidating the
relationship between liquid crystalline behaviour and chemical constitution, the accepted principle was that the molecule has to be rod-like in
shape for thermotropic mesomorphism to occur, but it has emerged during
the last decade that compounds composed of disc-like molecules may also
exhibit stable mesophases. (310) Some typical discotic molecules are shown
in fig. 6.1.1. Generally speaking, they have flat (or nearly flat) cores with six
or eight (or sometimes four) long chain substituents, commonly with ester
or ether (but rarely more complex) linkage groups. Available experimental
evidence indicates that the presence of these side chains is crucial to the
formation of discotic liquid crystals. The mesophases whose structures are
clearly identified fall into two distinct categories, the columnar and the
nematic. A smectic-like (lamellar) phase has also been reported but the
precise arrangement of the molecules in each layer is not yet fully
understood.
The basic columnar structure is as illustrated in fig. 1.1.8(a); it is
somewhat similar to the hexagonal phase of soap-water and other
lyotropic systems (fig. 1.2.2). However, a number of variants of this
structure have been found. Fig. 6.1.2 presents the different two-dimensional lattices of columns that have been identified; here the ellipses
denote discs or, more precisely, cores (216) that are tilted with respect to the
column axis. Table 6.1.1 gives the space groups of the columnar structures
formed by some derivatives of triphenylene. (These are planar space groups
that constitute the subset of the 230 space groups when symmetry elements
relating to translations along one of the axes, in this case the column axis,
are absent.)
388

6.1 Liquid crystalline structures

389

RO
RO

I = Cu

(g)

M = Cu

(h)

Fig. 6.1.1. Examples of disc-shaped mesogens: (a) hexa-n-alkanoates of benzene,(3)


(b) hexakis ((4-octylphenyl)ethynyl)benzene,(11) (c) hexa-n-alkanoates of scylloinositol,(12) (d) hexa-n-alkanoates of triphenylene and hexa-n-alkoxytriphenylene, (1314) (e) hexa-n-alkyl and alkoxybenzoates of triphenylene, (1516) (/) hexa-nalkanoates of truxene, (1718) (g) bis(3,4-nonyloxybenzoyl)methanato copper(II),(19)
(h) octasubstituted metallophthalocyanine.(20)
High resolution X-ray studies have been reported on the columnar
phases of a few compounds. The measurements were made on very well
oriented monodomain samples obtained by preparing freely suspended
liquid crystal strands, typically about 200 jum in diameter and 1.5-2 mm

6. Discotic liquid crystals

390

Table 6.1.1. Columnar structures formed by some derivatives of


triphenylene^
R

Space group

C5HUO
C 7 H 15 O
C 8 H 17 O
C U H 23 COO

P6 2/m 2/m
P6 2/m 2/m
P6 2/m 2/m

C 7 H 15 COO
QH^-O-C.H.-COOf

PVa

P6 2/m 2/m
P^/a
P 2x/a
C2/m

Lattice
parameters (A)
a=
a=
a=
a=
a=
a=
a=
a=

18.95
22.2
23.3
44.9, b
26.3
37.8,/?
51.8, 6
30.7, 6

= 26.4
= 22.2
= 32.6
= 28.4

Temperature
(C)
^80
^80
^80
117
105
100
165
185

f The C 6 H 4 groups are para substituted.

(b)

id)

(e)

Fig. 6.1.2. (a)-(e) Columnar phases of disc-shaped molecules.(6) Plan views of the
two-dimensional lattice; ellipses denote discs that are tilted with respect to the
column axis: (a) hexagonal (P6 2/m 2/m); (b) rectangular (P2 1 /a); (c) oblique
(P^; (d) rectangular (P 2 /a); (e) rectangular face-centred, tilted columns (C2/m).
(/) The nematic phase.
long, with the column axis parallel to the axis of the strand/ 2 2 6) The results
of these studies are summarized below.
(a) The correlation length of the two-dimensional lattice of the
hexagonal columnar phase of (C 1 3 H 2 7 COO) 6 -truxene (fig. 6.1.1 ( / ) ) is
greater than 4000 A (or approximately 200 columns), the lower limit being

6.1 Liquid crystalline structures

391

set by the instrumental resolution. (25) Long-range order of the twodimensional lattice is, in fact, to be expected theoretically (see 6.3.1).
Within each column, on the other hand, the flat molecular cores form an
oriented, one-dimensional liquid, i.e., they are orientationally ordered but
translationally disordered. In contrast the flexible hydrocarbon chains are
highly disordered and produce a nearly isotropic scattering pattern. This
striking difference between the ordering of the cores and the tails is also
seen in the X-ray scattering from the columnar mesophases of triphenylene
compounds (fig. 6.1.1 (d)), but is not so apparent in hexa-n-octanoate of
benzene (fig. 6.1.1 (a)), which has a much smaller core. (2324)
Deuterium NMR spectroscopy of the disco tic phase of hexa-n-hexyloxy
triphenylene has led to similar conclusions. (27) Spectra of two selectively
deuterated isotopic species, one in which all aromatic positions are
substituted and the other in which only the a-carbon side chains are
substituted, bring out the difference between the order parameters of the
cores and the tails. Fig. 6.1.3 gives the quadrupole splittings of the
aromatic and the a-aliphatic deuterons versus temperature in the mesophase region. It is seen that the rigid core is highly ordered, the
orientational order parameter s ranging from 0.95 to 0.90, whereas the
a-aliphatic chains are in a disordered state.
These studies emphasize the fact that any realistic theory of the statistical
mechanics of discotic phases cannot treat the molecules as rigid discs, but
has to take into account the conformational degrees of freedom of the
hydrocarbon chains.
(b) Triphenylene hexa-n-dodecanoate exhibits two columnar phases,
designated as D h and D r , the subscripts h and r standing for hexagonal and
rectangular respectively. The D h -D r transition, which is weakly first order,
is associated with a small distortion of the lattice, consistent with a
herringbone arrangement in the rectangular structure with only the core of
the molecule tilted with respect to the column axis. However, high resolution synchrotron X-ray studies on monodomain discotic strands (2324)
have established that the tilt of the molecular core persists in the D h
phase as well, except that the tilts in neighbouring columns are now
rotationally uncorrelated, i.e., they are free to assume different azimuthal
angles with reference to the column axis (neglecting possible orientational
short-range order). Thus the D h -D r transition may be looked upon as an
orientational order-disorder transition.
(c) Hexa-hexyl thiotriphenylene (fig. 6.1.1 (d) with R = SC 6H13) shows
a transition from an hexagonal 'ordered' phase to an hexagonal 'disordered' one. The former is a phase in which there is regularity in the

392

6. Discotic liquid crystals

0.40

10

rc-r(c)
Fig. 6.1.3. Quadrupole splittings for the aromatic v and a-aliphatic VQ deuterons
of deuterated hexa-n-hexyloxytriphenylene (THE6) as functions of temperature
(T Tc) in the mesophase region, where Tc is the mesophase-isotropic transition
point. The open circles correspond to measurements on neat THE6-ard 6 and
THE6-ad 12 separately, while the filled circles correspond to a 2:1 mixture of the
two isotopic species. The scale on the upper right-hand side gives the orientational
order parameter of the aromatic part. The curve at the bottom gives the ratio of the
quadrupole splittings for the a-aliphatic and aromatic deuterons (Goldfarb, Luz
and Zimmermann(27)).

stacking of the triphenylene cores in each column, and the latter one in
which the column is liquid-like. X-ray studies using freely suspended
strands,(26) have shown that in the ordered phase there is a helicoidal
stacking of the triphenylene cores within each column, the helical period
being incommensurate with the intermolecular spacing.(28) In addition, a
three-column superlattice develops as a result of the frustration caused by
molecular interdigitation in triangular symmetry. It may be argued that if
there is no intercolumn interaction true long-range translational order

393

6.1 Liquid crystalline structures


.OR

RO
OC

R0-

R =C9H19COOR
R

=C12H25-

-C-N
0
.CO

RO

-OR

OR

-coo-

(b)
(a)
Fig. 6.1.4. (a) Cyclotricatechylene hexaesters, the cores of which are cone-shaped
(Malthete and Collet(29)), (b) hexa-(/?-n-dodecyloxybenzoyl) derivative of macrocyclic polyamines which is hollow at the centre (Lehn, Malthete and Levelut(33)).
Both (a) and (b) show columnar mesophases; the latter mesophase has been
described as 'tubular'.
cannot exist within a column because of the Peierls-Landau instability.
The existence of a regular periodicity in the stacking in each column
therefore implies that neighbouring columns must be in register. Thus,
ordered columnar phases, of which several have been reported, can
probably be compared with the highly ordered smectic phases of rod-like
molecules, e.g., smectics B, E, G, H, etc., which possess three-dimensional
positional order. However, more high resolution studies are necessary
before general conclusions can be drawn about the true nature of these
phases.
Columnar mesophases are also formed when the flat core of the
molecule is replaced by a conical one (2932) as in the cyclotricatechylene
hexaesters (fig. 6.1.4(a)). With macrocyclic molecules, which are hollow at
the centre (fig. 6.1.4(6)), the columns are in the form of tubes; these
mesophases have been described as ' tubular \ (33)
The nematic phase (N D) is exhibited by relatively few compounds;
examples are hexakis((4-octylphenyl)ethynyl)benzene (fig. 6.1.1(6)) and
the hexa-n-alkyl and alkoxybenzoates of triphenylene (fig. 6.1.1 (e)). The
N D phase has an orientationally ordered arrangement of the discs with no
long-range translational order (fig. 6.1.2(/)). Unlike the usual nematic of
rod-like molecules, N D is optically negative, the director n now representing
the preferred axis of orientation of the disc normal. The properties of this
phase will be discussed in greater detail in 6.5. A twisted nematic (or
cholesteric) phase, with the helical axis normal to the director, has also
been identified.(34)

394

6. Discotic liquid crystals

The hexa-n-alkanoates of truxene (fig. 6.1.1 (/)) show an unusual


sequence of transitions. For the higher homologues the phase sequence on
cooling is: isotropic->hexagonal columnar->rectangular columnar^
discotic nematic-> reentrant hexagonal columnar^crystal/ 1 7 ' 1 8 ) It has
been suggested that the truxene molecules are probably associated in pairs,
and that these pairs break up at higher temperatures, which might explain
this extraordinary behaviour. However, this conjecture still remains to be
proved.
i?w(/?-n-decylbenzoyl)methanato-copper(II), which is similar to the
molecule depicted in fig. 6.1.1 (g), but with four chains instead of eight, has
been reported to exhibit a smectic-like lamellar mesophase. (357) A tilted
smectic C type of structure has been proposed, (36) but the disposition of the
molecules in each layer does not appear to have been resolved. It is worth
noting that these copper complexes were the first paramagnetic mesogens
to be synthesized.
Molecules which combine the features of the rod and the disc may be
expected to form new types of mesophases. (38) An example is the biaxial
nematic phase reported in thermotropic systems (see 6.6). Malthete
ettf/.(39>40)have prepared an interesting series of mesogens shaped like
stick insects called ' phasmids' (fig. 6.1.5 (a)). Some of them form columnar
mesophases; the structure proposed for the hexagonal phase is shown
schematically in fig. 6.1.5(6).

6.2 Extension of McMillan's model of smectic A to discotic liquid


crystals
Transitions between the columnar (D) and the nematic (N D) phases occur
in some compounds. Available data for a few homologous series of
compounds indicate that the lower members of the series show only the N D
phase, the next few members show both D and N D phases, the temperature
range of the latter phase diminishing with increasing chain length, while for
the higher members D transforms directly into the isotropic (I) phase.
Broadly the trend is reminiscent of the behaviour of the smectic
A-nematic-isotropic sequence of transitions in systems of rod-like
molecules. This suggests that one should be able to give a qualitative
description of the D-NDI transitions by extending McMillan's mean field
model of smectic A (see 5.2) so that the density wave is now periodic in
two dimensions. (413)
The hexagonal phase can be described by a superposition of three
density waves with wavevectors

395

6.2 Extension of McMillan's model


H25C,2C>

12H25

Fig. 6.1.5. (a) Hexa-n-alkoxy terepthal-to-(4-benzoyloxyaniline) and (b) the


structure of the hexagonal columnar mesophase formed by this compound. The
mesophase has been described as 'phasmidic'. (Malthete, Levelut and Tinh.(39))
4TT

V 3d
R

V3

C = A + B,
where d is the lattice constant.
The appropriate single particle potential, which depends on the
orientation of the short axis of the molecule as well as the position r of its
centre of mass, may be written in the mean field approximation as
V^x, y9 cos 0) = -Vo P2(cos 8) {s + a<r[cos (A r) + cos (B r) + cos (C r)]},
(6.2.1)
retaining only the leading terms in the Fourier expansion. Here Vo is the

396

6. Disco tic liquid crystals

interaction energy which determines the nematic-isotropic transition, a is


the McMillan parameter given by
2exp[-(27rro/V3</)2],

(6.2.2)

r0 being the range of interaction which is of the order of the size of the
aromatic core,

= </(cos0)>

(6.2.3)

is the usual orientational order parameter, and


a = |<[cos (A r)+ cos (B r)+ cos (C r)] i>2(cos 9)}

(6.2 A)

is an order parameter coupling the orientational and the translational


ordering. The angular brackets represent the statistical average over the
distribution function derived from the potential (6.2.1), the spatial
integrations being carried out over a primitive cell of the hexagonal lattice.
This form of the potential ensures that the energy of the molecule is
minimum when the disc is centred in the column with its plane normal to
the z axis.
The free energy can then be calculated using standard arguments:

(6.2.5)
Numerical calculations yield three possible solutions to the equations:
(1) s = a = 0
(isotropic phase),
(2) s 4= 0, a = 0 (nematic phase),
(3) s 4= 0,(7 41 0 (hexagonal columnar phase).
For any given values of a and T, the stable phase is the one which
minimizes the free energy (6.2.5). Interpreting a as a measure of the chain
length as in McMillan's model, the phase diagram as a function of a (fig.
6.2.1) reflects the observed trends for a homologous series of compounds.
The temperature range of the nematic phase decreases with increasing a,
and for a > 0.64 the D phase transforms directly to the isotropic phase. It
should be noted that since the wavevectors of the hexagonal lattice satisfy
the relation A + B C = 0, the columnar-nematic transition is always of
first order: a Landau expansion of the free energy contains a nonvanishing cubic term in the order parameter o.
The theory has been extended using a variational principle to solve the
problem with the full potential rather than the one truncated to the first
Fourier component.(44) This method, which is closely analogous to the
extension of McMillan's model by Lee ettf/.,(45)leads to some qualitative

6.2 Extension of McMillan's model

397

Isotropic
1.0

I0-8

Nematic

Columnar

I
0.4

I
0

0.2

0.4
0.6
Model parameter (a)

J
0.8

Fig. 6.2.1. Theoretical plot of the reduced transition temperature against the model
parameter a showing the hexagonal, nematic and isotropic phase boundaries. All
the transitions are of first order.
improvements in the phase diagram for a homologous series. A simpler
mean field version of this theory has been presented which yields essentially
similar results.(46)
For the rectangular lattice, the nature of the phase diagram depends on
the axial ratio b/a.m) When b/a is only slightly different from y/3, the
phase diagram is similar to that for the hexagonal structure. As the
asymmetry of the lattice is increased, one of the density waves disappears
before the other to give rise to a smectic (or lamellar) phase, which, in turn,
transforms to the N D phase at a higher temperature. Evidence of a smecticlike disco tic phase has been reported, (357) but whether this corresponds to
the lamellar phase predicted by theory is a matter for further study.
As mentioned in 6.1, transitions are observed between different types of
columnar phases. In particular, the transition from the rectangular (D r ) to
the hexagonal (Dh) structures has been established experimentally to be an
orientational order-disorder transition. The tilts of the molecular cores
with respect to the column axis in neighbouring columns are ordered or
correlated to give rise to a rectangular packing with a herringbone
structure in the D r phase, whereas they are free to assume different

398

6. Discotic liquid crystals

azimuthal angles with reference to the column axis in the D h phase.


Theoretical models have been developed to describe this transition. (479)
6.3 The columnar liquid crystal: applications of the continuum theory
6.3.1 Fluctuations
We now discuss the fundamental question of fluctuations in the columnar
phase. Let us suppose that the liquid-like columns are along the z axis and
that the two-dimensional lattice (assumed to be hexagonal) is parallel to
the xy plane. The two basic deformations in such a structure are (i) the
curvature deformation (or bending) of the columns without distortion of
the lattice and (ii) lattice dilatation (or compression) without columnar
curvature. There can also be coupling between the two types of distortion
but, as shown by Kleman and Oswald,(50) the coupling term merely rescales
the bend elastic constant of the columns. We shall consider only the
vibrations of the lattice in its own plane. The free energy may be written

2 \ 9x

dy )

2 \_\ dx

dy

where B and D are the elastic constants for the deformation of the twodimensional lattice in its own plane; ux and uy are the displacements along
x and y at any lattice point, and /r33 is the Frank constant for the curvature
deformation (bending) of the columns. (In the notation of the standard
crystal elasticity theory, B = | ( c n + c12) and D = \{clx c12).) We neglect
here the splay and twist deformations because they give rise to a distortion
of the lattice and involve considerable energy. (51) We also neglect any
contributions from the surface of the sample. Writing the displacement u
in terms of its Fourier components,
H(r) = X>(q)exp(iq-r);
Q

substituting in (6.3.1), we get, in the harmonic approximation

and from the equipartition theorem


<ul)=kBT/(
where Bo = B + 2D,k9 = 2k3S and qL =

(6.3.2)

6.3 The columnar liquid crystal

399

The mean square displacement at any lattice point is given by

where L is the length of the columns, L the linear dimension of the lattice
in the xy plane and d its periodicity. Assuming that U > L
<W2> = [kB T/4BXd$\ [l-(d/L%

(6.3.3)

(5>52>53)

The structure is
where X = (ko/Bo)* is a characteristic length.
therefore stable as L -> oo. As is well known from the classical work of
Peierls(54) and Landau (55) , the two-dimensional lattice itself is an unstable
system with <w2> diverging as lnL. ( 5 6 5 7 ) Hence the curvature elasticity of
the liquid-like columns stabilizes the two-dimensional order in the
columnar liquid crystal. This result was first proved by Landau, who
observed: ' Thus bodies having such a structure could in theory exist, but
it is not known whether they do in fact exist in Nature.' (58)
6.3.2 X-ray scattering
The ' structure factor' for the intensity of X-ray scattering may be written
as
5(K)=
The second exponential term on the right-hand side is the familiar
Debye-Waller factor exp( W). Now

where p = (x2+y2)* and U is the confluent hypergeometric Kummer


function.(59) For z > (Ap)*9 the second term of (6.3.4) varies as
(6.3.5)

400

6. Discotic liquid crystals

Thus, in contrast to smectic A and the two-dimensional lattice, for which


the displacement-displacement correlation is of logarithmic form, the
columnar liquid crystal gives the usual Bragg reflexions. As noted in 6.1,
this is borne out by high resolution X-ray experiments on well oriented
monodomain discotic strands.(25)
So far it has been assumed that the length (Z/) of the liquid columns
is much larger than L. One may similarly consider the opposite situation
Z/ <^ L. In this case, it turns out that for the bounded sample
const.

for very small L

In Z/

for larger L.

If the surfaces are free <V> will have additional terms, which may be
analogous to those for a two-dimensional lattice. In any case, it is clear that
the mean square fluctuations of the lattice as well as the Debye-Waller
factor may be expected to show a certain dependence on the linear
dimensions of the sample.
6.3.3 Light scattering
The mean square amplitude of the orientational fluctuations of the director
and the corresponding intensity of light scattering may be worked out in a
similar fashion. For example, when the incident and scattered beams are
both polarized in the plane perpendicular to the optic axis (or column axis)
the scattered intensity is given by
Tco*
where q = (ql + qffi and co is the angular frequency of the light.(41) The
scattering is similar to that from smectic A in certain geometries (see
5.3.4), but is, of course, small compared to that from a nematic.
6.3.4 Mechanical instabilities
If the elasticity theory of the columnar phase is developed to the next
higher order, one finds that ux x, uxy and uz z of the linear theory have to
be replaced as follows:
U

x,x^Ux,x

\Ul,x~\Ul,z~Ux,zUz,x

Ux, y ~> I K , y + Uy x) - \{UX xUyx + Ux yUyy

+ Ux z + Uy z)

6.3 The columnar liquid crystal

401

where ux x = dux/dx, ux y = dux/dy, etc. As a result, a lattice strain like ux x


gets coupled to ux z which represents the tilt of the columns. This type of
coupling is present in smectic A as well and gives rise to an undulation
instability which has been studied in detail (see 5.3.3). An equivalent type
of instability may be expected to occur in the columnar liquid crystal as
well and can be analysed exactly as in the case of smectic A. (50) The
columns are parallel to the glass plates, and the separation (d) between the
plates is increased (see fig. 5.3.4). If ux = ax, (a being positive) then at a
critical value a = ac given by

kS3

a spatially periodic distortion should set in with a wavevector qz given by

Another type of instability arises from the coupling between uz z and ux z


or uy z. This may be described as a buckling instability. The analysis is very
similar except that a compressive stress is applied parallel to the columns.
At a critical value of the stress given by

the columns buckle. This has actually been observed, (60) but there is a
significant discrepancy between theory and experiment which has not been
explained satisfactorily.
6.3.5 Acoustic wave propagation
The basic hydrodynamic equations for the hexagonal columnar phase

(
l T+ 8^

402

6. Discotic liquid crystals


dt
dK
At
dK
6e

where
8
^2

^2
A

^2
+

^2
8

2 = heat density; 9 = volume strain; p = density; Vi = components of the


velocity; KpK = thermal conductivity parallel and perpendicular to the
column axis; ut = components of the column displacement; P = pressure;
e = energy density expressible as F-\~\A62 + C6(ux x-\-uy y), where F is
given by (6.3.1), A is the bulk modulus and C a coupling constant; C =
permeation coefficient; = thermomechanical coefficient; rj ijkl = viscosity
coefficients.
In the absence of a temperature gradient, and assuming that permeation
is negligible, i.e., u = Vi9 and that damping is weak, it turns out that under
adiabatic conditions there are three propagating acoustic modes for any
arbitrary direction not coinciding with an axis of symmetry. Of these, one
is the familiar longitudinal wave arising from density fluctuations, the
velocity (Fx) of which is practically independent of the direction of
propagation. Another is a transverse wave which is exactly analogous to
the second sound of smectic A (see 5.3.6). The velocities Vl and V2 of these
two waves are given by the roots of the equation

where y/ is the angle between the direction of propagation and the column
axis. We have ignored the coupling between the lattice and the curvature
deformation of the columns. The third mode is a transverse wave whose
polarization is orthogonal to that of second sound. This wave propagates

6.4 Defects in the columnar liquid crystal

403

Fig. 6.3.1. Dependence of the sound velocities on the polar angle y/ in the columnar
liquid crystal.
because the two-dimensional lattice can sustain a shear. Its velocity for any
arbitrary direction is given by
K3 = (/)//>)* sin <p.
Fig. 6.3.1 depicts the variation of the velocities of the three waves with
polar angle. This angular dependence can be studied by Brillouin
scattering, as has been demonstrated for smectic A (see fig. 5.3.11). In the
present case, there should be one, two or three pairs of Brillouin
components depending on the orientation, but no experiments have yet
been reported.

6.4 Defects in the columnar liquid crystal


The symmetry elements in the hexagonal structure composed of liquid-like
columns are shown in fig. 6.4.1. The director n is parallel to the column
axis; a (or equivalently a' or a") is the lattice vector of the two-dimensional
hexagonal lattice; L2, T2 and 92 are two-fold axes of symmetry; L 3 is a
three-fold axis and L6 a six-fold axis. Any point on L2 or L6 is a centre of
symmetry. Any planes normal to n and the planes (L6, T2) and (L6,62) are
planes of symmetry. Defects in such a structure have been investigated in
detail by Bouligand(61) and by Kleman and Oswald.(50)

6. Discotic liquid crystals

404

Fig. 6.4.1. The hexagonal columnar phase with the symmetry elements in the
structure.
Fig. 6.4.2 gives photographs of the columnar liquid crystal appearing in
the isotropic phase as the sample is cooled slowly. The growth pattern
reflects the hexagonal symmetry of the columnar mesophase.
6.4.1 Dislocations
The Burgers vector b of a dislocation in the columnar structure is normal
to the columnar axis n. In general
b = /a + ma',
where / and m are positive or negative integers. For edge dislocations, b is
perpendicular to the dislocation line L. Two types of edge dislocations are
possible: longitudinal edge dislocations (L parallel to n, figs. 6.4.3(a) and
(b)) and transverse edge dislocations (L perpendicular to n, figs. 6.4.3 (c)
and (d)). For screw dislocations b is parallel to L (figs. 6.4.3 (e) and (/)). A
hybrid composed of screw and edge dislocations is shown in fig. 6.4.3 (g).
Longitudinal edge dislocations
In this case, the dislocation line L is along n, the z axis. Let the Burgers
vector be along y. Minimization of the free energy (6.3.1) gives
{B + D)

9 (%
ox\ax

oy

-DA(%*-%*)
=0
oy\ox
oy J

6.4 Defects in the columnar liquid crystal

405

Fig. 6.4.2. The discotic liquid crystal appearing in the isotropic phase when a
sample is cooled very slowly. The growth pattern is diagnostic of the hexagonal
symmetry of the columnar structure: (a) Queguiner, Zann and Dubois (62), (b)
Bouligand(61).

406

6. Discotic liquid crystals


b= a

,M

Fig. 6.4.3. Dislocations in the columnar phase: (a) and (Z?) longitudinal edge
dislocations; (c) and (d) transverse edge dislocations; (e) and (/) screw dislocations ; (g) a hybrid of screw and edge dislocations. It should be noted that the
Burgers vector b = a for (a), (c), (e) and (g), and b = a a' for (b), (d) and (/).
(Bouligand.(61))
and

These are exactly like the equations for edge dislocations in crystals.(63) The
solutions are
B y2\

6.4 Defects in the columnar liquid crystal

407

where 0 = tan" 1 (y/x). The energy of the dislocation is

where Ec and rc are the energy and radius of the core. The radial and
angular components of the force of interaction between two edge
dislocations are
Db1b2
fr = 2n(\-v)r'

/ TTrV s i n 2 a '
2n(\ v)r
where v is Poisson's ratio.
Transverse edge dislocations
The dislocation line is now normal to n. Let it be along the x axis, so that
the only component of the displacement that needs to be considered is u y.
Minimization of the free energy yields
dy2

33

dz* '

which is exactly of the same form as that for edge dislocations in smectic A
(see 5.4.2). The solution is
b

and the energy


where X = [k33/(B + D)]*. The properties of these dislocations are essentially the same as those of smectic A edge dislocations.
Screw dislocations
Let the dislocation line be along x. The relevant component of the
displacement is therefore ux, and the equation of equilibrium becomes
X

8/ ~

1^

33

8z4 '

6. Disco tic liquid crystals

408
(a)

(b)

Oo>o O nO

oo
o o oo oo
o oo ooo
o o o oo

o
0

o o S o o
o

o o oo
o ooo
o o oo

(c)

Fig. 6.4.4. Disclinations in the columnar phase: (a) and (b) n/3 and n/3
longitudinal wedge disclinations about the sixfold axis L6; (c) and (d) n transverse
wedge disclinations about the binary axes T2 and 62 respectively; (e) n transverse
wedge disclination leading to the formation of walls; (/) two n disclinations at right
angles to each other, one about T2 and the other about 92. (Bouligand.(61))
which is again of the same form as the equation for an edge dislocation in
smectic A. The solution is
b b
uxx = 7 +
4

and the energy

An

6.4 Defects in the columnar liquid crystal

409

Fig. 6.4.5. Photograph indicating the presence of two n disclinations at right angles
to each other as depicted in fig. 6.4.4 (/). (Oswald.(64))

where in this case X = (k33/D)k In contrast, the screw dislocation in smectic


A has no self energy (apart from the core) under the same approximation.
Thus the energies and interactions of screw dislocations in the columnar
phase are entirely different from those of their counterparts in smectic A or
the crystal.
6.4.2 Disclinations
Longitudinal wedge disclinations are the standard crystal disclinations in
a hexagonal lattice. The rotation vector is L 6 or L3 or L2, parallel to the
columns. Two examples of such defects are shown in figs. 6.4.4 (a) and (b).
The lattice gets compressed near a positive disclination and stretched near
a negative one. Generally, most of these disclinations have prohibitively
large energies, and hence they occur as unlike pairs at the core of a
dislocation.
In transverse wedge disclinations the rotation vector is 62 or T2, normal
to n. Examples of n disclinations are shown in figs. 6.4.4(c), (d) and (e).
Two transverse wedge disclinations may occur in association as shown
in fig. 6.4.4 (/). The angle between the rotation axis may be 90, 60 or 30.
Such defects have been observed experimentally. (64) Fig. 6.4.5 presents a
photograph which may be interpreted as arising from two n disclinations

6. Discotic liquid crystals

410

(a)

Fig. 6.4.6. Developable domains in the columnar phase, (a) The developable
surface is degenerated into a straight line S common to the planes P. The columnar
axes C form coaxial circles about S. (b) The developable surface D is a cylinder. The
columns (or the layers in the case of smectic A) are a set of parallel and equispaced
involutes of a circle, (c) A Reimann surface generated by half-tangents to a helix.
The columns are normal to the half-tangents. (Bouligand.(61))

parallel to the glass slides and at right angles to each other as depicted in
fig. 6.4.4 (/).
The symmetry of the columnar phase also permits the occurrence of
twist disclinations in the hexagonal lattice and of hybrids consisting of a
twist disclination in the hexagonal lattice and a wedge disclination in the
director field. According to Bouligand these defects are not likely to exist.

6.4.3 Developable domains


We shall now consider the most general deformations that can occur
without distorting the hexagonal lattice and involving only the bending of
the columns. The problem reduces to one of finding physical planes P

6.5 The discotic nematic phase

411

perpendicular to the columns, the envelope of these planes defining the


'developable' surface D. We give below some specific examples/ 61 ' 65)
(i)

If the developable surface degenerates into a straight line, then the


columns form a set of coaxial circles around this line (fig. 6.4.6(a)).
The line need not necessarily be a twofold axis, 02 or 7L The structure
of this defect is similar to the 2n disclinations of smectics.
(ii) If the developable surface is a cylinder, then the columns are along the
involutes of a circle in any plane normal to the axis of the cylinder (fig.
6.4.6 (b)). This defect is similar to a disclination of unit strength. It is
seen that this is an obvious generalization of the solutions for smectic
A. Structures similar to this have been observed by Oswald.
(iii) If the developable surface is a Reimann surface generated by halftangents to a helix, then the columns are normal to the half-tangents
and are involutes of the helix (fig. 6.4.6 (c)). This defect has the
features of a dislocation and a disclination, and can therefore be
described as a dispiration.

6.5 The discotic nematic phase


Theoretical calculations had shown, even before the discovery of disco tics,
that a transition from the isotropic to the nematic phase is possible, in
principle, in an assembly of plate-like particles.(66"9) On the experimental
side, Brooks and Taylor(70) observed that there is a mesophase transformation at high temperatures during the carbonization of certain
graphitizable organic materials such as petroleum and coal tar pitches. The
transformation is an irreversible reaction which proceeds from the
isotropic melt to a nematic type of mesophase with increasing temperature.
The mesophase, referred to as the 'carbonaceous' phase, is a complex
multicomponent system composed of large plate-like polynuclear aromatic
molecules having a wide range of molecular weights (with a mean
molecular weight around 2000) and has a transient existence. We shall not,
however, discuss the carbonaceous phase any further, but confine our
attention to stable nematic liquid crystals formed by relatively simple
compounds.
Fig. 6.1.2(/) gives a schematic illustration of the structure of the discotic
nematic (N D ): in contrast to the classical nematic of rod-shaped molecules,
the director n now represents the preferred orientation of the short
molecular axis (or the disc norm~ ,. The symmetry of the two kinds of
nematics is the same, and identical types of defects - the schlieren texture,

412

6. Disco tic liquid crystals

Fig. 6.5.1. Schlieren texture in a discotic nematic film. (Destrade et al.a5))


umbilics, e t c . - a r e seen in both cases. Fig. 6.5.1 is a photograph of a
typical schlieren pattern exhibited by a thin N D film when viewed under a
polarizing microscope.
The N D phase is optically and diamagnetically negative. However, its
dielectric anisotropy Ae may be positive or negative depending on the detailed molecular structure. For example, hexa-n-heptyloxybenzoate of triphenylene (fig. 6.1.1 (V)) and hexa-n-dodecanoyloxytruxene (fig. 6.1.1 (/))
are both dielectrically positive in the nematic phase/ 71 ' 72) This is because
while the contribution of the electronic polarizability of these disc-shaped
molecules to Ae is negative, the permanent dipoles associated with the six
ester linkage groups make a stronger positive contribution, so that Ae > 0
for the two compounds. On the other hand, hexakis ((4-octylphenyl)ethynyl)benzene (fig. 6.1.1 (&)), which does not have any permanent dipoles,
does, in fact, show negative Ae, as expected/ 73 ' 74)
Very few quantitative measurements of the physical properties have
been reported. The Frank constants for splay and bend have been
determined using the Freedericksz method.(71"5) Interestingly the values are
of the same order as for nematics of rod-like molecules. The twist constant
k22 has not yet been measured. The fact that the diamagnetic anisotropy is
negative makes it somewhat more difficult to measure these constants by

6.5 The discotic nematic phase

(a)

413

(b)

Fig. 6.5.2. Flow alignment of the director in nematic liquid crystals, (a) For rodshaped molecules the alignment angle 9 with respect to the flow direction lies
between 0 and 45, while (b) for disc-shaped molecules it lies between 90 and
-45 (or equivalent^ between 90 and 135).

the Freedericksz technique. However, by a suitable combination of electric


and magnetic fields it is possible, in principle, to determine all three
constants.
The hydrodynamic equations of the classical nematic (3.1) are
applicable to the N D phase as well. There are six viscosity coefficients (or
Leslie coefficients) which reduce to five if one assumes Onsager's reciprocal
relations. A direct estimate of an effective value of the viscosity of N D from
a director relaxation measurement(71'73) indicates that its magnitude is
much higher than the corresponding value for the usual nematic.
We have already seen in 3.6.4 that in ordinary nematics of rod-like
molecules, the Leslie coefficients ju2 and ju3 are both negative and \ju2\ > |//3|.
Under planar shear flow, the director assumes an equilibrium orientation
60 given by
tan 2 0o =
where 60 lies between 0 and 45 with respect to the flow direction (fig.
6.5.2(a)). In practice 00 is usually a small angle. In certain nematics, it is
found that //3 > 0 at temperatures close to the nematic-smectic A transition
point. Under these circumstances there is no equilibrium value of #0, and
in the absence of an orienting effect due to the walls or a strong external
field, the flow becomes unstable.(76)
It has been suggested that in the N n phase, the disc-like shape of the
molecule may have a significant effect on ju2 and //3.(77'78) The stable
orientation of the director under planar shear will now be as shown in fig.
6.5.2(b). Thus it can be argued that both ju2 and jus should be positive, and
the flow alignment angle 60 should lie between 45 and 90. It then
follows that when ju2 < 0, the director tumbles and the flow becomes

414

6. Discotic liquid crystals

Fig. 6.6.1. Two examples of molecules which exhibit the biaxial nematic phase:
(a) (bis-1 -(/?-n-decylbiphenyl)-3-(/?-ethoxyphenyl) propane-1,3-dionatocopper
(H))<8i,82) (hereafter referred to as complex A); (b) l,12-te(pentakis((4-pentylphenyl)ethynyl)pheny loxy)dodecane.(83)

unstable. As yet, however, no experimental studies have been carried out


to verify any of these ideas.
6.6 The biaxial nematic phase

The biaxial nematic (Nb) phase was first identified by Yu and Saupe(79) in
a ternary amphiphilic system composed of potassium laurate, 1-decanol
and D 2 O. In such systems the constituent units are molecular aggregates,
called micelles, whose size and shape are sensitive to the temperature and
concentration; the N b phase was found to occur over a range of
temperature/concentration. There are obvious advantages in having a
single-component, low-molar-mass thermotropic N b phase. The sugges-

6.6 The biaxial nematic phase

415

Fig. 6.6.2. Conoscopic figure demonstrating the biaxiality of the nematic phase of
complex A. (8284)

tion was made(38) that a convenient way of achieving this would be by


'bridging the gap between rod-like and disc-like mesogens', i.e., by
preparing a mesogen that combines the features of the rod and the disc.
This has proved to be useful and the N b phase has been observed in some
relatively simple compounds.(80) Two examples are given in fig. 6.6.1. The
biaxiality of the nematic phase was established by optical observations,
supported by X-ray evidence (826) (figs. 6.6.2 and 6.6.3).
The difference between the structures of the uniaxial and biaxial nematics
is illustrated schematically in fig. 6.6.4. The N b phase is depicted here as an
orthorhombic fluid whose preferred molecular orientation is described by
an orthonormal triad of director fields. (In principle, nematics of lower
symmetry are possible, but none of the N b phases identified to date have
been reported to be other than orthorhombic.) The structure, therefore,
gives rise to an additional pair of diffuse (liquid-like) X-ray diffraction
peaks(87-9) (fig. 6.6.3(6)).
A number of important ideas concerning the N b phase have been
discussed theoretically - molecular statistical and phenomenological
theories/ 69 ' 90103) continuum theories, (10417) topological theories of defects/118-25* etc. For example, Saupe(104) and Kini(109) who used different
theoretical approaches, have both concluded that the incompressible
orthorhombic nematic has 12 curvature elastic constants (excluding three
which contribute only to the surface torque) and 12 viscosity coefficients.

6. Discotic liquid crystals

416

(i)

-12 - 6
(a)

(ii)

12 - 2 4 - 1 8 - 1 2 - 6
26 (deg)

-6

12

-24 -18 -12


20 (deg)

(ii)

(i)

-12

12

18

24

-6

12

18

24

Fig. 6.6.3. Raw microdensitometer scans of the X-ray intensity plotted against
diffracting angle (26) for magnetically aligned nematic samples: (a) the uniaxial
nematic phase of 80CB at 77 C; (b) the biaxial nematic phase of complex A at
168.5 C, (i) meridional scan (parallel to / / ) , (ii) equatorial scan (perpendicular to
H). M represents the diffraction peaks from the mylar film which covered the
windows of the sample holder and heater assembly. (8586)

Again, remarkable predictions have been made on the basis of the


homotopy theory regarding the properties of defects in the N b phase. The
usual law of coalescence of defects (see 3.5.1) breaks down; the
combination rule is now non-Abelian. Moreover, there can be an
entanglement of disclination lines, which, in turn, can give rise to what

6.6 The biaxial nematic phase

(a)

417

(b)

Fig. 6.6.4. Schematic diagram of the molecular order in (a) the uniaxial nematic
phase and (b) the biaxial nematic phase.

Toulouse(118) describes a s ' topological rigidity'. These and other ideas have
yet to be investigated experimentally. The availability of the N b phase in
simple thermotropic systems is likely to make it conveniently possible to
test some of these predictions.

References

Chapter 1
1 F. Reinitzer, Monatsch Chem., 9, 421 (1888).
2 O. Lehmann, Z. Physikal. Chem., 4, 462 (1889). For a brief historical
account of the early work see, H. Kelker and R. Hatz, Handbook of Liquid
Crystals, Verlag Chemie, Weinheim, Chapter 1 (1980).
3 D. Demus, H. Demus and H. Zaschke, Flussige Kristalle in Tabellen, VEB
Deutscher Verlag fur Grundstoffindustrie, Leipzig (1974).
D. Demus and H. Zaschke, Flussige Kristalle in Tabellen, Vol. II, VEB
Deutscher Verlag fur Grundstoffindustrie, Leipzig (1984).
4 G. Friedel, Ann. Physique, 18, 273 (1922).
5 I. G. Chistyakov and W. M. Chaikowsky, Mol. Cryst. Liq. Cryst., 7, 269
(1969).
6 A. de Vries, Mol. Cryst. Liq. Cryst., 10, 219 (1970); in Proceedings of the
International Liquid Crystal Conference, Bangalore, December 1973,
Pramana Supplement-1, p. 93.
7 G. W. Stewart, Discussions of Faraday Soc, p. 982 (1933).
8 A. Saupe, Angew. Chem. Intnl. Edn., 1, 97 (1968).
9 A. D. Buckingham, G. P. Ceasar and M. B. Dunn, Chem. Phys. Letters, 3,
540 (1969).
10 A. J. Leadbetter, in Thermotropic Liquid Crystals (ed. G. W. Gray), p. 1,
Wiley, Chichester (1987).
11 P. S. Pershan, Structure of Liquid Crystal Phases, World Scientific,
Singapore (1988).
12 R. J. Birgeneau and J. D. Litster / . de Physique Lettres, 39, 399 (1978).
R. Pindak, D. E. Moncton, S. C. Davey and J. W. Goodby, Phys. Rev. Lett.
46, 1135(1981).
13 S. Diele, P. Brand and H. Sackmann, Mol. Cryst. Liq. Cryst., 17, 163 (1972)
A. Tardieu and J. Billard, / . Phys. (Paris) Colloq. 37, C3-79 (1976); G.
Etherington, A. J. Leadbetter, X. J. Wang, G. W. Gray and A. Tajbaksh,
Liquid Crystals, 1, 209 (1986).
14 H. Sackmann and D. Demus, Mol. Cryst. Liq. Cryst., 21, 239 (1973).
15 D. Demus and L. Richter, Textures of Liquid Crystals, Verlag Chemie, New
York (1978).
16 G. W. Gray and J. W. Goodby, Smectic Liquid Crystals, Leonard Hill,
London (1984).

418

References: Chapter 1

419

17 K. Herrmann, Z. Krist., 92, 49 (1935).


The problem has been re-examined recently applying group theoretical
methods by N. Boccara, Ann. Physique, 76, 72 (1973).
18 J. D. Bernal and I. Fankuchen, / . Gen. PhysioL, 25, 111 (1941).
19 A. J. Mabis, Acta CrySt., 15, 1152 (1962).
20 S. Chandrasekhar, B. K. Sadashiva and K. A. Suresh, Pramana, 9, 471
(1977).
21 A. M. Levelut, / . Chim. Phys., 88, 149 (1983).
C. Destrade, P. Foucher, H. Gasparoux, N. H. Tinh, A. M. Levelut and J.
Malthete, Mol. Cryst. Liquid CrySt., 106, 121 (1984).
S. Chandrasekhar, Advances in Liquid Crystals, Vol. 5 (ed. G. H. Brown) p.
47, Academic, New York (1982).
S. Chandrasekhar, Phil. Trans. Roy. Soc. London, A309, 93 (1983).
S. Chandrasekhar and G. S. Ranganath, Rep. Prog. Phys., 53, 57 (1990).
22 A. M. Giroud-Godquin and J. Billard, Mol. Cryst. Liquid Cryst., 66, 147
(1981).
K. Ohta, H. Muroki, A. Takagi, K. I. Hatada, H. Ema, I. Yamamoto and
K. Matsuzaki, Mol. Cryst. Liquid Cryst., 140, 131 (1986).
A. C. Ribeiro, A. F. Martins and A. M. Giroud-Godquin, Mol. Cryst.
Liquid Cryst. Lett., 5, 133 (1988).
23 A. Blumstein (ed.) Liquid Crystalline Order in Polymers, Academic, New
York (1978).
H. Finkelmann, J. Koldehoff, H. Ringsdorf, Angew. Chem. Int. Ed. Engl, 17
935 (1978).
S. G. Kostromin, V. V. Sinitzyn, R. V. Talroze, V. P. Shibaev, N. A. Plate,
Makromol. Chem., Rapid Commun., 3, 809 (1982).
H. Ringsdorf, A. Schneller, Makromol. Chem., Rapid Commun., 3, 557
(1982).
A. Ciferri, W. R. Krigbaum and R. B. Meyer (eds.) Polymer Liquid Crystals,
Academic, New York (1982).
E. T. Samulski, Physics Today, 35, 40 (1982).
Proceedings of the International Conference on Liquid Crystal Polymers,
Bordeaux, France, Mol. Cryst. Liquid Cryst., Parts A & B, 153 & 155
(1988).
24 O. Herrmann-Schonherr, J. H. Wendorff and H. Ringsdorf, Makromol
Chem., Rapid Commun., 7, 97 (1986).
25 O. Herrmann-Schonherr, J. H. Wendorff, H. Ringsdorf & P. Tschirner,
Makromol. Chem., Rapid Commun., 7, 791 (1986).
M. Ebert, O. Herrmann-Schonherr, J. H. Wendorff, H. Ringsdorf and P.
Tschirner, Makromol. Chem., Rapid Commun., 9, 445 (1988).
26 H. Ringsdorf, R. Wiistefeld, E. Zerta, M. Ebert and J. H. Wendorff, Angew.
Chem. Int. Ed. Engl., 28, 914 (1989).
H. Ringsdorf and R. Wiistefeld, in Molecular Chemistry for Electronics (eds.
P. Day, D. C. Bradley and D. Bloor). The Royal Society, London, p. 23
(1990); Phil. Trans. Roy. Soc, A330, 95 (1990).
27 V. Luzzati and A. Tardieu, Ann. Rev. of Phys. Chem., 25, 79 (1974).
A. E. Skoulios and V. Luzzati, Acta Cryst., 14, 278 (1961).
A. S. C. Lawrence, Mol. Cryst. Liquid Cryst., 1, 1 (1960).
S. Friberg, J. Am. Oil Chem. Soc, 48, 578 (1971).
P. Ekwall, L. Mandell and K. Fontell, Mol. Cryst. Liquid Cryst., 8, 157
(1969).
F. B. Rosevear, / . Soc. Cosmet. Chem., 19, 581 (1968).

420

28
29
30
31
32
33
34
35
36
37
38
39

40
41
42
43

1
2
3
4
5
6
7
8
9
10

References: Chapter 2
P. A. Winsor, Mol. Cryst. Liquid Cryst., 12, 141 (1971).
G. H. Brown, J. W. Doane and V. D. Neff, A Review of the Structure and
Physical Properties of Liquid Crystals, Butterworths, London (1971).
C. R. Safinya, Lecture Notes for Nato ASI on 'Phase Transitions in Soft
Condensed Matter', Geil, Norway, April 4-14, 1989 (Plenum) (to be
published).
V. Luzzati, A. Tardieu, T. Gulik-Krzywickyi, E. Rivas and F. Reiss-Husson
Nature, 220, 485 (1968).
K. D. Lawson and T. J. Flautt, / . Am. Chem. Soc, 89, 5489 (1967).
C. L. Khetrapal, A. C. Kunwar, A. S. Tracey and P. Diehl, in NMR-Basic,
Principles and Progress, Vol. 9 (eds. P. Diehl, E. Fluck and R. Kosfild),
Springer-Verlag, Berlin (1975).
P. A. Winsor, Chem. Rev., 68, 1 (1968).
J. M. Seddon, Biochim. et Biophys. Ada., 1031, Pl-69 (1990).
C. Robinson, Trans. Faraday Soc, 52, 571 (1956).
D. Chapman, Faraday Soc. Symp., No. 5, 163 (1971).
A. Beguin, J. Billard, F. Bonamy, J. M. Buisine, P. Cuvelier, J. C. Dubois
and P. Le Barny, Mol. Cryst. Liquid Cryst., 115, 9 (1984).
G. W. Gray (ed.) Thermotropic Liquid Crystals, Critical Reports on Applied
Chemistry Vol. 22, Wiley, Chichester (1987), and references therein.
H. Arnold, Z. Phys. Chem. (Leipzig), 226, 146 (1964).
G. W. Gray, in Liquid Crystals and Ordered Fluids, Vol. 2 (eds. J. F.
Johnson and R. S. Porter), p. 617, Plenum (1974).
G. W. Gray, K. J. Harrison and J. A. Nash, Proceedings of the International
Liquid Crystal Conference, Bangalore, December 1973, Pramana Supplement
I, p. 381.
D. Vorlander, Z. Phys. Chem., A126, 449 (1927).
W. L. Mcmillan, Phys. Rev., A6, 936 (1972).
K. Bergmann and H. Stegemeyer, Z. Naturforsch., 34a, 251 (1979).
J. W. Goodby, G. W. Gray, A. J. Leadbetter and M. A. Mazid, Proceedings
of the Conference on Liquid Crystals of One- and Two-dimensional Order,
Garmisch-Partenkirchen, January 1980 (eds. W. Helfrich and G. Heppke),
Springer-Verlag, p. 3 (1980).
Chapter 2
J. D. Bernal and D. Crowfoot, Trans. Faraday Soc, 29, 1032 (1933).
W. R. Krigbaum, Y. Chatani and P. G. Barber, Ada Cryst., B26, 97 (1970).
J. L. Galigne and J. Falgueirettes, Ada Cryst., B24, 1523 (1968).
J. A. Pople and F. E. Karasz, J. Phys. Chem. Solids, 18, 28 (1961).
F. E. Karasz and J. A. Pople, / . Phys. Chem. Solids, 20, 294 (1961).
S. Chandrasekhar, R. Shashidhar and N. Tara, Mol. Cryst. Liquid Cryst.,
10, 337 (1970).
S. Chandrasekhar, R. Shashidhar and N. Tara, Mol. Cryst. Liquid Cryst.,
12, 245 (1972).
J. E. Lennard-Jones and A. F. Devonshire, Proc Roy. Soc, A169, 317
(1939); ibid., A170, 464 (1939).
See, for example, A. R. Ubbelohde, Melting and Crystal and Structure,
Clarendon, Oxford (1965).
See, for example, R. Fowler and E. A. Guggenheim, Statistical
Thermodynamics, Cambridge University Press, Cambridge (1965).
R. H. Wentorf, R. J. Buehler, J. O. Hirschfelder and C. F. Curtiss, / . Chem.
Phys., 18, 1484 (1950).

References: Chapter 2

421

11 L. M. Amzel and L. N. Becka, / . Phys. Chem. Solids, 30, 521 (1969).


See also G. W. Smith, Advances in Liquid Cryst., 1, 189 (1975).
12 D. Demus and R. Rurainski, Z. Phys. Chem. (Leipzig), 253, 53 (1973).
13 E. Mclaughlin, A. Shakespeare and A. R. Ubbelohde, Trans. Faraday Soc,
60, 25 (1964).
14 H. Arnold, Z. Phys. Chem. (Leipzig), 226, 146 (1964).
15 W. Maier and A. Saupe, Z. Naturforsch., 15(a), 287 (1960).
16 E. Bauer and J. Bernamont, / . Phys. Radium, 7, 19 (1936).
17 J. Mayer, T. Waluga and J. A. Janik, Phys. Letters, 41A, 102 (1972).
18 M. J. Press and A. S. Arrott, Phys. Rev., A8, 1459 (1973).
19 N. V. Madhusudana, R. Shashidhar and S. Chandrasekhar, Mol. Cryst.
Liquid Cryst., 13, 61 (1971).
20 A. P. Kapustin and N. T. Bykova, Sov. Phys. - Crystallogr., 11, 297 (1966).
21 S. Chandrasekhar and R. Shashidhar, Mol. Cryst. Liquid Cryst., 16, 21
(1972).
22 S. Chandrasekhar, S. Ramaseshan, A. S. Reshamwala, B. K. Sadashiva, R.
Shashidhar and V. Surendranath, in Proceedings of the International Liquid
Crystals Conference, Bangalore, December 1973, Pramana Supplement I,
p. 117.
R. Shashidhar and S. Chandrasekhar, / . de Physique, 36, Cl-49 (1975).
23 L. Onsager, Ann. N.Y. Acad. Sci., 51, 627 (1949).
24 R. Zwanzig, / . Chem. Phys., 39, 1714 (1963).
25 J. P. Straley, / . Chem. Phys., 57, 3694 (1972).
26 J. E. Mayer and M. G. Mayer, Statistical Mechanics, John Wiley, New York
(1940).
27 J. E. Mayer, Handbuch der Physik, Vol. 12, p. 73, Springer-Verlag, Berlin
(1958).
28 L. K. Runnels and C. Colvin, / . Chem. Phys., 53, 4219 (1970).
29 P. J. Flory, Proc. Roy. Soc, A234, 73 (1956).
P. J. Flory and G. Ronca, Mol. Cryst. Liquid Cryst., 54, 289 (1979).
30 J. P. Straley, Mol. Cryst. Liquid Cryst., 22, 333 (1973).
31 M. A. Cotter, J. Chem. Phys., 66, 1098 (1977).
Also in The Molecular Physics of Liquid Crystals (eds. G. R. Luckhurst and
G. W. Gray) London and New York: Academic Press, p. 169 (1979).
32 H. Reiss, H. L. Frisch and J. L. Lebowitz, / . Chem. Phys., 31, 369 (1959).
33 M. A. Cotter and D. E. Martire, / . Chem. Phys., 52, 1902 (1970); ibid., 52,
1909 (1970); ibid., 53, 4500 (1970).
34 G. Lasher, / . Chem. Phys., 53, 4141 (1970).
K. M. Timling, / . Chem. Phys., 61, 465 (1974).
35 K. L. Savithramma and N. V. Madhusudana, Mol. Cryst. Liquid Cryst., 62,
63 (1980).
36 F. C. Andrews, / . Chem. Phys., 62, 272 (1975).
37 J. Vieillard-Baron, Mol. Phys., 28, 809 (1974).
38 D. W. Rebertus and K. M. Sando, / . Chem. Phys., 67, 2585 (1977).
I. Nezbeda and T. Boublik, / . Phys., B28, 353 (1978).
39 B. Barboy and W. M. Gelbart, / . Chem. Phys., 71, 3053 (1979).
B. M. Mulder and D. Frenkel, Mol. Phys., 55, 1193 (1985).
40 U. P. Singh and Y. Singh, Phys. Rev., A33, 2725 (1986).
A. Perera, G. N. Patey and J. J. Weiss, J. Chem. Phys., 89, 6941 (1988).
J. Colot, X. G. Wu, H. Xu and M. Baus, Phys. Rev., A38, 2022 (1988).
41 S. D. Lee, / . Chem. Phys., 87, 4972 (1987); ibid., 89, 7036 (1988).
42 D. Frenkel and B. M. Mulder, Mol. Phys., 55, 1171 (1985).

422

43
44
45
46
47
48

49
50
51
52
53
54

55
56
57

58
59
60
61

References: Chapter 2
D. Frenkel, B. M. Mulder and J. P. McTague, Phys. Rev. Letters., 52, 287
(1984).
The approximate symmetry between the phase diagrams for rod-like and
disc-like particles was also noted by K. L. Savithramma and N. V.
Madhusudana, Mol. Cryst. Liquid Cryst., 74, 243 (1981); 90, 35 (1982), who
used the scaled particle theory.
A. Wulf and A. G. De Rocco, / . Chem. Phys., 55, 12 (1971).
G. I. Agren and D. E. Martire, / . Chem. Phys., 61, 3959 (1974).
R. Alben, / . Chem. Phys., 59, 4299 (1973).
G. I. Agren and D. E. Martire, J. de Physique, 36, Cl-141 (1975).
J. P. Straley, Phys. Rev., A10, 1881 (1974).
W. M. Gelbart and B. Barbory, Mol. Cryst. Liquid Cryst., 55, 209 (1979).
M. Born, Sitz d. Phys. Math., 25, 614 (1916).
W. Maier and A. Saupe, Z. Naturforsch., 13a, 564 (1958); ibid., U*, 882
(1959); ibid., 15a, 287 (1960).
For an introduction to intermolecular forces, see J. O. Hirschfelder, C. F.
Curtiss and R. B. Bird, Molecular Theory of Gases and Liquids, Part 3,
Wiley, NY, Chapman and Hall, London (1954).
V. Tsvetkov, Acta Physicochim. (USSR), 16, 132 (1942).
A. Saupe and W. Maier, Z. Naturforsch., 16a, 816 (1961).
P. Chatelain, Bull. Soc. Franc. Miner. Crist., 78, 262 (1955).
S. Chandrasekhar and N. V. Madhusudana, / . de Physique, 30, C4-24
(1969).
N. V. Madhusudana, 'A statistical theory of the nematic phase', thesis,
Mysore University, India (1970).
H. S. Subramhanyam and D. Krishnamurti, Mol. Cryst. Liquid Cryst., 22,
239 (1973).
I. Haller, H. A. Huggins, H. R. Lilienthal and T. R. Mcguire, / . Phys.
Chem., 11, 950 (1973).
Y. Poggi, G. Labrunie and J. Robert, C. R. Acad. Sci., B277, 561 (1973).
C. C. Huang, R. S. Pindak and J. T. Ho, / . de Physique Lettres, 35, L-185
(1974).
J. Cheng and R. B. Meyer, Phys. Rev., A9, 2744 (1974).
S. A. Shaya and H. Yu, / . de Physique, 36, Cl-59 (1975).
J. W. Emsley (ed.) Nuclear Magnetic Resonance of Liquid Crystals, NATO
ASI Series C-141 (1985).
B. Cabane and W. G. Clark, Phys. Rev. Letters, 25, 91 (1970); Sol. State
Commun., 13, 129 (1973).
J. C. Rowell, W. D. Phillips, L. R. Melby and M. Panar, J. Chem. Phys., 43,
3442 (1965).
W. D. Phillips, J. C. Rowell and L. R. Melby, / . Chem. Phys., 41, 2551
(1964).
A. Saupe and G. Englert, Phys. Rev. Letters, 11, 462 (1963).
A. Carrington and G. R. Luckhurst, Mol. Phys., 8, 401 (1964).
A. Saupe, Angew. Chem. Intnl. Edn., 1, 97 (1968); Angew. Chemie, 80, 99
(1968).
A. Saupe, in Magnetic Resonance (eds. C. K. Coogan, N. S. Ham, S. N.
Stuart, J. R. Pitbrow and G. V. H. Wilson), Plenum, New York (1970).
L. C. Snyder, / . Chem. Phys., 43, 4041 (1965).
A. D. Buckingham and K. A. McLauchlan, in Progress in NMR
Spectroscopy, Vol. 2, (eds. J. M. Emsley, J. Feeney and L. H. Sutcliffe),
Pergamon, Oxford (1967).

References: Chapter 2

423

62 G. R. Luckhurst, Quart. Rev., 22, 179 (1968); Mol. Cryst. Liquid Cryst., 21,
125 (1973).
63 G. H. Brown, J. W. Doane and V. D. Neff, A Review of the Structure and
Physical Properties of Liquid Crystals, Butterworths, London (1971).
64 P. Diehl and C. L. Khetrapal, in NMR - Basic Principles and Progress, Vol.
1 (eds. P. Diehl, E. Fluck and R. Kosfeld), Springer-Verlag, Berlin (1969).
65 S. Chandrasekhar and N. V. Madhusudana, Applied Spectroscopy Reviews,
6, 189(1972).
66 J. R. McColl and C. S. Shih, Phys. Rev. Letters, 29, 85 (1972).
67 H. Gasparoux, B. Regaya and J. Prost, C R. Acad. ScL, 272B, 1168 (1971).
68 R. Alben, J. R. McColl and C. S. Shih, Sol. State Commun., 11, 1081 (1972).
69 A. Saupe, Z. Naturforsch., 19a, 161 (1964).
70 M. A. Cotter, Mol. Cryst. Liquid Cryst., 39, 173 (1977).
See also B. Widom, J. Chem. Phys., 39, 2808 (1963).
71 S. Chandrasekhar and N. V. Madhusudana, Mol. Cryst. Liquid Cryst., 17, 37
(1972).
72 S. Chandrasekhar and N. V. Madhusudana, Acta Cryst., A27 303 (1971).
S. Chandrasekhar, N. V. Madhusudana and K. Shubha, Faraday Soc.
Syrnp., no. 5, 26 (1971).
73 R. L. Humphries, P. G. James and G. R. Luckhurst, / . Chem. Soc, Faraday
Trans. 2,68, 1031 (1972).
R. L. Humphries and G. R. Luckhurst, Mol. Cryst. Liquid Cryst., 26, 269
(1974).
74 S. Marcelja, / . Chem. Phys., 60, 3599 (1974).
75 G. R. Luckhurst, in Recent Advances in Liquid Crystalline Polymers (ed.
L. L. Chapoy), Elsevier, London, p. 105 (1986).
76 B. Deloche, B. Cabane and D. Jerome, Mol. Cryst. Liquid Cryst., 15, 197
(1971).
77 R. G. Horn, / . de Physique, 39, 199 (1978).
R. G. Horn and T. E. Faber, Proc. Roy. Soc. {London), A368, 199 (1979).
78 S. Chandrasekhar and N. V. Madhusudana, Mol. Cryst. Liquid Cryst., 24,
179 (1973).
79 G. R. Luckhurst, C. Zannoni, P. G. Nordio and C. Segre, Mol. Phys., 30,
1345 (1975).
J. D. Bunning, D. A. Crellin and T. E. Faber, Liquid Crystal, 1, 37 (1986).
B. Bergersen, P. Palffy-Muhoray and D. A. Dunmur, Liquid Crystal, 3, 347
(1988).
80 S. Jen, N. A. Clark, P. S. Pershan and E. B. Priestley, Phys. Rev. Letters, 31,
1552(1973).
S. Jen, N. A. Clark, P. B. Pershan and E. B. Priestley, / . Chem. Phys., 66,
4635 (1977).
81 K. Miyano, / . Chem. Phys., 69, 4807 (1978).
82 S. N. Prasad and S. Venugopalan, / . Chem. Phys., 75, 3035 (1981).
83 G. W. Gray, Molecular Structure and the Properties of Liquid Crystals,
Academic, London (1962).
84 P. J. Flory, Statistical Mechanics of Chain Molecules, Interscience, New
York (1969).
85 R. L. Humphries, P. G. James and G. R. Luckhurst, Symp. Faraday Soc,
no. 5, 107 (1971).
86 P. Palffy-Muhoray, D. A. Dunmur, W. H. Muller and D. A. Balzarini, in
Liquid Crystals and Ordered Fluids, Vol. 4 (ed. A. C. Griffin and J. F.
Johnson), Plenum, New York, p. 615 (1984).

424

References: Chapter 2

87 C. Casagrande, M. Veyssie and H. Finkelmann, / . de Physique Lettres., 43,


L-671 (1982).
88 R. Pratibha and N. V. Madhusudana, Mol. Cryst. Liquid Cryst. Letters, 1,
111 (1985).
89 W. Maier and G. Meier, Z. Naturforsch., 16a, 262 (1961).
90 W. Maier and G. Meier, Z. Naturforsch., 16a, 470 (1961).
91 M. Schadt, J. Chem. Phys., 56, 1494 (1972).
92 C. J. F. Bottcher, Theory of Electric Polarization, Elsevier, Amsterdam
(1973).
93 L. Onsager, J. Am. Chem. Soc, 58, 1486 (1936).
94 For a review of experimental studies on dielectric dispersion, see S.
Chandrasekhar and N. V. Madhusudana, in Topics in Molecular
Interactions (ed. W. J. Orville-Thomas, H. Ratajezak and C. N. R. Rao),
Elsevier, Amsterdam and Indian Academy of Sciences, Bangalore, p. 139
(1985).
95 W. Maier and G. Meier, Z. Naturforsch., 16a, 1200 (1961).
96 A. Axmann, Z. Naturforsch., 21a, 290 (1966).
97 G. Meier and A. Saupe, Mol. Cryst., 1, 515 (1966).
A. J. Martin, G. Meier and A. Saupe, Faraday Soc. Symp. no. 5, 119 (1971).
98 A. Saupe, Z. Naturforsch., 15a, 810 (1960).
99 A. Saupe, Z. Naturforsch., 15a, 815 (I960).
100 I. Haller and J. D. Litster, in Liquid Crystals, Vol. 3 (eds. G. H. Brown
and M. M. Labes), Gordon & Breach, New York, p. 85 (1972).
101 A. Saupe and J. Nehring, / . Chem. Phys., 56, 5527 (1972).
102 R. G. Priest, Phys. Rev., A7, 720 (1973).
103 J. P. Straley, Phys. Rev., A8, 2181 (1973).
104 A. Poniewierski and J. Stecki, Mol. Phys., 38, 1931 (1979).
105 S. D. Lee, Phys. Rev., A39, 3631 (1989).
106 J. Stecki and A. Poniewierski, Mol. Phys., 41, 1451 (1981).
107 W. M. Gelbart and A. Ben-Shaul, / . Chem. Phys., 11, 916 (1982).
108 D. A. Dunmur and W. H. Miller, Chem. Phys. Lett., 86, 353 (1982).
109 K. Singh and Y. Singh, Phys. Rev., A4, 548 (1986); ibid., A35, 3535 (1987).
110 P. Chatelain, Ada Cryst., 1, 315 (1948); Bull. Soc. Franc. Miner. Crist., 11,
353 (1954).
111 B. W. Van der Meer, F. Postma, A. J. Dekker and W. H. de Jeu, Mol.
Phys., 45, 1227 (1982).
112 E. Govers and G. Vertogen, Liquid Crystal, 2, 31 (1987).
113 For up-to-date reviews on the subject see M. A. Cotter, Proceedings of the
International Liquid Crystal Conference, Bangalore, December 1982, Mol.
Cryst. Liquid Cryst., 97, 29 (1983).
M. A. Cotter, Phil. Trans. Roy. Soc, A309, 127 (1983).
114 J. G. J. Ypma and G. Vertogen, Phys. Letts., A60, 212 (1977).
115 W. Wagner, Mol. Cryst. Liquid Cryst., 75, 169 (1981).
116 W. M. Gelbart and B. A. Baron, / . Chem. Phys., 66, 207 (1977).
117 M. A. Cotter, / . Chem. Phys., 66, 4710 (1977).
118 P. Palffy-Muhoray and B. Bergersen, Phys. Rev., A6, 2704 (1987).
119 J. Zadoc-Kahn, Ann. Physique {Series 2), 6, 455 (1936).
120 V. Tsvetkov, Ada Physicochim. (USSR), 19, 86 (1944).
121 N. A. Tolstoi and L. N. Fedotov, / . Exp. Theor. Phys. (USSR), 17, 564
(1947).
122 V. N. Tsvetkov and E. I. Ryumtsev, Sov. Phys. - Crystallogr., 13, 225
(1968).

References: Chapter 2

425

123 B. Cabane and W. G. Clark, Phys. Rev. Lett., 25, 91 (1970).


R. Blinc, D. E. O'Reilly, E. M. Peterson, G. Lahajnar and I. Levstek, Sol.
State Commun., 6, 839 (1968).
S. K. Ghosh, E. Tettamanti, P. L. Indovina, Phys. Rev. Letters, 29, 638
(1972).
See also G. H. Brown, J. W. Doane and V. D. Neff, A Review of the
Structure and Physical Properties of Liquid Crystals, Butterworths, London
(1971).
124 G. Foex, Trans. Faraday Soc, 29, 958 (1933).
125 P. G. de Gennes, Mol. Cryst. Liquid Cryst., 12, 193 (1971).
126 L. D. Landau and E. M. Lifshitz, Statistical Physics, Part I, 3rd edn,
Pergamon, Oxford (1980).
127 See for example, J. W. Beams, Rev. Mod. Phys., 4, 133 (1932).
128 T. W. Stinson and J. D. Litster, Phys. Rev. Letters, 25, 503 (1970).
129 N. V. Madhusudana and S. Chandrasekhar, in Liquid Crystals and Ordered
Fluids, Vol. 2 (eds. J. F. Johnson and R. S. Porter), Plenum, New York, p.
657 (1974).
130 B. R. Ratna, M. S. Vijaya, R. Shashidhar and B. K. Sadashiva in
Proceedings of the International Liquid Crystals Conference, Bangalore,
December 1973, Pramana Supplement I, p. 69.
B. R. Ratna and R. Shashidhar, Pramana, 6, 278 (1976).
131 T. W. Stinson, J. D. Litster and N. A. Clark, J. de Physique, 33, Cl-69
(1972).
132 G. Chu, C. S. Bak, F. L. Lin, Phys. Rev. Letters, 28, 1111 (1972).
133 T. W. Stinson and J. D. Litster, Phys. Rev. Letters, 30, 688 (1973).
134 P. Martinoty, S. Candau and F. Debeauvais, Phys. Rev. Letters, 27, 1123
(1971).
135 N. V. Madhusudana and S. Chandrasekhar, Sol. State Commun., 13, 377
(1973); Pramana, 1, 12 (1973).
136 N. V. Madhusudana and S. Chandrasekhar, in Proceedings of the
International Liquid Crystals Conference, Bangalore, December 1973,
Pramana Supplement I, p. 57.
137 N. V. Madhusudana, K. L. Savithramma and S. Chandrasekhar, Pramana,
8, 22 (1977).
138 H. A. Bethe, Proc. Roy. Soc, 149, 1 (1935).
139 T. S. Chang, Proc. Camb. Phil. Soc, 33, 524 (1937).
140 T. J. Krieger and H. M. James, / . Chem. Phys., 22, 796 (1954).
141 K. L. Savithramma, in 'Theoretical studies on order and phase transitions
in Liquid Crystals', thesis, Mysore University, India (1982).
142 J. G. J. Ypma and G. Vertogen, Sol. State Commun., 18, 475 (1976); Phys.
Letters, 60A, 212; 61A, 45; 61A, 125 (1977).
J. G. J. Ypma, G. Vertogen and H. T. Koster, Mol. Cryst. Liquid Cryst.,
37, 57 (1976).
143 P. Sheng and P. J. Wojtowicz, Phys. Rev., A14, 1883 (1976).
144 J. G. J. Ypma, 'A molecular statistical description of nematic liquid
crystals', thesis, University of Groningen (1977).
145 J. S. Smart, Effective Field Theories of Magnetism, Saunders, Philadelphia
(1966).
146 B. Strieb, H. B. Callen and G. Horwitz, Phys. Rev., 130, 1798 (1963).
147 H. N. W. Lekkerkerker, J. Debruyne, R. Van der Haegen and R. Luyckx,
Physica (Utrecht), 94A, 465 (1978).

426

References: Chapter 3

148 R. Van der Haegen, J. Debruyne, R. Luyckx and H. N. W. Lekkerkerker,


/ . Chem. Phys., 73, 2467 (1980).
149 G. Ballensiefen and D. Wagner, Phys. Stat. Sol, 12, 613 (1965).
150 H. J. F. Jansen, G. Vertogen and J. G. J. Ypma, Mol Cryst. Liquid Cryst.,
38, 87 (1977).
151 G. R. Luckhurst and P. Simpson, Mol. Phys., 47, 251 (1982).
152 M. J. Bradshaw and E. P. Raynes, Mol. Cryst. Liquid Cryst. Letters, 72, 73
(1981); Mol. Cryst. Liquid Cryst., 91, 145 (1983).
For a full list of references to dielectric studies on strongly polar materials
see S. Chandrasekhar, Plenary Lecture at the Tenth International Liquid
Crystal Conference, York, July 15-21, 1984; Mol. Cryst. Liquid Cryst.,
124, 1 (1985).
153 A. J. Leadbetter, R. M. Richardson and C. N. Colling, J. de Physique, 36,
Cl-37 (1975).
154 C. A. Croxton and S. Chandrasekhar, in Proceedings of the International
Liquid Crystals Conference, Bangalore, December 1973, Pramana
Supplement /, p. 237
155 A. Ferguson and S. J. Kennedy, Phil. Mag., 26, 41 (1938).
156 S. Krishnaswamy and R. Shashidhar, in Proceedings of the International
Liquid Crystals Conference, Bangalore, December 1973, Pramana
Supplement I, p. 247.
157 S. Krishnaswamy and R. Shashidhar, Mol. Cryst. Liquid Cryst., 35, 253
(1976), ibid., 38, 353 (1977).
158 S. Krishnaswamy, Proceedings of the International Liquid Crystal
Conference, Bangalore, 1979 (ed. S. Chandrasekhar), Hey den, London, p.
487 (1980).
159 M. G. J. Gannon and T. E. Faber. Phil Mag., 37, 117 (1978).
160 See for example, Statistical Mechanics of the Liquid Surface by C. A.
Croxton, John Wiley, Chichester (1980).
161 D. Langevin, / . de Physique, 33, 249 (1972).
D. Langevin and M. A. Bouchiat, / . de Physique, 33, 101 (1972); ibid., 33,
Cl-77 (1972).
162 P. S. Pershan, A. Braslau, A. H. Weiss and J. Als-Nielsen, Phys. Rev., A35,
4800 (1987).
163 B. M. Ocko, A. Braslau, P. S. Pershan, J. Als-Nielsen and M. Deutsch,
Phys. Rev. Letters, 57, 94 (1986).
1
2
3
4
5
6
7
8

Chapter 3
C. W. Oseen, Trans. Faraday Soc, 29, 883 (1933).
H. Zocher, Trans. Faraday Soc, 29, 945 (1933).
F. C. Frank, Disc. Faraday Soc, 25, 19 (1958).
A. Anzelius, Uppsala Univ. Arsskr., Mat. Och Naturvet, 1, (1931).
J. L. Ericksen, Arch. Rational Mech. Anal, 4, 231 (I960).
J. L. Ericksen, Trans. Soc. Rheol, 5, 23 (1961).
F. M. Leslie, Quart. J. Mech. Appl Math., 19, 357 (1966); Arch. Rational
Mech. Anal, 28, 265 (1968); Advances in Liquid Crystals, Vol. 4 (ed. G. H.
Brown), p. 1, Academic, New York (1979).
P. C. Martin, O. Parodi and P. S. Pershan, Phys. Rev., A6, 2401 (1972).
J. D. Lee and A. C. Eringen, / . Chem. Phys., 54, 5027 (1971); ibid., 55, 4504
4509 (1971); ibid., 58, 4203 (1973).
H. Schmidt and J. Jahnig, Annals of Physics, 71, 129 (1972).
For a full list of references to papers on continuum theories, see review by

References: Chapter 3

9
10
11
12
13
14
15
16
17
18
19
20
21
22

23
24
25

26

27

28
29
30

427

J. D. Lee and A. C. Eringen in Liquid Crystals and Ordered Fluids, Vol. 2


(eds. J. F. Johnson and R. S. Porter), p. 315, Plenum, New York (1974).
See, for example, C. Truesdell and W. Noll, The Non-Linear Field Theories
of Mechanics, Handbuch der Physik, Vol. 3/3, Springer- Verlag, Berlin
(1965).
L. Onsager, Phys. Rev., 37, 405 (1931); ibid., 38, 2265 (1932).
O. Parodi, / . de Physique, 31, 581 (1970).
C. Truesdell, Rational Thermodynamics, McGraw-Hill, New York (1969).
W. W. Beens and W. H. de Jeu, / . de Physique, 44, 129 (1983).
See, for example, J. F. Nye, Physical Properties of Crystals, Oxford
University Press, Oxford (1957).
J. L. Ericksen, Phys. Fluids, 9, 1205 (1966).
See, for example W. L. Ferrar, Algebra, A Text Book of Determinants,
Matrices and Algebraic Forms, Oxford University Press, Oxford (1941).
J. L. Ericksen, Arch. Rational Mech. Anal., 10, 189 (1962).
V. Freedericksz and V. Tsvetkov, Phy. Z. Soviet Union, 6, 490 (1934).
See also, V. Freedericksz and V. Zolina, Trans. Faraday Soc, 29, 919 (1933).
A. Saupe, Z. Naturforsch., 15a, 815 (1960).
H. Gruler, T. J. Scheffer and G. Meier, Z. Naturforsch, 27a, 966 (1972).
See also, I. Haller, / . Chem. Phys., 57, 1400 (1972).
P. G. de Gennes, Mol. Cryst. Liquid Cryst., 12, 193 (1971).
N. V. Madhusudana, P. P. Karat and S. Chandrasekhar, in Proceedings of
the International Liquid Crystal Conference, Bangalore, December 1973,
Pramana Supplement I (ed. S. Chandrasekhar), p. 225.
P. P. Karat, 4 Electric and Magnetic field effects in Liquid Crystals', thesis,
University of Mysore, India, 1977.
P. E. Cladis, Phys. Rev. Letters, 31, 1200 (1973).
P. Chatelain, Bull. Soc. Franc. Miner. Crist., 66, 105 (1943).
D. Berreman, Phys. Rev. Letters, 28, 1683 (1972).
J. L. Janning, Appl. Phys. Letters, 21, 173 (1972).
E. Guyon, P. Pieranski and M. Boix, Lett, in Applied and Eng. Science, 1, 19
(1973).
A. Rapini and M. Papoular, / . de Physique, 30, C4-C5 (1971); for a
discussion of the anchoring properties and alignment of liquid crystals, see
E. Guyon and W. Urbach, in Electrooptic Non-emissive Displays (eds. A. R.
Kmetz and F. K. Von Willisen), p. 121, Plenum, New York (1976).
J. Cognard, Mol. Cryst. Liquid Cryst., Suppl. Ser. I, 1 (1982).
B. Ya Zel'dovich, N. V. Tabiryan and Yu S. Chilmgaryan, Sov. Phys.
JETP, 54, 32 (1981).
S. D. Durbin, S. M. Arakelian and Y. R. Shen, Phys. Rev. Lett., 47, 1411
(1981).
C. J. Gerritsma, W. H. de Jeu and P. Van Zanten, Phys. Lett., 36A, 389
(1971).
J. Bradshaw, D. G. McDonnell and E. P. Raynes, Mol. Cryst. Liquid Cryst.,
70,289(1981).
J. Bradshaw, E. P. Raynes, J. D. Bunning and T. E. Faber, / . de Physique,
46, 1513 (1985).
C. Z. Van Doom, Phys. Lett., 42A, 537 (1973).
F. M. Leslie, Mol. Cryst. Liquid Cryst., 12, 57 (1970).
M. Schadt and W. Helfrich, Appl. Phys. Lett., 18, 127 (1971).
S. L. Arora and J. L. Fergason, in Proceedings of the International Liquid

428

31

32
33
34

35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62

References: Chapter 3
Crystals Conference, Bangalore, December 1973, Pramana Supplement I (ed.
S. Chandrasekhar), p. 520.
G. W. Gray, K. J. Harrison and J. A. Nash, Electron. Lett., 9, 130 (1973); in
Proceedings of International Liquid Crystal Conference, Bangalore, December
1973, Pramana Supplement I (ed. S. Chandrasekhar) p. 381.
See also G. W. Gray (ed.) Thermotropic Liquid Crystals, Wiley, Chichester
(1987).
T. J. Scheffer, SID Seminar Lecture Notes, 7.1 (1986).
P. M. Alt and P. Pleshko, IEEE Trans. Electron Devices, ED-21, 146 (1974).
C. M. Waters, V. Brimmell and E. P. Raynes, in Proceedings of the Third
International Display Research Conference, Kobe, Japan (1983), p. 396; Proc.
SID 25/4, 261 (1984).
C. M. Waters, E. P. Raynes and V. Brimmell, Mol. Cryst. Liquid Cryst., 123,
303 (1985).
E. P. Raynes and C. M. Waters, Displays, April 1987, p, 59.
T. J. Scheffer and J. Nehring, Appl. Phys. Lett., 45, 1021 (1984).
W. R. Hoffner and D. W. Berremann, J. Appl. Phys., 53, 8599 (1982).
E. P. Raynes, Mol. Cryst. Liquid Cryst. Lett., 4, 1 (1986); 4, 69 (1987).
D. W. Berremann, Phil. Trans. Roy. Soc, A309, 203 (1983).
V. G. Taratula, A. J. Hurd and R. B. Meyer, Phys. Rev. Lett., 55, 246
(1985).
For a concise review of Freedericksz transition in polymer nematics, see U.
D. Kini, Mol. Cryst. Liquid Cryst., 153, 1 (1987).
F. Lonberg and R. B. Meyer, Phys. Rev. Lett., 55, 718 (1985).
U. D. Kini, / . de Physique, 47, 693 (1986).
C. Oldano, Phys. Rev. Lett., 56, 1098 (1986).
W. Zimmerman and L. Kramer, Phys. Rev. Lett., 56, 2655 (1986).
E. Miraldi, C. Oldano and A. Strigazzi, Phys. Rev. A 34, 4348 (1986).
J. Nehring and A. Saupe, / . Chem. Phys., 54, 337 (1971).
H. P. Hinov and A. I. Derzhanski, / . de Physique Colloq., France 40, C3-308
(1979).
T. V. Korkishko, V. G. Chigrinov, R. V. Galiulin, A. A. Sonin and N. A.
Tikhnomirova, Sov. Phys. Cristallogr., 32, 263 (1987).
C. Oldano and G. Barbero, Phys. Lett., 110A, 213 (1985); / . de Physique
Lett., 46,451 (1985).
G. Barbero and A. Strigazzi, Liquid Crystals, 5, 693 (1989).
C. Oldano, presented at the Int. Topical Meeting on Optics of Liquid
Crystals, Torino, Italy, 14-20, Oct. 1988.
H. P. Hinov, Mol. Cryst. Liquid Cryst., 148, 197 (1987).
A. I. Derzhanski and H. P. Hinov, Phys. Lett., A56, 465 (1976).
M. I. Barnik, L. M. Blinov, T. V. Korkishko, B. A. Umansky and V. G.
Chigrinov, Mol. Cryst. Liquid Cryst., 99, 53 (1983).
S. Faetti and V. Palleschi, / . de Physique, 46, 415 (1985).
N. V. Madhusudana and R. Pratibha, Mol. Cryst. Liquid Cryst., 179, 207
(1990).
S. Faetti, Mol. Cryst. Liquid Cryst., 179, 217 (1990).
G. Barbero, N. V. Madhusudana and C. Oldano, / . de Physique, 50, 2263
(1989).
O. Lehmann, Ann. Physik, 2, 649 (1900).
G. Friedel, Ann. Physique, 18, 273 (1922).
J. Nehring and A. Saupe, / . Chem. Soc, Faraday Trans. II, 68, 1 (1972).
M. Kleman, Points, Lines and Walls, Wiley, Chichester (1983).

References: Chapter 3
63
64
65
66
67

429

M. Kleman, Rep. Prog. Phys., 52, 555 (1989).


S. Chandrasekhar and G. S. Ranganath, Adv. Phys., 35, 507 (1986).
W. F. Brinkman and P. E. Cladis, Physics Today, 35(5), 48 (1982).
J. Friedel, Dislocations, Pergamon, Oxford (1967).
See R. de Wit, Fundamental Aspects of Dislocation Theory (eds. J. A.
Simmons, R. de Wit and R. Bullough), Spl Pub. No. 317, National Bureau
of Standards, 1, 715(1970).
68 A. Saupe, Mol. Cryst. Liquid Cryst., 21, 211 (1973).
69 C. M. Dafermos, Quart. J. Mech. Appl. Math., 23, 549 (1970).
70 J. L. Ericksen, Liquid Crystals and Ordered Fluids (eds. J. F. Johnson and
R. S. Porter), p. 189, Plenum, New York (1970).
71 H. Imura and K. Okano, Phys. Lett., 42A, 405 (1973).
72 G. S. Ranganath, Mol. Cryst. Liquid Cryst., 34, 71 (1976).
73 J. D. Eshelby, Phil. Mag., 42A, 359 (1980).
74 G. S. Ranganath, Mol. Cryst. Liquid Cryst., 97, 77 (1983).
75 P. E. Cladis and M. Kleman, / . de Physique, 33, 591 (1972).
76 R. B. Meyer, Phil. Mag., 27, 405 (1973).
77 C. Williams, P. Pieranski and P. E. Cladis, Phys. Rev. Lett., 29, 90 (1972).
78 J. Friedel and P. G. de Gennes, C. R. Acad. Sci., Paris, B268, 257 (1969).
79 A. M. J. Spruijt, Solid State Comm., 13, 1919 (1973).
80 J. Nehring, Phys. Rev., XI, 1737 (1973).
81 G. S. Ranganath, Mol. Cryst. Liquid Cryst., 87, 187 (1982).
82 E. Dubois-Violette and O. Parodi, J. Phys., 30, C4-57 (1969).
83 S. Candau, P. Le Roy and F. Debeauvais, Mol. Cryst. Liquid Cryst., 23, 283
(1973).
84 M. J. Press and A. S. Arrott, / . de Physique, 36, Cl-177 (1975).
85 N. V. Madhusudana and K. R. Sumathy, Mol. Cryst. Liquid Cryst. Lett.,
92, 179 (1983).
86 G. E. Volovik, Sov. Phys. JETP, 58, 1159 (1983).
87 R. B. Meyer, Mol. Cryst. Liquid Cryst., 16, 355 (1972).
88 G. S. Ranganath, Mol. Cryst. Liquid Cryst., 40, 143 (1977).
89 W. Helfrich, Phys. Rev. Lett., 21, 1518 (1968).
90 F. Brochard, / . de Physique, 33, 607 (1972).
L. Leger, Mol. Cryst. Liquid Cryst., 24, 33 (1973).
91 A. Rapini, / . Phys., Paris, 34, 629 (1973).
92 V. P. Mineev and G. E. Volovik, Phys. Rev. B, 18, 3197 (1978).
93 V. P. Mineev, Soviet Scientific Reviews, Section A, Phys. Rev., Vol. 2 (ed. by
I. M. Khalatnikov), Harwood Academic Publishers, London, p. 173 (1980).
94 P. B. Sunil Kumar and G. S. Ranganath, Mol. Cryst. Liquid Cryst., 177, 131
(1989).
95 U. D. Kini and G. S. Ranganath, Mol. Cryst. Liquid Cryst., 38, 311 (1977).
96 I. E. Dzyaloshinskii, Sov. Phys. JETP, 31, 773 (1970).
97 R. Nityananda and G. S. Ranganath, Proceedings of the International
Conference on Liquid Crystals, Bangalore, 1979 (ed. S. Chandrasekhar), p.
205, Heyden, London (1980).
98 G. S. Ranganath, Proceedings of the International Conference on Liquid
Crystals, Bangalore, 1979 (ed. S. Chandrasekhar), p. 213, Heyden, London
(1980).
99 J. L Ericksen, Liquid Crystals and Ordered Fluids (ed. J. F. Johnson and
R. S. Porter), Plenum, New York, p. 181 (1970).
100 F. C. Frank, Phil. Mag., 42, 809, 1014 (1951).

430

References: Chapter 3

101 For a review of viscometric studies on the different mesophase types see
R. S. Porter, E. M. Barrall and J. F. Johnson, / . Chem. Phys., 45, 1452 (1966).
102 M. Miesowicz, Nature, 158, 27 (1946).
103 F. M. Leslie, G. R. Luckhurst and H. J. Smith, Chem. Phys. Lett., 13, 368
(1972).
104 V. Tsvetkov, Acta Physicochim. (USSR), 10, 557 (1939).
105 H. Gasparoux and J. Prost, / . de Physique, 32, 65 (1971).
106 P. G. de Gennes, / . de Physique, 32, 789 (1971).
107 F. J. Bock, H. Kneppe and F. Schneider, Liquid Crystals, 3, 217 (1988).
108 R. J. Atkin, Arch. Rational Mech. Anal, 38, 224 (1970).
109 J. L. Ericksen, Trans. Soc. Rheol, 13, 9 (1969).
110 J. Fisher and A. G. Frederickson, Mol. Cryst. Liquid Cryst., 6, 255 (1969).
111 H. C. Tseng, D. K. Silver and B. A. Finlayson, Phys. Fluids, 15, 1213
(1972).
112 U. D. Kini and G. S. Ranganath, Pramana, 4, 19 (1975).
113 B. A. Finlayson, in Liquid Crystals and Ordered Fluids, Vol. 2 (eds. J. F.
Johnson and R. S. Porter), p. 211, Plenum, New York (1974).
114 Ch. Gahwiller, Mol. Cryst. Liquid Cryst., 20, 301 (1973).
115 F. J. O'Neill, Liquid Crystals, 1, 271 (1986).
116 K. Skarp, T. Carlsson, S. T. Lagerwall and B. Stebler, Mol. Cryst. Liquid
Cryst., 66, 199(1981).
T. Carlsson and K. Skarp, ibid, 78, 157 (1981).
117 T. Carlsson, 'Theoretical investigations of the flow properties of nematic
liquid crystals', thesis, Chalmers University of Technology, Goteborg
(1984).
118 P. G. de Gennes, Phys. Letters, 41A, 479 (1972).
119 F. M. Leslie, Mol. Cryst. Liquid Cryst., 1, 407 (1969).
120 P. Pieranski and E. Guyon, Phys. Letters, A 49, 237 (1974).
121 P. Martinoty and S. Candau, Mol. Cryst. Liquid Cryst., 14, 243 (1971).
122 P. J. Barratt and F. M. Leslie, / . de Physique, 40, C3-73 (1979).
123 P. Pieranski, F. Brochard and E. Guyon, / . de Physique, 34, 35 (1973).
See also F. Brochard, Mol Cryst. Liquid Cryst., 23, 51 (1973).
124 L. Leger, Sol. State Comm., 10, 697 (1972); Mol. Cryst. Liquid Cryst.,
24, 33 (1973).
125 M. G. Clark and F. M. Leslie, Proc. Roy. Soc, A361, 463 (1978).
126 C. Z. Van Doom, / . de Physique, 36, CI-261 (1975).
127 C. Z. Van Doom, / . Appl Phys., 46, 3738 (1975); C. Z. Van Doom, C. J.
Gerritsma and J. J. de Klerk in The Physics and Chemistry of Liquid
Crystal Devices (ed. G. Sprokel), p. 95, Plenum Press, New York (1980).
D. W. Berreman, / . Appl. Phys., 8, 101 (1975).
128 P. Chatelain, Acta Cryst., 4, 453 (1951).
129 P. G. de Gennes, C R. Acad. Scl, 266, 15 (1968).
130 Orsay Liquid Crystals Group, / . Chem. Phys., 51, 816 (1969).
131 Orsay Liquid Crystals Group, Phys. Rev. Letters, 22, 1361 (1969).
L. Leger-Quercy, ' Etude experimentale des fluctuations thermiques
d'orientation dans un cristal liquide nematique par diffusion inelastique de
la lumiere', thesis, University of Paris, France (1970).
132 L. D. Landau and E. M. Lifshitz, Statistical Physics, Part I, 3rd edn.,
Pergamon, Oxford (1980).
See also, I. L. Fabelinskii, Molecular Scattering of Light, Plenum, New
York (1968).
133 W. Kast, Angew. Chemie, 67, 592 (1955).

References:

Chapter 3

431

134 E. F. Carr, / . Chem. Phys., 38, 1536 (1963); ibid., 39, 1979 (1963); ibid., 42,
738 (1965); Advances in Chemistry Series, 63, 76 (1967); Mol. Cryst. Liquid
Cryst., 1, 253 (1969).
135 V. N. Tsvetkov and G. M. Mikhailov, Acta Physicochim. (USSR), 8, 77
(1938).
136 Y. Bjornstahl, Z. Phys. Chem., A175, 17 (1933).
137 V. Naggiar, Ann. Physique, 18, 5 (1943).
138 R. Williams, / . Chem. Phys., 39, 384 (1963).
139 G. H. Heilmeier, L. A. Zanoni and L. A. Barton, Proc. IEEE, 56, 1162
(1968).
140 A. P. Kapustin and L. K. Vistin, Kristallografiya, 10, 118 (1965).
G. Elliot and J. G. Gibson, Nature, 205, 995 (1965).
141 P. A. Penz, Phys. Rev. Lett., 24, 1405 (1970).
142 Orsay Liquid Crystals Group, Mol. Cryst. Liquid Cryst., 12, 251 (1971).
143 P. A. Penz, Mol. Cryst. Liquid Cryst., 15, 141 (1971).
144 H. Gruler and G. Meier, Mol. Cryst. Liquid Cryst., 16, 299 (1972).
145 W. Helfrich, / . Chem. Phys. 51, 4092 (1969).
For a concise review of electric field effects see W. Helfrich, Mol. Cryst.
Liquid Cryst., 21, 187 (1973).
146 E. Dubois-Violette, P. G. de Gennes and O. Parodi, / . de Physique, 32, 305
(1971).
147 I. W. Smith, Y. Galerne, S. T. Lagerwall, E. Dubois-Violette and G.
Durand, / . de Physique, 36, Cl-237 (1975).
148 See, for example, L. D. Landau and E. M. Lifshitz, Fluid Mechanics,
Pergamon, Oxford (1966).
149 P. A. Penz and G. W. Ford, Phys. Rev., A6, 414 (1972).
150 W. Greubel and U. Wolff, Appl. Phys. Lett., 19, 213 (1971).
151 L. K. Vistin, Sov. Phys. - Crystallogr., 15, 514 (1970).
152 E. Dubois-Violotte, G. Durand, E. Guyon, P. Manneville and P. Pieranski,
Solid State Phys., Suppl. 41, 147 (1978).
153 W. J. A. Goossens, Adv. in Liquid Crystals, Vol. 3 (ed. G. H. Brown), p. 1,
Academic, New York (1978).
154 See L. M. Blinov, Electro and Magneto-Optics of Liquid Crystals, Nauka,
Moscow (1978).
155 For a concise review of instabilities in nematics and cholesterics, see S.
Chandrasekhar and U. D. Kini in Polymer Liquid Crystals, (eds. A. Cferri,
W. R. Krigbaum and R. B. Meyer), Chapter 8, Academic, New York
(1982).
156 E. Dubois-Violette, C. R. Hebd. Seanc Acad. Sci., Ser. B 273, 923-6
(1971).
157 E. Guyon and P. Pieranski, C R. Hebd. Seanc Acad. Sci., Ser. B 274, 656
(1972).
158 P. Pieranski and E. Guyon, Solid State Commun., 13, 435 (1973).
159 P. Pieranski and E. Guyon, Phys. Rev., A 9, 404 (1974).
160 P. Manneville and E. Dubois-Violette, J. de Physique, 37, 285 (1976).
161 F. M. Leslie, / . de Physique, D 9, 925 (1976).
162 P. G. de Gennes, in Molecular Fluids (eds. R. Balian and G. Weill),
Gordon & Breach, New York p. 373 (1976).
163 E. Dubois-Violette, E. Guyon, I. Janossy, P. Pieranski and P. Manneville,
/ . Mec, 16, 733 (1977).
164 S. Chandrasekhar, Hydrodynamic & Hydromagnetic Stability, Clarendon
Press, Oxford (1961).

432

References: Chapter 4

165 H. Hervet, F. Rondelez and W. Urbach, Proceedings of the International


Liquid Crystal Conference, Bangalore, 1979 (ed. S. Chandrasekhar), p. 263,
Heyden, London (1980).
166 E. Dubois-Violette, E. Guyon and P. Pieranski, Mol. Cryst. Liquid Cryst.,
26, 193 (1974).
167 P. Pieranski, E. Dubois-Violette and E. Guyon, Phys. Rev. Lett., 30, 736
(1973).
168 H. N. W. Lekkerkerker, / . de Physique Lett., 38, 277 (1977).
169 E. Guyon, P. Pieranski and J. Salan, C. R. Acad. ScL, B 287, 41 (1978).
170 E. Guyon, P. Pieranski and J. Salan, / . Fluid. Mech., 93, 65 (1979).
171 R. B. Meyer, Phys. Rev. Lett., 22, 918 (1969).
172 J. Prost and J. P. Marcerou, 7. de Physique, 38, 315 (1977).
173 J. P. Marcerou and J. Prost, Mol. Cryst. Liquid Cryst., 58, 259 (1980).
174 I. Dozov, Ph. Martinot-Lagarde and G. Durand, / . de Physique Lettres, 43
L-365 (1982).
175 J. Prost and P. S. Pershan, J. Appl. Phys., 47, 2298 (1976).
176 I. Dozov, I. Penchev, Ph. Martinot-Lagarde and G. Durand, Ferroelectrics
Lett., 2, 135 (1984).
177 D. Schmidt, M. Schadt and W. Helfrich, Z. Naturforsch. 27A, 277 (1972).
178 N. V. Madhusudana and G. Durand, / . Phys. Lett., 46, L-195 (1985).
179 C. Hilsum and F. C. Saunders, Mol. Cryst. Liquid Cryst., 64, 25 (1980).
180 S. Hirata and T. Tako, Jpn. J. Appl. Phys. 20, L459 (1981).
181 R. Ribotta, A. Joets and L. Lei, Phys. Rev. Lett., 56, 1595 (1986).
182 W. Zimmermann and L. Kramer, Phys. Rev. Lett., 55, 402 (1985).
183 N. V. Madhusudana, V. A. Raghunathan and K. R. Sumathy, Pramana J. Phys. 28, L-311 (1987).
184 N. V. Madhusudana and V. A. Raghunathan, Mol. Cryst. Liquid Cryst.
Lett., 5, 201 (1988).
185 V. A. Raghunathan and N. V. Madhusudana, Pramana - J. Phys., 31,
L-163 (1988).
186 W. Thorn, W. Zimmermann and L. Kramer, Liquid Crystals, 4, 309 (1989).
187 L. Kramer, E. Bodenschatz, W. Pesch, W. Thorn and W. Zimmermann,
Liquid Crystals, 5, 699 (1989).
188 G. Barbero, I. Dozov, J. F. Palierne and G. Durand, Phys. Rev. Lett., 56,
2054 (1986).
189 G. Durand, in Incommensurate Crystals, Liquid Crystals and Quasicrystals
(eds. J. E. Scott and N. A. Clark), p. 235, Plenum Press, New York (1988).
190 S. Faetti and V. Palleschi, Phys. Rev., A 30, 3241 (1984).
191 When the dielectric anisotropy is very large the Freedericksz deformation
can set in discontinuously (i.e., as a first order transition) in certain
geometries; see S. M. Arakelyan, A. S. Karayan and Yu. S. Chilingaryan,
Dokl. Akad. Nauk. SSR, 275, 52 (1984).
B. J. Frisken and P. Palffy-Muhoray, Phys. Rev., A 39, 1513 (1989).
U. D. Kini, Liquid Crystals, 8, 745 (1990).
Chapter 4
M. C. Mauguin, Bull. Soc. Franc. Miner. Crist., 34, 71 (1911).
C. W. Oseen, Trans. Faraday Soc, 29, 883 (1933).
H. de Vries, Acta Cryst., 4, 219 (1951).
H. Poincare, Theorie Mathematique de la Lumiere, Vol. 2, Paris, Chap 12
(1982).
5 R. C. Jones, / . Opt. Soc. Am., 31, 488 (1941); ibid., 31, 493 (1941).

1
2
3
4

References: Chapter 4

433

6 See, for example, G. N. Ramachandran and S. Ramaseshan, Handbuch der


Physik, Vol. 25/1, Crystal Optics, Springer-Verlag, Berlin (1961).
7 S. Chandrasekhar and K. N. Srinivasa Rao, Acta Cryst., A24, 445 (1968).
8 R. C. Jones, / . Opt. Soc. Am., 32, 486 (1942).
9 F. Abeles, Ann. Physique, 5, 777 (1950).
10 S. Chandrasekhar, G. S. Ranganath, U. D. Kini and K. A. Suresh, Mol.
Cryst. Liquid Cryst., 24, 201 (1973).
11 Analogous expressions for a twisted pile of crystal plates were derived by: L.
Sohncke, Math. Ann., 9, 504 (1876); F. Pockels, Lehrbuch der Kristalloptik,
p. 289, Teubner, Leipzig (1906); R. C. Jones, J. Opt. Soc. Am., 31, 500
(1941).
12 C. Robinson, Tetrahedron, 13, 219 (1961).
13 R. Cano and P. Chatelain, C R. Acad. Sci., 259, 352 (1964).
14 S. Chandrasekhar, G. S. Ranganath and U. D. Kini, Mol. Cryst. Liquid
Cryst. Lett., 3, 163 (1986).
15 E. P. Raynes and R. J. A. Tough, Mol. Cryst. Liquid Cryst. Lett., 2, 141
(1985).
16 R. Nityananda and U. D. Kini, in Proceedings of the International Liquid
Crystals Conference, Bangalore, December 1973, Pramana Supplement I, p.
311.
17 S. Chandrasekhar, Trieste Lectures on Polymers, Liquid Crystals and LowDimensional Solids, 1980 (eds. N. H. March and M. P. Tosi), Chapter 9,
Plenum, New York (1984).
18 G. S. Ranganath, S. Chandrasekhar, U. D. Kini, K. A. Suresh and S.
Ramaseshan, Chem. Phys. Lett., 19, 556 (1973).
19 E. Sackmann and J. Voss, Chem. Phys. Lett., 14, 528 (1972).
20 G. S. Ranganath, K. A. Suresh, S. R. Rajagopalan and U. D. Kini, in
Proceedings of the International Liquid Crystals Conference, Bangalore,
December 1973, Pramana Supplement I, p. 353.
21 S. Chandrasekhar, G. S. Ranganath and K. A. Suresh, in Proceedings of the
International Liquid Crystals Conference, Bangalore, December 1973,
Pramana Supplement I, p. 341.
22 R. Nityananda, U. D. Kini, S. Chandrasekhar and K. A. Suresh,
in Proceedings of the International Liquid Crystals Conference, Bangalore,
December 1973, Pramana Supplement I, p. 325.
23 C. G. Darwin, Phil. Mag., 27, 315, 675 (1914); ibid., 43, 800 (1922).
24 I. S. Gradshteyn and I. M. Ryzhik, Tables of Integrals, Series and Products,
p. 27, Academic, New York (1965).
25 R. Dreher, G. Meier and A. Saupe, Mol. Cryst. Liquid Cryst., 13, 17 (1971).
26 S. Chandrasekhar and J. Shashidhara Prasad, Mol. Cryst. Liquid Cryst., 14,
115(1971).
Similar measurements have also been reported by J. C. Martin and R. Cano,
C. R. Acad. Sci., B278, 219 (1974).
27 G. Borrmann, Physik. Z., 42, 157 (1941).
P. P. Ewald, Rev. Mod. Phys., 37, 46 (1965).
28 K. A. Suresh, Mol. Cryst. Liquid Cryst., 35, 267 (1976).
29 S. N. Aronishidze, V. E. Dmitrienko, D. G. Khostariya and G. S. Chilaya,
JETP Lett., 32, 17(1980).
30 E. I. Kats, Sov. Phys.-JETP, 32, 1004 (1971).
31 R. Nityananda, Mol. Cryst. Liquid Cryst., 21, 315 (1973).
32 G. Joly, ' Contribution a l'Etude de la Propagation de Certaines

434

33
34
35
36

37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52

53
54
55
56
57
58
59
60
61
62

References: Chapter 4
Ondes electromagnetiques dans les Piles de Reusch', thesis, University of
Science and Technology, Lille, France (1972).
M. Aihara and H. Inaba, Optics Commun., 3, 77 (1971).
A. S. Marathay, / . Opt. Soc. Am., 61, 1363 (1971).
R. Nityananda, ' Study of Some Novel Optical Effects in Periodic
Structures', thesis, Bangalore University, India (1975).
See, for example, M. Born and E. Wolf, Principles of Optics, Pergamon,
Oxford (1959).
An analytical solution in the two-wave approximation has been given by
V. A. Belyakov and V. E. Dmitrienko, Sov. Phys. - Solid State, 15, 1811
(1974).
V. E. Dmitrienko and V. A. Belyakov, Sov. Phys. - Solid State, 15, 2365
(1974).
D. Taupin, / . de Physique, 30, C4-32 (1969).
D. W. Berreman and T. J. Scheffer, Phys. Rev. Lett., 25, 577 (1970).
E. Sackmann, S. Meiboom, L. C. Snyder, A. E. Meixner and R. E. Dietz,
/ . Am. Chem. Soc, 90, 3567 (1968).
See, for example, R. W. Wood, Physical Optics, 3rd edn., Macmillan, New
York (1934).
S. Chandrasekhar and J. Shashidhara Prasad, in Physics of the Solid State
(eds. S. Balakrishna, M. Krishnamurthi and B. R. Rao), p. 77, Academic,
London (1969).
P. E. Cladis, A. E. White and W. F. Brinkmann, / . de Physique, 40, 325
(1979).
J. Rault, Phil. Mag., 28, 11 (1973).
J. Rault, Sol State Commun., 9, 1965 (1971).
J. Rault, Phil Mag., 30, 621 (1974).
P. G. de Gennes, C. R. Hebd Seanc. Acad. Sci., Paris, 266B, 571 (1968).
T. J. Scheffer, Phys. Rev., A 5, 1327 (1972).
C. Caroli and E. Dubois-Violette, Sol State Commun., 7, 799 (1969).
Y. Bouligand and M. Kleman, / . de Physique, 31, 1041 (1970).
C. Robinson, J. C. Ward and R. B. Beevers, Disc. Faraday Soc, 25, 29
(1958).
Y. Bouligand, Defects and Textures in Liquid Crystals in Dislocations in
Solids, Vol. 5 (ed. F. R. N. Nabarro) p. 299, North Holland, Amsterdam
(1980).
D. Demus and L. Richter, Textures of Liquid Crystals, Verlag Chemie, New
York, 1978.
G. W. Gray and J. W. Goodby, Smectic Liquid Crystals, Leonard Hill,
Glasgow, 1984.
G. S. Ranganath, Mol. Cryst. Liquid Cryst., 94, 285 (1983).
Orsay Liquid Crystal Group, / . de Physique, 30, C4-38 (1969).
M. Kleman and J. Friedel, / . de Physique, 30, C4-43 (1969).
F. M. Leslie, Proc Roy. Soc, A307, 359 (1968).
P. C. Martin, O. Parodi and P. S. Pershan, Phys. Rev., A 6, 2401 (1972).
T. C. Lubensky, Mol Cryst. Liquid Cryst., 23, 99 (1973).
O. Lehmann, Ann. Physik, 4, 649 (1900).
N. V. Madhusudana and R. Pratibha, Mol. Cryst. Liquid Cryst. Lett., 5, 43
(1987).
F. C. Frank and M. H. L. Pryce, 1958, quoted by C. Robinson, J. C. Ward
and R. Beevers (reference 50).
C. Robinson, Trans. Faraday Soc, 52, 571 (1956).

References: Chapter 4

435

63 H. Imura and K. Okano, Phys. Lett., A42, 403 (1973).


64 P. G. de Gennes in Molecular Fluids (eds. R. Balian and G. Weill), p. 373,
Gordon & Breach, New York (1976).
65 R. S. Porter, E. M. Barrall and J. F. Johnson, / . Chem. Phys., 45, 1452
(1966).
66 S. Candau, P. Martinoty and F. Debeauvais, C R. Acad. ScL, B277, 769
(1973).
67 W. Helfrich, Phys. Rev. Lett., 23, 372 (1969).
68 U. D. Kini, G. S. Ranganath and S. Chandrasekhar, Pramana, 5, 101 (1975).
69 F. M. Leslie, Mol. Cryst. Liquid Cryst., 7, 407 (1969).
70 U. D. Kini, / . de Physique, 40, C3-62 (1979).
71 U. D. Kini, Pramana, 14, 463 (1980).
72 S. Bhattacharya, C. E. Hong and S. V. Letcher, Phys. Rev. Lett., 41, 1736
(1978).
73 J. J. Wysocki, J. Adams and W. Haas, Phys. Rev. Lett., 20, 1024 (1968).
H. Baessler and M. M. Labes, Phys. Rev. Lett., 21, 1791 (1968).
74 P. G. de Gennes, Sol. State Commun., 6, 163 (1968).
75 R. B. Meyer, Appl. Phys. Lett., 12, 281 (1968).
76 R. B. Meyer, Appl. Phys. Lett., 14, 208 (1969).
77 G. Durand, L. Leger, F. Rondelez and M. Veyssie, Phys. Rev. Lett., 22, 227
(1969).
The experiment has been done using electric fields by F. J. Kahn, Phys. Rev.
Lett. 24, 209 (1970).
78 F. M. Leslie, Mol. Cryst. Liquid Cryst., 12, 57 (1970).
79 W. Helfrich, / . Chem. Phys., 55, 839 (1971).
80 F. Rondelez and J. P. Hulin, Sol. State Commun., 10, 1009 (1972).
81 F. Rondelez and H. Arnould, C R. Acad. ScL, B273, 549 (1971).
82 C. Gerritsma and P. Van Zanten, Phys. Lett., A37, 47 (1971).
See also, T. J. Scheffer, Phys. Rev. Lett., 28, 593 (1972).
83 F. Rondelez, 'Contribution a l'Etude des Effets de Champ dans les Cristaux
Liquides Nematiques et Cholesteriques', thesis, University of Paris (1973).
84 J. P. Hurault, / . Chem. Phys., 59, 2068 (1973).
85 H. Arnould-Netillard and F. Rondelez, Mol. Cryst. Liquid Cryst., 26, 11
(1974).
86 G. H. Heilmeier and J. E. Goldmacher, Appl. Phys. Lett., 13, 132 (1968).
87 W. E. Haas, J. E. Adams, G. A. Dir and C. W. Mitchell, Proceedings of the
SID 14/4, Fourth quarter, 121 (1973).
88 J. Cheng and R. B. Meyer, Phys. Rev., A 9, 2744 (1974).
89 P. G. de Gennes, Mol. Cryst. Liquid Cryst., 12, 193 (1971).
90 F. Reinitzer, Montash Chem., 9, 421 (1888).
91 O. Lehmann, Z. Phys. Chem., 56, 750 (1906).
92 J. H. Flack, P. P. Crooker, D. L. Johnson and S. Long in Liquid Crystals
and Ordered Fluids, Vol. 4 (eds. A. C. Griffin and J. F. Johnson), p. 901,
Plenum, New York (1984).
93 D. S. Rokhsar and J. P. Sethna, Phys. Rev. Lett., 56, 1727 (1986).
R. M. Hornreich and S. Shtrikman, Phys. Rev., A 24, 635 (1981); Phys. Rev.
Lett., 59, 68 (1987).
V. M. Filev, JETP Lett., 43, 677 (1986).
94 H. Onusseit and H. Stegemeyer, Z. Naturforsch., (a) 36, 1083 (1981).
95 P. E. Cladis, P. Pieranski and M. Joanicot, Phys. Rev. Lett., 52, 542 (1984).
96 P. E. Cladis, T. Garel and P. Pieranski, Phys. Rev. Lett., 57, 2841 (1986).

436

References: Chapter 4

97 S. Meiboom, J. P. Sethna, P. W. Anderson and W. F. Brinkman, Phys. Rev.


Lett., 46, 1216 (1981).
98 D. W. Berreman, in Liquid Crystals and Ordered Fluids, Vol. 4 (eds. A. C.
Griffin and J. F. Johnson), p. 925, Plenum, New York (1984).
99 S. A. Brazovskii and S. G. Dmitriev, Sov. Phys.-JETP, 42, 497 (1976).
100 S. A. Brazovskii and V. M. Filev, Sov. Phys.-JETP, 48, 573 (1978).
101 R. M. Hornreich, M. Luban and S. Shtrikman, Phys. Rev. Lett., 35, 1678
(1975).
102 H. Grebel, R. M. Hornreich and S. Shtrikman, Phys. Rev., A 30, 3264
(1984).
103 For a comprehensive review of the Landau theory of the blue phases, see,
R. M. Hornreich and S. Shtrikman, Mol. Cryst. Liquid Cryst., 165, 183
(1988).
104 P. P. Crooker, Mol. Cryst. Liquid Cryst., 98, 31 (1983).
V. A. Belyakov and V. E. Dmitrienko, Sov. Phys. USP., 28, 535 (1985).
H. Stegemeyer, Th. Blumel, K. Hiltrop, H. Onusseit and F. Porsch, Liquid
Crystals, 1, 1 (1986).
T. Seideman, Rep. Prog. Phys., 53, 659 (1990).
105 P. N. Keating, Mol. Cryst. Liquid Cryst., 8, 315 (1969).
106 See, for example, C. Kittel, Introduction to Solid State Physics, 3rd edn,
p. 184, John Wiley, New York (1968).
107 J. L. Fergason, N. N. Goldberg and R. J. Nadalin, Mol. Cryst., 1, 309
(1966).
108 J. L. Fergason, Scientific American, 211, 77 (1964).
109 O. S. Selawry, H. S. Selawry and J. F. Holland, Mol. Cryst., 1, 495 (1966).
M. Gautherie, / . de Physique, 30, C4-122 (1969).
W. E. Woodmansee, Appl. Optics, 7, 1721 (1968).
110 J. Hansen, J. L. Fergason and A. Okaya, Appl. Opt., 3, 987 (1964).
F. Keilmann, Appl. Opt., 9, 1319 (1970).
R. D. Ennulat and J. L. Fergason, Mol. Cryst. Liquid Cryst., 13, 149
(1971).
111 K. Iizuka, Electronics Lett., 5, 26 (1969).
112 P. Pollmann and H. Stegemeyer, Chem. Phys. Lett., 20, 87 (1973).
113 S. Chandrasekhar and B. R. Ratna, Mol. Cryst. Liquid Cryst., 35, 109
(1976).
114 P. H. Keyes, H. T. Weston and W. B. Daniels, Phys. Rev. Lett., 31, 628
(1973).
115 R. Shashidhar and S. Chandrasekhar, / . de Physique, 36, Cl-49 (1975).
116 G. Friedel, Ann. Physique, 18, 273 (1922).
117 J. E. Adams and W. E. L. Haas, Mol. Cryst. Liquid Cryst., 15, 27 (1971).
118 J. E. Adams, W. E. L. Haas and J. J. Wysocki, Phys. Rev. Lett., 22, 92
(1969).
119 T. Nakagiri, H. Kodama and K. K. Kobayashi, Phys. Rev. Lett., 27, 564
(1971).
120 F. D. Saeva and J. J. Wysocki, / . Am. Chem. Soc, 93, 5928 (1971).
121 A. D. Buckingham, G. P. Ceaser and M. B. Dunn, Chem. Phys. Lett., 3,
540 (1969).
H. Stegemeyer, K. J. Mainusch and E. Steinger, Chem. Phys. Lett., 8, 425
(1971).
122 F. D. Saeva, Mol. Cryst. Liquid Cryst., 23, 171 (1973).
123 W. J. A. Goossens, Mol. Cryst. Liquid Cryst., 12, 237 (1971).

References: Chapter 5

437

124 B. W. Van der Meer, G. Vertogen, A. J. Dekker and J. G. J. Ypma,


/ . Chem. Phys., 65, 3935 (1976).
J. P. Straley, Phys. Rev., A 14, 1835 (1976).
125 A. Wulf, J. Chem. Phys., 59, 1487 (1973).
126 A. G. Priest and T. C. Lubensky, Phys. Rev., A 9, 893 (1974).
127 W. J. A. Goossens, / . de Physique, 40, C3-158 (1979).
128 D. Coates and G. W. Gray, Mol. Cryst. Liquid CrySt., 24, 163 (1973).
129 Interatomic Distances Supplement, Special publication No. 18, p. 5, The
Chemical Society, London (1965).
Chapter 5
1 H. Sackmann and D. Demus, Mol. Cryst. Liquid Cryst., 21, 239 (1973).
For a discussion of the principles underlying the calculation of phase
diagrams for liquid crystalline mixtures see M. Domon and J. Billard, in
Proceedings of the International Liquid Crystals Conference, Bangalore,
December 1973, Pramana Supplement I, p. 131.
2 P. S. Pershan, Structure of Liquid Crystal Phases, World Scientific Lecture
Notes in Physics, Vol. 23 (World Scientific, Singapore) 1988.
3 A. J. Leadbetter, in Thermotropic Liquid Crystals (ed. G. W. Gray), p. 1,
Wiley, Chichester (1987).
4 S. Diele, P. Brand and H. Sackmann, Mol. Cryst. Liquid Cryst., 17, 163
(1972).
5 A. Tardieu and J. Billard, / . de Physique Colloq., 37, C3-79 (1976).
6 G. Etherington, A. J. Leadbetter, X. J. Wang, G. W. Gray and A.
Tajbakhsh, Liquid Crystals, 1, 209 (1986).
7 W. L. McMillan, Phys. Rev., A 4, 1238 (1971).
8 J. G. Kirkwood and E. Monroe, / . Chem. Phys., 9, 514 (1941).
9 K. Kobayashi, Mol. Cryst. Liquid Cryst., 13, 137 (1971).
10 R. B. Griffiths, Phys. Rev., B 7, 545 (1973).
11 W. L. McMillan, Phys. Rev., A 6, 936 (1972).
12 R. G. Priest, Mol. Cryst. Liquid Cryst., 37, 101 (1976).
13 F. T. Lee, H. T. Tan, Y. M. Shih and C. W. Woo, Phys. Rev. Lett., 31, 1117
(1973).
14 L. Senbetu and C. W. Woo, Phys. Rev., A 17, 1529 (1978).
15 M. D. Lipkin and D. W. Oxtoby, / . Chem. Phys., 79, 1939 (1983).
16 J. Katriel and G. F. Kventsel, Phys. Rev., A 28, 3037 (1983).
17 G. F. Kventsel, G. R Luckhurst and H. B. Zwdie, Mol. Phys., 56, 589
(1985).
18 A. Kloczkowki and J. Stecki, Mol. Phys., 55, 689 (1985).
19 D. A. Badalyan, Sov. Phys. Crystallogr. 27, 10 (1982).
20 W. Wagner, Mol. Cryst. Liquid Cryst., 98, 247 (1983).
M. Nakagawa and T. Akahane, / . Phys. Soc, Japan, 54, 69 (1985).
21 C. D. Mukherjee, B. Bagchi, T. R. Bose, D. Ghose, M. K. Roy and M. Saha,
Phys. Lett., 92A, 403 (1982).
22 J. W. Doane, R. S. Parker, B. Cvikl, D. L. Johnson and D. L. Fishel, Phys.
Rev. Lett., 28, 1694 (1972).
23 B. Cabane and W. G. Clark, Sol. State Commun., 13, 129 (1973).
24 P. G. de Gennes, / . de Physique, 30, C4-65 (1969).
25 L. D. Landau and E. M. Lifshitz, Statistical Physics, Part I, 3rd edn,
Pergamon, Oxford (1980).
26 R. Schaetzing and J. D. Litster, Advances in Liquid Crystals, Vol. 4 (ed.
G. H. Brown) Academic Press, New York p. 147 (1979).

438

References:

Chapter 5

27 J. Als-Nielsen, R. J. Birgeneau, M. Kaplan, J. D. Litster and C. R. Safinya,


Phys. Rev. Lett., 39, 1668 (1977).
28 A. Caille, C. R. Acad. ScL, Paris, 274B, 891 (1972).
29 Y. Imry and L. Gunther, Phys. Rev., B 3, 3939 (1971).
L. Gunther, Y. Imry and J. Lajzerowicz, Phys. Rev., A 22, 1733 (1980).
B. Jancovici, Phys. Rev. Lett., 19, 20 (1967).
J. Als-Nielsen, in Symmetries and Broken Symmetries in Condensed Matter
Physics (ed. N. Boccara) p. 107, IDSET-Paris (1981).
J. D. Litster, in Proceedings of the Conference on Liquid Crystals of One- and
Two-dimensional Order, Garmisch-Partenkirchen, FRG (eds. W. Helfrich
and G. Heppke), p. 65, Springer-Verlag, Berlin (1980).
30 J. Als-Nielsen, J. D. Litster, R. J. Birgeneau, M. Kaplan, C. R. Safinya, A.
Lindegaard-Andersen and S. Mathiesen, Phys. Rev., B 22, 312 (1980).
31 M. Delaye, R. Ribotta and G. Durand, Phys. Lett., 44A, 139 (1973).
32 N. A. Clark and R. B. Meyer, App. Phys. Lett., 22, 493 (1973).
33 G. Durand, in Proceedings of the International Liquid Crystals Conference,
Bangalore, December 1973, Pramana Supplement I, p. 23.
34 J. D. Litster, R. J. Birgeneau, M. Kaplan, C. R. Safinya and J. Als-Nielsen,
in Ordering in Strongly Fluctuating Condensed Matter Physics, NATO ASI
Vol. 50 (ed. T. Riste) p. 357, Plenum Press, New York (1980).
35 G. Durand, C. R. Acad. Sci., 275B, 629 (1972).
36 R. Ribotta, G. Durand and J. D. Litster, Sol. State Commun., 12, 27 (1973).
37 N. A. Clark and P. S. Pershan, Phys. Rev. Lett., 30, 3 (1973).
38 R. Ribotta, D. Salin and G. Durand, Phys. Rev. Lett., 32, 6 (1974).
39 P. C. Martin, O. Parodi and P. S. Pershan, Phys. Rev., A 6, 2401 (1972).
40 W. Helfrich, Phys. Rev. Lett., 23, 372 (1969).
41 K. Sakamoto, R. S. Porter and J. F. Johnson, Mol. Cryst. Liquid CrySt., 8,
443 (1969).
42 U. D. Kini, G. S. Ranganath and S. Chandrasekhar, Pramana, 5, 101 (1975).
43 A. E. Lord, Phys. Rev. Lett., 29, 1366 (1972).
44 K. Miyano and J. B. Ketterson, Phys. Rev. Lett., 31, 1047 (1973).
45 See, for example, J. F. Nye, Physical Properties of Crystals, Clarendon,
Oxford (1957).
46 See, for example, F. Brochard and P. G. de Gennes, in Proceedings of the
International Liquid Crystal Conference, Bangalore, December 1973,
Pramana Supplement I, p. 1.
47 Y. Liao, N. A. Clark and P. S. Pershan, Phys. Rev. Lett., 30, 639 (1973).
48 G. Grinstein and R. Pelcovits, Phys. Rev. Lett., 47, 856 (1981); Phys. Rev.,
A 26, 915(1982).
49 G. F. Mazenko, S. Ramaswamy and J. Toner, Phys. Rev. Lett., 49, 51
(1982); Phys. Rev., A 28, 1618 (1983).
50 For an excellent summary of the theory and its implications see J. Prost,
Advances in Physics, 33, 1 (1984).
51 R. Kubo, Rep. Prog. Phys., 29, 255 (1966).
52 S. Ramaswamy and J. Toner, Phys. Rev., A 28, 3159 (1983).
53 S. T. Milner and P. C. Martin, Phys. Rev. Lett., 56, 77 (1986).
54 S. Bhattacharya and J. B. Ketterson, Phys. Rev. Lett., 49, 997 (1982).
J. L. Gallani and P. Martinoty, Phys. Rev. Lett., 54, 333 (1985); ibid., 54,
1736E (1985); ibid., 55, 638E (1985).
C. Baumann, J. P. Marcerou, J. Prost and J. C. Rouillon, Phys. Rev. Lett.,
54, 1268 (1985).
55 G. Friedel, Ann. Physique, 18, 273 (1922).

References:

Chapter 5

439

56 W. H. Bragg, Nature, 133, 445 (1934).


See also N. H. Hartshorne and A. Stuart, Crystals and the Polarizing
Microscope, 2nd edn, Edward Arnold, London (1950).
57 M. Kleman, / . de Physique, 38, 1511 (1977).
58 CH. S. Rosenblatt, R. Pindak, N. A. Clark and R. B. Meyer, / . de Physique,
38, 1105(1977).
59 S. Chandrasekhar and N. V. Madhusudana, Acta Cryst. A26, 153 (1970).
60 F. Grandjean, C. R. Acad. ScL, 166, 165 (1917).
61 S. Chandrasekhar, Mol. Cryst., 2, 71 (1966).
62 P. G. de Gennes, C R. Hebd. Seanc. Acad. ScL, Paris, 275B, 939 (1972).
63 P. S. Pershan, J. Appl. Phys., 45, 1590 (1974).
64 M. Kleman and C. E. Williams, / . de Physique Lettres, 35, L-49 (1974).
65 J. Friedel, in Dislocations, Pergamon, Oxford (1967).
66 C. E. Williams and M. Kleman, / . de Physique Lettres, 35, L-33 (1974).
67 J. Friedel and M. Kleman, in Fundamental Aspects of Dislocation Theory
(eds. J. A. Simmons, R. de Wit and R. Bullough) Special Publication No.
317, Washington DC: National Bureau of Standards, Vol. 1, p. 607.
68 P. G. de Gennes, Sol. State Commun., 10, 753 (1972).
69 P. G. de Gennes, Mol. Cryst. Liquid Cryst., 21, 49 (1973).
70 S. Chandrasekhar and R. Shashidhar, in Advances in Liquid Crystals, Vol. 4
(ed. G. H. Brown), p. 83, Academic Press, New York (1979).
71 H. Gruler, Z. Naturforsch., 28a, 474 (1973).
72 See, for example, E. A. Lynton, Superconductivity, 2nd edn, Methuen,
London (1964).
73 F. Brochard, / . de Physique, 34, 411 (1973).
F. Jahnig and F. Brochard, / . de Physique, 35, 301 (1974).
See also F. Jahnig, in Proceedings of the International Liquid Crystals
Conference, Bangalore, December 1973, Pramana Supplement I, p. 31.
74 W. L. McMillan, Phys. Rev., A 7, 1419 (1973).
75 L. Cheung, R. B. Meyer and H. Gruler, Phys. Rev. Lett., 31, 349 (1973).
76 W. L. McMillan, Phys. Rev., A 9, 1720 (1974).
77 R. Mahmood, I. Khan, C. Gooden, A. Baldwin, D. L. Johnson and M. E.
Neubert, Phys. Rev., A 32, 1286 (1985).
78 B. I. Halperin, T. C. Lubensky and S. K. Ma, Phys. Rev. Lett., 32, 292
(1974).
79 T. C. Lubensky and J. H. Chen, Phys. Rev., B 17, 366 (1978).
80 C. Dasgupta and B. I. Halperin, Phys. Rev. Lett., 47, 1556 (1981).
81 S. G. Dunn and T. C. Lubensky, / . de Physique, 42, 1201 (1981).
82 J. Thoen, H. Marynissen and W. Van Dael, Phys. Rev., A26, 2886
(1982); Phys. Rev. Lett., 52, 204 (1984).
83 M. A. Anisimov, P. E. Cladis, E. E. Gorodetskii, D. A. Huse, V. E. Podneks,
V. G. Taratuta, W. Van Saarloos, V. P. Voronov, Phys. Rev., 41, 6749
(1990).
84 T. C. Lubensky, / . Chim. Phys., 80, 31 (1983).
85 J. D. Litster, in Light Scattering near Phase Transitions (eds. H. Z.
Cummins and A. P. Levanyuk) Chapter 10, North-Holland, Amsterdam
(1983).
86 M. A. Anisimov, Mol Cryst. Liquid Cryst., 162A, 1 (1988).
87 H. E. Stanley, Introduction to Phase Transitions and Critical Phenomena,
Clarendon, Oxford (1971).
S. K. Ma, Modern Theory of Critical Phenomena, Benjamin Cummings,
Reading, Mass. (1976).

440

References: Chapter 5

P. Pfeuty and G. Toulouse, Introduction to the Renormalization Group and to


Critical Phenomenon, John Wiley & Sons, London (1977).
88 J. D. Litster and R. J. Birgeneau, Physics Today, 35, 26 (1982).
89 C. W. Garland, M. Meichle, B. M. Ocko, A. R. Kortan, C. R. Safinya, L. J.
Yu, J. D. Litster and R. J. Birgeneau, Phys. Rev., A 27, 3234 (1983).
90 B. M. Ocko, R. J. Birgeneau, J. D. Litster and M. E. Neubert, Phys. Rev.
Lett., 52, 208 (1981).
91 K. K. Chan, 'X-ray Scattering Study of Nematic and Smectic A Critical
Behaviour in Liquid Crystals', thesis, Harvard University, 1984.
92 K. W. Evans-Lutterodt, J. W. Chung, B. M. Ocko, R. J. Birgeneau, C.
Chiang, C. W. Garland, J. W. Goodby and Nguyen Huu Tinh, Phys. Rev.,
A 36, 1387(1987).
93 C. W. Garland, G. Nounesis and K. J. Stine, Phys. Rev., A 39, 4919 (1989).
94 C. R. Safinya, R. J. Birgeneau, J. D. Litster and M. E. Neubert, Phys. Rev.
Lett., 47, 668 (1981).
95 D. Brisbin, R. De Hoff, T. E. Lokhart and D. L. Johnson, Phys. Rev. Lett.,
43, 1171 (1979).
96 J. D. Litster, J. Als-Nielsen, R. J. Birgeneau, S. S. Dana, D. Davidov, F.
Garcia-Golding, M. Kaplan, C. R. Safinya and R. Schaetzing, / . de
Physique, 40, C3-309 (1979).
97 D. Djurek, J. Baturic-Rubcic and K. Franulovic, Phys. Rev. Lett., 33, 1126
(1974).
98 R. J. Birgeneau, C. W. Garland, G. B. Kasting and B. M. Ocko, Phys. Rev.,
B 24, 2624(1981).
99 G. B. Kasting, K. J. Lushington and C. W. Garland, Phys. Rev., B 22, 321
(1980).
100 B. M. Ocko, R. J. Birgeneau, J. D. Litster and M. E. Neubert, Phys. Rev.
Lett., 52, 208 (1981).
101 D. Davidov, C. R. Safinya, M. Kaplan, S. S. Dana, R. Schaetzing, R. J.
Birgeneau and J. D. Litster, Phys. Rev., B 19, 1657 (1979).
102 J. C. Le Guillou and J. Zinn-Justin, Phys. Rev. Lett., 39, 95 (1977); Phys.
Rev., B21, 3976(1980).
103 C. Bagnuls and C. Bervillier, Phys. Rev., B 32, 7209 (1985).
104 D. R. Nelson and J. Toner, Phys. Rev., B 24, 363 (1981).
105 A. R. Day, T. C. Lubensky and A. J. McKane, Phys. Rev., A 27, 1461
(1983).
106 J. Toner, Phys. Rev., B 26, 462 (1982).
107 H. Birecki, R. Schaetzing, F. Rondelez and J. D. Litster, Phys. Rev. Lett.,
36, 1376 (1976).
M. Fisch, L. Sorensen and P. Pershan, Phys. Rev. Lett., 48, 943 (1982).
M. Fisch, P. Pershan and L. Sorensen, Phys. Rev., A 29, 2741 (1984).
108 H. J. Fromm, J. de Physique, 48, 647 (1987).
M. E. Lewis, T. Khan, H. Vithana, A. Baldwin, D. L. Johnson and M. E.
Neubert, Phys. Rev., A 38, 3702 (1988).
109 H. K. M. Vithana, V. Surendranath, M. Lewis, A. Baldwin, K. Eidner, R.
Mahmood and D. L. Johnson, Phys. Rev., A 41, 2031 (1990).
110 T. C. Lubensky, / . de Physique, 36, Cl-151 (1975).
111 P. Kassubek and G. Meier, Mol. Cryst. Liquid Cryst., 8, 305 (1969).
112 R. S. Pindak, C. C. Huang and J. T. Ho, Solid State Commun., 14, 821
(1974).
113 G. Sigaud, F. Hardouin, M. F. Achard and H. Gasparoux, / . de Physique,
40, C3-356 (1979).

References: Chapter 5

441

114 J. Prost, / . de Physique, 40, 581 (1979).


115 J. Prost and P. Barois, / . Chim. Phys., 80, 65 (1983).
116 F. C. Frank and J. H. Van der Merwe, Proc. Roy. Soc, London, A198, 216
(1949); A200, 125(1949).
117 B. R. Ratna, R. Shashidhar and V. N. Raja, Phys. Rev. Lett., 55, 1476
(1985); 56, 269 (E) (1986).
P. Das, K. Ema, C. W. Garland and R. Shashidhar, Liquid Crystals, 4, 581
(1989).
118 B. R. Ratna, R. Shashidhar, V. N. Raja, G. Heppke and N. H. Tinh,
presented at the 12th International Liquid Crystal Conference, Freiburg
(1988).
R. Shashidhar and B. R. Ratna, Liquid Crystals, 5, 421 (1989).
119 S. Kumar, Li Chen and V. Surendranath, Phys. Rev. Lett., 67, 322 (1991).
120 P. Barois, J. Prost and T. C. Lubensky, / . de Physique, 46, 391 (1985).
121 R. Shashidhar, B. R. Ratna, S. Krishna Prasad, S. Somasekhara and G.
Heppke, Phys. Rev. Lett., 59, 1209 (1987).
Y. H. Jeong, G. Nounesis, C. W. Garland and R. Shashidhar, Phys. Rev.,
A 40, 4022 (1989).
J. Prost, J. Pommier, J. C. Rouillon, J. P. Marcerou, P. Barois, M.
Benzekri, A. Babeau and H. T. Nguyen, Phys. Rev., B 42, 2521 (1990).
122 Y. Park, T. C. Lubensky, P. Barois and J. Prost, Phys. Rev., A 37, 2197
(1988).
123 K. K. Chan, P. S. Pershan, L. B. Sorensen and F. Hardouin, Phys. Rev.
Lett., 54, 1694 (1985); Phys. Rev., A 34, 1420 (1986).
124 J. Wang, 'The Frustrated Smectic Phases in Liquid Crystals', thesis,
University of Pennsylvania, 1985.
125 F. Hardouin, A. M. Levelut, M. F. Achard and G. Sigaud, / . Chim. Phys.,
80, 53 (1983).
126 Y. Shapira in Multicritical Phenomena (eds. R. Pynn and A. Skjeltrop), p.
35, Plenum Press, New York (1983).
127 V. N. Raja, R. Shashidhar, B. R. Ratna, G. Heppke and Ch. Bahr, Phys.
Rev., A 37, 303 (1988).
K. Ema, G. Nounesis, C. W. Garland and R. Shashidhar, Phys. Rev., A 39,
2599 (1989).
128 H. Rohrer, A. Aharony and S. Fishman, / . Magn. Magn. Mater., 15, 396
(1980).
129 P. E. Cladis, Phys. Rev. Lett., 35, 48 (1975).
130 P. E. Cladis, R. K. Bogardus, W. B. Daniels and G. N. Taylor, Phys. Rev.
Lett., 39, 720 (1977).
131 N. H. Tinh, / . Chim. Phys., 80, 83 (1983).
132 N. H. Tinh, F. Hardouin, C. Destrade and A. M. Levlutt, J. de Physique
Lettres, 43, L33 (1982).
133 R. Shashidhar, B. R. Ratna, V. Surendranath, V. N. Raja, S. Krishna
Prasad and C. Nagabhushana, / . de Physique Lettres, 46, L445 (1985).
V. N. Raja, B. R. Ratna, R. Shashidhar, G. Heppke, Ch. Bahr, J. F.
Marko, J. O. Indekeu and A. N. Berker, Phys. Rev. A 39, 4341 (1989).
V. N. Raja, presented at the 13th AIRAPT International Conference on
High Pressure Science and Technology, Bangalore, October 1991.
134 S. Chandrasekhar and N. V. Madhusudhana, Molecular Interactions and
Dynamics (eds. W. J. Orville-Thomas, H. Ratajczak and C. N. R. Rao),
Elsevier, Amsterdam, Indian Academy of Sciences, Bangalore, Vol. 4, p.
139(1985).

442

References: Chapter 5

135 P. E. Cladis, in Proceedings of the International Conference on Liquid


Crystals, Bangalore 1979 (ed. S. Chandrasekhar), p. 105, Heyden, London
(1980).
136 L. Longa and W. H. de Jeu, Phys. Rev., A 26, 1632 (1982).
137 A. N. Berker and S. J. Walker, Phys. Rev. Lett., 47, 1469 (1981).
J. O. Indekeu and A. N. Berker, / . de Physique, 49, 353 (1988).
138 G. R. Luckhurst and B. A. Timimi, Mol. Cryst. Liquid Cryst. Lett., 64, 253
(1981).
T. R. Bose, C. D. Mukherjee, M. K. Roy and M. Saha, Mol. Cryst. Liquid
Cryst., 126, 197 (1985).
139 N. V. Madhusudana and Jyotsna Rajan, Liquid Crystals, 1, 31 (1990).
140 For a critical discussion of some of the theoretical models, see S.
Chandrasekhar, Proceedings of the 10th International Liquid Crystal
Conference, York 1984; Mol. Cryst. Liquid Cryst., 124, 1 (1985).
141 F. Dowell, Phys. Rev., A 28, 3526 (1983); ibid., A 31, 3214 (1985); ibid.,
A 36, 5046(1987).
142 G. Pelzl, S. Diele, I. Latif, W. Weissflog and D. Demus, Crystal Res. Tech.,
17, K78 (1982).
S. Diele, G. Pelzl, I. Latif and D. Demus, Mol. Cryst. Liquid Cryst. Lett.,
92, 27 (1983).
143 J. W. Park, C. S. Bak and M. M. Labes, J. Am. Chem. Soc, 97, 4398
(1975).
144 C. S. Oh, Mol. Cryst. Liquid Cryst., 42, 1 (1977).
145 A. C. Griffin, T. R. Britt, N. W. Buckley, R. F. Fisher, S. J. Havens and
D. W. Goodman, Liquid Crystals and Ordered Fluids, Vol. 3 (eds. J. F.
Johnson and R. S. Porter), p. 61, Plenum Press, New York, (1978).
146 B. Engelen and F. Schneider, Z. Naturforsch., 33A, 1077 (1978).
B. Engelen, G. Heppke, R. Hopf and F. Schneider, Ann. Phys., 3, 403
(1978).
147 L. J. Yu and M. M. Labes, Mol. Cryst. Liquid Cryst., 54, 1 (1979).
148 M. Domon and J. Billard, / . de Physique, 40, C3-413 (1979).
149 K. P. L. Moodithaya and N. V. Madhusudana, in Proceedings of the
International Conference on Liquid Crystals, Bangalore, December 1979,
(ed. S. Chandrasekhar), p. 297, Heyden, London (1980).
150 F. Schneider and N. K. Sharma, Z. Naturforsch, 36A, 62 (1981); ibid., 36A,
1086(1981).
151 K. Araya and Y. Matsunaga, Mol. Cryst. Liquid Cryst., 67, 153 (1981).
152 G. Derfel, Mol. Cryst. Liquid Cryst., 82, 277 (1982).
153 J. Szabon and S. Diele, Crystal Res. Tech., 17, 1315 (1982).
154 K. W. Sadowska, A. Zywocinski, J. Stecki and R. Dabrowski, / . de
Physique, 43, 1673 (1982).
W. Waclawek, R. Dabrowski and A. Domagala, Mol. Cryst. Liquid Cryst.,
84, 255 (1982).
155 A. Boij and P. Adomenas, Mol. Cryst. Liquid Cryst., 95, 59 (1983).
156 E. Chino and Y. Matsunaga, Bull. Chem. Soc, Japan, 56, 3230 (1983).
157 N. K. Sharma, G. Pelzl, D. Demus and W. Weissflog, Z. Phys. Chem.
{Leipzig), 261, 579 (1980).
158 J. W. Goodby, T. M. Leslie, P. E. Cladis and P. L. Finn, Proceedings of the
ACS Symposium on Liquid Crystals and Ordered Fluids, Las Vegas 1982
(eds. A. C. Griffin and J. F. Johnson), p. 89, Plenum, New York (1984).
159 N. V. Madhusudana, V. A. Raghunathan and M. Subramanya Raj Urs,
Mol. Cryst. Liquid Cryst., 106, 161 (1984).

References: Chapter 5

443

160 K. A. Suresh, Mol. Cryst. Liquid Cryst., 97, 417 (1983).


161 P. E. Cladis and S. Torza, / . Appl. Phys., 46, 584 (1975).
H. P. Hinov, / . de Physique, 42, 307 (1981).
N. V. Madhusudana and B. S. Srikanta, Mol. Cryst. Liquid Cryst., 99, 375
(1983).
162 A. A. Abrikosov, Zh. Eksp. Teor. Fiz., 32, 1442 (1957); Sov.
Phys.-JETP,
5, 1174(1957).
163 S. R. Renn and T. C. Lubensky, Phys Rev., A 38, 2132 (1988).
164 J. W. Goodby, M. A. Waugh, S. M. Stein, E. Chin, R. Pindak and J. S.
Patel, Nature {London), 337, 449 (1989); / . Am. Chem. Soc, 111, 8119
(1989).
165 G. Srajer, R. Pindak, M. A. Waugh, J. W. Goodby and J. S. Patel, Phys.
Rev. Lett., 64, 1545 (1990).
166 J. M. Kosterlitz and D. G. Thouless, / . Phys., 6, 1181 (1973).
167 B. I. Halperin and D. R. Nelson, Phys. Rev. Lett., 41, 121 (1978).
168 R. J. Birgeneau and J. D. Litster, / . de Physique Lettres, 39, L-399 (1978).
169 A. J. Leadbetter, J. C. Frost and M. A. Mazid, / . de Physique Lettres, 40,
L-325 (1979).
170 R. Pindak, D. E. Moncton, S. C. Davey and J. W. Goodby, Phys. Rev.
Lett., 46, 1135(1981).
171 R. Bruinsma and D. R. Nelson, Phys. Rev., B 23, 402 (1981).
172 J. D. Brock, A. Aharony, R. J. Birgeneau, K. W. Evans-Lutterodt, J. D.
Litster, P. M. Horn, G. B. Stephenson and A. R. Tajbakhsh, Phys. Rev.
Lett., 57, 98 (1986).
J. D. Brock, D. Y. Noh, B. R. McClain, J. D. Litster, R. J. Birgeneau,
A. Aharony, P. M. Horn and J. C. Liang, Z. Phys., B 74, 197 (1989).
173 C. C. Huang, in Theory and Applications of Liquid Crystals (eds. J. L.
Ericksen and D. Kinderlehrer), p. 185, Springer-Verlag, Berlin (1987).
174 L. Richter, D. Demus and H. Sackmann, Mol. Cryst. Liquid Cryst., 71, 269
(1981).
D. Demus, S. Diele, M. Klapperstuck, V. Link and H. Zaschke, Mol.
Cryst. Liquid Cryst., 15, 161 (1971).
P. A. C. Gane, A. J. Leadbetter and P. G. Wrighton, Mol. Cryst. Liquid
Cryst., 66, 247(1981).
175 S. A. Brazovskii, Zh. Eksp. Teor, Fiz., 68, 175 (1975); Sov. Phys. JETP,
41, 85 (1975).
176 J. Swift, Phys. Rev., A 14, 2274 (1976).
177 J. J. Benattar, F. Moussa and M. Lambert, J. Chim. Phys., 80, 99 (1983).
178 D. R. Nelson and B. I. Halperin, Phys. Rev., B 21, 5312 (1980).
179 V. N. Raja, S. Krishna Prasad, D. S. Shankar Rao, J. W. Goodby and
M. E. Neubert, Ferroelectrics, 121, 235 (1991).
180 M. Hosino, H. Nakano and H. Kimura, / . Phys. Soc, Japan, 46, 1709
(1979).
181 A. M. Somoza and P. Tarazona, Phys. Rev. Lett., 61, 2566 (1988).
182 A. Wulf, Phys. Rev., A 11, 365 (1975).
183 W. L. McMillan, Phys. Rev., A 8, 1921 (1973).
184 D. Cabib and L. Benguigui, J. Phys., 38, 419 (1977).
185 B. W. Van der Meer and G. Vortogen, J. Phys., 40, C3-222 (1979).
186 R. G. Priest, / . Phys., 36, 437 (1975); / . Chem. Phys. 65, 408 (1976).
187 M. Matsushita, / . Phys. Soc, Japan, 50, 135 (1981).
188 W. J. A. Goossens, / . Phys., 46, 1411 (1985).

444

References: Chapter 5

189 W. J. A. Goossens, Europhys. Lett., 3, 341 (1987); Mol. Cryst. Liquid


Cryst., 150, 419 (1987).
190 A. Saupe, Mol. Cryst. Liquid Cryst., 7, 59 (1969).
191 J. Nehring and A. Saupe, / . Chem. Soc. Faraday Trans. 2, 68, 1 (1972).
192 Orsay Liquid Crystals Group, Sol. State Commun., 9, 653 (1971).
193 Y. Galerne, J. L. Martinand, G. Durand and M. Veyssie, Phys. Rev. Lett.,
29, 562 (1972).
194 H. Brand and H. Plenier, J. de Physique, 41, 553 (1980); 45, 563, (1984).
P. Schiller, Wiss. Z. Univ. Halle, 34, 61 (1985).
195 F. M. Leslie, I. W. Stewart and N. Nakagawa, presented at the 13th
International Liquid Crystal Conference, Vancouver, Canada, 1990, Mol.
Cryst. Liquid Cryst., 198, 443 (1991).
196 Y. Bouligand and M. Kleman, / . de Physique, 40, 79 (1979).
197 C. Allet, M. Kleman and P. Vidal, J. de Physique, 39, 181 (1978).
198 L. Bourdon, J. Sommeria and M. Kleman, J. de Physique, 43, 77 (1982).
199 R. B. Meyer, B. Stebler and S. T. Lagerwall, Phys. Rev. Lett., 41, 1393
(1978).
S. T. Lagerwall and B. Stebler, Proceedings of the International Conference
on Liquid Crystals, Bangalore (ed. S. Chandrasekhar), p. 223, Heyden,
London (1979); Phys. Bull, 31, 349 (1980).
200 M. Delaye, / . de Physique, 40, C3-350 (1979).
Y. Galerne, Phys. Rev., A 24, 2284 (1981).
201 C. R. Safinya, M. Kaplan, J. Als-Nielsen, R. J. Birgeneau, D. Davidov,
J. D. Litster, D. L. Johnson and M. E. Neubert, Phys. Rev., B 21, 4149
(1980).
202 C. C. Huang and J. M. Viner, Phys. Rev., A 25, 3385 (1982).
203 R. J. Birgeneau, C. W. Garland, A. R. Kortan, J. D. Litster, M. Meichle,
B. M. Ocko, C. Rosenblatt, L. J. Yu and J. W. Goodby, Phys. Rev., A 27,
1251 (1983).
For a concise review, see J. D. Litster, Phil. Trans. Roy. Soc, London,
A 309, 145 (1983).
204 R. Ribotta, R. B. Meyer and G. Durand, / . de Physique Lettres, 35, 161
(1974).
205 G. S. Ranganath, Mol. Cryst. Liquid Cryst., 94, 285 (1983).
206 S. T. Lagerwall and B. Stebler, in Ordering in Strongly Fluctuating
Condensed Matter Systems (ed. T. Riste), p. 383, Plenum, New York
(1980).
207 D. L. Johnson, J. Chim. Physique, 80, 45 (1983).
208 D. L. Johnson, D. Allender, R. De Hoff, C. Maze, E. Oppenheim and R.
Reynolds, Phys. Rev., B 16, 470 (1977).
G. Sigaud, F. Hardouin and M. F. Achard, Sol. State Commun., 23, 35
(1977).
209 R. Shashidhar, B. R. Ratna and S. Krishna Prasad, Phys. Rev. Lett., 53,
2141 (1984).
210 D. Brisbin, D. L. Johnson, H. Fellner and M. E. Neubert, Phys. Rev. Lett.
50, 178 (1983).
211 R. Shashidhar, A. N. Kalkura and S. Chandrasekhar, Mol. Cryst. Liquid
Cryst. Lett., 82, 311 (1982).
212 G. Sigaud, Y. Guichard, F. Hardouin and L. G. Benguigui, Phys. Rev.,
A 26, 3041 (1982).
213 R. Shashidhar, S. Krishna Prasad and S. Chandrasekhar, Mol. Cryst.
Liquid Cryst., 103, 137 (1983).

References: Chapter 5

445

214 S. Somasekhara, R. Shashidhar and B. R. Ratna, Phys. Rev., A 34, 2561


(1986).
215 R. M. Hornreich, M. Luban and S. Shtrikman, Phys. Rev. Lett., 35, 1678
(1975).
216 J. H. Chen and T. C. Lubensky, Phys. Rev., 14, 1202 (1976).
217 C. R. Safinya, R. J. Birgeneau, J. D. Litster and M. E. Neubert, Phys. Rev.
Lett., 47, 668 (1981).
218 C. R. Safinya, L. J. Martinez-Miranda, M. Kaplan, J. D. Litster and R. J.
Birgeneau, Phys. Rev. Lett., 50, 56 (1983).
219 L. J. Martinez-Miranda, A. R. Kortan and R. J. Birgeneau, Phys. Rev.
Lett., 56, 2264 (1986).
220 R. De Hoff, R. Biggers, D. Brisbin and D. L. Johnson, Phys. Rev., A 25,
472 (1982).
221 M. A. Anisimov, V. P. Voronov, A. O. Kulkov and F. Kholmurdov, J. de
Physique, 46, 2137 (1983).
222 C. W. Garland and M. E. Huster, Phys. Rev., A 35, 2365 (1987).
223 S. Witanachi, J. Huang and J. T. Ho, Phys. Rev. Lett., 50, 594 (1983).
224 L. Solomon and J. D. Litster, Phys. Rev. Lett., 56, 2268 (1987).
225 J. Huang and J. T. Ho, Phys. Rev. Lett., 58, 2239 (1987); Phys. Rev., A 42,
2449 (1990).
226 G. Grinstein and J. Toner, Phys. Rev. Lett., 51, 2386 (1983).
227 T. C. Lubensky, Mol. Cryst. Liquid Cryst., 146, 55 (1987).
228 X. Wen, C. W. Garland and M. D. Wand, Phys. Rev., A 42, 6087 (1990).
229 W. Helfrich and C. S. Oh, Mol. Cryst. Liquid Cryst., 14, 289 (1971).
230 D. W. Berreman, Mol. Cryst. Liquid Cryst., 22, 175 (1973).
231 R. Nityananda, Pramana, 2, 35 (1974).
232 R. B. Meyer, L. Liebert, L. Strzelecki and P. Keller, / . de Physique Lettres,
36, L-69 (1975).
233 S. Garoffand R. B. Meyer, Phys. Rev. Lett., 38, 848 (1977).
234 Ch. Bahr and G. Heppke, Mol. Cryst. Liquid Cryst., 4, 31 (1986).
235 R. Shashidhar, B. R. Ratna, G. G. Nair, S. Krishna Prasad, Ch. Bahr and
G. Heppke, Phys. Rev. Lett., 61, 547 (1988).
236 S. Dumrongrattana and C. C. Huang, Phys. Rev. Lett., 56, 464 (1986).
C. C. Huang and S. Dumrongrattana, Phys. Rev., A 34, 5020 (1986).
237 B. I. Ostrovskii, A. Z. Rabinovich, A. S. Sonin and B. A. Strukov, Sov.
Phys.-JETP, 47, 912 (1978).
238 Ch. Bahr and G. Heppke, Liquid Crystals, 2, 825 (1987).
239 S. M. Khened, S. Krishna Prasad, B. Shivkumar and B. K. Sadashiva, / . de
Physique II, 1, 171 (1991).
240 E. P. Pozhidayev, M. A. Osipov, V. G. Chigrinov, V. A. Baikalov, L. M.
Blinov and L. A. Beresnev, Sov. Phys.-JETP, 67, 283 (1988).
241 P. Pieranskii, E. Guyon and P. Keller, / . de Physique, 36, 1005 (1975).
242 A. Jakli, L. Bata, A. Buka, N. Eber and I. Janossy, / . de Physique Lettres,
46, L-759 (1985).
243 N. Eber, L. Bata and A. Jakli, Mol. Cryst. Liquid Cryst., 142, 15 (1987).
A. Jakli and L. Bata, Liquid Crystals, 1, 105 (1990).
244 L. J. Yu, H. Lee, C. S. Bak and M. M. Labes, Phys. Rev. Lett., 36, 388
(1976).
245 L. M. Blinov, L. A. Beresnev, N. M. Shtykov and Z. M. Elashvili, / . de
Physique, 40, C3-269 (1979).
246 A. D. L. Chandani, Y. Ouchi, H. Takezoe, A. Fukuda, K. Terashima, K.

446

247
248
249
250
251
252
253
254
255
256
257
258

1
2
3
4
5
6
7
8

9
10
11
12
13
14
15

References: Chapter 6
Furukawa, A. Kishi, Jap. J. AppL Phys., 28, LI261 (1989), and references
therein.
R. B. Meyer, Mol. Cryst. Liquid Cryst., 40, 33 (1977).
S. A. Pikin and V. L. Indenbom, Sov. Phys. USP, 21, 487 (1978).
B. Zeks, Mol. Cryst. Liquid Cryst., 114, 259 (1984).
T. Carlsson, B. Zeks, A. Levstik, C. Filipic, I. Levstik and R. Blinc, Phys.
Rev., A 36, 1484 (1987).
R. Blinc, B. Zeks, M. Copic, A. Levstik, I. Musevic, I. Drevensek,
Ferroelectrics, 104, 159 (1990).
N. A. Clark and S. T. Lagerwall, AppL Phys. Lett., 36, 899 (1980).
T. P. Rieker, N. A. Clark, G. S. Smith, D. S. Parmar, E. B. Sirota and
C. R. Safinya, Phys. Rev. Lett., 59, 2658 (1987).
K. Skarp and M. Handschy, Mol. Cryst. Liquid Cryst., 165, 439 (1988).
D. Armitage, J. I. Thackara and W. D. Eades, Ferroelectrics, 85, 291
(1988).
S. Matsumoto, H. Hatoh and A. Murayama, Liquid Crystals, 5, 1345
(1989).
J. Dijon in Liquid Crystal Applications and Uses (ed. B. Bahadur), World
Scientific, Singapore (1990).
G. Andersson, I. Dahl, W. Kuczynski, S. T. Lagerwall, K. Skarp and B.
Stebler, Ferroelectrics, 84, 285 (1988).
E. E. Gorodetskii and V. E. Podneks, presented at the 13th International
Liquid Crystal Conference, Vancouver, 22-27 July, 1990.

Chapter 6
O. Lehmann, Z. Physikal. Chem., 4, 462 (1889).
L. Gattermann and A. Ritschke, Ber. Deut. Chem. Ges., 23, 1738 (1890).
D. Vorlander, Ber. Deut. Chem. Ges., 39, 803 (1906).
S. Chandrasekhar, B. K. Sadashiva and K. A. Suresh, Pramana, 9, 471
(1977).
For reviews on discotics, see refs. 4-9 below.
S. Chandrasekhar, Advances in Liquid Crystals, Vol. 5 (ed. G. H. Brown), p.
47, Academic Press, New York (1982).
S. Chandrasekhar, Phil. Trans. Roy. Soc, London, A309, 93 (1983).
A. M. Levelut, / . Chim. Phys., 88, 149 (1983).
C. Destrade, P. Foucher, H. Gasparoux, N. H. Tinh, A. M. Levelut and J.
Malthete, Mol. Cryst. Liquid Cryst., 106, 121 (1984).
A. J. Leadbetter, in Thermotropic Liquid Crystals, Critical Reports on
Applied Chem., Vol. 22 (ed. G. W. Gray), p. 1, John Wiley & Sons,
Chichester (1987); P. S. Pershan, Structure of Liquid Crystal Phases, World
Scientific, Singapore (1988).
S. Chandrasekhar and G. S. Ranganath, Rep. Prog, in Physics, 53, 57 (1990)
A more or less complete bibliography on the subject up to 1984 has been
compiled by H. P. Hinov, Mol. Cryst. Liquid Cryst., 136, 221 (1986).
B. Kohne and K. Praefcke, Chimia, 41, 196 (1987).
B. Kohne and K. Praefcke, Angew. Chem., 96, 70 (1984).
J. C. Dubois, Annals Phys., 3, 131 (1978).
C. Destrade, M. C. Mondon, J. Malthete, / . de Physique, 40, C3-17 (1979).
C. Destrade, M. C. Bernaud, H. Gasparoux, A. M. Levelut and N. H. Tinh,
Proceedings of the International Conference on Liquid Crystals, Bangalore,
December 1979 (ed. S. Chandrasekhar), p. 29, Heyden, London (1980).

References: Chapter 6

447

16 N. H. Tinh, H. Gasparoux and C. Destrade, Mol. Cryst. Liquid Cryst., 68,


101 (1981).
17 C. Destrade, J. Malthete, N. H. Tinh and H. Gasparoux, Phys. Lett., 78A,
82 (1980).
18 N. H. Tinh, J. Malthete and C. Destrade, / . de Physique Lettres, 42, L-417
(1981).
19 A. M. Giroud-Godquin, M. M. Gauthier, G. Sigoud, F. Hardouin and
M. F. Achard, Mol. Cryst. Liquid Cryst., 132, 35 (1986).
For a comprehensive review of the chemistry of metallomesogens, see, A. M.
Giroud-Godquin and P. M. Maitlis, Angew. Chem. Int., Ed. Engl., 30, 375
(1991).
20 C. Piechocki, J. Simon, A. Skoulios, D. Guillon and P. Weber, / . Am. Chem.
Soc, 104, 5245 (1982).
21 F. C. Frank and S. Chandrasekhar, / . de Physique, 41, 1285 (1980).
22 D. H. Van Winkle and N. A. Clark, Phys. Rev. Lett., 48, 1407 (1982).
23 C. R. Safinya, K. S. Liang, W. A. Varady, N. A. Clark, and G. Andersson
Phys. Rev. Lett., 53, 1172 (1984).
24 C. R. Safinya, N. A. Clark, K. S. Liang, W. A. Warady and L. Y. Chiang,
Mol. Cryst. Liquid Cryst., 123, 205 (1985).
25 E. Fontes, P. A. Heiney, M. Ohba, J. N. Haseltine and A. B. Smith, Phys.
Rev., A 37, 1329(1988).
26 E. Fontes, P. A. Heiney and W. H. De Jeu, Phys. Rev. Lett., 61, 1202 (1988).
27 D. Goldfarb, Z. Luz and H. Zimmermann, J. de Physique, 42, 1303 (1981).
28 A. M. Levelut, P. Oswald, A. Ghanem and J. Malthete, J. de Physique, 45,
745 (1984).
See also, A. M. Levelut, / . de Physique Lettres, 40, L 81 (1979).
29 J. Malthete and A. Collet, Nouv. J. Chem., 9, 151 (1985).
30 A. M. Levelut, J. Malthete and A. Collet, J. de Physique, 47, 351 (1986).
31 H. Zimmerman, R. Poupko, Z. Luz and J. Billard, Z. Naturforsch., 40a, 149
(1985).
32 Lin Lei, Wuli (in Chinese) 11, 171 (1982).
33 J. M. Lehn, J. Malthete and A. M. Levelut, / . Chem. Soc. Chem. Commun.,
1794 (1985).
34 C. Destrade, N. H. Tinh, J. Malthete and J. Jacques, Phys. Lett., 79A, 189
(1980).
J. Malthete, C. Destrade, N. H. Tinh and J. Jacques, Mol. Cryst. Liquid
Cryst. Lett., 64, 233 (1981).
35 A. M. Giroud-Godquin and J. Billard, Mol. Cryst. Liquid Cryst., 66, 147
(1981).
36 K. Ohta, H. Muroki, A. Takagi, K. I. Hatada, H. Ema, I. Yamamoto and
K. Matsuzaki, Mol. Cryst. Liquid Cryst., 140, 131 (1986).
37 A. C. Ribeiro, A. F. Martins and A. M. Giroud-Godquin, Mol. Cryst.
Liquid Cryst. Lett., 5, 133 (1988).
38 S. Chandrasekhar, Plenary Lecture, X International Liquid Crystals
Conference, York, 15-21 July 1984; Mol. Cryst. Liquid Cryst., 124, 1 (1985).
39 J. Malthete, A. M. Levelut and N. H. Tinh, / . de Physique Lettres, 46, L-875
(1985).
40 N. H. Tinh, C. Destrade, A. M. Levelut and J. Malthete, / . de Physique, 41,
553 (1986).
41 E. I. Kats, Sov. Phys. JETP, 48, 916 (1978).
42 G. E. Feldkamp, M. A. Handschy and N. A. Clark, Phys. Lett. 85A, 359
(1981).

448

References:

Chapter 6

43 S. Chandrasekhar, K. L. Savithramma and N. V. Madhusudana, ACS


Symposium on Ordered Fluids and Liquid Crystals, Vol. 4 (eds. J. F. Johnson
and A. C. Griffin), p. 299, Plenum Press, New York (1982).
44 D. Ghose, T. R. Bose, C. D. Mukherjee, M. K. Roy, and M. Shah, Mol
Cryst. Liquid Cryst., 138, 379 (1986).
45 F. T. Lee, H. T. Tan, Y. M. Shih and C. W. Woo, Phys. Rev. Lett., 31, 1117
(1973).
46 D. Ghose, T. R. Bose, M. K. Roy, C. D. Mukherjee and M. Shah, Mol.
Cryst. Liquid Cryst., 154, 119 (1988).
47 E. I. Kats and M. I. Monastyrsky, J. Phys., 45, 709 (1984).
48 Y. F. Sun and J. Swift, Phys. Rev., A 33, 2735, 2740 (1986).
49 D. Ghose, T. R. Bose, M. K. Roy, M. Shah and C. D. Mukherjee, Mol.
Cryst. Liquid Cryst., 173, 17 (1989).
50 M. Kleman and P. Oswald, J. de Physique, 43, 655 (1982).
51 J. Prost and N. A. Clark in Proceedings of the International Conference on
Liquid Crystals, Bangalore, December 1979 (ed. S. Chandrasekhar), p. 53,
Heyden, London (1980).
52 G. S. Ranganath and S. Chandrasekhar, Curr. ScL, 51, 605 (1982).
53 V. G. Kammensky and E. I. Kats, Preprint of the Landau Inst. ofTheor.
Phys., Moscow (1982).
54 R. E. Peierls, Helv. Phys. Ada. Supplement, 7, 81 (1934).
55 L. D. Landau, in Collected Papers of L. D. Landau (ed. D. Ter Haar), p.
210, Gordon & Breach, New York (1967).
56 B. Jancovici, Phys. Rev. Lett., 19, 20 (1967).
57 L. Gunther, Y. Imry and I. Lajzerowicz, Phys. Rev., A 22, 1733 (1980).
58 L. D. Landau and E. M. Lifshitz, Statistical Physics, Part I, 3rd edn.,
p. 435, Pergamon Press, Oxford (1980).
59 M. Abramovitz and I. A. Stegun, in Handbook of Mathematical Functions,
p. 504, Dover, New York (1965).
60 M. Cagnon, M. Gharbia and G. Durand, Phys. Rev. Lett., 53, 938 (1984).
61 Y. Bouligand, / . de Physique, 41, 1297, 1307 (1980).
62 A. Queguiner, A. Zann and J. C. Dubois, Proceedings of the International
Liquid Crystal Conference, Bangalore, December 1979 (ed. S.
Chandrasekhar) p. 35, Heyden, London, (1980).
63 See for example, J. Friedel, Dislocations, Pergamon, Oxford (1967).
64 P. Oswald, / . de Physique Lettres, 42, L171 (1981).
65 M. Kleman, / . de Physique, 41, 317 (1980).
66 A. Isihara, J. Chem. Phys., 19, 114 (1951).
67 L. K. Runnels and C. Colvin, in Liquid Crystals, Vol. 3 (eds. G. H. Brown
and M. M. Labes), p. 299, Gordon & Breach, New York (1972).
68 R. Alben, Phys. Rev. Lett., 30, 778 (1973).
69 J. P. Straley, Phys. Rev., A 10, 1881 (1974).
70 J. D. Brooks and G. H. Taylor, Carbon, 3, 185 (1965).
71 B. Mourey, J. N. Perbet, M. Hareng and S. Le Berre, Mol. Cryst. Liquid
Cryst., 84, 193 (1982).
72 V. A. Raghunathan, N. V. Madhusudana, S. Chandrasekhar and C.
Destrade, Mol. Cryst. Liquid Cryst., 148, 77 (1977).
73 G. Heppke, H. Kitzerow, F. Oestreicher, S. Quentel and A. Ranft, Mol.
Cryst. Liquid Cryst. Lett., 6, 71 (1988).
74 G. Heppke, A. Ranft and B. Sabaschus, Mol. Cryst. Liquid Cryst. Lett., 8,
17(1991).

References: Chapter 6

449

75 T. Warmerdam, D. Frenkel and J. J. R. Zijlstra, / . de Physique, 48, 319


(1987).
76 See reveiw by S. Chandrasekhar and U. D. Kini, 'Instabilities in low
molecular weight nematic and cholesteric liquid crystals' in Polymer Liquid
Crystals (eds. A. Ciferri, W. R. Krigbaum and R. B. Meyer), Chapter 8,
Academic Press, New York (1982).
77 T. Carlsson, / . de Physique, 44, 909 (1983).
78 G. E. Volovik, JETP Lett., 31, 273 (1980).
79 L. J. Yu and A. Saupe, Phys. Rev. Lett., 45, 1000 (1980).
80 See review by K. Praefcke, B. Kohne, B. Gundogan, D. Singer, D. Demus,
S. Diele, G. Pelzl and U. Bakowsky, Mol. Cryst. Liquid Cryst., 198, 393
(1991).
81 S. Chandrasekhar, B. K. Sadashiva, S. Ramesha and B. S. Srikanta,
Pramana, J. Phys., 27, L713 (1986).
82 S. Chandrasekhar, B. K. Sadashiva, B. R. Ratna and V. N. Raja, Pramana,
J. Phys., 30, L49 (1988).
83 K. Praefcke, B. Kohne, D. Singer, D. Demus, G. Pelzl and S. Diele, Liquid
Crystals, 7, 589 (1990).
84 S. Chandrasekhar, B. R. Ratna, B. K. Sadashiva and V. N. Raja, Mol.
Cryst. Liquid Cryst., 165, 123 (1988).
85 S. Chandrasekhar, V. N. Raja and B. K. Sadashiva, Mol. Cryst. Liquid
Cryst. Lett., 7, 65 (1990).
86 S. Chandrasekhar, in 'Sir Charles Frank, an Eightieth Birthday Tribute' (eds.
R. G. Chambers, J. E. Enderby, A. Keller, A. R. Lang and J. W. Steeds) p.
254, Adam Hilger, Bristol (1991).
87 Similar results have been obtained by, M. Ebert, O. Herrmann-Schonherr,
and J. H. Wendorff, Makromol. Chem. Rapid Commun., 9, 445 (1988), in
their X-ray study of the N b phase of a sanidic discotic polymer. Evidence of
biaxiality in certain nematic polymers has also been reported by Hessel and
Finkelmann(88) and by Windle et a/.<89>
88 F. Hessel and H. Finkelmann, Polymer Bull., 15, 349 (1986).
89 A. H. Windle, C. Viney, R. Golombok, A. M. Donald and G. R. Mitchell,
Faraday Disc. Chem. Soc, 79, 55 (1985).
90 M. J. Freiser, Phys. Rev. Lett., 24, 1041 (1970).
91 C. S. Shih and R. Alben, / . Chem. Phys., 57, 3055 (1971).
R. Alben, / . Chem. Phys., 59, 4299 (1973).
92 R. Alben, Phys. Rev. Lett., 30, 778 (1973).
93 P. B. Vigman, A. I. Larkin and V. M. Filev, Sov. Phys. - JETP, 41, 944
(1976).
94 Y. Rabin, W. E. McMullen, W. M. Gelbart, Mol. Cryst. Liquid Cryst., 89,
67 (1982).
95 A. Saupe, P. Boonbrahm and L. J. Yu, / . Chim. Phys., 80, 7 (1983).
96 D. L. Johnson, D. Allender, R. De Hoff, C. Maze, E. Oppenheim and R.
Reynolds, Phys. Rev., B 16, 470 (1977).
97 P. H. Keyes, Proceedings of the Eighth Symposium on Thermophysical
properties - Thermophysical Properties of Fluids, Vol. 1 (ed. Jan V. Sengers),
p. 419, The American Soc. of Mech. Engineers, New York (1981).
98 R. G. Calfish, Z. Y. Chen, A. Berker and J. M. Deutch, Phys. Rev., A 30,
2562 (1984).
99 Z. Y Chen and J. M. Deutch, / . Chem. Phys., 80, 2151 (1984).
E. A. Jacobsen and J. Swift, Mol. Cryst. Liquid Cryst., 87, 29 (1982).

450

References: Chapter 6

100 J. W. Doane, NMR of Liquid Crystals (ed. J. W. Emsley), p. 441, Riedel,


Dordrecht (1985).
101 S. R. Sharma, P. Palffy-Muhoray, B. Bergersen and D. A. Dunmur, Phys.
Rev. A23, 3752 (1985).
102 D. W. Allender and M. A. Lee, Mol. Cryst. Liquid Cryst., 110, 331 (1984).
D. A. Allender, M. A. Lee and N. Hafiz, Mol. Cryst. Liquid Cryst., 124, 45
(1985).
103 T. C. Lubensky, Mol. Cryst. Liquid Cryst., 146, 55 (1987).
104 A. Saupe, J. Chem. Phys., 75, 5118 (1981).
105 M. Liu, Phys. Rev., A 24, 2720 (1981).
106 H. Brand and H. Pleiner, Phys. Rev., A 24, 2777 (1981).
107 E. A. Jacobsen and J. Swift, Mol. Cryst. Liquid Cryst., 78, 311 (1981).
108 G. E. Volovik and E. I. Kats, Zh. Eksp. Teor. Fiz., 81, 240 (1981).
109 U. D. Kini, Mol. Cryst. Liquid Cryst., 108, 71 (1984).
U. D. Kini, Mol. Cryst. Liquid Cryst., 112, 265 (1984).
110 E. Govers and G. Vertogen, Phys. Rev., A 30, 1998 (1984).
111 A. Chaure, / . Eng. ScL, 23, 797 (1985).
112 U. D. Kini and S. Chandrasekhar, Physica, 156A, 364 (1989).
113 U. D. Kini and S. Chandrasekhar, Mol. Cryst. Liquid Cryst., 179, 27
(1990).
114 D. Baalss, Z. Naturforsch., 45a, 7 (1990).
115 A. V. Kaznacheev and A. S. Sonin, Sov. Phys. Crystallogr., 33, 149 (1988).
116 L. G. Fel, Sov. Phys. Crystallogr., 34, 737 (1989); ibid., 35, 148 (1990).
117 D. Monselesan and H. R. Trebin, Phys. Stat. Sol. (b), 155, 349 (1989).
118 G. Toulouse, / . de Physique Lettres, 38, L-67 (1977).
119 V. Poenaru and G. Toulouse, / . de Physique, 8, 887 (1977).
120 G. Toulouse and M. Kleman, J. de Physique Lettres, 37, L149 (1976).
121 G. E. Volovik and V. P. Mineev, Sov. Phys.-JETP Lett., 24, 561 (1976).
122 N. D. Mermin, Rev. Mod. Phys., 51, 591 (1979).
123 V. P. Mineev, Sov. Scientific Reviews, Section A, Phys. Reviews, Vol. 2 (ed.
I. M. Khalatnikov), p. 173, Harwood Academic Publishers, London (1980)
124 H. R. Trebin, Adv. Phys., 31, 195 (1982).
125 M. Kleman, Points, Lines and Walls, Wiley, New York, (1983).

Index

anisaldazine
crystal structure, 18
orientational order parameter, 27
surface tension, 81
antiferroelectric short range order in
nematics, 75-80
antiparallel correlations in smectic A,
350-8
see also near neighbour correlations,
smectic A polymorphism
applications
liquid crystal display devices (LCD):
active matrices, 111; colour displays,
111; dye displays, 51, 112; dynamic
scattering mode, 178, 288; multiplexed
displays, 111; storage mode, 288;
supertwisted nematic (STN), 111-12;
surface-stabilized ferroelectric LCD,
380, 386-7; twisted nematic (TN),
106-11; twisted nematic, dynamical
characteristics of, 161-2; twisted
nematic, optical bounce and reverse
twist effects in, 112-13, 167
optical modulator, 381, 387
thermography, 296-7
tunable colour filter, 387
p-azoxyanisole, see PAA
p-azoxyphenetole, see PAP
backflow, 162-7
Bethe approximation, 71
biaxial nematics, 3, 378, 414-17, 449
(reference 87)
biaxiality:
in nematics, see biaxial nematics
in smectic C, 362-3
in smectic C*, 379
molecular, 37, 47-8, 298
bicritical point, 354
binary nematic mixtures, 51

nematic-nematic coexistence, 51
order parameter of individual
components, 51
blue phases, 5-6, 16
Landau theory of, 295-6
optical Kossel diagrams, 292
single crystals, 292
unit cells of blue phases I and II as cubic
networks of disclinations, 294
bond orientational order, 6, 360-2, 3 6 3 ^ ,
see also hexatic phases
Borrmann effect in cholesterics, 232-8
Bragg-Williams approximation, 20
Brillouin scattering, 325-6, 402, 404
Carr-Helfrich instability, 181-2, 286-7
CBOOA
correlation lengths, 343, 347
critical exponents associated with the
smectic A-nematic transition, 347
damping rate of undulation mode, 321-2
Helfrich deformation, 316
influence of twist distortion on the
smectic A-nematic transition
temperature, 305
layer compressibility, 315
orientational order parameter, 309
penetration depth, 315-16, 322
splay and bend constants, 344
chevron patterns, 178, 180, 387
cholesteric pitch
compensated mixtures of right- and lefthanded cholesterics, 218, 248, 276-7
critical divergence of, 349
dependence on
composition (in mixtures) 297-8;
magnetic field strength normal to the
helical axis, 277-80; pressure, 297;
temperature, 296-7
twist per unit length, 97

451

452

Index

cholesteryl nonanoate:
blue phases, 16
Borrmann effect, 236
critical divergence of pitch, 349
molecular structure, 16
transitions, 16
Clausius-Clapeyron equation, 27
conductivity:
electrical, 177, 181
thermal, 202-5
consistency relation
Chang, 72
Krieger-James, 72
Maier-Saupe, 44
continuum theory
of cholesterics:
coefficients of thermal conductivity,
260; distortion by external fields,
277-88; elastic free energy density, 97,
284-5; energetics of defects, 248-58;
flow along the helical axis
(permeation), 270-3; flow normal to
the helical axis, 274-6;
thermomechanical coupling, 258-67;
torque induced by electric field
(electromechanical coupling), 264-7;
torque induced by heat flux (the
Lehmann rotation phenomenon),
262-3; viscosity coefficients, 260
of columnar liquid crystals:
acoustic wave propagation, 401-2;
elastic free energy density, 397;
energetics of defects, 402-8;
fluctuations, 398; light scattering, 400;
mechanical instabilities, 400-1;
Peierls-Landau instability, 398; X-ray
scattering, 399
of nematics (including discotic
nematics):
backflow and kickback effects, 162-7;
distortions due to external fields,
98-117; elastic free energy density, 57,
97, 116; electrohydrodynamic
instabilities, 177-95; energetics of
defects, 117-44; Ericksen-Leslie
equations, 85-94; flexoelectricity,
205-12; flow properties, 144-59,
413-14; Freedericksz effect, static
theory, 98-106; Freedericksz effect,
dynamics of, 161-7; hydrodynamic
instabilities, 155-7, 195-201,413-14;
light scattering, 162-77;
Oseen-Zocher-Frank elasticity
equations, 94-7; Parodi's relation, 94;
periodic distortions in highly
anisotropic media, 113-15; reflexion
of shear waves, 159-61; summary of

equations, 97-8; thermal instability


201-5; twisted nematic and
supertwisted nematic devices 106-15;
viscosity coefficients, 92-3
of smectic A:
breakdown of conventional
hydrodynamics, 325; damping rate of
undulation mode, 322; elastic free
energy density, 310; energetics of
defects, 327-39; fluctuations, 317-19;
light scattering, 317-19; mechanical
instability, 31416; Peierls-Landau
instability 312-14; permeation, 321;
ultrasonic propagation, 323-5;
viscosity coefficients, 321
of smectic C:
elastic free energy density, 367;
fluctuations, 367; light scattering, 367
convection
oscillatory, 205
stationary, 202-5
critical end point, 355
critical point, 353
curvature elasticity, see elastic constants
4-cyanobenzylidene-4/-octyloxyaniline, see
CBOOA
cyanobiphenyls (4/-n-alkyl-4cyanobiphenyl) nCB
5CB (pentyl compound):
antiparallel local order, 80;
applications, 110; dielectric anisotropy
79; molecular structure, 15; surface
tension, 81; transitions, 15; X-ray
studies, 79
6CB:
electric and magnetic birefringence, 66
7CB:
applications, 110; X-ray studies, 79
8CB:
critical exponents associated with the
smectic A-nematic transition, 347;
molecular structure, 15; transitions, 15
9CB:
critical exponents associated with the
smectic A-nematic transition, 347
12CB:
surface-induced smectic order in the
isotropic phase, 84; X-ray specular
reflectivity from free surface, 84
80CB (octyloxycompound):
critical exponents associated with the
smectic A-nematic transition, 347;
PT diagram, 356; reentrant
behaviour at elevated pressures, 356;
surface-induced smectic order in
nematic phase, 83; X-ray specular
reflectivity from free surface, 83

Index
cybotactic groups, 2, 4, 83
Darwin's theory of X-ray diffraction, see
optical properties of cholesterics
defects
in cholesterics:
X (chi)-edge disclinations, 252; / (chi)screw disclinations, 249-51;
disclination pairs, 250-1; dislocations,
254-8; edge dislocations, 254-7;
fingerprint textures, 254-6;
Grandjean-Cano pattern, 257-8;
lattice disclinations, 252^4;
pincements, 254-6; screw dislocations,
254
in columnar liquid crystals, 402-11
developable domains, 40911;
disclinations, 408-9; dislocations,
403-8; longitudinal dislocations,
406-7; longitudinal wedge
disclinations, 408; screw dislocations,
408; transverse dislocations, 407;
transverse wedge disclinations, 408
in nematics:
angular forces between disclinations,
140-3; Brochard-Leger walls, 136;
consequences of elastic anisotropy,
13943; core structure, 1434; director
field in the neighbourhood of, 119-20;
disclination loops, 127-8; energies and
interactions, 120-2; Helfrich walls,
135; non-singular structures, 123-6;
schlieren textures in discotic nematics,
412; singular points, 129-30; solitons,
137-9; threads and schlieren textures,
7, 117; twist disclinations, 126-8;
umbilics, 136-7; wedge disclinations,
117-28
in smectic A:
disclinations, 328-9; edge dislocations,
333-7; energies and interactions,
335-7; focal conic textures, 327-33;
screw dislocations, 338
in smectic C:
core structure 3 7 3 ^ ; disclinations in
the odirector field, 368-9; edge
dislocations, 369-70; toric domains,
329
dielectric anisotropy, theory of, 51-7, 76-8
dielectric dispersion, 56
4,4'-di-n-alkoxyazoxybenzenes
dielectric dispersion, 56
odd-even effect, 48-51
structure, 50
see also PAA, PAP
director, definition of, 85-6
disclinations, see defects

453

discotic liquid crystals


columnar structures, 10, 390
columnar phases of conical, macrocyclic,
phasmidic molecules, 3934
discotic lamellar phase, 9, 394, 397
discotic nematic, 10, 393, 411-14
disc-shaped molecules, 389
extension of McMillan's model to
discotics, 394-7
high resolution X-ray studies on freely
suspended strands, 389-93
ordering of the cores and chains, 390-2
reentrant behaviour, 394
structure and classification of
mesophases, 8-10, 388-93
twisted nematic (or cholesteric), 393
see also continuum theory, defects,
discotic nematics, discotic polymers
discotic, nematics, 10, 390, 393, 411-14
diamagnetic anisotropy, 412
dielectric anisotropy, 412-13
director, 390, 412
Frank (elastic) constants, 413
hydrodynamical properties, 413-14
optical anisotropy, 412
discotic polymers, 10-12
columnar nematic, 11-12
columnar phase, 11-12
sanidic nematic, 11-12
dispersion forces, 41-2
donors and acceptors, 12, 358
Dupin cyclides, 329, see focal conic
textures
dynamic scattering, 178, 288, see also
applications
elastic constants
nematics:
critical divergence of, 3 4 2 ^ ;
dependence on order parameter,
59-60; experimental determination of
splay, twist and bend (or Frank)
constants, 98-106, 172; permanent
splay and twist, 96, 117
smectic A:
critical divergence of twist, bend and
layer compressibility constants, 343,
348; experimental determination of
layer compressibility, 315, 322
electric birefringence (Kerr effect), 55,
63-6
electric field induced distortions
cholesterics:
Carr-Helfrich instability, 286-7;
conduction and dielectric regimes,
288; square grid pattern, 282, 286-8;
storage mode, 288

454

Index

nematics:
Freedericksz transition, 106; principle
of the twisted and supertwisted
nematic devices, 110-13, see also
electrohydrodynamic instabilities,
flexoelectricity
smectic C*:
electroclinic effect, 380, 387; the
surface-stabilized ferroelectric display
device, 386-7; unwinding of helix,
380
electroclinic effect, 380, 382
application as optical modulators and
tunable colour filters, 387
electrohydrodynamic instabilities in
nematics
Carr-Helfrich mechanism, 181^1
chevron pattern, 178, 180
conduction and dielectric regimes, 178,
183
dynamic scattering, 178
flow and orientation patterns, 182
Helfrich's theory, 181-6
influence of flexoelectricity, 210-11
threshold for DC and AC excitation,
186-7, 187-94
Williams domains, 178-82
electromechanical coupling coefficient in
cholesterics, 266-7
electron spin resonance, 39, 146
enantiotropic transition, definition of, 14
end chains, their contribution to
mesophase stability
hard rods with semiflexible tails, 37
influence of chain length on
the columnar-nematic-isotropic phase
sequence, 394-7; the smectic
A-nematic-isotropic phase sequence,
306
odd-even effect in homologous series,
48-51
ordering of the cores and chains in the
columnar mesophase, 391-2
stiffening of the chains as a mechanism
for reentrance, 357-8
Ericksen-Leslie theory, 85-94
equal areas principle, 22-4, 37, 73
ferroelectric liquid crystals, 378-87
applications, 386-7
experimental studies on the polarization,
tilt angle, pitch, dielectric properties,
381^
generalized Landau theory, 385-6
liquid crystalline ferroelectrics as
improper ferroelectrics, 385
rotational viscosities, 384

soft mode and Goldstone mode, 382-3


structure of smectic C* and the origin of
ferroelectricity, 379-80
flexoelectricity, 205-11
contribution of flexoelectricity to
electrohydrodynamic instability,
210-11
determination of the flexoelectric
coefficients, 205-9
flow birefringence, 69-71
flow properties
of cholesterics:
flow along helical axis, 271-3; flow
normal to the helical axis, 274-7;
Helfrich's model of permeation,
2701; non-Newtonian behaviour,
268; oscillatory variation of apparent
viscosity with pitch, 276-7; secondary
flow, 274
of nematics:
backflow and kickback effects, 162-7;
flow alignment in discotic nematics,
413-14; flow alignment in ordinary
nematics, 150, 157; instability in
cybotactic nematics near the smectic A
transition point, 157; Miesowicz's
experiment, 144-5; Poiseuille flow,
148-52; shear flow, 152-7; transverse
pressure and secondary flow, 157-9;
Tsvetkov's experiment, 145-8; velocity
and orientation profiles, 152-3, 164,
166; viscosity (or Leslie) coefficients,
97-8; see also hydrodynamic
instabilities
of smectic A:
non-Newtonian behaviour, 319;
Helfrich's model of permeation, 321
fluctuation-dissipation theorem, 174
fluctuations
director fluctuations
in columnar liquid crystals, 400; in
nematics, 167-70; in smectic A,
317-19; in smectic C, 367
lattice fluctuations in the columnar
phase, 397-8
layer fluctuations in smectic A, 312-14
Peierls-Landau instability, 312-14, 398
pretransitional fluctuations
near the cholesteric-isotropic
transition, 289-93; near the nematicisotropic transition, 66-8; near the
smectic A-cholesteric transition, 349;
near the smectic A-nematic transition,
342; near the smectic C-nematic
transition, 378; near the smectic C smectic A transition, 371
Frank constants, see under elastic constants

Index
Freedericksz effect
in nematics
dynamical theory, 161-7; static
theory, 98-101
in smectic C, 365
free energy of elastic deformation, see
continuum theory
free standing monodomain smectic films,
studies on, 360-4
freely suspended monodomain discotic
strands, studies on, 389-93
Ginsburg criterion, 371
Ginsburg-Pitaevskii equation, 373
Goldstone mode in ferroelectric (smectic
C*) liquid crystals, 382-4
Grandjean-Cano pattern, 219, 257-8, 349
hard particle theories
Andrew's method, 37
Flory's lattice model, 36
hard rods with semiflexible tails, 37
Onsager's theory, 31
scaled particle theory, 36-7
stiffening of the tails as a mechanism for
reentrance, 357-8
summary of theories, 37, 411
Zwanzig's model, 31-6
Helfrich deformation in
cholesterics (square grid pattern), 281-9
columnar liquid crystals, 400
smectic A, 314-15
Helfrich's model of permeation, 270-1,
321
Helfrich's theory of electrohydrodynamic
instabilities in nematics, 181-7
Helfrich walls, 135
hexatic phases, 6
normal and tilted types, 301
X-ray studies of the degree of bondorientational ordering, 360^4
homeotropic orientation, 6, 106
homogeneous (or planar) orientation, 6,
104-5
orienting effect of grooves, 104-5
hybrid models: hard rods with a
superposed attractive potential, 60-1
hydrodynamics of liquid crystals
general formulation by Martin, Parodi
andPershan, 319,367-8
hydrodynamics of nematics, cholesterics,
smectics and discotics, see continuum
theory, flow properties
induced discotic phases, 12
induced smectic phases, 358
isotherms. 2 2 ^ . 37. 73

455

Landau-de Gennes model, 61-3


Landau-Ginsburg free energy, 342
Landau-Ginsburg parameter, 358
laser light beating spectroscopy, 177, 319,
321,367
Lennard-Jones potential, 21, 22
Leslie coefficients, 97-8, 260
Lifshitz point, 3745
light scattering
columnar liquid crystals, 400
nematics
dependence on temperature, 59-60;
depolarization, 171; determination of
elastic constants, 172, 347;
determination of viscosity coefficients,
176; eigenmodes and power spectrum
172-7; intensity and angular
dependence, 170-2; orientational vs
density fluctuations, 171; scattering
from the free surface, 82-3; scattering
from the isotropic phase, 66-8
smectic A
contribution of the undulation modes,
317-22; determination of the damping
rate, 320-2
smectic C
angular dependence of the intensity
and damping rate, 366-7
liquid crystal display devices, see
applications
living systems, 14
lyotropic systems:
aerosol-OT-water system, 12
plant virus preparations, 8
poly-y-benzyl-glutamate in organic
solvents, 13, 113-14,217-18
sufractant-water compositions, soaps, 8,
12
types of molecular packing, 12-14
magnetic birefringence (Cotton-Mouton
constant), 62-7
magnetic coherence length, 101, 135-7,
194, 278, 312
magnetic field effects in
cholesterics
cholesteric-nematic (unwinding)
transition, 277-80; square grid
pattern, 281-6; see also electric field
induced distortions
nematics:
dynamic effects, 161-7; effect on
electrohydrodynamic, hydrodynamic
and thermal instabilities, 179, 194,
196, 201, 205; rotating magnetic field,
144-8; static effects, 98-104, 106-9,
152-5

456

Index

magnetoclinic effect, 371-2


Maier-Saupe theory of nematics, 38-51,
70-1
applications
binary mixtures, 51; dependence of
elastic constants on the order
parameter, 57-60; dielectric
anisotropy, 51-6; odd-even effect,
47-51
extension of the theory
to cholesterics, 298; to discotics,
394-7; to smectic A (McMillan's
model), 302-9
limitations of the theory, 47-9, 70-71
Mauguin-Oseen-de Vries model, 237^41
MBBA
backflow effects, 165
disclination patterns, 122
electrohydrodynamic distortions, 180,
183^
flexoelectric coefficients, 207-9
flow birefringence, 71
heats of transition, 26
light scattering studies
in nematic phase, 60; in the isotropic
phase, 67-8
magnetic birefringence, 64
mixtures with cholesterics, dependence of
pitch on composition, 298
molecular orientation at free surface,
83
oblique rolls in electrohydrodynamic
patterns, 211
roll instability threshold, 198
thermal instability, 202, 205
torque in a rotating magnetic field, 148
transverse pressure and secondary flow,
158
viscosity (or Leslie) coefficients, 155
volume change at transition, 26
McMillan's model of smectic A, 302-9
melting of
inert gas crystals, Lennard-Jones and
Devonshire model, 17-19
molecular crystals, Pople-Karasz model,
19-27
mesomorphic groups, Hermann's
classification, 8
mesomorphism, thermotropic and
lyotropic, 1
metallomesogens, 389, 414
4-methoxybenzylidene-4/-butylaniline, see
MBBA
Miesowicz's viscosity coefficients, 144-5,
155
miscibility criterion, 300
molecular field, 38

molecular flexibility, 37, 41, 48-52; see also


end chains
molecular statistical theories:
cholesterics, 298
discotics, 394-7
nematics, 17-61, 71-80
smectic A, 302-9
smectic C, 364
monotropic transition, 14, 16, 379
multicritical point, nematic-smectic
A-smectic C (NAC), 374-8
Chen-Lubensky model, 375-8
high resolution X-ray calorimetric and
light scattering studies, 378
topology of phase diagram in the
temperature-concentration and
pressure-temperature planes, 374, 376
near neighbour correlations
antiparallel (or antiferroelectric)
correlation, 75-8
Bethe approximation, 71
correlation lacking a centre of inversion,
289
Krieger-James approximation, 72-3
short range order parameter, 75
nematic-isotropic transition
Bethe approximation, 71-80
effect of pressure, 27-30, 46-7
Landau-de Gennes model, 61-3
latent heat, 15-16,26,47
Maier-Saupe theory, 38-48
molecular statistical theories, 17^8
pressure-induced mesomorphism, 28-30
short range order effects, 61-70
volume change, 26, 45
nematic liquid crystals, 1-3, 8-10
nematics as solvents, 39, 146
nuclear magnetic resonance (NMR)
studies, 39, 61, 308-9, 392
odd-even effect, 48-51
Onsager's reciprocal relations, 69, 94
Onsager's theory of dielectric polarization,
52
Onsager's theory of the hard rod fluid, 31
optical properties of cholesterics
absorbing systems, 220-2, 232-6
analogy with Darwin's dynamical theory
of X-ray diffraction, 222-232
Borrmann effect, 232-6
equivalence of the continuum and
dynamical theories, 241-5
exact solution of the wave equation: the
Mauguin-Oseen-de Vries model,
237^1
oblique incidence, 245-7

Index
propagation along the helical axis,
213^5
propagation normal to the helical axis,
247-8
rotatory power, reflectivity and circular
dichroism as functions of wavelength,
pitch and sample thickness, 217-20,
227-32, 239^1
optical textures
cholesterics:
droplets with a /-line of strength 2,
442; fingerprint texture, 255-6;
Grandjean-Cano pattern, 258; helical
configurations of disclination pairs,
143, 249-51
columnar liquid crystals:
texture showing two orthogonal n
disclinations, 410
nematics:
disclination loops, 127-8; schlieren
textures {structures a noyaux), 6, 30,
117-20, 365, 412; singular points,
129-32; threads, 7-8; umbilics,
136-8
smectic A:
batonnets, 30, 333; fan-shaped and
polygonal textures, 9, 331; focal conic
textures, 9, 30, 327-33; stepped drop
(goutte a gradins), 332, toric domains,
329
smectic C:
edge dislocations, 370; toric domains,
329
order parameter
complex order parameter
smectic A, 340; smectic C, 370
orientational order parameter
curves, 40, 45, 75, 77, 303-5, 308-9;
definition, 3841; experimental
determination of <P2>, 39-^0, 308-9;
<.P4> from polarized Raman studies,
48-9; short range order parameter,
745, 77; variation near an interface,
82,211
positional order parameter in melting
theory, 23-4
translational order parameter
amplitude of one-dimensional density
wave (smectic A), 303-5, 307, 340;
density wave in two dimensions
(columnar phase), 395
orientation distribution function in MBBA
derived from Raman studies, 49
Oseen's equation of motion for the
director, 87
Oseen-Zocker-Frank elasticity equations,
94-7

457

PAA
comparison of properties with hard rod
and hybrid model calculations, 37,
60
dielectric constants, 53
dielectric dispersion, 56
dipole moment, 55
electrohydrodynamic distortions, 178
flow birefringence, 69-71
heat of transition and volume change,
26
Leslie coefficients, 150
light scattering studies, 59, 66-8, 83, 172,
176
lines of constant order parameter vs. In V
and In T, 46
magnetic and electric birefringence, 62-5
Miesowicz coefficients, 146
molar volume at the transition, 44
molecular orientation at the free surface,
83
molecular structure, 15
orientational order parameter, 27, 40
PAA -I- cholesteryl nonanoate, Borrmann
effect in, 236;
Poiseuille flow, 150-3
P-T diagram and other pressure studies,
28,46
surface tension, 81
torque in a rotating magnetic field
(Tsvetkov's experiment), 148
transitions, 15
twist elastic constant, 104
see also 4,4/-di-n-alkoxyazoxybenzenes
PAP
heat of transition and volume change,
26
light scattering studies, 59
Miesowicz coefficients, 146,
orientational order parameter, 27
twist elastic constant, 104
see also 4,4/-di-n-alkoxyazoxybenzenes
Parodi's relation, 93, 98
penetration depth, 314, 322, 342, 398
phase diagrams, 28-9, 291, 295, 306, 353-6,
376, 397
planar structure, see homogeneous
orientation
plastic crystals, 22
Poiseuille flow, 148-52, 157-9, 268-9,
270-3
polymorphism in
smectic A, 350-2
smectic C, 364-5
thermotropics, 14-16
pressure studies, 27-9, 46-7, 297, 356, 372,
374-6

458

Index

pretransition effects in the vicinity of


cholesteric-isotropic transition:
anomalous optical rotation in the
isotropic phase, 289-93;
coherence lengths, 293
nematic-isotropic transition:
Bethe approximation, 71-80;
coherence length, 68; flow
birefringence, 69-71; Landau-de
Gennes model, 61-3; light scattering
studies, 66-8; magnetic and electric
birefringence, 63-6; see also near
neighbour correlations
smectic A-cholesteric transition:
analogy with the behaviour of a
superconductor in a magnetic field,
348-9; critical divergence of the pitch,
349-50
smectic A-nematic transition:
critical divergence of twist and bend
elastic constants, 343; critical
divergence of viscosity coefficients,
345; critical exponents, 347; high
resolution X-ray, light scattering,
specific heat and other studies, 347;
Landau-Ginsburg description, 341-2;
longitudinal and transverse coherence
lengths, 343
smectic C-nematic transition:
Chen-Lubensky model, 3748; critical
divergence of the splay, twist and bend
constants, 378
smectic C-smectic A transition:
Landau mean field description, 370-2;
transition induced by layer
compression, 372; variation of tilt
angle, specific heat, magnetoclinic
effect near the transition, 370-2
smectic C*-smectic A transition:
critical variation of the frequencies,
dielectric strengths and viscosities
associated with the soft mode and
Goldstone mode, 384; generalized
Landau theory, 386; temperature
dependence of tilt angle, polarization,
pitch, electroclinic effect, dielectric
constant, 381-3
Raman scattering studies, determination of
the orientational order parameter
<^4>, 48-9
reentrant phenomenon:
in binary mixtures of polar compounds,
355
in discotics, 394
in non-polar materials, 358
in pure polar compounds at elevated
pressures, 356

molecular models, 357


multiple reentrance, 357
rotational transitions in crystals, 19
second sound, 324-6, 402, 404
shear flow, 69, 152-7, 195-201, 274-7, 383
shear waves, reflexion of, 159-61
short-range order, see near neighbour
correlations, pretransition effects
smectic A, 6, 301, 341
absence of true long range translational
order, 321
high resolution X-ray scattering studies,
313-14
polymorphic forms of smectic A and
transitions between them, 351-5,
see also under continuum theory, defects,
optical textures
smectic A* or the twist grain boundary
phase:
example of a compound which shows
this phase, 359
structure of smectic A*, 359
X-ray and optical studies, 359-60
smectic A-nematic transition:
complex order parameter, 340
influence of twist and bend distortions
on the transition, 358-9
McMillan's molecular model, 302-9
phenomenological theory, 340-1
review of theoretical situation and its
comparison with experiment, 347
tricritical point, 309, 341
see also under pretransition effects
smectic C:
molecular models, 364
optical biaxiality, 363
polymorphic forms in polar systems, 365
structure and symmetry, 362
temperature dependence of tilt angle 363
tilt and bond orientational order, 3 6 3 ^
see also under continuum theory, defects,
textures
smectic C-smectic A transition:
complex order parameter and the helium
analogy, 370
Landau mean field theory, 370-1
see also under pretransition effects
smectic C*:
applications
optical modulators, 386-7; surface
stabilized ferroelectric LCDs, 386-7;
tunable colour filters, 386-7
description of the structure and the
origin of ferroelectricity, 378-80
director modes, the soft mode and the
Goldston mode, 382
electric field effects, 380

Index
electroclinic effect, 380
measurements of the tilt angle,
spontaneous polarization, pitch,
dielectric constant and dispersion,
381-3
optical properties, 379
rotational viscosity measurements,
382-3
smectic C*-smectic A transition,
generalized Landau theory, 386; see
also under pretransition effects
smectic liquid crystals: structural
classification of smectics A-K, and the
chiral modifications of some of them,
301
stepped drop (goutte a gradins), 332
supertwisted nematic device, 111-12, see
also under applications, continuum
theory
surface-stabilized ferroelectric liquid crystal
device, 380, 3867, see also under
applications
surface studies:
interferometric study of the free surface
of a smectic, the stepped drop, 332-3
light scattering and optical reflectivity
from free surface 82-3
surface-induced orientational order, 82
surface-induced smectic order 83^4
surface tension measurements, 81-2
variation of the order parameter in the
liquid-vapour transition zone, 82, 211
X-ray reflectivity from free surface, 83^4
tetracritical point, 378
thermography, 2967
threshold
cholesteric-nematic transition, 280
electrohydrodynamic instabilities, 187,
194,211
Freedericksz transition, 99-101
Helfrich deformation in smectics, 315
hydrodynamic instabilities, 196, 201
mechanical instabilities in columnar
liquid crystals, 400-1
smectic C-smectic A transition through
layer compression, 373
square grid pattern in cholesterics, 281
thermal instability, 204
twisted nematic cell, 109-10
translational order parameter, see under
order parameter
tricritical point, 309, 341, 347-8, 371
triple point, 28-9
Tsevetkov's experiment, 145-8
twisted nematic device, 106-15, see also
applications, and under continuum
theory

459

twist grain boundary (TGB) phase, see


smectic A*
Ultrasonic propagation:
angular variation of Brillouin scattering,
326, 404
in columnar phase, 401-2
in smectic A, 323-5
second sound, 324, 402
undulation mode in
smectic A, 317, 322
smectic C, 367
virial coefficients, 34
viscosity coefficients (see also flow
properties)
cholesterics
Leslie coefficients, 260; viscosity
coefficients coupling thermal and
mechanical effects, 260, 262-3
columnar liquid crystals, 401-2
nematics:
experimental determination of, 144-8,
155-7, 161-2, 177; Leslie coefficients,
97-8; Miesowicz coefficients, 145;
twist viscosity and its critical
divergence, 147, 161, 177, 345
smectic A
permeation, 321; viscosity coefficients,
321
viscous stress tensor, 92, 97, 260, 321
viscous torque, 93, 147, 1 6 3 ^ , 183, 270,
345
volume change at transition, 26, 45
Williams domains, 178-80, see also
electrohydrodynamic instabilities
X-ray studies
absence of true long range translational
order in smectic A, 313
amplitude of density wave in smectic A
and pretransitional smectic-like order,
307-8
correlation length of two-dimensional
lattice in the columnar phase, 390
diffraction from
biaxial nematic, 416; cybotactic
nematic, 4; ordinary nematic, 3
evidence of interdigitated antiparallel
ordering in 5CB and 7CB, 79-80
grain size in TGB phase, 360
hexagonal-rectangular transition in
columnar phase, 391
hexatic ordering, 360-2
identification of
polymorphic forms of smectic A, 350;
reentrant phases, 355-7

460
location of
critical end point, 355; critical point
353; tricritical point, 354
longitudinal and transverse correlation
lengths and susceptibility near the
smectic A-nematic transition, 343,
346-7
ordering of cores and tails in columnar
phase, 391
reflectivity from free surface showing
surface induced smectic order.

Index
structure factor and Debye-Waller
factor for the columnar structure, 399
structures of columnar phases, 390-3
studies in the vicinity of the NAC
multicritical point, 377-8
tilt angle measurements in smectics C
and C*, 372, 381
XY model, 347, 361, 370
Zwanzig's model, 31-6

You might also like