0% found this document useful (0 votes)
779 views357 pages

Introduction To Liquid Crystals

Uploaded by

ADITYA AGRAWAL
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
779 views357 pages

Introduction To Liquid Crystals

Uploaded by

ADITYA AGRAWAL
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 357

INTRODUCTION TO

LIQUID CRYSTALS
INTRODUCTION TO
LIQUID CRYSTALS

Edited by
E. B. Priestley
Peter J. Wojtowicz
Ping Sheng
RCA Laboratories
Princeton, New Jersey

PLENUM PRESS • NEW YORK AND LONDON


Library of Congress Cataloging in Publication Data
Main entry under title:
Introduction to liquid crystals.
Includes bibliographical references and index.
1. Liquid crystals. I. Priestley, E. B., 1943- II. Wojtowicz, Peter 1., 1931-
III. Sheng, Ping, 1946-
QD923J57 548'.9 75-34195
ISBN-13: 978-1-4684-2177-4 e-ISBN-13: 978-1-4684-2175-0
001: 10.1007/978-1-4684-2175-0

© 1974, 1975 RCA Laboratories


Princeton, New Jersey
Softcover reprint of the hardcover 1st 1974
Plenum Press, New York is a division of Plenum Publishing Corporation
227 West 17th Street, New York, N.Y. 10011
United Kingdom edition published by Plenum Press, London
A Division of Plenum Publishing Company, Ltd.
Davis House (4th Floor), 8 Scrubs Lane, Harlesden, London, NWlO 6SE, England
All rights reserved
No part of this book may be reproduced, stored in a retrieval system, or transmitted,
in any form or by any means, electronic, mechanical, photocopying, microftlming,
recording or otherwise, without written permission from the Publisher
Preface

The existence of liquid crystals has been known for nearly a centu-
ry; yet it is only in the last ten years that their unique optical, electri-
cal, electro-optic, and thermal properties have been exploited to any
significant extent in such technological applications as digital d~­
plays and thermography. Digital watches equipped with liquid-crys-
tal displays (LCD's) have recently made their debut in the electronic
watch market, and the large-scale use of LCD's in a variety of other
applications requiring reliable, low-power digital displays is immi-
nent. There is good reason to believe that liquid crystals will be the
first electro-optic materials to find widespread commercial use. Apart
from applications, liquid crystals are unique among the phases of
matter. Lurking beneath their garish display of color and texture is a
great complexity of physical and chemical interaction that is only
now beginning to unfold in the face of a decade-old resurgence in all
aspects of liquid~rystal research. RCA Laboratories has participated
in this resurgence from its beginning in the early 1960's and at
present maintains active liquid-crystal programs both in basic re-
search and in device engineering.
In view of the widespread interest in liquid crystals at RCA Labo-
ratories, an in-house weekly seminar devoted to the subject of liquid
crystals was organized in the fall of 1973. The resulting lectures were
subsequently published in three issues of the RCA Review and, with
the incorporation of much additional material, eventually grew into
the present volume.
The book is intended as a tutorial introduction to the science and
technology ofliquid crystals. We believe it will serve as a useful prim-
er for those interested in the physics of liquid crystals and those
using or contemplating the use of liquid crystals in practical devices.
The book is not meant to be a review of the entire field of liquid crys-

v
vi PREFACE

tals, and no attempt has been made to include exhaustive compila-


tions of literature references. Emphasis has been given to areas gen-
erally ignored in other texts; specific topics emphasized include the
statistical mechanics of the molecular theory and various aspects of
device fabrication:
The eighteen chapters in the present volume can be divided into
four groups. Chapters 1 and 2 give a brief introduction to the struc-
tural and chemical properties of liquid crystals. The next eight chap-
ters develop both the microscopic statistical theories (Chapters 3-7)
and the macroscopic continuum theories (Chapters 8-10) of liquid
crystals. Chapters 11 through 17 treat various aspects of device appli-
cations and related considerations, including packaging; optical, elec-
tro-optic, and electro-chemical effects; addressing techniques; and
the use of liquid crystals in optical waveguides. The final chapter pre-
sents an overview of lyotropic liquid crystals with particular empha-
sis on the crucial role of the hydrophobic effect in their stability.
Important contributions to the success of this venture were made
by several individuals. We wish to express our sincere appreciation
for their kind efforts. Roger W. Cohen made the initial suggestion for
the seminar series and has given continued encouragement and sup-
port during the preparation of the book. George D. Cody was most in-
strumental in arranging for the publication of the lectures in the RCA
Review and had the foresight to suggest that they be incorporated in
the present volume. Ralph F. Ciafone provided outstanding editorial
support during publication of the lectures in the RCA Review and
also assumed substantial editorial responsibility for many of the
tasks involved in putting the lectures together as a book. Mrs. Doro-
thy C. Beres was especially helpful in typing the manuscripts for the
majority of the chapters. Thanks are also due each lecturer-author
whose contributions were invaluable in making the seminar series
and this resulting volume complete.

E. B. Priestley
Peter J. Wojtowicz
Ping Sheng
Contents

Chapter 1
Liquid Crystal Mesophases • E. B. Priestley
1. Mesophases ................................................. 1
1.1 Disordered Crystal Mesophases .................................. 2
1.2 Ordered Fluid Mesophases ...................................... 2
2. Types of Liquid Crystals ........................................ 3
2.1 Thermotropic Liquid Crystals ..................................... 3
2.2 Lyotropic Liquid Crystals ........................................ 3
3. Classification According to Molecular Order ......................... 4
3.1 Nematic Order ............................................... 4
3.2 Cholesteric Order ............................................. 5
3.3 Smectic Order ................................................ 7
4. Polymorphism in Thermotropic Liquid Crystals ....................... 9
5. Molecular Structure of Thermotropic Mesogens ..................... 10
6. Properties of Ordered Fluid Mesophases .......................... . 12

Chapter 2
Structure- Property Relationships in Thermotropic Organic Liquid
Crystals. Aaron W. Levine
1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2. Organic Mesophases ......................................... 16
3. General Structural Features of Mesogens .......................... 17
4. Effects of Structure on Mesophase Thermal Stability ................. 18
5. Homologous Series .......................................... . 22
6. Materials for Device Applications ................................ 24
7. Summary ................................................. . 27

Chapter 3
Introduction to the Molecular Theory of Nematic Liquid Crystals •
Peter J. Wojtowicz
1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2. Symmetry and the Order Parameter .............................. 32

vii
viii CONTENTS

3. The Molecular Potential ........................................ 34


4. The Orientational Distribution Function ............................ 35
5. Thermodynamics of the Nematic Phase ........................... 37
6. Fluctuations at Tc ............................................ 41

Chapter 4
Generalized Mean Field Theory of Nematic Liquid Crystals •
Peter J. Wojtowicz
1. Introduction ................................................. 45
2. The Pair Interaction Potential .................................... 46
3. The Mean Field Approximation .................................. 47
4. Statistical Thermodynamics .................................... 51
5. Nature of the Parameters UL . . • • . • • • • . . • • • . . • • . . • • . . . • • . • • • • • • . . 52
6. The Need for Higher Order Terms in V1 . . . . . • . . • • . . • • • . . . • • • • • • • • . . 54

ChapterS
Hard Rod Model of the Nematic-Isotropic Phase Transition • Ping Sheng
1. Introduction ................................................. 59
2. Derivation of Onsager Equations ................................. 61
3. Solution of Onsager Equations in a Simplified Case ................... 66

Chapter 6
Nematic Order: The Long Range Orlentatlonal Distribution Function •
E. B. ,Priestley
1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
2. The Orientational Distribution Function ............................ 72
3. Macroscopic Definition of Nematic Order .......................... 74
4. Relationship Between Microscopic and Macroscopic Order Parameters ... 75
5. Experimental Measurements .................................... 77
5.1 Measurements of (P2(COS 8) Based on Macroscopic Anisotropies ...... 77
5.2 Measurements of (P2(cos 8) Based on Microscopic Anisotropies ....... 78
6. Experimental Data ............................................ 79

Chapter 7
Introduction to the Molecular Theory of Smectlc-A Liquid Crystals •
Peter J. Wojtowicz
1. Introduction ................................................. 83
2. Symmetry. Structure and Order Parameters ........................ 84
3. Phase Diagrams ............................................. 87
4. The Molecular Potential ........................................ 88
5. Statistical Thermodynamics .................................... 91
6. Numerical Results ............................................ 93
7. Improved Theory ............................................. 96
8. The Possibility of Second-Order Transitions ......................... 99
Appendix .................................................. 100
CONTENTS ix

ChapterS
Introduction to the Elastic Continuum Theory of Liquid Crystals •
Ping Sheng
1. Introduction ................................................ 103
2. The Fundamental Equation of the Continuum Theory of liquid Crystals ... 104
3. Applications of the Elastic Continuum Theory ...................... 110
3.1 Twisted Nematic Cell ......................................... 110
3.2 Magnetic Coherence Length .........................,.......... 112
3.3 Freedericksz Transition ...................................... . 115
3.4 Field-Induced Cholesteric-Nematic Transition ..................... 120
4. Concluding Remarks ......................................... 125
Appendix .................................................. 126

Chapter 9
Electrohydrodynamlc InstabllHles In Nemallc Liquid Crystals •
Dietrich Meyerhofer
1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
2. Nature of the Instability and the Balance of Forces .................. 131
3. Dielectric Response ......................................... 131
4. Hydrodynamic Effects ........................................ 132
5. The Boundary Value Problem in the Conduction Regime .............. 134
6. The Torque Balance Equation .................................. 136
7. Numerical Results and Comparison with Experiment ................. 139
8. Range of Applicability ........................................ 140

Chapter 10
The Landau-de Gennes Theory of Liquid Crystal Phase TransHlons •
Ping Sheng and E. B. Priestley
1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143
2. Derivation of the Fundamental Equations of the Landau--{je Gennes Theory
....................................................... 145
2.1 The Partition Function ........................................ 145
2.2 The Landau Expansion ....................................... 150
2.3 Generalization of the Landau Expansion to Liquid Crystals ............ 153
3. Thermodynamic Properties of Liquid Crystal Phase Transitions ......... 165
4. Fluctuation Phenomena ....................................... 168
4.1 Homophase Fluctuations in the Isotropic Phase ..................... 189
4.2 Heterophase Fluctuations ..................................... 182
5. Observation of Fluctuations Using Light Scattering .................. 189
6. Magnetic Birefringence and the Paranematic Susceptibility ............ 193
Appendix A ................................................ 195
Appendix B ................................................ 198

Chapter 11
Introducllon to the Opllcal Properties of Cholesteric and Chlral Nemallc
Liquid Crystals • E. B. Priestley
1. Introduction ................................................ 203
2. Maxwell's Equations ......................................... 205
x CONTENTS

3. Discussion ................................................. 211


4. Conclusion ................................................ 215
Appendix A ................................................ 216
Appendix B ................................................ 216

Chapter 12
Liquid-Crystal Displays-Packaging and Surface Treatments •
L. A. Goodman
. 1. Introduction ................................................ 219
2. Packaging ................................................. 219
3. Electrodes ................................................. 220
4. Surface Orientation .......................................... 222
5. Influence of Packaging on Surface Orientation ..................... 230
6. Summary ................................................. 231

Chapter 13
Pressure Effects In Sealed Liquid-Crystal Cells • Richard WIlliams
1. Introduction ................................................ 235
2. Effect of Temperature Change ................................. 237
3. Effect of Glass Thickness ..................................... 238
4. The Case of a Rigid Container .................................. 239

Chapter 14
Liquid-Crystal Displays-Electro-optic Effects and Addressing Techniques •
L. A. Goodman
1. Introduction ................................................ 241
2. Electro-optic Phenomena ..................................... 242
2.1 Field-Induced Birefringence .................................... 242
2.2 Twisted Nematic Effect ....................................... 245
2.3 Guest-Host Effect ........................................... 248
2.4 Cholesteric-to-Nematic Transition ............................... 249
2.5 Dynamic Scattering .......................................... 251
2.6 Storage Mode .............................................. 255
2.7 Transient Response ......................................... 258
3. Display-Related Parameters ................................... 259
3.1 Display Life ................................................ 259
3.2 Temperature Dependence .................................... 260
4. Addressing Techniques ....................................... 261
4.1 Matrix Addressing ........................................... 261
4.2 Beam Scanning ............................................ 270
5. Summary ................................................. 273

Chapter 15
Liquid-Crystal Optical Waveguides • D. J. Channln
1. Introduction ................................................ 281
2. Guided Optical Waves ........................................ 282
3. Phase Matching and Coupling .................................. 286
CONTENTS xi

4. Scattering ................................................. 287


5. Liquid Crystal Waveguides ..................................... 288
6. Conclusions ............................................... 294

Chapter 16
The Electro-optic Transfer Function in Nematic Liquids • Alan Sussman
1. Introduction ................................................ 297
2. Geometrical Considerations in Optical Measurements ................ 299
3. Field Effects-Negative Dielectric Anisotropy ...................... 303
4. Field Effects-Positive Dielectric Anisotropy ....................... 305
5. Hydrodynamic Effects-Diffraction by Domains .................... 308
6. Dynamic Scattering .......................................... 312
7. Photoconductor Control ....................................... 314

Chapter 17
Electrochemistry in Nematic Liquid-Crystal Solvents • Alan Sussman
1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
2. Equilibrium Properties of Bulk Solutions ........................... 320
3. Electrochemical Reactions .................................... 328

Chapter 18
Lyotropic Liquid Crystals and Biological Membranes: The Crucial
Role of Water • Peter J. Wojtowicz
1. Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
2. Lyotropic Liquid Crystals ...................................... 334
2.1 Constituents of Lyotropics ..................................... 335
2.2 Micelles ................................................... 336
2.3 Structure of Lyotropics ....................................... 338
3. Biological Membranes ........................................ 340
3.1 Constituents of Membranes .................................... 341
3.2 Structures of Membranes ..................................... 342
4. Interaction of Amphiphilic Compounds with Water ................... 344
4.1 Solubility of Hydrocarbons in Water .............................. 345
4.2 Solubility of Ionized Species in Water ............................ 346
4.3 Aggregation of Amphiphilic Compounds .......................... 347
5. Conclusion ................................................ 349

Appendix ................................................... 351


Index ...................................................... 353
Liquid Crystal Mesophases

E. B. Priestley
RCA laboratories, Princeton, N. J. 08540

1. Mesophases

Everyday experience has led to universal familiarity with substances


that undergo a single transition from the solid to the isotropic liquid
phase. The melting of ice at Oac to form liquid water is perhaps the
most common such phase transition. There are, however, many
organic materials that exhibit more than a single transition in passing
from solid to liquid, thereby necessitating the existence of one or
more intermediate phases. It is not surprising that the molecular
ordering in these intermediate phases, known as "mesophases", lies
between that of a solid and that of an isotropic liquid. The partial
ordering of the molecules in a given mesophase may be either trans-
lational or rotational, or both. Clearly, translational order can be
realized regardless of molecular shape, whereas rotational order has
2 CHAPTER 1

meaning only when the constituent molecules are nonspherical (elon-


gated). Thus, there is good reason to expect molecular structure to
be an important factor in determining the kind and extent of ordering
in any particular mesophase.
Two basically different types of mesophases have been observed.
First, there are those that retain a 3-dimensional crystal lattice, but
are characterized by substantial rotational disorder (i.e., disordered
crystal mesophases), and second, there are those with no lattice, which
are therefore fluid, but nevertheless exhibit considerable rotational
order (i.e., ordered fluid mesophases). Molecular structure is in fact
important and, generally speaking, molecules comprising one of these
two types of mesophase are distinctly different in shape from molecules
comprising the other. Indeed, with the possible exception of some
polymorphous smectic materials, there are no known substances that
show both disordered crystal and ordered fluid mesophases.

1.1 Disordered Crystal Mesophases

Disordered crystal mesophases are known as "plastic crystals".' In


most cases plastic crystals are corpposed of "globular" (i.e. essentially
spherical) molecules, for which the barriers to rotation are small rela-
tive to the lattice energy. As the temperature of such a material is
raised, a point is reached at which the molecules become energetic
enough to overcome these rotational energy barriers, but not sufficiently
energetic to break up the lattice. The result is a phase in which the
molecules are translationally well ordered but rotationally disordered,
i.e., a disordered, or plastic crystal. Further increase in the tempera-
ture wiII result eventually in the molecules becoming energetic enough
to destroy the lattice, at which point a transition to the isotropic liquid
occurs. Perhaps the most striking property of plastic crystal meso-
phases is the ease with which they may be deformed under stress. It
is this softness or "plasticity" that gives these mesophases their name.
A detailed discussion of plastic crystals is outside the scope of this
chapter. However, those interested in pursuing the subject further are
directed to Ref. [lJ which provides an excellent elementary review of
the properties of plastic crystals.

1.2 Ordered Fluid Mesophases

Ordered fluid mesophases are commonly called "liquid crystals'''-6 and


are most often composed of elongated molecules. In these mesophases,
the molecules show some degree of rotational order (and in some cases
LIQUID CRYSTAL MESOPHASES 3

partial translational order as well) even though the crystal lattice has
been destroyed. Lack of a lattice requires that these mesophases be
fluid; they are, however, ordered fluid phases. It is this simultaneous
possession of liquid-like (fluidity) and solid-like (molecular order)
character in a single phase that makes liquid crystals unique and gives
rise to so many interesting properties.
In what follows, the various ordered fluid mesophases are described,
with particular attention being given to the nature of the molecular
ordering in each case. Also, some of the consequences of simultaneous
liquid-like and solid-like behavior in a single phase are discussed
qualitatively. Later chapters will deal at length with many subjects we can
consider only briefly here.

2. Types of Liquid Crystals

Two types of liquid crystal mesophases must be differentiated, viz.


thermotropic and lyotropic. Thermotropic liquid crystals are of in-
terest both from the standpoint of basic research and also for applica-
tions in electro-optic displays, temperature and pressure sensors, etc.
Lyotropic liquid crystals, on the other hand, are of great interest
biologically and appear to play an important role in living systems.

2.1 Thermotropic Liquid Crystals

The term "thermotropic" arises because transitions involving these


mesophases are most naturally effected by changing temperature. Ma-
terials showing thermotropic liquid crystal phases are usually organic
substances with molecular structures typified by those of cholesteryl
nonanoate and N- (p-methoxybenzylidene) -p'-n-butylaniline (MBBA)
shown in Fig. 1. Axial ratios of 4-8 and molecular weights of 200-500
gm/mol are typical for thermotropic liquid crystal meso gens. In this
type of liquid crystal, every molecule participates on an equal basis
in the long range ordering.

2.2 Lyotropic Liquid Crystals

Solutions of rod-like entities in a normally isotropic solvent often form


liquid-crystal phases for sufficiently high solute concentration. These
anisotropic solution mesophases are called "lyotropic liquid crystals"!·9
Although the rod-like entities are usually quite large compared with
typical thermotropic liquid-crystal meso gens, their axial ratios are
seldom greater than -15. Deoxyribonucleic acid (DNA), certain
viruses (e.g., tobacco mosaic virus (TMV)), and many synthetic poly-
4 CHAPTER 1

peptides all form lyotropic mesophases when dissolved in an appropriate


solvent (usually water) in suitable concentration. The conformation of
most of these materials is quite temperature sensitive, i.e. the rods
themselves are rather unstable with respect to temperature changes.
This essentially eliminates the possibility of thermally inducing phase
transitions involving lyotropic mesophases. A more natural parameter
which can be varied to produce such transitions is the solute concen-
tration. The principal interaction producing long range order in lyo-
tropic liquid crystals is the solute-solvent interaction; solute-solute
interactions are of secondary importance. To a good approximation,
then, only the rod-like entities (solute) participate in the long range
ordering.

CHOLESTERYL NONANOATE

(0)

N - (P - METHOXY BENZYLIDENE) - p'- BUTYLANILINE (MBBA)

( b)

Fig. 1-Examples of molecular structures that give rise to thermotropic


mesophases.

3. Classification According to Molecular Order


With the distinction between thermotropic and lyotropic mesophases in
mind, we proceed to the classification of these mesophases using a
scheme based primarily upon their symmetry. This scheme, first pro-
posed by Friedel in 1922,10 distinguishes three major classes-the
nematic, the cholesteric, and the smectic.

3.1 Nematic Order


The molecular order characteristic of nematic liquid crystals is shown
LIQUID CRYSTAL MESOPHASES 5

schematically in Fig. 2. Two features are immediately apparent from


the figure:
(1) There is long range orientational order, i.e., the molecules tend
to align parallel to each other.
(2) The nematic phase is fluid, i.e., there is no long range correla-
tion of the molecular center of mass positions.
In the state of thermal equilibrium the nematic phase has symmetry
co/mm and is therefore uniaxial. The direction of the principal axis
n (the director) is arbitrary in space.

A
n

1
Fig. 2-Schematic representation of nematic order.

3.2 Cholesteric Order

Fig. 3 shows the equilibrium structure of the cholesteric phase. As


in the nematic phase, lack of long range translational order imparts
fluidity to the cholesteric phase. On a local scale, it is evident that
cholesteric and nematic ordering are very similar. However, on a
n
larger scale the cholesteric director follows a helix of the form

nilJ = cos (qoz + <p)


ny = sin (qoz + <p)
nz = 0

where both the direction of the helix axis z in space and the magnitude
of the phase angle rf> are arbitrary. Thus the structure of a cholesteric
liquid crystal is periodic with a spatial period given by
6 CHAPTER 1

7T
L=--.
Iqol
The sign of qo distinguishes between left and right helicies and its
magnitude determines the spatial period. When L is comparable to
optical wavelengths, the periodicity results in strong Bragg scattering
of light. If the wavelength of the scattered light happens to be in the
visible region of the spectrum, the cholesteric phase will appear
brightly colored.

-
x
"n

Fig. 3-Schematic representation of cholesteric order.

It is interesting to note that a nematic liquid crystal is really


nothing more than a cholesteric with qo = 0 (infinite pitch). In fact,
the two are subclasses of the same family, the distinction being
whether the equilibrium value of qo is identically zero, or finite. If
the constituent molecules are optically inactive, i.e., are superimposabie
on their mirror image, then the mesophase will be nematic. If, on the
other hand, the constituent molecules are optically active, i.e., are not
superimposable on their mirror image, then the mesophase will be
cholesteric (except if the molecule and its mirror image are present
in precisely equal amounts, i.e. a "racemic" mixture, in which case
the mesophase will again be nematic).
Finally, a comment on nomenclature is in order. Cholesteric liquid
LIQUID CRYSTAL MESOPHASES 7

crystals get their name historically from the fact that the first
materials that were observed to exhibit the characteristic helical
structure were esters of cholesterol. It would be useful to continue
identifying as "cholesteric" those mesophases whose constituent mole-
cules are derivatives of cholesterol, and to use the term "chiral
nematic" to identify mesophases formed by optically active, non-
steroidal molecules.

3.3 Smectic Order

As many as eight smectic phases have been tentatively identified;


however, except for three that have been reasonably well characterized,
considerable uncertainty still exists about the exact nature of the
molecular ordering in these phases.
We will discuss only the three best understood smectic phases, the
smectic A, C, and B phases. All three appear to have one common
feature, viz. one degree of translational ordering, resulting in a layered
structure. As a consequence of this partial translational ordering, the
smectic phases are much more viscous than either the nematic or
cholesteric phase.

Smectic A Order

Within the layers of a smectic A mesophase the molecules are aligned


parallel to the layer normal and are uncorrelated with respect to center
of mass position, except over very short distances. Thus, the layers
are individually fluid, with a substantial probability for inter-layer
diffusion as well. The layer thickness, determined from x-ray scatter-
ing data, is essentially identical to the full molecular length. At
thermal equilibrium the smectic A phase is optically uniaxial due to
the infinite-fold rotational symmetry about an axis parallel to the layer
normal. A schematic representation of smectic A order is shown in
Fig. 4(a).

Smectic COrder

Smectic C order is depicted in Fig. 4(b). X-ray scattering data from


several smectic C phases indicates a layer thickness significantly less
than the molecular length. This has been interpreted as evidence for
a uniform tilting of the molecular axes with respect to the layer
normal. The fact that the smectic C phase is optically biaxial is further
evidence in support of a tilt angle. Tilt angles of up to 45° have been
8 CHAPTER 1

observed and in some materials the tilt angle has been found to be
temperature dependent. As in the smectic A phase, the layers are
individually fluid and inter-layer diffusion can occur, although most
likely with somewhat lower probability.

~
11111 \" 1" 1'1'1 ' I' \11 \
, 1'/"" \',"",, u' 1
11',1,'\111111 U III
(0)

111 ;;/; fI!


II I ;g I /III/; I
"n
1/11 / 1///
$/111111/
( b)

Fig. 4-Schematic representation of two types of smectic order: (a) smectic


A order; (b) smectic Corder.

Smectic BOrder

In addition to the layered structure, x-ray scattering data indicate


ordering of the constituent molecules within the layers of the smectic
B phase. Hence, the layers are no longer fluid, in contrast to the
smectic A and C phases. However, the mechanical properties of the
smectic B phase are quite different from those that would be expected
for a material having full 3-dimensional order; thus, we are forced to
conclude that the ordering, whatever its detailed nature, cannot be of
the sovt familiar in solids. It has been suggested that the smectic B
phase may in fact be a plastic crystaI. l1 If so, this would be very
interesting as it would provide the first opportunity to investigate both
disordered-crystal and ordered-fluid mesophases in a s'ingle material
LIQUID CRYSTAL MESOPHASES 9

(for any polymorphous smectic substance having a smectic B phase).


This model is not the only one possible for the smectic B phase, how-
ever. It could also be that this phase is composed of a collection of
2-dimensionalsolid layers coupled by very weak forces, such that the
layers could slip over one another quite easily.ll Although,:in principle,
these two models are expel"imentally distinguishable, there are no data
that differentiate between them at present. Smectic B phases can be
either bIaxial or uniaxial depending upon whether or nQlt there is a
finite tilt angle of the sort discussed above.

4. Polymorphism in Thermotropic Liquid Crystals

Many thermotropic materials have been observed to pass through more


than one mesophase between the solid and isotropic liquid phases.
Such materials are said to be "polymorphous". One can predict the
order of stability of these mesophases on a scale of increasing tem-
perature simply by utilizing the fact that raising the temperature of
any material results in progressive destruction of molecular order.
Thus, the more ordered the mesophase, the closer in temperature it lies
to the solid phase. From the description of the various types of order
given in Sec. 3, we can immediately draw the following conclusions:
(1) For a material having nematic and smectic trimorphous phases,
the order of mesophase stability with increasing temperature
will be
solid ~ smectic B ~ smectic C ~ smectic A ~ nematic ~ isotropic
This order of stability is incomplete agreement with experimental
observation.
(2) For a material having nematic and/or smectic phases, but not all
those listed in (1), the order of stability can be obtained from
that shown in (1) by simply deleting those phases not present.
This is also confirmed experimentally.
(3) For materials having both cholesteric and smectic mesophases,
the order is identical to that shown in (1) except the word
"nematic" is replaced by "cholesteric". The following have been
experimentally observed:
(a) solid ~ cholesteric ~ isotropic
(b) solid ~ smectic A ~ cholesteric ~ isotropic.

Finally, we note that there are no known examples of polymorphism


involving both nematic and cholesteric mesophases (recall the dis-
10 CHAPTER 1

cussion of molecular symmetry in Sec. 3.2). Of course, in the presence


of an external electric or magnetic field, a cholesteric liquid crystal
can be forced into a nematic structure. Such field induced distortions
have been studied extensively.

5. Molecular Structure of Thermotropic Mesogens

At present there is no way of predicting with certainty whether or


not a given molecule will exhibit liquid-crystal mesophases. However,
the presence of common structural features in the majority of thermo-
tropic liquid-crystal mesogens makes possible certain generalizations
regarding the types of molecules most likely to show liquid-crystalline
behavior. The two structural features that appear essential are (1) the
constituent molecules must be elongated and (2) they must be rigid.
It should be borne in mind that these are only generalizations, and
exceptions do exist.
We have already encountered one group of structurally related
compounds that often form liquid-crystal phases, viz. the substituted
cholesterols. It is evident from Fig. 1 (a) that these molecules are
elongated and quite rigid, thereby satisfying the two criteria above.
As mentioned earlier, all mesophases produced by molecules of this
type have finite pitch due to the optical activity of the parent cholesterol
molecule.
As a second group, we consider molecules whose structures can
be represented schematically as

LINKAGE GROUP

with n = 0, 1, or 2. Many of these molecules exhibit nematic and


smectic mesophases and, if optically active, may give rise to chiral
nematic mesophases. It is apparent even from this crude representa-
tion that the requirement of elongated molecules is satisfied by this
structure. The need for rigidity is satisfied, in general, by restricting
the linkage groups to those containing multiple bonds. Fig. 5 shows
three nematogenic molecules that serve to illustrate the structure for
n = 0, 1, and 2, respectively. In the p-pentyl-p'-cyanobiphenyl (PCB)
molecule, the biphenyl is considered as a single aromatic group. Thus
n = 0, there are no linkage groups, and the substituents are -C5Hu
and -CN. The p-azoxyanisole (P AA) molecule illustrates the case for
n = 1. Its two phenyl aromatic groups are linked by an azoxy group
LIQUID CRYSTAL MESOPHASES 11

and both substituents are CH30-. Notice the double bond between
the nitrogens of the linkage group. An example of the n = 2 case is
provided by the 2,6-di- (p-methoxybenzylideneamino) -naphthalene mole-
cule. The aromatic groups are phenyl, naphthyl and phenyl respec-
tively, with two -CH = N - linkage groups and two CH30- substitu-
ents. Again the linkage groups contain double bonds to provide the
requisite rigidity. Brown et al' list many other aromatic linkage,
and substituent groups that can be combined in the manner indicated
to form potentially mesomorphic molecules.

C5 H11 --@-@- CN
P- PENTYL -p!.. CYANOBIPHENYL (PCBI

°
CH30B-N=~ -@-OCH 3
P- AZOXYANISOLE (PAAI

0-0-CH"N~ ~
CH
3 ---v:::.r ~N=HC~OCH3

2,6-01-( P- METHOXY BENZYLIOENEAMINO)-NAPHTHALENE

Fig. 5-Examples of themotropic nematogens.

A word of caution is in order at this point. This purely "mechan-


ical" approach to "organic synthesis" is not to be taken too seriously.
It has many obvious shortcomings, not the least of which is its total
disregard for the laws of chemical combination. In spite of this, one
can very quickly write down a great many structures that do not
violate any chemical principles and that have a good chance of showing
liquid crystal mesophases. However, one should not be at all surprised
by the appearance of the occasional "dud".
Finally, there are other classes of materials, such as the substituted
monocarboxylic acids, that also form liquid-crystal mesophases. How-
ever, they represent relatively few of the known thermotropic meso-
12 CHAPTER 1

gens. For more details concerning the relationship of molecular struc-


ture to liquid crystallinity and for a discussion of these other classes
of mesomorphic materials, the reader is referred to the following
chapter12 and references therein.

6. Properties of Ordered Fluid Mesophases

The combination of molecular order and fluidity in a single phase


results in several remarkable properties unique to liquid crystals. By
now, it is quite evident that the constituent molecules of liquid crystal
mesophases are structurally very anisotropic. Because of this shape
anisotropy, all the molecular response functions, such as the electronic
polarizability, are anisotropic. The long range order in the liquid-
crystal phases prevents this molecular anisotropy from being com-
pletely averaged to zero, so that all the macroscopic response functions
of the bulk material, such as the dielectric constant, are anisotropic as
well. We have, therefore, a flexible fluid medium whose response to
external perturbations is anisotropic.
Consider the dielectric tensor £a(3' In the uniaxial nematic phase
(chosing the z-axis parallel to the nematic axis), £a(3 has the form

o
£1.
0)
0
o £11

with an anisotropy defined by .:l£ = £11 - £ 1.' Here £11 and £1. refer to
the dielectric constant parallel and perpendicular to the nematic axis,
respectively. The response of a nematic liquid crystal to an external
electric field depends on both the sign and magnitude of .:l£. If.:l£ is
positive, the lowest energy state in the presence of the field will be
that in which the nematic axis lies parallel to the field. For a negative
.:l£, the lowest energy state will be that in which the nematic axis is
perpendicular to the field. Due to the fluid nature of the phase, the
field strength necessary to cause such realignment of the nematic axis
is not very large. This ability to control the orientation of the
nematic axis by means of weak external fields is the basis for several
applications of nematic liquid crystals in optical display devices.
Cholesteric liquid crystals have also been put to practical use. In
this case, it is primarily the sensitivity of the pitch to changes in
temperature, pressure, etc. that are of interest. Recall that when the
pitch of a cholesteric is equal to an optical wavelength, Bragg scatter-
ing occurs. It is evident that by choosing a cholesteric of appropriate
pitch, changes in temperature and pressure can be monitored by means
LIQUID CRYSTAL MESOPHASES 13

of the accompanying color change of the material. Cholesteric liquid


crystals are especially useful in applications where large-area tem-
perature or pressure profiles must be determined.
Largely because of their high viscosity, smectic liquid crystals have
not come into widespread use. While it seems unlikely they will be
useful in any application where speed of response is important, they
may be valuable as storage media!"

References
1 J. G. Aston, "Plastic Crystals," in Physics and Chemistry of the Organic Solid State,
pp. 543-583, ed. by D. Fox, M. M. Labes, and A. Weissberger, Interscience Pub., N.Y.,
N.Y. (1963).
2 G. H. Brown, J. W. Doane, and V. D. Neff, A Review of the Structure and Physical
Properties of Liquid Crystals, CRC Press, Cleveland, Ohio (1971).
3 G. Durand and J. D. Litster, "Recent Advances in Liquid Crystals," in Annual Reviews
of Materials Science, Vol. 3, pp. 269-292, ed. by R. A. Huggins, Annual Reviews, Inc.,
Palo Alto, Calif. (1973).
4 I. G. Chistyakov, "Liquid Crystals," Sov. Phys. Usp., Vol. 9, p. 551 (1967).
5 A. Saupe, "Recent Results in the Field of Liquid Crystals," Angew. Chem. In!. Ed.
(English), Vol. 7, p. 97 (1968).
6 G. W. Gray, Molecular Structure and the Properties of Liquid Crystals, Academic
Press, N.Y., N.Y. (1962).
7 A. S. C. Lawrence, "Lyotropic Mesomorphism in Lipid-Water Systems," Mol. Cryst.
Liquid Crys!., Vol. 7, p. 1 (1969).
8 P. A. Winsor, "Binary and Multicomponent Solutions of Amphiphilic Compounds,"
Chem. Rev., Vol. 68, p. 1 (1968).
9 P. Ekwall, L. Mandell, and K. Fontell, "Solubilization in Micelles and Mesophases
and the Transition from Normal to Reversed Structures," Mol. Crys!. Liquid Crys!., Vol.
8, p. 157 (1969).
10 G. Friedel, "Les etats Mesomorphes de la Matiere," Ann. de Physique, Vol. 18, p.
273 (1922).
11 P. G. de Gennes, "Some Remarks on the Polymorphism of Smectics," Mol. Cryst.
Liquid Crys!., Vol. 21, p. 49 (1973).
12 A. W. Levine, "Structure-Properly Relationships in Thermotropic Liquid Crystals,"
Chapter 2.
13 F. J. Kahn, "IR-Laser-Addressed Thermo-Optic Smectic Liquid-Crystal Storage Dis·
plays," Appl. Phys. Lett., Vol. 22, p. 111 (1973).
Structure-Property Relationships in Thermotropic
Organic Liquid Crystals

Aaron W. Levine
RCA Laboratories, Princeton, N. J. 08540

1. Introduction

The phenomenon of thermotropic liquid crystallinity has been known i


at least since 1888. Since the early observations by Reinitzer l,2 and
Lehmann 3 of unusual melting behavior in certain organic compounds,
mesogenic compounds have been both actively sought and incidentally
discovered. Several relatively recent discoveries 4- 9 have shown the
technological utility of organic mesophases. Substantial impetus has
thus been provided for the systematic investigation of the relation-
ships between molecular structure and liquid crystallinity.
The literature of organic liquid crystals has been reviewed several
times. IO- 13 Reviews of those properties and compounds usable in some
electro-optical devices have also appeared. 14 The intention of this chapter
is to present a brief summary of thermotropic mesophasic behavior in
organic compounds as related to molecular structure. Recent activity
in this field will be indicated in an attempt to show current trends in
theory and experimentation. The cited reviews and the more recent
symposia reprints and abstracts I5 are commended to the interested
readers.
15
16 CHAPTER 2

2. Organic Mesophases

Organic liquid crystals may be broadly classified as smectic, nematic.


or cholesteric. These states are similar in that they exist between the
fully crystallized solid state and the isotropic melt and differ in the
extent of lattice order preserved in the mesophase. In the smectic
phase, molecules are constrained to be parallel with their neighbors
in layers, and translational motion between these layers is of low proba-
bility. In the nematic phase, molecules are able to translate in any
direction with respect to their immediate neighbors but are still
constrained to be parallel with them along the "nematic director".
When an otherwise nematogenic molecule also possesses chirality, a
twisting of the nematic director occurs and the resulting structure is
termed cholesteric. This phase, so named because it is frequently
encountered in derivatives of cholesterol, possesses some unique prop-
erties as a result of its twist but can, within limits, be considered as
a nematic phase for purposes of structural arguments. Fig. 1 illus-
trates these three types of ordering.

SM(CTIC

11111111111111111111111111111
1i1l[1[ [I[I [II I[ [111[111[11
11[1[111111111111111111111111

Fig. l-Molecular ordering in various mesophases.

The molecules in Fig. 1 are illustrated as being rods. When one


considers that the various mesophases are differentiated from each
other, and from the isotropic liquid, by the degree of freedom of mole-
cular motion, it is reasonable to hypothesize that intermolecular at-
tractive forces play an important role in mesophase stability. For a
given level of molecular attractions, long range ordering of mole-
cules would clearly be encountered more often as the molecules become
less spherical in shape. Thus, one would anticipate that molecular
associations capable of withstanding temperatures higher than the
STRUCTURE-PROPERTY RELATIONSHIPS 17

crystalline melting point, i.e., mesophases, would be more frequently


observed for molecules of rod-like shape. In addition to a molecular
shape that is considerably longer than it is wide (linear molecules):
mesophasic thermal stability is favored by molecular rigidity, per-
manent dipoles within the molecules, and a high level of molecular
polarizability. Nearly all mesogenic compounds contain multiple bonds
along their long axes, <aromatic nuclei, and either polar or long-chain
terminal groups.

DI POLAR

LINEAR

POLAR IZABLE

®-0- N =CH ---@- OCH 3 @-N=CH-@- OCH 3

MESOMORPHIC NOT MESOMORPHIC

Fig. 2-Some examples of mesogenic and non-mesogenic compounds illus-


trating some general structural requirements.

Fig. 2 illustrates these points with some well-known mesogens and


closely related compounds whose melting behavior is normal. Note
that, in this chapter, the term melting point is used to refer to the
solid-to-mesomorphic (or isotropic) transition while the abbreviations
C, S, N, Ch, and I will be used when referring to transitions involving
the crystalline, smectic, nematic, cholesteric, and isotropic states,
respectively. Thus, the melting point of a nematogen refers to the
C-N transition while N-I will describe the nematic clearing transition.

3. General Structural Features of Mesogens

The vast majority of compounds exhibiting liquid crystalline prop-


erties may be regarded as possessing a central linkage and end groups
(structure 1). When the central linkage is small, such as -C=N - or
-C=C-, aromatic nuclei are nearly always present. Mesogens con-
taining both carbocyclic and heterocyclic aromatic residues are known
18 CHAPTER 2

Terminal Group ~ Central Group ~ Terminal Group

Structure 1

and, in a sufficiently long molecule, substitution other than at para


positions is permissible. Central groups that require no highly polariz-
able aromatic residues are exemplified by the u:;-steroid ring system
and the carboxylic acid dimers. Fig. 3 shows the more common central
linkages.

0··· HO
-c, OH ···0"-"c-
~
- CH=CH- -C:C-

acid dim .. olefin acelylene


- CH= N- -CH= N-
o
1!.5 -sle,oid ozom.lhin. nil,one
(Schiff's bose I

-N=N- -C-O-
n
o
ozo ule,

Fig. 3-Common central linkages of mesogenic compounds of Structure 1.

Terminal groups vary widely in chemical nature. Fig. 4 presents


those most frequently encountered. The terminal group frequently
determines which mesophase will be observed. This wiII be discussed
in detail in the section on homologous series.

4. Effect of Structure on Mesophase Thermal Stability

Of the transitions observed upon melting a liquid crystalline material,


the melting point is least correlated with molecular structure. This is
so because the thermal stability of the solid state is principally in-
fluenced by attractive forces in the crystal lattice. Since these short-
range associations are only secondarily related to molecular structure,
the melting point is not usually influenced in an obvious manner by
molecular structure.
The forces responsible for the mesophase, however, are primarily
dipole-dipole (both permanent and induced) and dispersion forces,
which, since the molecules in the mesophase or melt are free to rotate,
STRUCTURE-PROPERTY RELATIONSHIPS 19

should be correlated with molecular structure. If the mesophase is to


survive a vibrational excitation that has already exceeded the lattice
energy, intermolecular polar attractions are required. In the follow-
ing considerations, the mesomorphic thermal stability will be used to
compare the relative efficiency of various structures in stabilizing a
mesophase. It is important to understand that we are refering here
to the maximum temperature at which a particular liquid crystalline

ALKYL - MAY BE BRANCHED

RO- ALKYLOXY: ALSO INTERNAL ETHERS


o
II
RO-C- CARBOA LKOXY

o
II
R-C-O- ALKYLCARBOXY
o
II
R-O-C-O- ALKYLCARBONATO

F, Ct, Br, I HALOGEN

- CN CYANO

NITRO

AMINO R MAY BE H

Fig. 4-Common terminal groups of mesogenic compound of Structure 1.

behavior is observed (i.e., the S-N, S-I, N -I, S-Ch, and Ch-I transi-
tions). Thermal stability is not to be confused with the mesomorphic
range, which is the difference between the isotropic transition and
melting temperatures.
For a mesophase to obtain, it is generally necessary that the ter-
minal groups of the molecules contain permanent dipoles. For example,
the N-I transition of 4-octyloxybenzoic acid!· is some 40° higher than
that of 4-nonylbenzoic acidI7 despite the essentially equivalent size
of the two molecules. Similarly, the 4-alkoxy-4'-cyanobiphenyls form
substantially more stable mesophases than 4-alkyl-4'-cyanobiphenyls.'8
Polarity of the termini, however, frequently gives rise to very strong
intermolecular attractions. In such cases, the melting points of the
compound may be raised so high that the mesophase cannot survive.
20 CHAPTER 2

For example, the end-group polarity of 6-hydroxy-2-naphthoic acid


(C-I=2500) is certainly greater than that of 6-methoxy-2-naphthoic
acid (C-N=206°; N-I=219°) .'9 In this case, hydrogen bonding is
thought to be responsible for the nonmesogenic property of the hydroxy
acid.
Frequently, more than one mesophase occurs upon heating a com-
pound. Under these conditions, not just the presence of polar inter-
actions but their direction is important. Thus, smectic states are
stabilized by multiple dipoles acting transverse to the molecular axis
(lateral attractions), while terminal attractions appear to be more
important in determining nematic thermal stability. It is often found
that the lower members of homologous series are purely or predo-
minantly nematic, while in the higher members, the smectic state
encroaches on, and eventually prevents observation of, the nematic.
The ratio of lateral to terminal attractions increases for an homologous
series due to increased dispersion forces and to increased shielding of
terminal dipolar attractions of, for example, a normal alkoxy or
acyloxy terminal group.
An additional difficulty in assessing the effects of end-group polarity
arises when compounds with very differently sized termini are com-
pared. Since it is usually difficult to change polarity substantially
without also changing the size of the polar substituent, great care is
necessary to predict mesophase stability. Thus, Brown and Shaw"
point out that N -I temperatures generally fall if the polarity of the
termini increase for a 'given molecular structure other than carboxylic
acid dimers. Aaron, Byron, and Gray 20 found, however, that for anils
of general structure 2, any substituent X provides a more thermally
stable nematic state than does X=H.

x -©- -©--©-
CH=N OCH 3

Structure 2

The polarity and polarizability of the central groups of mesogenic


compounds are more clearly correlated with thermal stability. Thus,
4-alkoxy-4'-biphenyl carboxylic acids 21 • form more stable mesophases
than corresponding 4-alkoxybenzoic acids.'· Similarly, esters of choles-
terol form generally more thermally stable mesophases than esters of
cholestanol. 22 These differ only by the presence of a double bond, as
shown in Fig. 5. Again, certain conjugated unsaturated aliphatic
carboxylic acids show liquid crystalline phases,"3 while normal aliphatic
STRUCTURE-PROPERTY RELATIONSHIPS 21

acids are not meso genic. For compounds based on central linkages
that allow approximate linearity, such as azo or azomethine, mesophase
thermal stability generally increases with increasing polarity or po!ar-
izability of the central linkage. Thus, N,N'-diarylazo nematogens clear
some 9 0 higher than corresponding Schiff bases!' Oxidation of both
classes of compounds (to azoxy and nitrone, respectively) raises the
N-J points by substantial amounts.

H
CHOLESTEROL CHOLESTANOL

H
EP I CHOLESTANOL CAPROSTANOL

Fig. 5-Ring structures for steroid derivatives discussed in text.

When the central linkage does not provide a lineal' molecule, how-
ever, mesophase stability suffers. Mesophases exhibited by derivatives
of phenyl benzoate, for example, are generally some 35 less stable
0

than those of correspondingly substituted azobenzenes.24 The effect of


linearity, or at least planarity, of molecules on mesophase stability is
further evidenced on comparison of trans-4-alkoxycinnamic acids" and
trans-stilbenes'6 with their corresponding cis isomers, which demon-
strates that severely nonlinear molecules show no mesomorphic prop-
erties. Again, derivatives of epicholestanol22 and caprostanol (Fig. 5)
are not known to be meso genic.
The geometrical anisotr~py of mesogenic compounds is certainly
their most obvious common feature. Since lateral attractions play an
important role in mesophase formation and stability, the effect of
molecular breadth has been studied rather extensively. Generally,
smectic states suffer more from molecular broadening than do nematics.
22 CHAPTER 2

Of course, when a substituent is added that changes both molecular


breadth and dipole attractions, a balancing of forces is observed, as
in the 5-substituted-6-alkoxy-2-naphthoic acids1' where broadening is
partially attenuated due to molecular geometry. Even more unusual
is the effect of addition of a 3-nitro group to 4-amino-4' nitro-p-
terphenyl which causes liquid crystallinity by changing terminal-group
interactions.27 A similar effect is known for stilbene derivatives'S and
several examples of Schiff bases exhibiting increased nematic thermal
stability as a result of 2-hydroxylation have been reported," By com-
paring the thermal stabilities of the cholesteric phase of 2-, 4-, and 2,
4-substituted benzoate esters of cholesterol, a marked effect of broaden-
ing is seen,'"30 This is perhaps surprising considering the size and
conformation of these compounds.
As the" previous discussion has indicated, despite the intuitive feel-
ing that mesophase thermal stability should be correlated with chemical
structure, the correlations are less than crystal clear. When large
changes are made in the structure of a molecule in order to study polar
effects, for example, other changes frequently occur that reinforce or
counterbalance the particular effect under study. Even so small a
change as reversing the termini of 4,4'-disubstituted benzylideneani-
lines causes some drastic changes in melting behavior.'"31

5. Homologous Series

One of the more common types of recent investigation into structure-


property relationships has been the synthesis of groups of molecules
differing from one another only by the number of methylene groups
in a terminal substituent. There appear to be two principal motiva-
tions for this type of study. First, the search for technologically use-
ful liquid crystals is usually directed toward materials of specific
nematic temperature ranges. The present lack of ability to predict
such ranges, coupled with the usual availability of general syntheses
for homologous series, often makes this type of investigation fruitful.
To date, few, if any, pure liquid crystals have been found that are
usable in display applications. Thus, all device manufacturing utilizes
mixtures of meso genic compounds. In many cases, these mixtures are
based on homologous series in order to obtain eutectic mixtures or to
guard against deleterious effects of material instabilities such as Schiff
base trans-substitution."
The second reason for studying homologous series is more funda-
mental. Most structure-property studies involving changes in polarity
of groups, polarizability, and molecular geometry, require substantial
STRUCTURE-PROPERTY RELATIONSHIPS 23

differences to be present among a collection of compounds. It is true,


for example, that a methyl, a cyano, and a nitro terminal group impart
very different polarity to a given molecule. The fact that these groups
are also quite different in size and shape, however, makes interpreta-
tion of structural effects difficult. In proceeding along an homologous
series, only very small changes are made at each step. It is thus hoped
that more specific conclusions would be possible when the effects on
mesophase behavior are correlated with molecular size and shape.
It was Gray who first recognized 1o that if the temperatures of a
mesophase-to-isotropic transition were plotted against the length of
the hydrocarbon chain being homologated, a smooth curve could be
drawn through the points. The same seemed true for mesophase-to-
mesophase transition temperatures. Of some seventy series available
at that time, seven general types of curves were found that could be
used to correlate all of the series. The curves differ from one another
in significant ways but some general conclusions are possible as enume-
rated below. It should be noted that subsequently reported homologous
series 33 also can be correlated by one or another of the general curves
of Gray, although, again, substantial differences in detail of the fit
among series that fit the same curve type are seen.
(1) The mesophase-to-isotropic transition temperature usually
falls with increasing chain length. A few exceptions to this rule are
known among compounds containing a biphenyl moiety which, due to
substitution, is restricted from adopting a coplanar conformation.
When the two aromatic nuclei are not coplanar, it is possible for the
termini of one molecule to interact strongly with one of the rings of
another molecule, causing additional attractions and a more thermally
stable nematic state as the terminus grows in length. A similar argu-
ment has been proposed for the appearance of mesomorphism in cer-
tain Schiff bases. 3'
(2) When more than one mcsophase is possible, the smectic state
increases in thermal stability, at the expense of the nematic or choles-
teric phase, as chain length increases. It is often seen that, while the
lower homologues are purely nematic, the higher are purely smectic
and the intermediate homologues exhibit both mesophases.
(3) Single smooth curves are usually not available to fit all of
the points. It is most often found that transition temperatures for
compounds with an odd number of carbon atoms in the chain are
correlated by one smooth curve, while homologues with even numbers
of carbon atoms fit another smooth curve. The transition temperature
difference between adjacent odd and even members of the series may
be only a few degrees or several tens of degrees depending on the
24 CHAPTER 2

particular series and the length of the chain. The odd and even
curves frequently converge at chain lengths of 8-10 carbon atoms but
this is not a necessary condition. This so-called "odd-even effect" is
usually justified in terms of a variation in conformation and, there-
fore, attractive forces as the methylene chain grows.

6. Materials for Device Applications

With the exception of cholesteric liquid crystal temperature indicators and


a few devices using smectic fluids,9,35 all liquid crystal displays rely on the
anisotropic properties of nematic liquids or mixtures of nematics and cho-
lesterics. In general, the material in a device is uniformly aligned in such a
way that, upon application of an electric field, the dielectric anisotropy of the
liquid crystal material causes it to become reoriented. Since light propagating
in the direction of the long molecular axes of nematic liquid crystals is unaf-
fected while light propagating in the transverse directions may be, the reori-
entation of molecules will be detectable if the original alignment was proper-
ly chosen. This is covered in detail in a number of papers 4,6-8,36-40 and is
summarized in Table 1. For the present purpose, it is only necessary to ap-
preciate the role of dielectric anisotropy in determining the suitability of var-
ious liquid crystals for particl,ilar types of displays.
Dielectric anisotropy is defined as the difference of the dielectric constants
in the directions parallel and perpendicular to the long molecular axis, i.e., a€
= €II - €.L' A linear molecule containing an on-axis polar group would have
positive dielectric anisotropy (a€ > 0). Since few mesogens are entirely linear,
most have a small perpendicular dielectric constant. Thus, in the absence of a
strong dipole along the molecular axis, Schiff bases, esters, stilbenes, acid di-
mers, etc., exhibit small negative dielectric anisotropy, in the range 0 to -1.
Molecules with substantial off-axis dipolar groups, such as a-substituted stil-
benes,41 exhibit correspondingly larger negative values of a€. Diaryl azo nem-
atogens are apparently more linear than corresponding diaryl Schiff bases
and frequently exhibit positive dielectric anisotropy.42 Even azoxy com-
O
t
pounds which would appear to have a large off-axis dipole (N=N) can occa-
sionally have 42 a€ > 0 although the more frequently exhibited property for
such compounds is negative dielectric anisotropy.
If a nematogen is substituted with a strongly dipolar moiety along its
major axis, positive anisotropy is ordinarily obtained. Thus, while 4-ethoxy-
benzylidene-4' -n- butyl aniline (EBBA) has weak negative anisotropy, 4-
ethoxybenzylidene-4' -cyano aniline (PEBAB) has a strongly positive dielec-
tric anisotropy. The cyano group (-C=N:) is the most commonly employed
substituent for this purpose, and most molecules bearing an on-axis CN sub-
stituent have a€ ~ + 10 or more. Other substituents can be used, and halogen,
~
:ll
C
~
C
Table 1-Summary of Important Nematic Liquid Devices :ll
m
Pola- -'a
rizers :ll
Original Observable Effect Re- Materials
o
"II
Device Type Phenomenon Alignment Field off ~ on quired Considerations Material Example m
:ll
-t
Dynamic Hydrodynamic Perpendicular Clear ~ scattering 0 Negative f:1e p-Azoxyanisole -<
scattering turbulence or parallel (white) conductivity :ll
required m
r
Deformation of Reorientation Perpendicular Change in 2 Negative f:1e MBBAa ~
aligned phases polarization oz
Reflective Production of Perpendicular Clear +-+ scattering 0 Negative f:1e MBBA+ en
J:
storage mode stable scattering or parallel either storable + cholesteric cholesteryl
centers with field = 0 oleate ~
Twisted Reorientation Parallel Change in 2 Positive f:1e MBBA+ PEBAB
nematic polarization
Guest-host Reorientation Parallel Color ~ colorless 1 Positive f:1e MBBA+ PEBAB
pleochroic + indophenol
dye added blue
Field-induced Cholesteric +-+ Scattering <--+ clear 0 Positive f:1e MBBA+ PEBAB
phase change nematic cholesteric + cholesteryl
added oleyl carbonate
a 4-Methoxybenzylidene-4' -n-butyl aniline.

"-l
U"1
26 CHAPTER 2

nitro, and dialkylamino have been tried. There is one report of stable nema-
togens with on-axis hydroxyl groups,43 but measurements of the anisotropy
were not made.
Dielectric anisotropy is an additive molar property. Thus, a small amount
of PEBAB [Ilf ~ 10] (about 10-15 mol %) dissolved in MBBA [Ilf ~ -0.2]
will provide a material suitable for twisted nematic devices. The threshold
will, of course, be higher for this mixture than for a pure positive one such as
4-pentyl-4' -cyanobiphenyl,44,45 where the dielectric anisotropy is much high-
er. There are other influences on the threshold voltage for liquid crystal cells,
principally the materials' elastic constants and, in the case of dynamic scat-
tering, material viscosity. The response times also are dependent upon elastic
constants, viscosity, and dielectric anisotropy. These factors are discussed at
length in a review by Goodman. 46
Probably the single most important consideration for a useful liquid crys-
tal material is its nematic (or cholesteric) temperature range. Since the de-
vices under consideration are operative only when a mesophase is present, a
reasonably wide range is required if a commercial product is to be made.
Most devices currently sold have nematic ranges of about 60-70°C and NI
transition temperatures of 60-70°C. While this is satisfactory for now, de-
vices of the future, particularly those intended for outdoor applications or for
use by the military, will require broader ranges. Since no single material yet
discovered has a satisfactory nematic range for device applications, all cur-
rently marketed devices contain mixtures of liquid crystals. Probably the
most common mixture is the eutectic of 4-methoxybenzylidene-4' -butylani-
line (MBBA) and 4-ethoxybenzylidene-4'-butyl aniline (EBBA).47 This eu-
tectic has been variously observed to have CN of _1O°C,48 0°C,47 and high-
er. 49 It does have a strong tendency to supercool, and is certainly usable over
the range 5-60°C. Many commercial dynamic scattering devices contain this
mixture, and modifications of it with both positive and negative dielectric
anisotropy additives are also in common use.
The problem of preparing new useful mixtures is a complex one, even if
consideration is given only to the nematic range. Trial-and-error attempts to
find eutectics in other homologous series or using compounds of differing
molecular structure have met with reasonable success. 14 ,33,44 With the grow-
ing number of liquid crystals available for making mixtures, combined with
the need for mixtures designed for specific purposes, an improved method of
obtaining eutectics has been sought. Some indication of the need for a meth-
od to predict the nematic ranges of mixtures is seen when one considers that,
from 100 individual compounds, 79,375,395 unique eutectic mixtures are pos-
sible, if only 2-, 3-, 4-, and 5-component mixtures are considered. Each of
these has an unending number of noneutectic compositions and it is thus
easy to see how the probability of finding the few required mixtures by trial
and error rapidly becomes infinitesimal, even for the most prolific laboratory
chemist.
STRUCTURE-PROPERTY RELATIONSHIPS 27

The relationship between clearing temperature of a nematic mixture and


the NI temperatures of its components is closely represented by Eq. [1],

[1 ]

where NI is in degrees centigrade and X represents mole fraction. There have


been occasional observations of significance deviations from this relation-
ship, but they are relatively few. Predicting the eN temperature of a mixture
is quite a bit less straightforward. To date, the best approximation reported
utilizes the Schroeder-Van Laar equation

in which the eutectic melting point, T E , is related to the mole fractions, Xi,
eN temperatures, Ti, and molar enthalpies of fusion, fJii, of the components
of the mixture, and the gas constant, R. For an N-component mixture, N
equations of the form of Eq. [2] may be written and solved simultaneously.
Several authors have reported 50-53 moderate to very good success in using
this method to predict eN temperatures of mixtures of up to five compo-
nents, provided the components are structurally similar.
An alternative method of solution54 applies the fact that the sum of the
mole fractions in any mixture must be unity. Making use of this relationship
and solving Eq. [2] for mole fraction, Eq. [3] is obtained:

The solution of Eq. [3] by the Newton-Raphson method55 is readily pro-


gramed and runs very quickly on a digital computer, provided the initial
guess for TE is chosen with care. This method allows rapid cursory evaluation
of the potential mixture utility of new compounds.
While this type of prediction is quite useful, it can fail badly when complex
mixtures of structurally different molecules are considered. Even in such
cases, however, the success rate in finding useful mixtures is much better
than that resulting from undirected experimentation. In addition, the meth-
od also provides data useful in improving the model upon which the approxi-
mation is based. It may be expected that, in the near future, a more exact
prediction of mixture properties will be possible.

7. Summary

Organic compounds that form mesophases upon melting are charac-


terized by being long, narrow, linear molecules. Both permanent
28 CHAPTER 2

dipoles and polarizable moieties are required. The thermal stability


of the mesophase or mesophases formed depends in large measure on
subtle structural, steric, and electronic effects in the central and ter-
minal groups. Structure-property relationships have been investigated
by systematic variation in polarity, polarizability, molecular size, and
nature of the central linkage and termini. Some general principles
have evolved but the field must not yet be considered fully understood.
Liquid crystal materials for device applications are mixtures, usually of eu-
~dctic composition. Preparation of such mixtures requires consideration of
dielectric anisotropy, viscosity, and nematic range as principal properties. A
method of estimating nematic ranges 'If mixtures is available, but requires
modification.

References
I F. Reinitzer, Monalsh., Vol. 9, p.421 (1888).
2 F. Reinitzer, "History of Liquid Crystals," Ann. Physik., Vol. 27, p. 213 (1908).
3 O. Lehmann, "Liquid Crystals," Ber" Vol. 41, p. 3774 (1908).
• G. H, Heilmeier, L. A. Zanoni, and L. A. Barton, "Dynamic Scattering: A New Electro-
optic Effect in Certain Classes of Nematic Liquid Crystals," Proc. IEEE, Vol. 56, p. 1162
(1968).
5 G. H. Heilmeier, J. A. Castellano, and L. A. Zanoni, "Guest-Host Interactions in Nematic
Liquid Crystals," Mol. Cryst. Liq. Cryst., Vol. 8, p. 293 (1969).
6 G. H. Heilmeier and J, E. Goldmacher, "A New Electric Field Controlled Reflective
Optical Storage Effect in Mixed Liquid Crystal Systems," Proc. IEEE, Vol. 57, p. 34
(1969).
7 G. H. Heilmeier and J, E. Goldmacher, "Electric-Field-Induced Cholesteric-Nematic
Phase Change in Liquid Crystals," J. Chem. Phys., Vol. 51. p. 1258 (1969).
8 M. Schadt and W. Helfrich, "Voltage Dependent Optical Activity of a Twisted Nemalic
Liquid Crystal," Appl. Phys. Lett., Vol. 18, p, 127 (1971).
9 F. J. Kahn, "IR-Laser-Addressed Thermo-Optic Smectic Liquid Crystal Storage Dis-
plays," Appl. Phys. Lett., Vol. 22, p, 111 (1973).
10 G, W. Gray, Molecular SIruclure and Ihe Properlies of Liquid Cryslals, Academic
Press, N. Y., N. Y. (1962).
11 G. H. Brown and W. G. Shaw, "The Mesomorphic State Liquid Crystals," Chem.
Rev., Vol. 57, p. 1049 (1957).
12 A, Saupe, "Recent Results in the Field of Liquid Crystals," Angew. Chem. (Inter-
national Ed. in English), Vol. 7, p. 97 (1968).

Properlies 0'
13 G, H. Brown, J. W. Doane, and V. D. Neff, A Review of Ihe SIruclure and Physical
Liquid Cryslals, CRC Press, Cleveland, Ohio (1971).
,. J. A. Castellano, "Liquid Crystals for Electro-Optical Application," RCA Rev., Vol. 33,
p. 296 (1972); L. T. Creagh, "Nematic Liquid Crystal Materials for Displays," Proc.
IEEE, Vol. 61, p, 814 (1973).
15 See, e.g., R. S. Porter and J. F. Johnson, Eds., Ordered Fluids and Liquid Cryslals,
Amer. Chem, Soc., Wash., D.C. (1967); G. H. Brown, G. J. Diene, and M. M. Labes,
Eds., Liquid Crystals, Gordon and Breach, N. Y., N. Y. (1966); J. F. Jollnson and R. S.
Porter, Eds., Liquid Cryslals and Ordered Fluids, Plenum Press, N. Y., N. Y. (1970).
16 G. W. Gray and B. Jones, "The Mesomorphic Transition Points of the p-n-Alkoxy.
benzoic Acids," J. Chem. Soc., p. 4179, Part IV 1953.
17 C. Weygand and R. Gabler, Z. Phys. Chem. (Leipzig), Vol. 46B, p. 270 (1940).
18 G, W. Gray, K, J. Harrison, and T. W. Nash, Abstracts of the 166th Natl. Meeting of
the A,C.S., Chicago, 1973, (Abstract CoII·142).
19 G. W. Gray and B. Jones, "Mesomorphism and Chemical Constitution. Part IV,"
J. Chem. Soc., p. 236, Part I 1955.
20 See Ref. (10), p. 183.
STRUCTURE-PROPERTY RELATIONSHIPS 29

21 G. W. Gray, B. Jones, and F. Marson, "Mesomorphism and Chemical Constitution.


Part VIII." J. Chem. Soc., p. 393, Part I 1957.
22 C. Wiegand, Z. Naturforsch, Vol. 4b, p. 249 (1949).
23 W. Maier and K. Markau, Z. Phys. Chem. (Frankfort), Vol. 28, p. 190 (1961).
24 L. K. Knaak, H. M. Rosenberg, M. P. Serve, "Estimation of Nematic-Isotropic Points
of Nematic Liquid Crystals," Mol. Cryst. Liq. Cryst., Vol. 17, p. 171 (1972).
25 G. W. Gray and B. Jones, "Mesomorphism and Chemical Constitution. Part II,"
J. Chem. Soc., p. 1467, Part II 1954.
26 W. R. Young, A. Aviram, and R. J. Cox, Angew. Chem. (International Ed. in English),
Vol. 10, p. 410 (1971) and references therein.
27 P. Culling, G. W. Gray, and D. Lewis, "Mesomorphism and Polymorphism in Simple
Derivatives of p-Terphenyl," J. Chem. Soc., p. 2693 (1960).
28 R. J. Cox, "Liquid Crystal Properties of Methyl Substituted Stilbenes," Mol. Cryst.
Liq. Cryst, Vol. 19, p. 111 (1972).
29 I. Teucher, C. M. Paleos, and M. M. Labes, "Properties of Structurally Stabilized Anil-
Type Nematic Liquid Crystals," Mol. Cryst. Liq. Cryst., Vol. 11, p. 187 (1970).
30 See Ref. (10), p. 194.
31 H. M. Rosenberg and R. A. Champa, "The Effect on Thermal Nematic Stability of
Schiff's Bases Upon Reversal of Terminal Substituents," Mol. Cryst. Liq. Cryst., Vol. 11,
p. 191 (1970).
32 H. Sorkin and A. Denny, "Equilibrium Properties of Schiff-Base Liquid-Crystal Mix-
tures," RCA Rev., Vol. 34, p. 308 (1973).
33 See, e.g., the following papers in Mol. Cryst. Liq. Cryst.: J. A. Castellano, M. T.
McCaffrey, and J. E. Goldmacher, "Nematic Materials Derived from p-Alkylcarbonato-p-
Alkoxyphenyl Benzoates," Vol. 12, p. 345 (1971); R. D. Ennulat and A. J. Brow ,1, "Meso-
morphism of Homologous Series," Vol. 12, p. 367 (1971); G. W. Gray and K. J. Harrison,
Vol. 13, p. 37 (1971); M. J. Rafuse and R. A. Soref, "Carbonate Schiff Base Nematic
Liquid Crystals: SyntheSiS and Electro-optic Properties," Vol. 18, p. 95 (1972); M. T.
McCaffrey and J. A. Castellano, "The Mesomorphic Behavior of Homologous p-Alkoxy-
p'-Acyloxyazoxybenzenes," Vol. 18, p. 209 (1972); and C. S. Oh, "The Effect of Hetero-
cyclic Nitrogen on the Mesomorphic Behavior of 4-Alkoxybenzylidene-2'-alkoxy-5'-
aminopyridenes," Vol. 19, p. 95 (1972).
34 V. I. Minkin, Y. A. Zhdanov, E. A. Madyantzeva, and Y. A. Ostroumov, "The Problem
of Acoplanarity of Aromatic Azomethines," Tetrahedron. Vol. 23, p. 3651 (1967); also
more recent calculations by H. B. Burgi and J. D. Dunitz, "Molecular Conformity of
Benzylide,1eanilines," Helv. Chim. Acta, Vol. 54, p. 1255 (1971).
35 C. Tani, "Novel Electro-Optical Storage Effect in a Certain Smectic Liquid Crystal," Appl. Phys.
Lett., Vol. 19, p. 241 (1971).
36 J. A. Castellano, E. F. Pasierb, G. H. Heilmeier, W. Helfrich, and M. T. McCaffrey, "Electronically
Tuned Optical Filter," Final Report to NASA, April 1970. Contract No. NAS12-638By.
37 M. F. Schiekel and F. Fahrenschon, "Deformation of Nematic Liquid Crystals with Vertical Orien-
tation in Electrical Fields," Appl. Phys. Lett. Vol. 19, p. 391 (1971).
36 L. T. Creagh, A. R. Kmetz, and R. A. Reynolds, "Performance Characteristics of Nematic Liquid
Crystal Display Devices," IEEE Trans. Electron Devices, p. 672 (1971).
39 F. J. Kahn, "Electric Field-Induced Orientational Deformation of Nematic Liquid Crystals: Tunable
Birefringence," Appl. Phys. Lett., Vol. 20, p. 199 (1972).
40 J. A. Castellano, E. F. Pasierb, C. S. Oh, and M. T. McCaffrey, "Electronically Tuned Optical Fil-
ters," Final Report to NASA, January 1972. Contract No. NAS1-10490, NTIS No. N72-20612.
41 W. R. Young, A. Aviram, and R. J. Cox, "New Non-Planar trans-Stilbenes Exhibiting Nematic
Phases at Room Temperature," Angew. Chern. Int. Ed. in English, Vol. 10, p. 410 (1971); W. H.
DeJeu and J. Van der Veen, "Instabilities in Electric Fields of a Nematic Liquid Crystal with Large
Negative Dielectric Anisotropy," Phys. Lett., Vol. 44A, p. 277 (1973).
42 W. H. DeJeu and T. W. Lathouwers, "Dielectric Properties of Some Nematic Liquid Crystals for
Dynamic Scattering Displays," Chem. Phys. Lett., Vol. 28, p. 239 (1974) and Ref. 5 therein.
43 J. E. Goldmacher and M. T. McCaffrey, "Nematic Mesomorphism in Benzylidene Anils Containing
a Terminal Alcohol Group," in J. F. Johnson and R. S. Porter, eds, Liquid Crystals and Ordered
Fluids, Plenum Press, N.Y., N.Y., 1970, p. 375.
44 G. W. Gray, K. J. Harrison, and J. A. Nash, "New Family of Nematic Liquid Crystals for Displays,"
Electron. Lett., Vol. 9, p. 130 (1973).
45 A. Ashford, J. Constant, J. Kirton, and E. P. Raynes, "Electro-Optic Performance of a New Room-
Temperature Nematic Liquid Crystal," Electron. Lett., Vol. 9, p. 118 (1973).
46 L. A. Goodman, "Liquid Crystal Displays," J. Vac. Sci. Tech., Vol. 10, p. 804 (1973); Also see
Chapter 14.
47 E. L. Strebel, "Eutectic Mixture of Para-Alkoxybenzylidene-Para-n-Alkyl Anilines," U.S. Patent
3,809,656, May 7, 1974.
48 L. T. Creagh, "Nematic Liquid Crystal Materials for Displays," Proc. IEEE, Vol. 61, p. 814 (1973).
30 CHAPTER 2

49 A. W. Levine, unpublished results.


50 J. A. Castellano, E. F. Pasierb, A. Sussman, C. S. Oh, D. Meyerhofer, M. T. McCaffrey, and R. N.
Friel, "Liquid Crystal Systems for Electro-Optical Storage Effects," Final Report to the AFML, Wright
Patterson AFB, December 1971. Contract No. F33615-70-C-1590. NTIS No. AD760-173.
51 U. Bonne, J. P. Cummings, J. Hicks, A. E. Johnson, B. L. Person, D. Saathoff, S. Scholdt, and J. L.
Stevens, "Properties and Limitations of Liquid Crystals for Aircraft Displays," Final Report to the
ONR, Department of the Navy, October 1972. Contract No. N00014-71-C-0262. NTIS No. AD751-
667.
52 E. C.-H. Hsu and J. F. Johnson, "Phase Diagrams of Binary Nematic Mesophase Systems," Mol.
Cryst. Llq. Cryst., Vol. 20, p. 177 (1973).
53 D. S. Holme, E. P. Raynes, and K. J. Harrison, "Eutectic Mixtures of Nematic 4'-Substituted 4-
Clanobiphenyls," J. Chem. Soc. Chem. Commun" p. 98 (1974).
5 A. W. Levine and R. J. Beshinske, unpublished results.
55 J. B. Scarborough, "Numerical Mathematical Analysis," Fifth Edition, The Johns Hopkins Press,
Baltimore, 1962, p. 199ff.
Introduction to the Molecular Theory of
Nematic Liquid Crystals

Peter J. Wojtowicz
RCA Laboratories, Princeton, N.J. 08540

1. Introduction

In this chapter we consider the very simplest approach to the molecular


theory of liquid crystals. We shaH approach the theory phenome-
nologicaHy, treating the problem of the existence of the nematic phase
as an order-disorder phenomenon. Using the observed symmetry of
the nematic phase we shaH identify an order parameter and then
attempt to find an expression for the orientational potential energy
of a molecule in the nematic liquid in terms of this order parameter.
Such an expression is easily found in the mean field approximation.
Once this is accomplished, expressions for the orientational molecular
distribution function are derived and the thermodynamic functions
simply calculated. The character of the transformation from nematic
liquid crystal to isotropic fluid is then revealed by the theory, and the
nature of the fluctuations near the transition temperature can be
explored.

31
32 CHAPTER 3

The results of this simple theory will be compared with experiment


and will be found to account qualitatively for many of the observed
features of nematic liquids. Though derived in a simple phenome-
nological way, this theory turns out to be equivalent to the well-known
pioneering theory of Maier and Saupe. 1 The relation of this theory to
more rigorous or complete approaches, as well as the required general-
izations ofthis theory will be considered in a subsequent chapter.2

2. Symmetry and the Order Parameter

The identification of the appropriate order parameter for nematic


liquid crystals is aided by a consideration of the observed structure
and symmetry of the phase. As in any liquid, the molecules in the
nematic phase have no translational order; i.e., the centers of mass of
the molecules are distributed at random throughout the volume of the
liquid. Experiments of many varieties, however, do demonstrate that
the nematic phase differs from ordinary liquids in that it is aniso-
tropic. The symmetry, in fact, is cylindrical; that is, there exists a
unique axis along which the properties of the phase display one set
of values, while another set of values is exhibited in all directions
perpendicular to this axis. The symmetry axis is traditionally referred
to as the "director". The optical properties of nematics provide an
example of how the cylindrical symmetry is manifest. For light
passing parallel to the director, optical isotropy is observed, while for
all directions perpendicular to the director, optical birefringence is
observed. Rays polarized parallel to the director have a different index
of refraction from those polarized perpendicular to the director.
Many experiments demonstrate that the anisotropy of nematics
arises because of the tendency of the rod-like molecules in the fluid
to align their long axis parallel to the director. This is shown sche-
matically in Fig. 1(a). The director is denoted by the symbol n; the
rod-like molecules are represented by the short lines. Note that at
finite temperatures, the thermal motion of the molecules prevents
perfect alignment with fJ.; the orientations of the molecules are in
fact distributed in angle, but with the director as the most probable,
or the most populated, direction.
If we look more closely at the orientation of a single molecule with
respect to the director, we find that the cylindrical symmetry of the
phase requires just a single order parameter to describe the structure.
In Fig. 1(b), we let the director lie along the z-axis of a fixed rec-
tangular coordinate system. The orientation of the rod-like molecule
can then be described using the three Eulerian angles shown. Because
MOLECULAR THEORY OF NEMATIC CRYSTALS 33

of the cylindrical symmetry, no order in the angles", (rotation about


the long molecuJar axis) or cp (rotation in the azimuthal direotion) is
permitted. If any angle", or cp was in some way preferred, a symmetry
lower than cylindrical would necessarily result. Thus, we are left with
the remaining angle 0 as the only one for which any degree of order
can exist. If any angle 0 is preferred, then the resultant symmetry is
indeed cylindrical; the most probable, or the prefered, angle in
nematics has been shown experimentally to be 0 = 0, the long molecular
axis parallel to n. If there is no preference for a particular 0, then
all such angles become equally probable and complete isotropy obtains;
this is the isotropic normal liquid phase .

.
n
.
n z

\ I '\ \, 1\
I \ I \ \ 8 /
/
I I I
\ I I\ I / I \
I I I \ \ \ /\

(0 ) (b)

Fig. 1- (a) Schematic representation of the structure of a nematic liquid


crystal. (b) The Euler angles required to describe the orientation
of a molecule in a nematic liquid.

The observed symmetry and structure of the nematic liquid has


enabled us to establish that a single order parameter will suffice to
describe the structure of the phase. Clearly, ordering in the polar
angle 0 distinguishes the nematic structure from the isotropic liquid.
However, 0 itself is not a convenient order parameter. By analogy
with ferromagnetism, one might expect that the projection of the
molecules along n, cosO, would be a natural order parameter. This is
not quite correct, however, since the electron spins in magnetism have
a definite polarity, the heads (north poles) being distinctly different
from the tails (south poles). In nematic liquids, the experiments on
most materials (there are apparently some exceptions") demonstrate
that the heads or tails of the molecules are not distinguished in the
nematic structure. That is, as many head ends point up as tail ends;
these, furthermore, are randomly distributed throughout the fluid.
Thus, to prevent the order parameter from describing situations in
which the nematic phase becomes polarized (as in a ferromagnet
34 CHAPTER 3

polarized with north poles up) we will use cos 2 (} rather than cos(} to
describe the structure. But now we must realize that it is not the
instantaneous value of cos 2 (} for a given molecule that we need. Rather,
we desire the average value of cos 2 (}, < cos 2 (} >, averaged over all
molecules in the liquid. When all the molecules are fully aligned
with ii, all () = 0 and < cos 2 (} > = 1. On the other hand, if the molecules
are randomly distributed in direction, all values of () are equally likely
and < cos 2 (} > = %.
By tradition, the order parameter in any order-disorder problem
is always taken such that it is unity in the perfectly ordered phase
and vanishes for the completely disordered phase. Examination of the
average values described above shows that the proper order parameter
for the nematic liquid crystal is

1
< P2 > = - (3 < cos 2 () > - 1). [1]
2

Clearly, < P2 > = 1 for the completely ordered nematic phase and
< P2 > = 0 for the disordered isotropic phase. The symbol < P 2 > is
used for the order parameter because we recognize the particular com-
bination in Eq. [1] to be the second-order Legendre polynominal,
P 2 (cos(}) = (3cos 2 () - 1) /2. We will continue to use this symbol in this
chapter and in the next' for its clarity when considering generaliza-
tions to the present theory. In the original theory' as well as in much
of the literature, the symbol S is used to represent < P 2 >. Values of
< p2 > between 0 and 1 describe degrees of ordering intermediate
between completely isotropic and completely ordered. It is the task
of order-disorder theory to (a) determine the temperature dependence
of < P 2 >, (b) calculate the thermodynamic and other properties in
terms of < p2 >, and (c) demonstrate the precise way in which the
transformation from finite < P 2 > to zero order occurs. We will now
examine these questions.

3. The Molecular Potential

The stability of the nematic liquid crystal results from interactions


between the constituent molecules. Without going into their nature
at this time, it is clear that there must exist interactions that cause
the molecules to prefer to align parallel to each other (and to the
mean direction of alignment, the director). In the spirit of the mean
field aproximation, we can attempt to mimic these intermolecular
interactions with an effective single-molecule potential function V.
MOLECULAR THEORY OF NEMATIC CRYSTALS 35

The potental V must have the correct orientation dependence; that is,
it should be a minimum when the molecule is parallel to the director
(parallel to the average direction of all the other molecules it interacts
with) and a maximum when the molecule is perpendicular to this
preferred condition. As we have seen above, the angular dependence
of - P 2 (cos ()) = - (3 cos2 () - 1) /2 is sufficient for this purpose. The
potential V representing the field of intermolecular interaction forces
should, furthermore, be a minimum when the phase is highly ordered
and should vanish when the phase becomes disordered. Thus, V should
be proportional to the degree of order, < P 2 >. Finally, V should
contain a factor v to describe the overall strength of the intermolecular
interactions; the differences between various materials will be ac-
counted for by allowing the strength v to vary from one substance
to another. Putting all the above together we arrive at mean field
approximation to the orientational potential energy function of a
single molecule:

V(cos (}) = -vP2 (cos (}) < P2 > . [2]

By mean field approximation we mean that the interactions between


individual molecules are represented by a potential of average force,
ignoring the fact that the individual behaviors and interactions of
molecules can be widely distributed about the average. This theory
therefore ignores fluctuations in the short-range order (mutual align-
ment of two neighboring molecules); the results of the theory will
then only be approximate in so far as near-neighbor fluctuation effects
have been neglected in the derivation of V, as well as in subsequent
calculations. The strengths and weaknesses of the mean field approxi-
mation are particularly well documented in the case of magnetism;4
similar considerations will apply to the case of nematic liquid crystals.

4. The Orientational Distribution Function

Having derived an approximate expression for the potential energy of


a single molecule in the nematic phase, we can now obtain the orienta-
tional distribution function. This function, which shall be denoted
by p(cos (}), describes how the molecules are distributed among the
possible directions about the director; it gives the probability of find-
ing a molecule at some prescribed angle () from n. With this function
we can compute the average values of various quantities of interest
pertaining to the nematic phase.
The rules of classical statistical mechanics give the orientational
distribution function in terms of the potential function V as:
36 CHAPTER 3

p(cos (J) = Z-l exp [-,8V(cos (J)],


[3]

Z = f
o
1

exp [-,8V(cos (J)]d(cos (J),

where Z is the single-molecule partition function and ,8 = l/kT (k is


Boltzmann's constant and T is the temperature). The integration over
all possible orientations of the molecule can be restricted to 0 ~ cos (J
~ 1, since both V and p are even functions of cos (J. As it stands,
however, Eq. [3] is still useless for the calculation of average values.
The reason is that the function V (and hence p) contains the order
parameter < P2 >, and this parameter is an as yet undetermined func-
tion of the temperature. Its temperature dependence, however, can be
determined as follows. < P 2 > is, afterall, just the average value of
the second Legendre function for a given molecule. Therefore, Eq. [3]
can be used formally to express this fact:

< P2 >= f
o
1

P 2 (cos {}) p(cos {}) d(cos {}). [4]

Expanding this equation to display its pertinent factors,

'f
1

P 2 (cos(J) exp [,8vP 2 (cos(J) <P2 >] d(cos{})


o

< P2 > = . [5]

f
1

exp [,8vP 2 (cos (J) < P 2 >] d(cos (J)


o

We now see in Eq. [5] a self-consistent equation for the determination


of the temperature dependence of < P 2 >. The order parameter < P 2 >
appears on both the left and right hand sides of the equation. For
every temperature T (or ,8) we can use a computer to obtain the value
(or values) of < P 2 > that satisfies the self-consistency equation. This
process has been accomplished and the results are depicted in Fig. 2.
< P 2 > = 0 is a solution at all temperatures; this is the disordered
phase, the normal isotropic liquid. For temperatures T below 0.22284
v/k, two other solutions to Eq. [5] appear. The upper branch tends
MOLECULAR THEORY OF NEMATIC CRYSTALS 37

to unity at absolute zero and represents the nematic phase, i.e., all
molecules trying to align with the director. The lower branch tends
to -lh at absolute zero and represents a phase in which the molecules
attempt to line up perpendicular to the director without azimuthal order.
This phase also has cylindrical symmetry but has not yet been seen
experimentally. That it ever will is rather unlikely, since we will see
later on that it is unstable with respect to the parallel-aligned nematic
phase. In the temperature range where Eq. [5] provides three solu-
tions, we require a criterion to determine which one of the three possi-
bilities actually exists. The laws of thermodynamics provide that the
observed solution, the stable phase, will be that one having the
minimum free energy.

,,
4 ,
,,
I

/
/
o ------------------------------------:;/---.---1

- 2

-4r ___ ----------- ---


-6[---~--- I 1 I I
I I
o 02 04 06 08 10 12 14 16 18 20 22 24
kT/v

Fig. 2-Temperature dependence of the order parameter obtained from


solving the self-consistency Equation, Eq. [5]. The stable equilib-
rium solutions are shown as the solid lines.

5. Thermodynamics of the Nematic Phase

The derivation of the free energy from the orientational distribution


function is straightforward. The free energy F is given by E-TS,
where E is the internal energy and S the entropy. The energy is com-
puted by taking the average value of the potential,

1 1
E=-N<V>=-N
2 2
f 1

v (cos ()) p (cos ()) d (cos ()) . [6] .


o
38 CHAPTER 3

The factor N comes in because we are dealing with a system of N


molecul~s. The factor 1/2 is required to avoid counting intermolecular
interactions twice (pair interactions have been approximated by a
single molecule potential V). The entropy is computed by taking the
average value of the logarithm of the distribution function:

N
S = -Nk < lnp > = - < V> + Nk InZ, [7J
T

where the quantities on the extereme right hand side have been defined
in Eqs. [6J and [3J. Combining Eqs. [6J and [7J,

1
F = -NkT InZ - - N < V> . [8J
2

At first sight this equation seems unusual with the presence of the
second term. The necessity for its existence can be verified im-
mediately in two ways. If we take the derivative (oF /0 < P2 » 7'
and set it equal to zero we regain Eq. [5J, the self consistency equa-
tion for < P 2 >. Thus, as required by thermodynamics, the self-
consistent solutions to our problem must be those that represent the
extrema of the free energy. Another verification of the correctness of
Eq. [8J comes from forming the derivative [df3F /df3J. Again, as
required by thermodynamics, Eq. [6J for the internal energy results.
The reason for the appearance of the second term in Eq. [8J is the
replacement of pair interactions by temperature-dependent single
molecule potentials."
The free energy for each of the three branches of the order para-
meter, Fig. 2, is computed from Eq. [8J by successive substitution of
the three different values of < P 2 > for each T. The free energy for
the < P 2 > = 0 branch is constant with T and equal to zero. The free
energy of the negative < P 2 > branch is negative but small in mag-
nitude. The free energy for the positive < P 2 > branch is negative
(with magnitude larger than that for the negative < P 2 > branch) up
to a temperature of 0.22019 v/k. The remainder of the positive < P 2 >
branch has positive values of F. Thus the stable phases to be observed
are as follows: from T = 0 to T = 0.22019 v/k the nematic phase is
stable. The order parameter decreases from unity to a minimum value
of 0.4289 at T = 0.22019 v/k. For temperatures above 0.22019 v/k the
isotropic phase with vanishing order parameter is stable (the stable
phases are shown as the solid lines in Fig. 2). At To = 0.22019 v/k we
MOLECULAR THEORY OF NEMATIC CRYSTALS 39

see that we have a first-order phase transition with the order para-
meter discontinuously changing from < P2 > = 0.4289 to < P 2 > = O.
To is usually called a critical temperature, although in much of the
liquid-crystal literature it is referred to as the clearing temperature.
Numerical values of the equilibrium order parameter < P 2 > for
various temperatures between zero and Tc = 0.22019 v/k have been
found on the computer and are presented in Table 1 as an aid to
anyone wishing to perform numerical calculations based on this or
the Maier-Saupe1 mean field theory.

Table i-Equilibrium values of the order parameter < Po > as a function


of temperature in the nematic range
kT/v <P2 > kT/v <P2 >
0 1.0000 .201 .6091
.01 .9899 .202 .6032
.02 .9794 .203 .5971
.03 .9687 .204 .5908
.04 .9576 .205 .5843
.05 .9461 .206 .5776
.06 .9342 .207 .5706
.07 .9218 .208 .5634
.08 .9089 .209 .5558
.09 .8953 .210 .5479
.10 .8810 .211 .5396
.11 .8657 .212 .5309
.12 .8493 .213 .5217
.13 .8315 .214 .5120
.14 .8210 .215 .5016
.15 .7910 .216 .4904
.16 .7655 .217 .4783
.17 .7372 .218 .4649
.18 .7041 .219 .4500
.19 .6644 .220 .4327
.20 .6148 .22019 .4289

We can get a better feel for Fig. 2 if we discuss it in terms of


some real temperatures. A typical nematic liquid crystal used in
device applications will have its clearing point To at about 50°C. Then
room temperature will correspond to about T = 0.2 v/k. Thus, the
degree of order at room temperature will be about < P 2 > = 0.615.
From room temperature upwards, < P 2 > will gradually decrease to
about 0.429 at the clearing point. Degrees of order larger than 0.75
to 0.8 are rarely if ever seen. Long before such order can be ac-
complished by lowering the temperature, the smectic phase and/or
the solid crystalline phase become more stable and spontaneously
appear. The general trend of the temperature dependence of < P 2 >
displayed in Fig. 2 is in qualitative agreement with experimental de-
terminations (the methods by which < P 2 > is measured are dis-
40 CHAPTER 3

cussed by E. B. Priestley in a separate lecture in this series). A more


detailed comparison for several real materials is deferred to the next
chapter.'
The first-order phase transition at T is very much like the phase
transition corresponding to the melting of a solid into its liquid phase.
That is, discontinuities in the volume, the internal energy (latent heat),
and the entropy of the system are observed. The present theory can-
not account for the small volume changes seen at T c ' since no distance
(volume) dependence of the potential V was assumed. We have
neglected this aspect in the present exercise, since the volume changes
actually observed over the short nematic ranges of most materials are
not very significant; theories that include the volume dependence are
available in the literature.,,6,7
The latent heat of the transition from nematic to isotropic liquid
can be calculated from Eq. [6] and the fact that < P2 > changes from
0.4289 to zero at To:

[91

The entropy of the transition can be computed from the formula


= fl.E/Tc so that
fl.S

[10]

The entropy change is a measure of the change in order at To and


conveniently comes out of the theory as a number without any para-
meters to be found or adjusted. First, we note that this is a very small
entropy change. Typical entropy changes for solid to liquid transitions
bf similar organic materials run around 25 caI/mol°R. Thus, the
nematic phase at Tc contains only a very slight amount of order,
certainly a lot less than a crystalline solid. The smallness of fl.S (T c)
is, of course, in keeping with the fact that the nematic phase possesses
order in only a single degree of freedom, the () angle of the individual
molecules. Secondly, we can compare Eq. [10] with some experimental
results; the agreement is quite satisfactory for such a simple theory.
As an example," the series of homologous 4,4'-d-n-alkoxyazoxybenzenes
from methyl to decyl displays entropy changes in the range from 0.3
to 1.9 caI/moloK, with most values being around 0.6 to 0.8.
The results obtained here and in the previous several sections are
completely equivalent to the mean field theory derived by Maier and
Saupe;' Eqs. [2], [3], [5], and [8] are identical to those presented in
this classic series of papers. Their approach is, of course, more
systematic than presented here, and the volume dependence of the
MOLECULAR THEORY OF NEMATIC CRYSTALS 41

potential is treated explicitly. In the next chapter" we will examine a


more systematic development of the theory and touch upon the volume
dependence of the potential.

6. Fluctuations at To
All nematic to isotropic liquid phase transitions have been observed
to be first order with a small volume change and a latent heat. The
present simple theory, as well as more sophisticated versions in the
literature, also display first-order changes in agreement with the
experiments. Many articles in the literature, however, refer to this
phase change as "nearly second order." Such remarks are not only
misleading but are entirely incorrect. The phase transition in nematics
cannot be anything but first order. The symmetry of the problem
demands it. Fig. 2 shows one way of looking at this: note that the
curves of < P2 > vs T for positive and negative < P 2 > are not sym-
metrical about < P 2 > = O. Because of this, the curves of positive
< P 2 > must intercept the abscissa with finite positive slope. In the
temperature range just above where this occurs, moreover, (positive)
< P 2 > is double valued. Thus, there is no way that a state of finite
< P 2 > can transform into a state of zero < P 2 > except discontinuous-
ly. In the case of ferromagnetism or anti ferromagnetism, on the other
hand, the curves of order parameter vs T are symmetrical about the
T-axis; the positive branch is the mirror image of the negative branch.
In addition, the free energies of these two branches are identical.
The order parameter curves then approach the abscissa symmetrically
and with infinite slope. In these cases transformations from states of
finite order to ones of vanishing order can occur continuously, and
second-order phase changes are allowed.
Why then did the phrase "nearly second order" arise in the
literature? One reason is that the latent heat and volume changes at
the nematic-isotropic transition are very small and closely approximate
the vanishing changes of second-order phase transitions. The most
prominent reason, however, is that many properties of the isotropic
liquid display critical behavior on cooling to temperatures close to Tc
(this phenomenon has been termed "pretransitional behavior" in the
liquid-crystal literature). That is, certain response functions (such
as the magnetically induced birefringence") are seen to diverge as Tc is
approached from above. The divergence of response functions at Tc
is one of the hallmarks of second-order phase changes.'o In ferro-
magnetism, as an example, the magnetic susceptibility diverges as
(T-Tc) -1.35 as the Curie point is approached from above. In the case
of the isotropic liquid above its clearing point, the divergence never
42 CHAPTER 3

actually occurs; the first-order phase change to the nematic phase


takes place just a few degrees before. Thus the origin of the remark
"nearly second order".
N ow, critical behavior and the divergence of response functions
are not usually seen at first-order phase transitions. No pretransi-
tional behavior is observed in any of the properties of a liquid about
to freeze into its crystalline solid. Why then is there pretransitional
behavior at the nematic-isotropic transition? My own feeling is that
the responsible factor is the small entropy of transition, I:!.S, Eq. [10].
The divergence of response functions occur because of the presence of
fluctuations. If in a disordered phase a fluctuation occurs such that a
very small region of the material rearranges itself into a locally
highly ordered structure, that region will have an enhanced response
to some external stimulus; the corresponding response function will
therefore increase in magnitude. If the fluctuations occur to a very
great extent (as they do when second-order phase transitions are
approached) then the response functions diverge. It is the small I:!.S
for nematics (compared to the I:!.S of the melting/freezing transition)
that permits' a substantial amount of fluctuations, giving rise to pre-
transitional behavior.
We can make this somewhat more quantitative. According to
Landau and Lifshitz,lI the probability of a fluctuation whose change
in entropy (order) is I:!.s is proportional to exp (I:!.s/k) ; here I:!.s is not
the molar entropy of the transition, the I:!.S of Eq. [10], but the
entropy of ordering of a small region of the material. If this small
region contains n molecules (which order to a degree normally dis-
played by a nematic just at T c ), then I:!.s can be related to I:!.S, and the
probability of such a fluctuation becomes proportional to

exp (-nI:!.S/R) , [11]

where R is the universal gas constant. The negative sign appears


because I:!.s describes fluctuations from disordered to ordered states
while I:!.S is concerned with the transition from ordered to disordered
phases. Consider now a fluctuation in which ten molecules in the
isotropic phase get together and become as ordered as the nematic
phase at or near its Te. We find from Eqs. [10] and [11] that the
probability of occurrence of such a fluctuation is proportional to about
e- 4 • For a fluctuation in liquid in which ten molecules get together
and become as ordered as the crystalline solid, on the other hand,
Eq. [11] plus the observation that I:!.S ,--,25 cal/moloK give the proba-
bility as ,--e- 125 • Thus we see that a significant fluctuation in the
MOLECULAR THEORY OF NEMATIC CRYSTALS 43

ordering of ten molecules is _e 120 more probable just above the iso-
tropic-nematic transition than it is just above a typical freezing point.
This, then, is why pretransitional behavior is observed in nematics but
not in the case of crystalline solids.

References

I W. Maier and A. Saupe, Z. Naturforschg., Vol. 14a, p. 882 (1959) and Vol. 15a, p.
287 (1960).
2 P. J. Wojtowicz, "Generalized Mean Field Theory of Nematic Liquid Crystals," Chapter 4.
3 R. Williams, "Optical-Rotary Power and Linear Electro-Optic Effect in Nematic
Liquid Crystals of p-Azoxyanisole," J. Chem. Phys., Vol. 50, p. 1324 (1969) and D.
Meyerhofer, A. Sussman, and R. Williams, "Electro-Optic and Hydrodynamic Prop-
erties of Nematic Liquid Films with Free Surfaces," J. Appl. Phys., Vol. 43, p. 3685 (1972).
4 J. S. Smart, Effective Field Theories of Magnetism, W. B. Saunders Co., Philadelphia,
Pa. (1966).
5 E. R. Callen and H. B. Callen, "Anisotropic Magnetization," J. Phys. Chem. SOlids,
Vol. 16, p. 310 (1960).
6 S. Chandrasekhar and N. V. Madhusudana, "Molecular Static~1 Theory of Nematic
Liquid Crystals," Acta Cryst., Vol. A27, p. 303 (1971).
7 R. L. Humphries, P. G. James, and G. R. Luckhurst, J. Chem. Soc., Faraday Trans. II,
Vol. 68, p. 1031 (1972).
8 G. H. Brown, J. W. Doane and D. D. Neff, A Review of the Structure and Physical
Properties of Liquid Crystals, p. 43, CRC Press, Cleveland, Ohio (1971).
9 T. W. Stinson and J. D. Litster, "Pretransitional Phenomena in the Isotropic Phase
of a Nematic Liquid Crystal," Phys. Rev. LeU., Vol. 25, p. 503 (1970).
10 H. E. Stanley, Introduction to Phase Transitions and Critical Phenomena, Oxford
University Press, N.Y., N.Y. (1971).
II L. D. Landau and E. M. Lifshitz, Statistical Physics, p. 344, Pergamon Press Ltd.,
London (1958).
Generalized Mean Field Theory of Nematic
Liquid Crystals

Peter J. Wojtowicz
RCA Laboratories, Princeton, N. J. 08540

1. Introduction

In the previous chapterl we examined a simple version of the molecular


theory of nematic liquid crystals. The problem was treated as an
order-disorder phenomenon with the solution based on a phenome-
nologically derived single-molecule orientational potential. The results
of this development were found to be equivalent to the well known
mean field theory of Maier and Saupe.2
In this chapter we will consider a more systematic approach to the
molecular theory of the nematic state. In deriving the theory we will
follow the development of Humphries, James, and Luckhurst. 3 We
start with a completely general pairwise intermolecular interaction
potential. After expanding in a series of appropriate spherical har-
monics we will systematically average the pair-interaction potential to
45
46 CHAPTER 4

obtain a generalized version of the single-molecule potential function


in the mean field approximation. The lVIaier-Saupe theory2 will result
from the retention of only the first term in the generalized potential.
More general versions of the theory result from the inclusion of
higher-order terms. In the final sections of the chapter we will discuss
the volume dependence of the interaction potential, examine the rea-
sons why higher-order terms in the potential appear to be required.
and consider the agreement between the theory and various experi-
ments.

2. The Pair Interaction Potential

The stability of the nematic liquid-crystal phase arises from the


existence of strong interactions between pairs of the constituent
molecules. As in normal liquids, the potential of interactions will have
attractive contributions to provide the cohesion of the fluid, and re-
pulsive contributions which prevent the interpenetration of the mole-
cules. In the case of the rod-like molecules of nematics, however, these
interactions are highly anisotropic. That is, the forces acting between
such molecules depend not only on their separation but also, and most
importantly, on their mutual orientations. From the symmetry and
structure of the nematic phase, we see that the rod-like molecules in
fact interact in a manner that favors the parallel alignment of neigh-
boring molecules.
A complete understanding of the nature of the intermolecular
interactions is not available. The precise mathematical form of the
pair potential is not known. In the absence of such detailed informa-
tion, we proceed by assuming a perfectly general form for the pair
potential and then deriving the theory systematically from it.
Many coordinates are required to describe the orientation dependent
interaction between a pair of asymmetric molecules. Fig. 1 (a) depicts
the situation; in addition to the parameter r which gives the separation
of the centers of gravity, we require the three Eulerian angles of each
of the two molecules. Since in nematics there appears to be no order-
ing of the molecules about their long axes, we can eliminate the angles
!fi by assuming the molecules to be axially symmetric. The angles OJ
and CPi can then be considered as polar angles with the intermolecular
vector r as the common polar axis. The intermolecular pair potential
V 12 is then a function of five coordinates:

[1]
MEAN FIELD THEORY OF NEMATIC CRYSTALS 47

Pople4 has demonstrated a very powerful expansion for the pair


potential between axially symmetric molecules:

V 12 = 471" L: UL1L2m(r) YL1m(81, CP1) YL2m *(8 2, CP2)' [2]


L1L 2m

where the YLm (8, cp) are the usual spherical harmonics. For axially
symmetric molecules and for the coordinates defined in Fig. lea), the
potential V 12 really depends only on r, 81, ()2' and the combination
CPl - CP2; thus, the appearance of only a single m index in Eq. [2].

t l :~ t2
.
n
I 8; I

if'
"

(ol
\~ ,
\
\

(b)

Fig. 1-The coordinate systems required to describe the interaction between


two asymmetric molecules: (a) the intermolecular vector r is the
mutual polar axis and (b) the director fi is the polar axis for each
molecule.

This expression is particularly convenient, because (a) it separates


the distance and orientation dependencies of the potential and (b) the
coefficients of the expansion, the ULIL2m(r) , are foundS to decrease
rapidly with increasing L1 a.nd L 2 • If the molecules have reflection
symmetry (or if we assume that we can neglect the ordering of heads
and tails of the molecules), then only terms with even Ll and L2 are
required in Eq. [2].

3. The Mean Field Approximation

A rigorous molecular theory of a fluid system based on a pairwise


interaction potential as complicated as Eq. [2] is impossibly difficult.
A simple but adequate approach is to derive a theory in the mean field
approximation. That is, we derive a single molecule potential that
serves to orient the molecule along the symmetry axis (the director)
of the nematic phase. The single-molecule potential represents (ap-
proximately) the mean field of intermolecular forces acting on a given
molecule. Mean field theories have been found capable of describing
the qualitative behavior of many different cooperative phenomena. The
48 CHAPTER 4

theories are not quantitatively correct, however, since intermolecular


short-range order and fluctuation effects are neglected. The properties
of mean field theories in the case of magnetism have been described
by Smart.6
In order to derive a mean field approximation to the potential, we
first have to express V 12 in terms of a polar coordinate system based
on the director, n, as the polar axis. The coordinate axes for the
molecules 1 and 2 must be rotated from that shown in Fig. 1 (a) to
that shown in Fig. 1 (b). The primed angles now describe the orienta-
tions of the molecules with respect to the new rotated coordinate
system. Mathematically, the rotation of the coordinate axes transforms
the spherical harmonics into the form

YLm(O, cp) = L: DpmLYLP(O', cp'), [3]


p

where the Dpm L are the elements of the Wigner rotation matrices. 7 In
the new coordinate system V 12 assumes the form

V 12 = 47T L: L: UL1L2m (r) Y L1P (O{, CP1') Y L2q* (0 2', cp/) (DpmJ,!) (D qmJ'2) 'I.".
L!L2m pq

[4]

To obtain the single molecule potential V 1 in the mean field ap-


proximation, it is necessary to take three successive averages of the
function V 12 • First, we average over all orientations of the inter-
molecular vector r. Next, we average V 12 over all orientations of
molecule 2. Finally, we average V 12 over all values of the inter-
molecular separation r. The combination of these three averaging
processes provides us with the potential energy of a given single
molecule as a function of its orientation with respect to the director.
This potential energy, moreover, is that experienced by the molecule
when subjected to the average force fields of neighboring molecules,
each averaged over all its possible positions and orientations-thus the
descriptive phrase, "mean field approximation".
The averaging of V 12 over all orientations of the intermolecular
vector r has an influence only on the Wigner rotation matrices:

[5]

where < > denotes the average value and the SkI are the usual Kron-
icker delta functions. The simple result expressed in Eq. [5] is valid
MEAN FIELD THEORY OF NEMATIC CRYSTALS 49

only if the distribution function for the intermolecular vectors is


spherically symmetric. Since the symmetry of the nematic phase is
cylindrical, this result will only be an approximation. A somewhat
different expression obtains if the cylindrical symmetry is taken into
account;3 we will return to this point later. Use of Eq. [5] in taking
the first average of V 12 gives

< V 12 > = 47T 2: U LLm(r) (2L + 1) -1 YLP(O{, cf>{) Y Lp * (02'cf>/).


Lpm [6]

The averaging of < V 12 > over all orientations of molecule 2 has


influence only on the spherical harmonics in 02',cf>2':

where PI (0 2') is the orientational molecular distribution function that


describes how molecule 2 is distributed among all possible directions
about the director (this function has been described in the previous
chapter 1 and will be considered further below). Because of the
cylindrical symmetry of the nematic phase, PI can be a function only
of the polar angle 0; no cf> dependence is permitted. Then, since PI is
not a function of cf>2', the cf>2' integrals vanish except for p = 0 and

[8]

< PL > = f
o
I

P L (cos 0) PI (cos 0) d (cos 0), [9]

where the P L are the Lth order Legendre polynomials and where Eq.
[9] is the Lth generalization of < P 2 >, the order parameter for the
nematic phase (as described in Ref. [1]). Use of Eqs. [8] and [9]
in taking the average of < V 12 > yields

«V12 » = L,ULLm(r)PL(cos 0 1') < PL > . [10]


Lm

The averaging of «V12 » over all values of the intermolecular


50 CHAPTER 4

separation r has influence only on the VCr) :

[11]

Here n 2 (r) is the molecular distribution function for the separation


of pairs of molecules and n is the number density of molecules in the
fluid. Use of Eq. [11] in taking the average of «V12 » gives

«< V 12 »> = .L: < V LLm (r) > PL(cos B1') < P L > . [12]
Lm

This is just the desired single-molecule potential in the mean field


approximation; we will write it as

V 1 (cos B) =.L: V L < P L > PL(cos B), [13]


L

VL =.L: < V LLm (r) > ,


m

where we have dropped the subscript and prime on the B. Going back
to the discussion following Eq. [2], we assume no preference between
the heads and tails of the molecules and restrict the L sum in Eq. [13]
to even values of L. The L = 0 term, moreover can be discarded, since
it is merely an additive constant. The first few terms of Eq. [13] are
then

V 1 (cos B) = V 2< P 2 > P 2 (cos B) + V 4 < P 4 > P 4 (COS B)


+ V6 < P 6 > P 6 (cos B) +. . . [14]

As shown in Ref. [3], if the true cylindrical symmetry of the distribu-


tion function for the orientation of the intermolecular vector were
taken into account, terms proportional to

[15]

would also appear in the single molecule potential V 1 • If only the first
term in Eq. [14] is retained, we obtain the mean field theory of Maier
and Saupe2 and the equivalent theory of the previous chapter' (v
= -V2 ). Since the original expansion on which V 1 is based (Eq. [2])
converges rapidly, retention of only the first term in V 1 should provide
a good approximation to the theory. Comparison with experiment
MEAN FIELD THEORY OF NEMATIC CRYSTALS 51

shows that this is indeed the case. The necessity for including higher-
order terms is quite apparent in some of the experiments, however,
and we will examine this point shortly.

4. Statistical Thermodynamics

As described in the previous chapter/ the orientational distribution


function corresponding to the single-molecule potential has the form

[16]

f
1

Zl = exp [ -f3VI (cos e) ]d(cos e),


o

where ZI is the single molecule partition function and f3 = l/kT (k is


Boltzmann's constant and T the temperature). The integration over
all possible orientations of the molecule can be restricted to 0 ~ cos e
~ 1, since VI and PI are even functions of cos e. Again, as it stands,
Eq. [16] is not complete and not yet useful in computing average
values. VI does after all contain the as yet unknown averages, < P L >.
The self-consistent determination of the temperature dependence of
the < P L > is realized by combining their definition (Eq. [9]) with
Eq. [16] :

f
1

PL(cos e) exp [-f3V I (cos e) ]d(cos e)


Q

< PL > = ---------------- [17]

f
o
I

exp [-f3V I (cos e) ]d(cos e)

There is one such self-consistency equation for each term L included


in the potential VI' Eq. [13]. In each of these equations one of the
< P L > appears on the left-hand side, and all the included < P L >
appear in the integrals on the right-hand side. The simultaneous
solution of these equations yields the temperature dependence of all
the < P L > originally included in the potential VI. In particular, the
solution for < P 2 > gives the temperature dependence of the tradi-
52 CHAPTER 4

tional order parameter of the nematic phase. One of the solutions


provided by the set of equations, Eq. [17], is < P L > = 0, all L; this
is the is.otropic normal liquid. To find which of all the possible solu-
tions is the physically observed stable solution, we must compute the
free energy and determine the solution giving the minimum free
energy.
The internal energy, entropy, and free energy are obtained in
exactly the same way as computed in the simpler theory of the
previous chapter:

1 1
E=-N<V1 >=-N L:UL <PL >2, [18]
2 2 L

N
S = -Nk < lnpl > =- L: U < P L L >2 + Nk In ZI' [19]
T L

1
F = -NkT In ZI - - NL: UL < PL >2. [20]
2 L

Just as in the case of the simpler theory,' the free energy is found to
include additional terms beyond the usually expected InZl term. The
form is correct, however, and arises because we have approximated
the action of the pair potential V 12 by the temperature-dependent
single-molecule potential V l' Note that setting the partial derivatives
(oF /0 < PI. > h, PI: to zero regains the required self-consistency
equations, Eq. [17]. Furthermore, testing Eqs. [18] and [20], we
see that they do satisfy the required thermodynamic identity, E
= (df3F/dfn·

5. Nature of the Parameters U L


The main physics of the problem, the character of the intermolecular
interactions, is contained in the parameters U L of the one molecule
potential VI' In spite of their central importance in the theory, how-
ever, not all that much is known about their magnitude or volume
dependence. Several attempts to estimate their properties have been
made, but very little in the way of definitive results has been obtained.
What is usually done in practice is to treat the UL as simple volume-
dependent functions with constants determined by fitting the theory
to experimental results.
The first attempt to elucidate the nature of the intermolecular
forces in nematics was performed by Maier and Saupe. 2 They did a
MEAN FIELD THEORY OF NEMATIC CRYSTALS 53

calculation of the London dispersion forces (induced dipole-induced


dipole) expected between asymmetric molecules. Since the polariza-
bilities of elongated molecules are anisotropic, the interactions involv-
ing these polarizabilities must likewise be anisotropic. Indeed, Maier
and Saupe found (within their approximations) an attractive inter-
action between pairs of molecules that led to the form

A
V1 = - - < P 2 > P 2 (cos ()). [21J
V2

Thus, U2 is determined from these calculations to be of the form


_A/V2, where A is a constant to be determined from experiment and
V is the mean molecular volume. The particular volume dependence
should not be surprising; the potential of dispersion forces (whether
between symmetric or asymmetric molecules) is expected to depend
on the intermolecular separation as 1.- 6 . On the average, then, this
translates into the above V-2 dependence.
Chandrasekhar and Madhusudana8 have also considered the calcu-
lation of the coeffidents U L required in V l' The first contribution that
these authors examined was the permanent dipole-permanent dipole
forces. These were shown to vary as r- 3 and provided a V-1 de-
pendence to U 1 • It was shown however, that this term vanished when
the pair potential V 12 is averaged over a spherical molecular distribu-
tion function. The authors thus discard this term and provide further
arguments for its neglect based on the empirical result that permanent
dipoles apparently play a minor role in providing the stability of the
nematic phase. The second contribution considered was the dispersion
forces based on induced dipole-induced dipole interactions and induced
dipole-induced quadrupole interactions. As mentioned above, the first
of these gives a V-2 dependent contribution, while the second provides
a contribution depending on V-BI3. The final contribution considered
was the repulsive forces that arise from the mutual impenetrability of
the molecules. Assuming (by analogy with the behavior of the better
understood, more simple molecules) that the repulsive potential varies
as r- 12 , this mechanism provides a contribution to the U L that varies
as V-4. All these contributions were then lumped together and an
average volume dependence was assumed; the authors chose to allow
the UL to have a V-3 dependence.
Humphries, James, and Luckhurst3 in their development of the
mean field theory decided to allow each of the U L to have the same
volume dependence, V-'Y, where y is a parameter to be determined by
54 CHAPTER 4

fitting theory to experiment. The use of an arbitrary parameter y


gives great latitude in reproducing experimental results and much
success can be expected. The procedure is, however, dangerous. In
the first place, the use of y in the theory provides still another adjust-
able parameter; with enough adjustable parameters, of course, any
experimental observation can be reproduced (but without learning
any physics). Second, the introduction of an arbitrary V-'Y dependence
can lead to abuses. There is one instance in the literature where
experimental data was force fit using a value of y equal to 10. Cer-
tainly, we know of no substantial attractive intermolecular potential
that varies as strongly as r- 30 •
Great care should be used in choosing and using various forms for
the U L in Eq. [13]. The simplest (and possibly least controversial)
procedure is to assume that the U L are simply constants independent
of the volume and to restrict the fitting of theory to experiment to a
narrow range of temperature near Te. Since the volume changes are
small over a short temperature interval, the approximation of constant
UL will be quite adequate and will keep the use of the theory sim'ple
and straightforward. The region near T e, moreover, is the most
interesting; it is here that < P2 > has its strongest temperature de-
pendence.'

6. The Need for Higher Order Terms in VI

As stated in Sec. 3, the retention of only the first term in Eq. [14]
leads to the mean field theory of Maier and Saupe 2 and the equivalent
theory of the previous chapter.' This version of the theory has been
shown to provide a good qualitative picture of the nematic phase and
its transition to the isotropic liquid. What, then, is it about the ex-
perimental facts that indicate the necessity of higher order terms
in VI?
In Fig. 2 we display (schematically) the comparison of several
experiments with the mean field theory, including only the term in U 2 .
We use T - Tc as the abscissa, so that the two materials depicted can
be conveniently compared on the same graph. The dashed line is the
theoretical curve computed from the mean field theory using only the
term in U 2 • We note that there is a striking difference in the proper-
ties of the two materials, P AA (para-azoxyanisole) and PAP (para-
azoxyphenetol). We note further that the simple theory cannot ac-
count for the order parameter of either material. Fig. 2 is, in fact,
clear evidence for the necessity of higher-order terms in the potential.
That the two substances have different values of Tc must mean that
MEAN FIELD THEORY OF NEMATIC CRYSTALS 55

the interaction potentials differ by at least a multiplicative constant.


However, the fact that the order parameter of PAP is everywhere
larger than that of P AA cannot be explained by allowing the interac-
tions to differ only by a multiplicative constant. Such a large differ-
ence in the values of < P2 > can only be accounted for by assuming
that the potentials are in some essential way different in the two
materials. This difference must be a difference in both the sign and
magnitude of the higher-order terms in VI'

1.0

.9

8
(P2>
.7

5 ....................
......... J
4

-40 -30 -20 -10 o


T-Tc

Fig. 2-Schematic representation of the temperature dependence of the


order parameter for PAA and PAP. The dotted line is the result
of the simple mean field theory.

The mean field theory with higher-order terms included is capable


of reproducing the experimental data of Fig. 2. The addition of the
term in U4 , Eq. [14], is found to be sufficient." The data on PAP can
be accounted for by having U4 = +0.12 U 2 , while the data on PAA are
fit by assuming U4 = -0.19 U 2 • In the case of PAP, U4 adds to the
ordering effects of U 2 and < P 2 > is raised above that predicted by the
simple theory. In the case of P AA, on the other hand, U4 decreases
the ordering effects of the U 2 term and a lowered < P 2 > results.
Similar results on these two materials have also been reported by
Chandrasekhar and Madhusudana. 8
The most dramatic evidence that the interaction potential must be
quite complicated comes from recent experiments by Jen, Clark,
Pershan, and Priestleylo on MBBA (N-(p'-methoxybenzylidene)-p-n
butylaniline). Using the Raman scattering technique these authors
were able to determine the temperature dependences of both < P 2 >
and < P 4 >. The experimental data are shown (with their error bars)
56 CHAPTER 4

as the dots in Fig. 2. As is readily apparent, the value of < P4 > be-
comes quite small and then turns negative (!) just before the transi-
tion temperature is reached. The simple version of the mean field
theory is just not capable of handling this situation; the mean field
theory with V 2 only is represented by the solid lines. Note in par-
ticular that < P 4 > is predicted to remain positive right up to the
transition point. The addition of the term in V 4 does not help very
much; the dashed lines represent the results 'O of the theory with V 4
= -0.55 V 2 • While < P 2 > has been brought into agreement with ex-
periment, < P 4 >, though reduced in magnitude, still remains quite
positive.
The physical meaning of a negative < P 4 > has been discussed by
Jen et al. 'O These authors have demonstrated that a negative < P 4 >
implies the existence of a broad orientational distribution function
PI (cos 0), Eq. [16]. In the simple mean field theory (V2 term only),
the distribution function is sharply peaked at 0 = 0; that is, the
probability of locating a molecule whose long axis is at an angle (}
with respect to the director is extremely high for angles near zero and
quite small when the angle 0 becomes appreciable. In the case of
MBBA, however, the probability for finite angles is almost as high as
for 0 = o. It therefore follows that the interaction potential VI re-
quired to describe the behavior of MBBA must be of a form which (at
high temperatures at least) provides a very shallow minimum at 0 = 0
and/or subsidiary minimum at some finite angle. In attempting to
reproduce the experimental results shown in Fig. 3, P. Sheng and I
have tried numerous potential functions that have this feature and
that are compatible with the systematic development, Eqs. [14] and
[15]. In addition to terms in V 2 and V 4 , Eq. [14], we have also in-
cluded terms in «P 2 >P4 +<P4 >P2 ), Eq. [15]. Though it has
been possible to reduce the magnitude of < P 4 > drastically, the de-
sired behavior as depicted in Fig. 3 has not been obtained.
Our most promising attempt at describing the experiments on
MBBA was realized by using the following single molecule potential:

V I (cos 0) =V < P 2 > P 2 (cos 0) + V' < f > f (cos 0),


[221
f (cos 0) = P 2 (cos 0) - 2P4 (cos 0) + P 6 (cos 0).

This potential is obtained by taking very special combinations of terms


from both Eqs. [14] and [15.] ThIS combination was purposely chosen
so that the part in V' gave a strong minimum in orientational potential
energy at an angle far from the director (0 = 52°, in fact). Even with
MEAN FIELD THEORY OF NEMATIC CRYSTALS 57

this potential function (specifically engineered to give the desired


result) we were unsuccessful in explaining the data on MBBA. For
certain values of U'/U we did get negative values of < P4 > of the
right magnitude allright; it just happened, however, that the nematic
phase became unstable with respect to other more complex structures.
The molecular origin of the negative < P 4 > thus remains a mystery.

.7 , - - - - - - - - - - - - - - - - ·

.5

.4

"'-
0.... 3

--- ---
V'

o
_I_ ._. ~:J----~--------,
·01
... j
-02
MBBA

Tc -T

Fig. 3-Temperature dependence of the parameters < P. > and < P. > for
the compound MBBA (data of Ref. [10]). The lines are theoretical
curves described in the text.

A final comment. It may be that the source of the negative <P4>


is not to be found in a complicated form of the potential V 1 but in
improving the mean field theory itself. As is well known from studies
of other cooperative phenomena, the mean field theory is only a first
approximation and theories that attempt to include short-range order
and fluctuation effects always show great improvements in reproducing
the details of experiments. Accordingly P. Sheng and 111 have derived
a theory of the nematic phase based on the well known "constant
coupling" theory of ferromagnetism (see Ref. [6]). Though short-
58 CHAPTER 4

range order is explicitely included and pair-potential functions in-


cluding terms through P 4 (cos ()12) have been used, the temperature
dependence of < P 4 > for MBBA has not been realized. The explana-
tion of the negative < P 4 > in MBBA, therefore, constitutes one of
the outstanding theoretical problems of the nematic liquid-crystal state.

References
1 P. J. Wojtowicz, "Introduction to the Molecular Theory of Nematic Liquid Crystals,"
Chapter 3.
2 W. Maier and A. Saupe, Z. Naturforschg., Vol. 14a, p 882 (1959), and Vol 15a, p. 287
(1960).
3 R. L. Humphries, P. G. James and G. R. Luckhurst, "Molecular Field Treatment of
Nematic Liquid Crystals," J. Chem. Soc., Faraday Trans. II, Vol. 68, p. 1031 (1972).
4 J. A. Pople, "The Statistical Mechanics of Assemblies of Axially Symmetric Molecules.
I. General Theory," Proc. Roy. Soc., Vol. A221, p. 498 (1954).
5 J. R. Sweet and W. A. Steele, "Statistical Mechanics of Linear Molecules. I. Potential
Energy Functions," J. Chem. Phys., Vol. 47, p. 3022 (1967).
6 J. S. Smart, Effective Field Theories of Magnetism, W. B. Saunders Co., Phila., Pa.
(1966).
7 M. E. Rose, Elementary Theory of Angular Momentum, J. Wiley and Sons, N. Y., N. Y.
(1957).
8 S. Chandrasekhar and N. V. Madhusudana, "Molecular Statistical Theory of Nematic
Liquid Crystals," Acta Cryst., Vol. A27, p. 303 (1971).
9 G. R. Luckhurst, unpublished notes.
10 S. Jen, N. A. Clark, P. S. Pershan, and E. B. Priestley, "Raman Scattering from a
Nematic Liquid Crystal: Orientational Statistics," Phys. Rev. Lett., Vol. 31, No. 26, p.
1552 (1973).
lIP. Sheng and P. J. Wojtowicz, to be published.
Hard Rod Model of the Nematic-Isotropic
Phase Transition

Ping Sheng
RCA Laboratories, Princeton, N. J. 08540

1. Introduction

In the previous chapters we have seen how an anisotropic, attractive


interaction between the molecules of the form P2(COS 812 ) can give rise
to a first-order nematic-isotropic phase transition. The origin of the
anisotropy lies in the fact that almost all the liquid-crystal molecules
are elongated, rod-like, and fairly rigid (at least in the central portion
of the molecule). It is clear, however, that besides the anisotropic
attractive interaction there must also be an anisotropic steric inter-
action that is due to the impenetrability of the molecules.
It is natural to ask what effect, if any, the steric interaction might
have on the nematic-isotropic phase transition. Onsager recognized
that a system of hard rods, without any attractive interaction, can
have a first-order transition from the isotropic phase to the anisotropic
phase as the density is increased. ' -s To see how this can come about,
we note that in a gas of hard rods there are two kinds of entropy_
One is the entropy due to the translational degrees of freedom, and the
other is the orientational entropy. In addition, there is a coupling
59
60 CHAPTER 5

between these two kinds of entropy that can be described as follows.


When two hard rods lie at an angle with respect to each other, the
excluded volume (the volume into which the center of mass of one
molecule cannot move due to the impenetrability of the other molecule)
is always larger than that when the two hard rods are parallel. This
fact can easily be illustrated in two dimensions. For simplicity let us
take two lines of length L. When they are perpendicular to each
other, the excluded area is L 2. When they are parallel, the excluded
area vanishes. This is shown in Fig. 1. It is only a short step to

.
£XCLUO£D
AREA ' 0
r

:.. . . . __
l-

. ... _-- .

(0 1 (bl

Fig. i-Excluded area of two lines. In (a), the excluded area between
two perpendicular lines is indicated by shade. In (b), the lines
are parallel, and the excluded area is zero.

visualize the same situation in three dimensions where the excluded


area is now replaced by excluded volume. The importance of excluded
volume is that the translational entropy favors parallel alignment of
the hard rods because this arrangement gives less excluded volume
and, therefore, more free space for the molecules to jostle around.
However, parallel alignment represents a state of low orientational
entropy. Therefore, a competition exists between the tendency to
maximize the translational entropy and the tendency to maximize the
orientational entropy.
In the limit of zero density the tendency to maximize the orienta-
tional entropy always wins because each molecule rarely collides with
another molecule, and the gain in excluded volume due to parallel align-
ment would only be a minimal addition to the already large volume of
space within which each molecule can move about. When the density
is increased, however, the excluded volume effect becomes more and
more important. We know that in the limit of tight-packing density,
the hard rods must be parallel. A transition between the isotropic and
the anisotropic states therefore must occur at some intermediate den-
sity. But this kind of general argument cannot tell us whether the
transition is smooth or abrupt. To be more precise, we will derive the
Onsager equations and actually solve them in a simplified case.
HARD ROD MODEL 61

2. Derivation of Onsager Equations

Consider a classical system of N hard rods in a volume V. The parti-


tion function Z, including the angular part, is 6

x f·. ·fd3NXd3NP exp {-[f


i~l
(_PO_i2
211
+ _C_P_"'i_-_PI/1_i_C_os_B_i_)2_
211 sin2Bi

[1]

Here k is the Boltzmann constant, T is the temperature, h is Plank's


constant, B, cp, if! are the usual Euler angles Cfollowing the notation
used in Goldstein, Classical Mechanics), Po, P"" PI/1 are the canonical
conjugate angular momenta, Iv 12 are the two principal moments of
inertia of a rigid rod, m is the mass of the rod, Vij is the interparticle
steric interaction, and Pa:' PY ' Pz are the linear momenta of the hard rod.
The kinetic energy part of the integral can be immediately integrated.
Integration over the translational momenta yields a term C27rmkT) 3N /2.
Integration over the angular momenta gives a term

N
C27rI 1 y27rI2 )N II sin Bi
i~1

Since nothing in the integrand depends on if!, this angle can also be
integrated to yield (27r) N. Grouping the constants and combining the

N
II sin Bi
i~1

factor with dNB yields

ZCN,V,kT) = - - - - - - - - - - - - -
N!h 6N
62 CHAPTER 5

[2]

where 'f/ = n3 /[kTI 12 / 3 (271'/ 2 ) 1/3]3/2 is a dimensionless constant (n


= h/27!') , ,\ = h/v'27!'mkT is the thermal wavelength, and n is the solid
angle. The angular integrals in Eq. [2] can be approximated to arbi-
trary accuracy by summations in the following manner. Divide the unit
sphere into K cells, each containing a solid angle t.w. The orientational
distribution of the rods is specified by the integers N a' which give the
number of rods oriented in the direction of ath angular cell. The
angular integrals can then be replaced by sum over all the possible
partitions {N1' N 2 , ••• N a, ... N K} with N1 + N2 + ... + NK = N,
multiplied by a factor [N!/(N]!N2! ... N K!)] for the number of
times the partition {N 1 , ••• N K } has been counted in the angular in-
tegrals' ;

(3a)

= L Z ({N a }, N, V, kT), [3b]


{Na}

where Z({N a}, N, V, kT) = __ (t._W_)_N_ _ f ...fd 3N X emp {- t Vii}


K i<ikT
'f/ N ,\3N IT (Nat)
a=1

Let us pick the maximum term in the summation

:L: Z({Na},N,V,kT) andlabelitZ({Na})max.


{N a }

Since the summation can have at most NK terms, we have the follow-
ing inequality:

Z({Na})max <ZeN, V,kT) <NKZ ({Na})ma",. [4]


HARD ROD MODEL 63

Taking the logarithm and dividing through by N, we have

1 1 1
-In Z ({N a} )ma., < -In Z <- (In Z ({N a } )maa: + KIn N). [5]
N N N

As N -,) 00, the term (lnN)/N approaches zero and (lnZ)/N can be
exaotly replaced by [lnZ({Na})max]/N. Therefore, our task can be re-
duced to the calculation of Z ({N a}, N, V, kT) as long as we remember
to maximize the result with respect to {N a} afterwards.
Now Z ({N a }, N, V, kT) can be put in the following form:

1
Z({NaJ,N, V,kT) = - -
'Y}N)...3N
K
(LlW)N J J••• d 3N X exp {
-
II (N a !)
a=l

[6]

Define Cf>ij == exp {- vi / (kT)} - 1. With Vij denoting the steric


repulsion, Cf>ij has the value zero when the two molecules are not in
contact but takes the value -1 when the center of mass of one mole-
cule enters the "excluded volume" of two hard rods. With this defini-
tion of Cf>ij we can rewrite the integrand as

[7]

The first term represents the behavior of the ideal gas, and the
second term is the first-order correction to this behavior due to steric
repulsion between the molecules. If we limit ourselves to densities
that are not too high, we can neglect higher-order terms, which rep-
resent the effects of excluded volume between more than two molecules.
With this crucial approximation, we have
64 CHAPTER 5

[8]

We note that <I>i; depends on the relative separation Xi; and the relative
orientation between two molecules. Therefore, a better way to label
the indices should be <I>all (Xi;), where a and f3 denote the angular
orientations of the two molecules and Xij the relative separation be-
tween their centers of mass. The summation over i and j can be
rewri tten as

As N ~ 00, the Ball term can be neglected. The integral of <I>all(x) is


just the negative of excluded volume Valx" Z({N a}, N, V, kT) is now
in the form

[10]

Taking the logarithm, dividing by N, and using Sterling's approxima-


tion InN a! = N alnNa' we obtain

In[Z({Na}, N, V, kT)] (,:lw) V K (Na Na Na )


- - - - - - - - - = In - I: - In - + - InN
N TJA 3 ~=1 N N N

1 [ N2 K NaNII ]
+-In 1 - - I : - - vall ex.
N 2V a,1I N N

(.:lW) V] K Na Na [ N2 K Na Nil ]1/N


=In [ - - - - I : -In-+ln 1 - - I : - - Vall ex.
TJA 3 N a=1 N N 2V a,1I N N

(.:lW) V] K Na Na [ N K Na Nil ]
= In [ - - - - I: - In - + In 1 - - I: -- - Vallex.
TJA 3 N a=l N N 2Va,1I N N

(.:lw) K Na Na P K NaNII
= I n - - - I: - I n - - - I : --Vallex .. [11]
TJA 3 p a=1 N N 2 a,1I N N
HARD ROD MODEL 65

Here p == N IV is assumed to be small.

Let us define an angular distribution function 1 such that N a


= N· I(n a) • (.!loo). Substituting NalN by IWa) and making the
replacement

LC~oo)~ / dO"
a

we can put Eq. [11] in the form

~ InZ( {Na}, N, V, kT) = -In TJ)..3 p - / dn/(n) In f(n)

- : // dndn'/(n)/(n') Vex. (0, - 0,'). [12]

The second term on the right-hand side of Eq. [12] represents the
orientational entropy, and the third term represents the effect of
excluded volume. In order to maximize the right hand side of Eq. [12]
with respect to 1(0,), we will use the Euler-Lagrange equation. This
equation says if one wants to maximize the integral

by varying the function y (x), the correct y(x) must satisfy the
equation

of d of
------=0.
oy
dx 0(::)
If in addition y(x) must satisfy the normalization condition

J
fJ
b

y(x)dx = 1, then the equation for y(x) becomes


66 CHAPTER 5

of d of
-------v==o,
oy
dx 0(::)
where v is the Lagrange multiplier. Applying the Euler-Lagrange
equation to Eq. [12] gives

In 1(0) +1+v+p f 1(0') Vex. (0 - 0') do' = O. [13]

Eqs. [12] and [13] are the Onsager equations. In principle, Eq. [13]
can be solved to get 1(0) as a function of p and v. The quantity v is
then determined by the normalization condition J I (O)dO = 1. When
there is more than one solution to Eq. [13], the results must be put
into Eq. [12] in order to select the one solution that maximizes the
right-hand side of Eq. [12]. Label that solution 10 (0). The free
energy is then

~ = In "'A3 p
NkT
+f do/o(o) lnlo(n) + ~ff
2
dOdo'lo(n)/o(o') Vex·(n - 0'). [14J

Other thermodynamic quantities can be obtained directly from Eq. [14].

3. Solution of Onsager Equations in a Simplified Case

In practice, Eq. [13] is a nonlinear integral equation and its exact


solution in the general case is difficult. Therefore, we will solve the
Onsager equations in a very simplified case. Let us constrain the
hard rods to point in only three orthogonal directions, say x, y, and z.
Choose z as the preferred direction. If the fraction of hard rods
pointing in the x-direction is r, the same fraction should point in the
y-direction, since x and y directions are equivalent by uniaxial sym-
metry of the nematic phase. The fraction pointing in the z-direction
is then 1 - 2r. The order parameter

S = f
1

P 2 (cos ()) I (cos ()) d (cos ())


HARD ROD MODEL 67

in this case is simply (1 - 3r). Eq. [12] can now be written

Zr(N, V, kT)
- - - - - = -In 'Y}A. 3 p - 2r lnr - (1 - 2r) In (1 - 2r)
N
p
- pr (2 - 3r) V ~ ex - - (1 - 4r + 6r2 ) V{X
2

'Y}A. 3 p 1 2
= -In - - - - (1 + 28) In (1 + 28) - - (1 - 8) In (1 - 8)
3 3 3

pCl - 8 2 ) p 1 + 28 2
V ~ex _--vllex [151
3 2 3

Fig. 2-Graphical solution of Eq. [16].

Here V ..L ex and vllex are the excluded volumes of two hard rods when
they are respectively, perpendicular and parallel to each other.
To maximize Zr (N, V, kT) IN, we merely differentiate with respect
to 8 and set the result equal to zero. This gives

1 +28
In = 2p [V ..L ex - V{X ] 8, [16]
1-8
68 CHAPTER 5

which is the equivalent of Eq. [13]. In order to calculate vex., we wiII


take the rods to have the shape of rectangular parallelepiped with
length L and width D. Then

(
1.0

075

0.50 I
1
I
0.25
\,
I
o ;,~-------------------------------

-025 :\
: '....
"-
-050

139 2 4 6 9 10

Fig. 3-Solution of Eq. [16] as a function of p[V"ex - Vllex]. Dark lines


indicate the branch that maximizes the right hand side of Eq. [15].

where V o = D2L is the volume of a parallelepiped and l = LID is the


ratio of length to breadth. Eq. [16] can be solved graphically by
plotting the right- and left-hand sides on the same graph and locating
the points of intersection. This is shown in Fig. 2. The results are
displayed in Fig. 3 as a function of p[V ~ ex - V{x]. Dark lines indi-
cate the stable solution. Label these values of S by So (p). The free
energy is

-F- = In [N'l)..3]
-- +-
1
(1 + 2So(p)) In (1 + 2S0(p))
NkT 3V 3
2
+- (1 - So (p)) In (1 - So (p ) )
3

NVo [I-(SO(P))2(
+2--
1 )
l+--2 +2 .
J [18]
V 3 l
HARD ROD MODE L 69

/.8 ). • 5
I!.P •. 07
liS •. 6443

ISOTROPIC
PHASE
/
14

13

21 22 2.3 2.4 2.5

.l
NVo

Fig. 4-Reduced pressure PV.lkT plotted as a function of reduced volume


VIV.N for l = 5.
The pressure is obtained by differentiating F with respect to V:
PVo Vo of
-=----=--+2 --
VoN (V oN)2[1- CS OCP)F ( 1 )
l+--2 +2 ,
1
kT kT oV V V 3 l
[19J
where we remember that aF /aSO(p) = 0 by definition of SO(p).
In Figs. 4 and 5, reduced pressure PVo/kT is plotted versus reduced
volume V / VoN for two values of l. It should be noted that there is a

PVo
IT
,( • 10
.45 I!.P •. 15
liS •. 7361
1\
1\
44 I \
I \
I \
I \
43 I \
I \

/
I \
I \
.42 I \
ANISOTROPIC I \
PHASE ,, I
.41 ,, I
I
,,
--_ .... ~

.40

4.5 50 5.5 6.0 65


V
NVo

Fig. 5-Reduced pressure PV.lkT plotted as a function of reduced volume


V IV.N for l = 10.
70 CHAPTER 5

van der Waals loop in each curve. Using the standard Maxwell
equal-area construction, we obtain the magnitudes of density and
order parameter discontinuities (Llp and LlS) across the transition.
Their values are typically Llp - 0.1 and LlS - 0.7, which are close to
the results obtained by the variational solution of the exact Onsager
equations. How well do these values agree with the experimental
results? Typical experimental values for Llp and LlS are - .01 and
.4, respectively. Therefore, there exists a large discrepancy between
the theoretical predictions and the experimental results.
What went wrong? The answer probably lies in the neglect of
short-range order. In real liquid crystals there are short-range at-
tractive interactions that favor parallel alignment of the molecules.
This means that parallel molecules prefer to lie close to each other
and form "bundles". In this kind of configuration, a molecule would
most probably interact with those molecules pointing in similar orien-
tations, thus reducing the excluded-volume effect. A lattice-gas model
calculation 7 that includes the short-range anisotropic attractive inter-
action as well as the steric repulsion has shown that "bundling"
would indeed soften the isotropic-anisotropic phase transition and de-
crease the values of Llp and LlS. Therefore, the joint consideration of
short-range order and steric repulsion seems to be the direction for
improving the hard-rod model.

References
1 L. Onsager, "The Effects of Shapes on the Interaction of Colloidal Particles," Ann.
N.Y. Acad. Sci., Vol. 51, p. 627 (1949).
2 R. Zwanzig, "First-Order Phase Transitions in a Gas of Long Thin Rods," J. Chem.
Phys., Vol. 39, p. 1714 (1963).
3 P. J. Flory, "Phase Equilibriums in Solutions of Rodlike Particles," Proc. Roy. Soc.
(London), Vol. A234, p. 73 (1956).
4 E. A. DiMarzio, "Statistics of Orientation Effects in Linear Polymer Molecules," J.
Chem. Phys., Vol. 35, p. 658 (1961).
5 G. Lasher, "Nematic Ordering of Hard Rods Derived from a Scaled Particle Treatment,"
J. Chem. Phys., Vol. 53, p. 4141 (1970).
6 R. H. Fowler, Statistical Mechanics, p. 62, Cambridge University Press, Cambridge,
England (1966).
7 P. Sheng, "Effects of Bundling in a Lattice Gas Model of Liquid Crystals," J. Chem.
Phys., Vol. 59, p.1942 (1973).
Nematic Order: The Long Range Orientational
Distribution Function

E. B. Priestley
RCA Laboratories, Princeton, N. J. 08540

1. Introduction

Earlier, l in Chapter 3, the existence of the nematic phase was presented as an


example of an order-disorder phenomenon. The symmetry and structure of
the nematic phase were used to identify the natural order parameter <P2 (cos
0», where P 2 (cos 0) is the second Legendre polynomial and the angular
brackets denote a statistical average over the orientational distribution func-
tion {(cos 0). The orientational potential energy of a single molecule was
shown, in the mean field approximation, to be

v (cos ()) = -v <P2 (cos ())> P 2 (cos ()) [1]

where v is a number that scales with the strength of the intermole-


cular interaction. Using the rules of classical statistical mechanics,
the theoretical orientational distribution function p (cos ()) was given
71
72 CHAPTER 6

in terms of the mean field potential V (cos 8) as

p (cos 8) = Z-l exp [-f3V (cos 0)] [2]

with

f
1

Z = exp [-f3V (cos 8)] d (cos 8).


o

In Eq. [2] Z is the single-particle partition function and f3 = l/kT,


where k is Boltamann's constant and T is the absolute temperature.
It is important at this point to distinguish between f (cos 8) and
p (cos 8). Given that the nematic phase is truly uniaxial, f (cos 8)
completely describes the long range orientational molecular order
and therefore can be regarded as the exact or true single-particle
distribution function. On the other hand, p (cos 8) is a theoretical
approximation to f (cos 8). The extent to which it represents a good
approximation depends upon how closely V (cos 8) approximates the
true potential of mean torque.
There is an obvious motivation for determining as much as pos-
sible experimentally about f (cos 8); it describes the long range
orientational ordering of the molecules which, of course, is the single
feature that distinguishes the nematic phase from the isotropic
phase. The average values of various quantities of interest pertaining
to the nematic phase could be computed if f ( cos 8) were known with
precision.

2. The Orientational Distribution Function

The main question we want to answer in this chapter is "What do


experiments tell us about the orientational distribution function ?".
To answer this question it is instructive to expand f (cos 8) in a
formal mathematical sense as'

2l +1
f (cos 8) = L: -- < PI (cos 8) > PI (cos 8), [3]
I 2
even

where the PI (cos 8) are the ltn even order Legendre polynomials.
Notice that the expansion has the correct symmetry and that f (cos 8)
is normalized. The coefficients <PI (cos 8) > are defined by
NEMATIC ORDER 73

<PI (COS ())> =J'P I (COS ()) I (COS ()) d (COS ()). [4]

-1

Explicitly, the first three coefficients are

<Po (cos ()) > = 1,


1
<P2 (cos ())> = - (3 <cos 2 (» -1), [5]
2
1
<P4 (cos ()) > = - (35 <cos 4 (» - 30 <cos 2 (» + 3).
8

There is no practical experimental means of measuring I (cos ()


directly. Rather, individual experiments each detect some property
of the nematic phase averaged over I (cos ()). Thus, experiments are
sensitive to I (cos ()) only through its influence on these averaged
properties. The most commonly measured such property is the tem-
perature dependence of <cos 2 (» and, hence, the temperature depend-
ence of the order parameter <P2 (cos B) >. Recent measurements of
<cos 4 (» as a function of temperature for two nematic materials' have
permitted the evaluation of <P 4 (cos ()) > over the entire nematic
temperature range as well. Therefore, the answer to the question
posed at the beginning of this section is that experiments measure
the coefficients in the expansion of I (cos ()), (Eq. [3]). In other words,
experiments measure different moments of the orientational distribu-
tion function. These experimentally determined moments can be
used to evaluate truncated series approximations to I (cos ()). It is
convenient to label these approximate distribution functions as IN (cos
()) where N is the Roman numeral corresponding to the number of
terms retained in the truncated series.
In the discussion so far, the constituent molecules of the nematic
phase have been assumed to be representable by simple rigid rods so
as to make the definition of a molecular axis unambiguous. The order
parameter <P2 (cos ()) > has been defined from a microscopic point of
view by treating it as a statistical average of individual molecular
behavior. In most cases this is an adequate description since real
nematogens most often do behave like simple rigid rods. However, if
the molecules do not behave like simple rigid rods, this microscopic
description of order is no longer adequate and we must find some other
74 CHAPTER 6

means for specifying the degree of order. For this, we turn to a con-
sideration of macroscopic response functions of the nematic phase and
show how a suitable order parameter can be defined using these
response functions.

3. Macroscopic Definition of Nematic Order

In deriving a macroscopic order parameter, we will use the diamagnetic


susceptibility as an example. However, any other macroscopic property,
e.g., refractive index or dielectric response, could be used as well.
-->
Consider the relationship between magnetic moment M and magnetic
->
field H

a,{3 = X,Y,Z • [6]

XafJ represents the a{3-component of the diamagnetic susceptibility


<--->
tensor X and, for static fields, XafJ = XfJa. In Eq. [6], the summation
convention over repeated indices is implied. For the uniaxial nematic
<--->
phase, we can write X in the diagonal, but completely general form

0)o , [7J
XII

where XII and X~ refer to the susceptibility parallel and perpendicular


<--->
to the symmetry axis, respectively. X itself is not a useful order para-
meter because it does not vanish in the isotropic phase, but has the
value

1
X= - (2x~ + XII)
3
[8]
1
=- X'Y'Y·
3

<---> <--->
However, if we extract the anisotropic part of X, viz Xa defined by

1
XaafJ = XafJ - - X'Y'Y SafJ' [9]
3
NEMATIC ORDER 75

which does vanish in the isotropic phase, we can use it to define a suit-
able order parameter.
Combining Eqs. [7] and [9] we find for the nematic phase

~ 2 (-1/2 0
Xa = -b.X 0 -1/2 [10]
3 0 0

where b.x = XII - x~· It should be noted that xa is temperature de-


pendent because the magnitude of ~x is temperature dependent. If
we denote the maximum possible diamagnetic anisotropy (i.e., the
anisotropy for a perfectly ordered nematic phase) as (~X)max' we can
define the desired order parameter as

3 Xaa(3
Qa(3=----- [11]
2 (~X) max

Notice that Qa(3 has values between 0 (isotropic phase) and 1 (per-
fectly ordered nematic phase) and that it has been derived without
any assumption about molecular rigidity. Qa(3 is a valid measure of
the order for any nematic liquid crystal. As mentioned earlier, Q a(3
could have been defined in terms of the anisotropy in any other macro-
scopic response function.
One might wonder if there is any relationship between Qa(3 and
<P2 (cos B». Indeed, in nematic phases whose constituent molecules
can be approximated by simple rigid rods, these two measures of the
long range order are related. The next section deals explicitly with
this relationship.

4. Relationship Between Microscopic and Macroscopic Order


Parameters

The diamagnetic susceptibility I; of a single rod-like molecule can be


described by two parameters 1;11 and I;~, representing the molecular
response to a magnetic field applied parallel and perpendicular to the
molecular axis, respectively. In tensor notation the magnetic sus-
ceptibility of the molecule is given by

[12]
76 CHAPTER 6

To find the macroscopic diamagnetic anisotropy we first rotate, by an


arbitrary rotation R(c/>, 0, t/;), where C/>, 0, and t/; are the Euler angles of the
rotation, and then perform a statistical average. In a uniaxial nematic,
the angles cp and rf; are isotropically distributed so averages over them
can be carried out explicitly. The orientational distribution function
f (cos B) describes the distribution of the angle B, and averages over B
are designated by angular brackets.
From the general formalism for rotation of second rank tensors'
it can be shown that

1 1
,,,,,,,ROT = - (2~..1. + 'II) + - (~II - , 1.) [3 (l-cos 2 B) cos 2rf; - (3 cos2 0-1)].
3 6
and [13]
1 1
'zzROT = - (2,..1. + 'II) + - ('II - 'J) (3 cos 2 B-1),
3 3

where 'il oT are components of the rotated molecular diamagnetic sus-


ceptibility tensor. The macroscopic diamagnetic anisotropy is then

[14]

where N is the number of molecules per cubic centimeter. Taking the


averages of the 'iloT and substituting into Eq. [14] gives

[15]

But N ('II - '..1.) is simply the macroscopic diamagnetic anistropy of


the perfectly ordered nematic phase, i.e.,

[16]

Thus,

[17]

Combining Eqs. [10], [11], and [17] yields

Qaf3 = <P2 (cos B» [


-1/2
0 -1/2
0 001 [18]
o 0 1

for a nematic liquid crystal composed of rod-like molecules.


NEMATIC ORDER 77

As noted at the beginning of Section 3, QafJ can be defined equally


well in terms of other macroscopic response functions such as the
~

dielectric tensor £. However, a theoretical relationship between ~£


(macroscopic measure of the order) and <P2 (cos ()> (microscopic
measure of the order) cannot be derived in the absence of certain ques-
tionable assumptions. This inability to establish a rigorous relationship
analogous to Eq. [14J between the dielectric anisotropy ~£ and the
polarizability ll'ij results from complicated depolarization effects caused
by the relatively large near-neighbor electrostatic interaction. Never-
theless, it has been established empirically5 that the analog of Eq. [17J
does hold, i.e., the macroscopic anisotropy does scale directly with the
microscopic order parameter <P2 (cos () >. Consequently, electrical
and optical anisotropy measurements continue to be used to measure
<P2 (cos ()> in nematics composed of rod-like molecules, even though
theoretical justification for this is presently lacking.

5. Experimental Measurements

Various experimental methods have been used to examine molecular


ordering in nematic liquid crystals". All the methods fall into one of
two groups-those that measure the anisotropy in some macroscopic
response function and those that measure <P2 (cos () > directly. As
we have already seen, those in the first group also permit a determina-
tion of <P2 (cos () > because of equations analogous to Eg. [17J that
relate the macroscopic anisotropy to the microscopic order. In general,
there is quite good agreement among the various measures of
<P2 (cos () > for materials that have been studied by one or more
methods from each group, indicating that nematogenic molecules can
be approximated rather well by hard rods. We turn now to a brief
description of the various experimental techniques and refer the in-
terested reader to more detailed discussions in each instance.

5.1 Measurements of <P2 (cos 9» Based on Macroscopic Anisotropies

Little need be said concerning this type of order parameter deter-


mination. It ,is evident from Eq. [17J that the only requirement dn
evaluating <P2 (cos ()> from macroscopic anisotropy measurements
is a knowledge of the maximum possible magnitude of the anisotropy
being measured. Clearly, this is the anisotropy that would be measured
in a perfectly ordered nematic phase and can be obtained by means of
a little algebra from the crystallin{! anisotropy. Diamagnetic, dielectric,
and optical anisotropy measurements have been used extensively to
78 CHAPTER 6

determine the temperature dependence of <P2 (cos () > for a variety


of nematic liquid crystals. Further details about the conduct of these
experiments and analysis of the data can be found in References [6]
and [7].

5.2 Measurements of <P 2 (cos 8» Based on Microscopic Anisotropies

In this section we group techniques that sense the anisotropy in


various properties of individual molecules as opposed to those discussed
above, which sense bulk anisotropies. It is impossible to measure the
anisotropy of a single molecule in a nematic phase; rather, these tech-
niques measure the statistical average (both temporal and spatial) of
the molecular anisotropy. The averaging process results in the mea-
sured anisotropy being proportional to <P2 (cos () > for most of these
techniques. For the technique based on Raman scattering measure-
ments, the anisotropy is related to both <P2 (cos () > and <P4 (cos () >.

(a) Magnetic Resonance Techniques

An excellent review of magnetic resonance methods has been given


recently by Brown et al,8 We simply point out that, for molecules hav-
ing nuclei with spin 1=1/2 (such as H), the dipole-dipole splitting
in the NMR spectrum of the nematic phase, due to a proximal pair
of such nuclei in the same molecule, is proportional to <P2 (cos ()>.
In addition, for molecules having nuclei with spin I ~ 1 (such as D and
14N), the quadrupole splitting can be used to determine <P2 (cos () >.
Finally, anisotropy in the Zeeman and hyperfine splitting observed in
the EPR spectrum of free radicals dissolved in nematic liquid crystals
also allows one to determine the order parameter, assuming the para-
magnetic guest aligns sympathatically with the host nematic material.

(b) Raman Scattering Technique

Details of the Raman scattering technique have been published else-


where." Successful utilization of the technique requires the existence
of a (preferably) strong, narrow, anisotropic Raman line associated
with an identifiable vibration of the nematogen. It is straight-
forward to show that the anisotropy in the Raman scattering from
such a vibration observed in the laboratory frame of reference con-
tains information about the statistical averages <P2 (cos () > and
<P4 (cos () >. This unique feature of the Raman technique, viz. the
ability to determine <P4> as well as <P 2 >, stems from the fourth
rank tensor nature of the scattering interaction.
NEMATIC ORDER 79

Having three moments <Po>, <P2 > and <P4 > of the orientational
distribution function allows us to evaluate fIII (cos 0) (See Sec. 2). In
the next section we present NMR, optical and magnetic anisotropy,
and Raman measurements of <P2 (cos 0», Raman measurements of
<P4 (cos 0», and representative plots of fIll (cos 0).

0.7 .---"'T""--,--.---"'T""--,--.------,

0.6

0.5

0.4

0.3

-- ---- --
/\

----- ------
o!
V
·0.2
/\
0....
V

0.1
1 • ...............

0

-0.1
· .. j
-0.2

-0.335
30 2S 20 15 10 S 0
Tc-T (OC)

Fig. I-Theoretical and experimental values of the nematic order para-


meters <P. (cos 9) > and <P. (cos 9) >: solid line, theoretical re-
sults of simple mean field theory9; dashed line, HJL theorylo;
crosses, Onsager-Lakatos theoryl3; filled circles, Raman measure-
ments; open circles, NMR data on partially deuterated MBBA;
squares, relative values obtained from measurements of the optical
anisotropy; and triangles, relative values obtained from measure-
ments of the diamagnetic anisotropy.? (Reprinted from Ref. [3].)

6. Experimental Data

In Fig. 1 we display the data of Reference [3] . Notice that the


<P2 > results from NMR, optical and diamagnetic anisotropy, and
Raman measurements all agree very well. This is reasonably com-
pelling evidence that all of these techniques, and the Raman technique
in particular, are indeed measuring <P2 (cos 0». The <P4 > data,
80 CHAPTER 6

which come only from the Raman measurements, are seen to become
negative near the nematic-isotropic phase transition. The physical
meaning and implications of negative <P4 > have been discussed by
Jen et ala and by P. Wojtowicz9 in Chapter 4.

3.0

.
o
2

o
o
o

o L-______~~o~o~o~~~====i-------------==~~==~
0000000

-0.31.0 0.8 0.6 0.4 0.2 1.0 0.8 0.6 0.4 0.2 0
cos 8 cos 8

Fig. 2-Plot of the theoretical and experimental truncated orientational


distribution function fIll (cos 0); solid line, HJL theory;10 circles,
Raman measurements. (Reprinted from Ref. [3].)

Fig. 2 shows a plot of fIlI (cos (j) for two temperatures using the
experimental Raman values of <P2 > and <P4 > in Eq. [3]. For com-
parison, we also plot fllI (cos (j) for the same two temperatures using
values of <P2 > and <P4 > calculated from the Humphries-James-
Luckhurst model 9,lO after the parameters were adjusted to obtain good
agreement with the <P2 > data. Note that fIll need not be positive
definite because of truncation errors. The principal result to see in
Fig. 2 is that the molecules have a stronger tendency to be tipped
away from the nematic axis than is predicted by mean field theory.
This tendency is strongest near the nematic-isotropic phase transition.
In addition to mean field theories, other statistical mechanical models
of nematic ordering have been developed. In particular there is the
On sager model",12 for the ordering of a system of hard rods as a
function of density. Using this model, Lakatos 13 has recently cal-
culated values of <PI (cos (j) > for all l. In Fig. 1 we also show these
results for <P 2 > and <P4 > at the transition density. The ratio of
<P4 >I<P2 > as a function of <P2 > is essentially the same as that
calculated from mean field theory and it also disagrees with the ex-
perimental results.
There has been considerable speculation3,9 regarding the origin of
NEMATIC ORDER 81

the discrepancy between experiment and mean field theory. As of this


writing, that discrepancy remains unexplained, though extensive work
is underway on both the experimental and theoretical fronts in an
attempt to better understand the exact nature of nematic ordering and
the single-particle orientational distribution function f (cos ().

References

1 P. J. Wojtowicz, "Introduction to the Molecular Theory of Nematic Liquid Crystals,"


Chapter 3.
o E. B. Priestley, P. S. Pershan, R. B. Meyer, and D. H. Dolphin, "Raman Scattering
from Nematic Liquid Crystals. A Determination of the Degree of Ordering," Vijnana
Parishad Anusandhan Patrika, Vol. 14. p. 93 (1971).
3 Shen Jen, N. A. Clark, P. S. Pershan, and E. B. Priestley, "Raman Scattering from a
Nematic Liquid Crystal: Orientational Statistics," Phys. Rev. Lett., Vol. 31, p. 1552 (1973);
E. B. Priestley and P. S. Pershan, "Investigation of Nematic Ordering Using Raman
Scattering," Mol. Cryst. Liquid Cryst., Vol. 23, p. 369 (1973); and E. B. Priestley and
A. E. Bell (to be published).
4 E. B. Priestley and P. J. Wojtowicz, unpublished results.
5 N. V. Madhusudana, R. Shashidhar, and S. Chandrasekhar, "Orientational Order in
Anisaldazine in the Nematic Phase," Mol. Cryst. Liquid Cryst., Vol. 13, p. 61 (1971).
6 A. Saupe and W. Maier, "Methods for the Determination of the Degree of Order in
Nematic Liquid·Crystal Layers," Z. Naturforschg., Vol. 16a, p. 816 (1961).
7 G. Sigaud and H. Gasparoux, J. Chern. Phys. Physicochim. BiOi., Vol. 70, p. 669
(1973); I. Haller, "Elastic Constants of the Nematic Liquid Crystalline Phase of
p-Methoxybenzylidene-p·n-Butlylaniline (MBBA)," J. Chem. Phys., Vol. 57, p. 1400 (1972);
P. I. Rose, Fourth International Liquid Crystal Conf., Kent State University, Kent, Ohio,
1972 (to be published) see also Ref. (5).
8 G. H. Brown, J. W. Doane, and V. D. Neff, A Review of the Structure and Physical
Properties 01 Liquid Crystals, The CRC Press, Cleveland, Ohio (1971).
9 P. J. Wojtowicz, "Generalized Mean Field Theory of Nematic Liquid Crystals,"
Chapter 4.
10 R. L. Humphries, P. G. James, and G. R. Luckhurst, "Molecular Field Treatment of
Nematic Liquid Crystals," J. Chem. Soc. Faraday Trans. II, Vol. 68, p 1031 (1972).
11 L. Onsager, "The Effects of Shapes on the Interaction of Colloidal Particles," Ann.
N.Y. Acad. Sci., Vol. 51, p. 627 (1949).
12 P. Sheng, "Hard Rod Model of the Nematic·lsotropic Phase Transition," Chapter 5.
13 K. Lakatos, J. Status Phys., Vol. 2, p. 121 (1970).
Introduction to the Molecular Theory of Smectic-A
Liquid Crystals

Peter J. Wojtowicz

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

Contributions to the theory of smectic-A liquid crystals have been


made by a number of investigators.1-5 In all cases the treatments are
an extension of the Maier-Saupe6 mean-field model of nematics ex-
amined in a previous chapter.7 Here we essentially follow the devel-
opment of McMillan. 3,4
The symmetry and structure of the smectic-A liquid crystals are
reviewed; the natural order parameters are identified. The relation-
ship of the smectic-A phase to the nematic (or cholesteric) and isotro-
pic phases in homologous series is also examined. The McMillan form
of the single molecule potential function is then deduced starting
from the Kabayashi form of the potentiap,2 and using the formal de-
velopment presented earlier.s The derivation of the statistical ther-
modynamics then follows, along with a presentation of McMillan's
numerical results and a comparison with experiment. Improvements
in the theory introduced by Lee et a1 5 are also considered. In the last
section, the important question of whether the smectic-A to nematic
(cholesteric) phase transition can ever be second order is examined.
83
84 CHAPTER 7

2. Symmetry, Structure and Order Parameters

An examination of the optical properties of smectic-A liquid crystals


shows that they have uniaxial symmetry. Just as in the nematic
phase, the smectic-A phase has a unique axis (again called the direc-
tor and denoted by n) along which the elongated rod-like molecules
tend to align. In addition, x-ray diffraction from smectic-A liquid
crystals displays one sharp ring demonstrating that this phase pos-
sesses one-dimensional translational periodicity. The structure is de-
picted in Fig. 1. The centers of mass of the molecules tend to lie on
planes perpendicular to the director. The spacing between planes, d,
is approximately a molecular length. There is no ordering of the cen-
ters of mass of the molecules within the planes.

"n
I/\\\I/I\'
\\ 1'/1\ I
\1(\1
\\I"(1 I'1 1/ \'\1\\\\\\I'\
1\\11\\\\1///11\ I
If"' 1\'\ 1"/\\ I'"
/I \1 \'\//1
J-y
x
Fig. 1-Schematlc representation of the structure of the smectic-A phase of liquid
crystals.

As in the nematics, the orientational order of the molecules is de-


scribed by the order parameter (P 2 (cosO», where P 2 is the second-
order Legendre polynomial, 0 is the angle between the long axis of the
molecule and the director, and < ) denotes the average value. A sim-
ple phenomenological deduction of this orientational order parameter
was presented in a previous chapter.7 The identification of the order
parameter required to describe the periodic layering of the molecules
is not as straightforward, however. For the smectic-A structure we
must examine the problem more formally.
In the case of nematics, Priestley 9 has described how the orienta-
tional distribution function could be expanded in a series of even-
order Legendre polynomials:

((roe.O) = L 2L +
2
1
(PL(cos8»)PL(cos8). [lJ
L( even)
MOLECULAR THEORY OF SMECTIC-A CRYSTALS 85

The traditional order parameter, (P 2 (cos 8» appears in the first non-


trivial term in the series. Succeeding terms contain the average values
of higher-order Legendre polynomials, which can be thought of as
order parameters of higher degree. The (PL ) thus describe features
of increasing subtlety in the orientational ordering, and many are
clearly required to give a good account of the true orientational dis-
tribution function.
How can this formal treatment of the distribution function (and
resulting order parameters) be generalized to include the smectic-A
structure? We find the clue in Kirkwood's treatment lO •ll of the melt-
ing of crystalline solids. In a crystal the density distribution function
(the translational molecular distribution function) is periodic in three
dimensions and can be expanded in a three-dimensional Fourier se-
ries. Kirkwood does this and then identifies the order parameters of
the crystalline phase as the coefficients in the Fourier series. For sim-
plicity let us consider a one-dimensionally periodic structure (such as
the smectic-A but with the orientational order suppressed for the mo-
ment). The distribution function, which describes the tendency of the
centers of mass of molecules to lie in layers perpendicular to the z-
direction, can be expanded in a Fourier series:

[2J

[3]

where d is the layer spacing and Eq. [3] expresses the ~ormalization
condition. Since the distribution is periodic we need only integrate
over a single period. We now find the coefficients, an by multiplying
both sides ofEq. [2] by cos (27rmzld) and integrating:

am = d 2fd cos(27rmz)
0 ~ f(z)dz. [4]

The integral on the right hand side is immediately recognized as the


definition of the average value, so that

am = <
d2 cos(27rmZ)
-d- . [5J

For the special case of m = 0, ao = lid. Combining these results we


obtain
86 CHAPTER 7

fez) = d1 + d~
2", ( cos(21T'nZ)
-d- cos(21T'nZ)
-d- . [6J

The coefficients in the series and hence the order parameters turn
out to be the average values of the cosine functions of the series (in
complete analogy to the situation in Eq. [1]). When the structure has
perfect periodic order, all the (cos(27rnz/d» have the value unity; for
the completely disordered system with all molecules randomly dis-
tributed in z, all (cos(27rnz/d) vanish. Again, many order parame-
ters are required to make a good approximation to the distribution
function, Eq. [6].
The smectic-A liquid crystals possess both orientational and trans-
lational order. The molecular distribution function must therefore
describe both the tendency of the molecules to orient along n and to
form layers perpendicular to n. The distribution function is thus a
function of both cosO and z, and can be expanded in a double series:

[7J
(even)

i lid
-1 0
f(cos(},z)dzd(cos(}) = 1. [8J

The coefficients ALn are found by multiplying both sides of Eq. [7] by
PK(cosO) cos(27rmz/d), integrating and recognizing the definition of
averages:

(X) == i1id
-1 0
Xf(cos(},z)dzd(cos(}). [9J

The results are


A"" = 1/2d,

Aon = ~( cos(21T'dnz ) ) , (n =F= 0),


[1OJ
2L + 1
ALa = 2d (PL(cos(}»,(L =F= 0),

ALIl = 2L 2~ I(PLCCOS(})COSC7rdnz),(L,n =F= 0).

In addition to the purely orientational and translational order pa-


rameters, the (PL(cosO) and (cos(27rnz/d) , we find the set of
mixed-order parameters, (PdcosO)cos(27rnz/d). These describe the
MOLECULAR THEORY OF SMECTlC·A CRYSTALS 87

correlation or coupling between the degrees of orientational and


translational order. The three order parameters of lowest degree in
Eq. [10] appear in all the published theories l - 5 of the smectic-A phase
and have been given special symbols:

rJ == (Picos()),
T == (cos(27rz/d», [11]
(f == (Picos() coS(27rz/d».

In the isotropic phase, TJ = T = (f = 0; in the nematic phase, TJ ~ 0, T =


(f = 0; in the smectic-A phase TJ ~ 0, T ~ 0, (f ~ 0. For perfect order

all three tend to unity. Part of the task of molecular theory is, of
course, to calculate the temperature dependence of these order pa-
rameters. Again we point out that although the three quantities of
Eq. [11] are sufficient to parametrize simple mean field models, a
good approximation to the true distribution function, f(cos(),z) re-
quires many terms in Eq. [7].

3. Phase Diagrams

Of special interest to the molecular theory are the transition temper-


atures at which the various liquid-crystal phases transform into each
other and into the isotropic fluid. The collection of such tempera-
tures in homologous series can be conveniently summarized in phase
diagrams such as those schematically depicted in Fig. 2. The phase
diagram for the homologous series of 4-ethoxybenzal-4-amino-n-
alkyl-a-methyl cinnamates is displayed in Fig. 2a. The regions of sta-
bility of the smectic-A, nematic and isotropic phases are shown. Fig.
2b depicts the phase diagram for the homologous series of the cho-
lesteryl esters of saturated aliphatic acids. Here the regions of stabili-
ty of the smectic-A, cholesteric, and isotropic phases are shown. For
the present purpose we can treat the cholesteric phase as thermody-
namically similar to the nematic; the terms in the free energy, which
differentiate between cholesteric and nematic structures, are very
small and may be neglected in this context. Three major features of
these diagrams should be noted: (a) the nematic (cholesteric) to iso-
tropic transition temperature TNI(TcI) decreases strongly with in-
creasing chain length, (b) the smectic-A to nematic (cholesteric) tran-
sition temperature T AN(T Ad first increases with chain length, then
stays constant or gently decreases, and (c) TNI(Tcy) and T AN(T Ad
are converging with increasing chain length, so that for sufficiently
88 CHAPTER 7

long chains the smectic-A phase transforms directly into the isotropic
fluid without passing through the nematic (cholesteric) phase.
Another experimental quantity of interest is the entropy of transi-
tion from smectic-A to nematic (cholesteric] structure. Just as in the

130 (a) (b) 100

_ 120 90
t
w 110 80
(J)

a:
::::J
I-
«
a: 100 70
w TNI
a.
:;;
w SMECTIC-A
I-
90 NEMATIC 60
z
0
1=

~
in
z 80 50
« SMECTIC-A
a:
I-

70 40
2 4 6 8 10 10 12 14 16 18
ALKYL CHAIN LENGTH

Fig. 2-(a) Schematic representation of the phase diagram of the homologous series
of 4-ethoxybenzol-4-amlno-n-alkyl-a-methyl clnnamates (after Ref. [3], data
of Ref. [12]). (b) Schematic representation of the phase diagram of the ho-
mologous series of cholesteryl esters of saturated aliphatic acids (after Ref.
[3], data of Ref. [13]).

case of the nematic to isotropic transition, the entropy changes are


very small, in keeping with the fact that only one degree of freedom
of the molecules (translational motion in the z-direction) is being in-
fluenced at T AN(T Ad. For both of the examples shown in Fig. 2 the
transition entropies range from about 0.2 to 1.2 cal/o mole, being low
for short chains and increasing with increasing chain length. 12 ,13

4. The Molecular Potential

The stability of the smectic-A structure is a direct consequence of the


interactions between the constituent molecules. Even though we have
virtually no detailed knowledge of their precise nature, we do know
that there must be both orientation and distance dependence in the
intermolecular pair potentials. That is, there must exist forces that
MOLECULAR THEORY OF SMECTIC-A CRYSTALS 89

cause the molecules to align parallel to each other and to form layers
perpendicular to the director. Kobayashi 1,2 has suggested a simple
form of pair interaction potential that contains the minimum neces-
sary features:

[12]

where r is the separation between the centers of mass of the mole-


cules and 812 is the angle between their long axes. The functional de-
pendence of U and Won r is not specified; U(r) represents the short
range central forces while W(r) describes the orientational forces due
to the anisotropic dispersion forces, quadrupole-quadrupole forces,
etc.
An exact statistical theory of smectics based on the pair potential,
Eq. [12], is extremely difficult to accomplish. Therefore we derive a
mean-field approximation to the theory. For this purpose we require
the mean-field version of the single molecule potential function. In a
previous chapter 8 this problem was examined for the case of the
nematic phase. A perfectly general form of V 12 was assumed and ex-
panded in a series of spherical harmonics. A new coordinate system
was then chosen such that the polar axes coincided with the director.
The single molecule potential was then obtained by averaging V 12
over all possible p.ositions and orientations of molecule 2 consistent
with the structure of the nematic phase. The resulting single mole-
cule potential had the form

[13]

where the constants (1: m ULLm(r) are the averages of the distance
dependent parts of the potential (averaged over all intermolecular
separations), and where the (PL ) are the averages of the Legendre
polynomials averaged with respect to the nematic orientational dis-
tribution function. The term in L = 0 is a constant and was discard-
ed; the term in L = 2 led to the Maier-Saupe6 version of the theory of
nematics.
The analogous process applied to the Kobayashi potential, Eq.
[12], in the case of the smectic-A structure is at the same time simpler
and more complex. It is simpler because Eq. [12] exhibits a simple an-
gular dependence and higher-order P L do not enter the calculation. It
is more complex, however, because we are now required to average
over the positons and orientations of the second molecule in a way
consistent with the smectic-A structure; that is, with a distribution
90 CHAPTER 7

function that depends on both angular and spatial coordinates, the


f(cosO,z) discussed in Section 2 above. Applying these averaging pro-
cedures to Eq. [12] we obtain the single molecule potential as

[14J

where the averages (U(r) and (W(r)P2) are functions of z, the po-
sition of the centers of mass of the molecule of interest with respect
to the layers, and where 0 is the angle between the axis of this mole-
cule and n. In obtaining Eq. [14] we have used the relation,
P 2(cosOd = P2(COSOl)P2(COS02) + terms in 'P2 - 1Pl. The terms in the
azimuthal angle 'P vanish in the averaging process since the smectic-A
phase has cylindrical symmetry. It has been customary in the treat-
ment l - 4 of the smectic-A phase to further simplify the potential by
expanding the position dependent terms in a Fourier series. Taking
U(r) as an example:

where U is the Fourier transform of U,


- - 1 roo .
V(s) = V(X I2,y12,;S) = 21r J o V(r12) COSSZI2dzI2'

Taking the average of U over the smectic-A distribution of molecule


1,

[15J

where the coefficient (Un) = (U(27rn/d» (2/Tr). The several steps re-
quired to obtain this form of Eq. [15] are presented in the Appendix.
Since there is no ordering of the molecules in the layers, the ( Un) are
just constants. The same considerations as above apply to the W(r)
portion of the potential. Retaining only the first few terms gives the
result

VI(cosO,z) = Vo + VITCOS (d21rZ) + ...


[15']
+ [ W o7] + WIITCOSC~Z) + ... ]PlCOSO),
where U 0, U 1. W 0, and WI are the Fourier coefficients of U and W,
and where 7], T, and IT are the order parameters, the average values in-
MOLECULAR THEORY OF SMECTlC·A CRYSTALS 91

troduced earlier, Eq. [11]. U o is a constant and can be discarded.


This version of the potential function is particularly instructive in
demonstrating the cooperative nature of the formation of the smec-
tic-A structure. The U 1 term shows the influence of the translational
order T in forcing the molecules into layers, the W 0 term shows the
influence of the orientational order parameter 1] in forcing the mole-
cules to align with fi, while the WI term shows how the degree of
translational order can influence the orientational order (and vice
versa) through the action of the mixed parameter u.
Specific forms of the functions U(r) and W(r) were chosen by
McMillan: 3,4

[16J
V(r) = oW(r),

where v and 0 are constants characterizing the strengths of the two


parts of the interaction, and where ro specifies the range of interac-
tion; ro is of the order of the length of the molecules. The Fourier
coefficients of Eq. [16] are

Wo = -v, VI = oWl>
[17J
WI = -va = -2v e- 1no/ d )'.
Substitution of Eq. [17] into Eq. [15'] gives the McMillan model of the
potential in the form

V\1(cosB,z) = -v{oaTcos(2;Z)
[18J
+ [11 + aucose;z) ]Picos 8 >}.
The functional dependence chosen in Eq. [16] makes it convenient to
discuss the variation of liquid-crystal behavior in homologous series.
Lengthening the alkyl chains in a series increases the spacing d. This
decreases the ratio ro/d and hence increases the parameter a intro-
duced in Eq. [17]. The properties of the short end of the homologous
series can thus be computed using small values of a while those at the
long end of the series will require larger values of a.

5. Statistical Thermodynamics

Having derived a particular form of the single molecule potential


function in the mean-field approximation, we are now in a position to
92 CHAPTER 7

calculate the thermodynamic properties of the model. According to


the rules of classical statistical mechanics the single molecule distri-
bution function corresponding to the potential function, Eq. [18], is

f.W<cosfJ,z) = Z-l exp[-/3V M(COSfJ,Z)J,


[19J
Z = ili d
exp[-/3V M(cosfJ,z)Jdzd(cosfJ),

where Z is the single molecule partition function and (3 = l/kT (k is


Boltzmann's constant and T the temperature). The integrations can
be restricted to 0 ~ cosO ~ 1 since VM is even in cosO, and 0 ~ z ~ d
since VM is periodic. The distribution function as it stands in Eq.
[19] is not as yet useful for computing average values or thermody-
namic quantities. The potential, VM, Eq. [18], contains the as yet un-
determined order parameters T/, T, and a. The self-consistent determi-
nation of the order parameters and their temperature dependence
can be realized by combining their definition, Eq. [11] with Eq. [19]:

TJ =
Jto Jrdp/cosfJ)fM(CosfJ,z)dzd(cosIJ),
o

T = fJ
0
l rd
o cos
(27l"Z)
d f ,w(coslJ,z)dzd(cosfJ), [20J

Jro Jro PicosfJ)cos (27rZ)


l d
(]" = d f M(cosIJ,z)dzd(cosfJ).

Each of the above equations contains one of the order parameters on


the left and all three order parameters in the integrals on the right.
The simultaneous solution of these three equations yields the tem-
perature dependence of the order parameters.
The set of self-consistent equations above admits a number of si-
multaneous solutions. In addition to the smectic-A, nematic, and iso-
tropic solutions, Eq. [20] also yields various less physical solutions.
To find which of the possible solutions represents the physically ob-
served states we must calculate the free energy and determine which
solution gives the minimum in this quantity at different tempera-
tures.
The energy, entropy, and free energy are obtained in exactly the
same way as described in earlier chapters 7 ,8 on the simpler theory of
the nematics:

[21J
MOLECULAR THEORY OF SMECTlC-A CRYSTALS 93

S = -Nk(lnf M)
[22]
Nv
= -1'(11 2 + aOT 2 + a0"2) + NklnZ,
1
F = -NkTlnZ + 2Nv(~ + aOT 2 + a0"2), [23]

where N is the number of molecules present. Just as in the case of the


mean-field theories of the nematics,7,8 the free energy is found to con-
tain an additional term beyond the usually expected InZ term. Again,
these terms arise because we have approximated a pair potential V 12
by a temperature-dependent single molecule potential VM. The form
of Eq. [23] can be verified to be correct, however. Setting the partial
derivatives of F with respect to TI, T, and u to zero regains the self-
consistency conditions, Eq. [20]. Further testing Eqs. [21] and [23] we
see that they do satisfy the thermodynamic identity, E = (a{3F/a{3).

6. Numerical Results

McMillan has obtained numerical solutions to Eqs. [20] through [23]


for several sets of the potential parameters () and u. In his first p~per3
he did not yet realize the need for the U (r) term in the potential, Eq.
[12], so that he was treating the case of () = O. In his subsequent
paper 4 he included the missing term. These two sets of results are
now examined.

1.0 r - - - - - - - - . - - - - - - r - - - - - - - - ,
Ir
W
I-
w
:E

i
<%
0.5
Ir
w
o
Ir
o a = 1.1

TEMPERATURE

Fig. 3-Temperature dependence of the order parameters for lj = 0 and several


values of a (after Ref. [3]).

When () = 0, the term in T drops out of the potential, Eq. [18], and
the theory now involves only the order parameters TI and u. The
translational order of the smectic-A phase is then described only by
the mixed-order parameter u. Fig. 3 shows the temperature depen-.
94 CHAPTER 7

dence of the order parameters for three representative values of a.


For a = 1.1 (long chain length) Tf and q change discontinuously at the
same temperature; both the translational and orientational order
vanish simultaneously. There is thus a first-order phase change from
the smectic-A structure directly into the isotropic fluid; no nematic
phase exists. For a = 0.85, Tf and q display discontinuities at TAN but
only q vanishes. Tf then vanishes discontinuously at a higher tempera-
ture TNI. The system thus displays a first-order transition from
smectic-A to nematic followed by another first-order transiton from
nematic to isotropic. For a = 0.6 (short chain length), q is seen to
vanish continuously at a temperature TAN. Tf shows a discontinuity in
slope at this temperature then vanishes discontinuously at a higher
temperature TNI. There is thus a second-order phase transition from
smectic-A to nematic followed by the usual first-order transition
from nematic to isotropic.
The collection of phase transition temperatures obtained from the
model with 0 = 0 and for numerous values of a is summarized in the
phase diagram shown in Fig. 4a. These results are to be compared
with the schematic phase diagram shown in Fig. 4b; this diagram is
representative of many real systems and displays the features dis-
cussed in Section 3. Note that while there is qualitative agreement
between the model and experiment, there are numerous significant
discrepancies in the overall behavior of the phase transition lines.
The entropy changes of the smectic-A to nematic phase transitions
have also been computed with 0 = 0 and for numerous values a. The
results are summarized in Fig. 5, where we have plotted !1S as a func-
tion of the ratio T AN/T NI ; small values of T AN/TNI correspond to
short chain length, large values of T AN/TNI to long chain length (see
Fig. 4a). The results of this model are shown as the solid line; the
points are experimental, taken from the homologous series described
in Fig. 2. Again, there is qualitative agreement in that the general
trend and order of magnitude are correct, but there are serious dis-
crepancies.
In his second paper4 McMillan included all the terms in the poten-
tial, Eq. [18] and studied the model for several values of 0 and a. He
was particularly interested in two substances, cholesteryl nonanoate
and cholesteryl myristate. The nonanoate calculations were made
using a ~ 0.41; 0 = 0 and 0.65; for the myristate, a ~ 0.45 was chosen
with 0 = 0 and 0.65. With 0 = 0.65 the temperature dependences of
the order parameters for both values of a were similar in appearance
to those shown in Fig. 3b. In both examples Tf looked like the Tf of Fig.
3b while T and q resembled the q of Fig. 3b. The model thus displayed
successive first-order phase changes-smectic-A to cholesteric fol-
MOLECULAR THEORY OF SMECTIC-A CRYSTALS 95

a
0.6 0.8 1.0

ISOTROPIC

NEMATIC
III SMECTIC-A
!::
z
:::l
In (0)
a::
<l (bl
III
W
a::
:::l
~
a::
w
11.
:;;
w
I-
z SMECTIC-A
0
t:
III
(c)

z
<l
a:: ISOTROPIC
I-

.38 .40 ~;I .42 .44

INCREASING CHAIN LENGTH-

Fig. 4-Schematic phase diagrams for liquid crystals: (a) theory of McMlllan,3 (b)
schematic representation of typical experimental diagrams and (c) theory of
Lee et al (after Ref. [5]).

3.0 r - - - - - , - - - - , - - - - - - ,

2.0
.,
a
.E
"-
au
1.0
(/)
<l

°85 .90 .95 1.0

Fig. 5-Entropy of transHlon plotted as a function of the ratio TANI TNI. Solid curve Is
theory of McMlllan, 3 dashed curve is theory of Lee et al5 and the pOints are
experlmental 12,13 for compounds of Fig. 2 (after Ref. [5]).
96 CHAPTER 7

lowed by cholesteric to isotropic in agreement with experimental ob-


servations. A quantitative test of the theory was attempted by com-
paring the temperature dependence of r2 with the measured x-ray
Bragg scattering intensity (to which it is proportional). The results
are shown in Fig. 6. In the case of the myristate, excellent agreement

CHOlESTERYl CHOlESTERYl
NONANOATE MYRISTATE
1.0r---------,----------,

~:=O
.~•....
11 =0.65 ~
/
EXP
.
...... ,..... "- ........
TEMPERATURE (oC)

Fig. 6-Comparlson of calculated and observed Bragg scattering Intensity (after Ref.
[4]).

with experiment is obtained using the value 0 = 0.65. For the nona-
noate, however, the agreement between theory and experiment is
poor, although the calculation with 0 = 0 is somewhat better than
with 0 = 0.65. Phase diagrams such as those shown in Fig. 4 were not
computed for finite values of o. This is unfortunate, since it is impor-
tant to know whether finite values of 0 in the potential, Eq. [18],
would have improved the appearance of the theoretical phase di-
agram, Fig. 4a.

7. Improved Theory

Motivated primarily by the disagreement between the theoretical and


experimental phase diagrams, Fig. 4 (a and b), Lee et al 5 have derived
a modified version of the mean-field theory of smectic-A liquid crys-
tals. The authors begin by adopting the McMillan form of the Kobay-
ashi pair potential, Eqs. [12] and [16]. The pair potential is then
treated intact without going into the process of deriving the equiva-
lent mean-field single molecule potential. Further, the pair potential
is not expanded in a truncated Fourier series as in the manner lead-
ing to Eq. [15'].
MOLECULAR THEORY OF SMECTlC·A CRYSTALS 97

The statistical thermodynamics is based on the variational princi-


ple. Guided by Eqs. [18] and [19], the following variational form was
chosen for the single molecule distribution function (of the i -th mole-
cule);

f(cos8"z,) = Z-l expl-I1V(cos8" z.)1

Z = £l£d expl-I1V(cosO"z,ldz,d(cos8.), [24J

V(cos8"z,) = -v[ aPz{cos8,) + bPZ<cosO.) cose~z,)


+ cocose~Z)J. [25J

where a, b, and c are variational parameters. The values of these pa-


rameters as a function of the temperature are determined by mini-
mizing the free energy. The free energy was taken to have the form

+ 1
'iN
2Jfl J
o
fdi1i d
o 0 0 f(cos8j,Zl)f(c0s82,Z2)V12(812>ZlZ)
X dz1d(cos81)dz 2d(cos8 2 ), [26J

where the first term comes from the entropy contribution (similar to
Eq. [22]), and where the second term is the energy E written in terms
of the average value of the pair potential. Although the energy is
computed from a pair potential (in contrast to the usual mean-field
definition, Eq. [21]) the free energy as written in Eq. [26] is still a
mean-field approximation. The reason for this is the use of two sin-
gle-molecule distribution functions in place of the proper (but un-
known) pair distribution function in calculating E from V 12. The
form of F is self-consistent, however, and, along with the variational
conditions (Eq. [27] below), does satisfy the required thermodynamic
identity, E = (a{3F/a{3).
The temperature dependence of the variational parameters a, b,
and c are obtained from minimization of the free energy, Eq. [26];

( aF)
aa = (aD
abj = (aD
acj = 0
. [27J
98 CHAPTER 7

Of the many possible solutions to these equations we take only those


giving the absolute minimum in F. The temperature dependence of
the order parameters 71, 7, and a are then calculated by combining the
equilibrium solutions of Eq. [27] with Eqs. [11], [24] and [25]. We
note in passing that if the pair potential V 12 (012,Z12) could be writ-
ten in (or approximated by) a separable form V 1 (OI,zd V 1 (02,Z2),
then 71 = a, a = b, and 7 = c.
Numerical calculations were performed 5 using 0 = 0.65 and numer-
ous values of the range parameter fo = 27rro/d, where ro is the range
of the interaction and d is the layer spacing (fo plays the same role in
parameterizing chain length as McMillan's a). The collection of
phase-transition temperatures obtained is displayed in Fig. 4c. The
agreement between the theoretical diagram and the representative
experimental phase diagram, Fig. 4b, is nothing less than remarkable.
All the features described in Section 3 and shown in Fig. 4b are in-
deed present in the theoretical results. Numerical calculations of the
entropy changes of the smectic-A to nematic phase transition were
also made. Fig. 5 shows the results, t,.S plotted vs. the ratio T AN/T NI ;
the result of this calculation is depicted as the dashed line. The
agreement with the experimental points is again remarkable. Note
further that this theory predicts that the entropy change, t,.S (Fig. 5)
will go to zero at TAN/TN! = 0.88; the first-order smectic-A to nemat-
ic phase change becomes second order. This feature is examined in
more detail in the next section.
The comparison of the results of the two models in Figs. 4 and 5
shows that Lee et al 5 have indeed made a very significant improve-
ment over the original considerations of McMillan. 3 Three important
changes were made in the theory: (a) the theory was based on a varia-
tional principle with E being computed from a pair potential V 12; (b)
the complete Kobayashi form of potential was used (finite 0); and (c)
the pair potential was not approximated by a truncated Fourier se-
ries. With all three improvements made simultaneously one can only
speculate as to which is the most important. (a) seems not to be a
major factor; the resulting theory is still a mean-field approach, and
using V 12 in E should only influence small details. (c) is certainly an
important factor. The influence of this improvement cannot be too
great, however, since the variational form for the potential of mean
force, Eq. [25], is in effect a truncated Fourier series of the true po-
tential. (b) seems to be the most important. The influence of the
purely translational term U in the Kobayashi potential, Eq. [12],
must be the over-riding factor that produced the spectacular im-
provements depicted in Figs. 4 and 5.
MOLECULAR THEORY OF SMECTIC-A CRYSTALS 99

8. The Possibility of Second-Order Transitions

The question of whether the normally first-order smectic-A to nema-


tic phase transition can be second order in some materials is some-
what controversial. All of the published mean-field theories l - 5 of the
smectic-A phase do exhibit second-order phase changes for certain
values of the potential parameters. In both McMillan's theory3 and
that of Lee et al,5 the second-order transition is predicted to occur at
that end of homologous series having short chain lengths. More spe-
cifically, these models predict the second-order changes to occur
when the ratio of transition temperatures TAN/TN! (or T Ac/Tcd is
at or below about 0.88 (see Fig_ 5).
The experimental situation is contradictory. For the homologous
series of 4-n-alkoxybenzylidene-4' -phenylazoanilines, Doane et aP4
have observed a possible second-order smectic-A to nematic transi-
tion at the extreme short end of the series. In the case of COC (cho-
lesteryl olyel carbonate), Keyes et aP5 found the first-order phase
transition to change to second order at a pressure of 2.66 kbar where
T AC/TCI = 0.88. The phase transitions of CBAOB (p-cyanobenzyli-
dene-amino-p-n-octyloxybenzene)16,17 and CBOOA (p-cyanobenzyli-
dene-p' -octyloxyaniline)18,19 have also been believed to be of second
order. Torza and Cladis,2o on the other hand, have concluded from
volumetric studies that the smectic-A to nematic phase transition is
unambiguously first order in CBOOA. Lin, Keyes, and Daniels 21 con-
cur in this conclusion based on high pressure studies of CBOOA. In
any case, we must be mindful of the fact that the experimental verifi-
cation of the order of the phase transition in these examples is most
difficult. The entropy and volume discontinuities are very small and
near critical fluctuations are present to complicate the interpretation.
(The nature of fluctuations in the nematic to isotropic transition has
been discussed in a previous chapter. 7)
The situation with respect to Landau-type phenomenological
theories is also contradictory. Drawing an analogy between the smec-
tic-A phase of liquid crystals and the superconducting phase of met-
als, de Gennes 22 ,23 has constructed a phenomenological theory from
which he concludes that the smectic-A to nematic phase transition
can be second order. Halperin and Lubensky,24 on the other hand,
have improved the analogy with superconductors and conclude that
the transition will always be at least weakly first order.
My own conjecture in this matter is that the smectic-A to nematic
phase transition is always first order. The basis of my argument is the
analogy between this transition and the melting of a crystalline solid.
When a crystal melts, three dimensional long-range translational
100 CHAPTER 7

order disappears; when a smectic-A goes nematic, one-dimensional


long-range translational order disappears.
Kirkwood lO,l1 has pointed out that the density distribution func-
tion of a crystalline solid (the translational molecular distribution
function) can be expanded in a three-dimensional Fourier series. The
coefficients in this series are then identified as the order parameters
of the crystalline phase. All these order parameters vanish discontin-
uously at the first-order melting point. Empirically, there are no sec-
ond-order melting transitions, nor do there seem to be any solid-liq-
uid critical points. Though not a proven fact (as far as I am aware), it
seems reasonable that crystal melting is always first order because all
of the order parameters cannot vanish simultaneously and contin-
uously before the free energy of the solid phase exceeds that of the
liquid phase.
In the smectic-A phase, the single-molecule distribution function,
Eqs. [7] and [10] can likewise be represented as a (one-dimensional)
Fourier series in which all the coefficients may be considered order
parameters. The disappearance of smectic-A order requires the si-
multaneous vanishing of all the order parameters. That they all can
vanish simultaneously and continuously before the free energy of the
smectic-A phase exceeds that of the nematic phase seems just as un-
likely here as in the (empirically verified) case of the crystalline sol-
ids. It seems clear to me that the reason the various theoretical treat-
ments mentioned above can exhibit second-order phase changes is
that an insufficient number of order parameters is included. In all the
treatments, either the potential, the potential of mean force, or the
distribution function are expressed in terms of highly truncated Fou-
rier series. Such truncation automatically limits the number of order
parameters. Small numbers of order parameters can then vanish si-
multaneously and continuously under certain conditions providing
the spurious second-order phase transitions.

Appendix

The average value of U(rlz}, averaged over the distribution of mole-


cule 1 is defined by

1, [28]

where we have temporarily suppressed the angular dependence of f


for clarity. Substituting the Fourier integral representation of U,
MOLECULAR THEORY OF SMECTIC-A CRYSTALS 101

[29]

where ([J(s) is the Fourer transform iJ averaged over all the (ran-
dom) positions of the molecules in the layers, and where the sine
terms which come from COSSZ12 = cos S(Z2 - zJ are omitted since
they will vanish in the averaging. The next step is to evaluate (cos
SZl) using the expansion, Eq. [6]:

= d2"
~
< (211"nz) J (211"nZl)
cos ~ (00 cos -d- cos SZI dZ b
o

= ~ <cose1l";z) )0 (s _ 2~n), [30]

where () is the usual delta function. Substitution of the result of Eq.


[30] into Eq. [29] gives

[31]

which is the desired result, Eq. [15].

References

1 K. K. Kobayashi, "Theory of Translational and Orientational Melting with Application to Liquid Crys-
tals." J. Phys. Soc. (Japan), 29, p. 101 (1970).
2 K. K. Kobayashi, "Theory of Translational and Orientatlonal Melting with Application to Liquid Crys-
tals," Mol. Cryst. Liq. Cryst., 13, p. 137 (1971).
3 W. L. McMillan, "Simple Molecular Model for the Smectlc-A Phase of Liquid Crystals," Phys. Rev.,
A., p. 1238 (1971).
4 W. L. McMillan, "X-ray Scattering from liquid Crystals. I. Cholesteryl Nonanoate and Myrlstate,"
Phys. Rev.• A8, p. 936 (1972).
5 F. T. Lee, H. T. Tan, Y. M. Shih. and C. W. Woo, "Phase Diagram for Liquid Crystals," Phys. Rev.
Lett., 31, p. 1117 (1973).
8 W. Maier and A. Saupa, Z. Naturforschg.• 1.a, p. 882 (1959) and 15a, p. 287 (1960).
7 P. J. Wojtowicz, "Introductlon to the Molecular Theory of Nematic Liquid Crystals," Chapter 3.
8 P. J. Wojtowicz, "Generalized Mean Field Theory of Nematic Liquid Crystals," Chapter 4.
9 E. B. Priestley, "Nematic Order: The Long Range Orientational Distribution Function," Chapter
6.
102 CHAPTER 7

10J. G. Kirkwood and E. Monroe, "Statistical Mechanics of Fusion," J. Chem. Phys., 9, p. 514
(1941).
11 J. G. Kirkwood, "Crystallization as a Cooperative Phenomenon," p. 67 in Phase Transformations
in Solids, ed. by R. Smoluchowski, J. Wiley and Sons, Inc., New York, (1951).
12 H. Arnold, Z. Physik Chem. (Leipzig) 239, p. 283 (1968); 240, p. 185 (1969).
13 G. J. Davis and R. S. Porter, "Evaluation of Thermal Transitions in Some Cholesteryl Esters of
Saturated Aliphatic Acids," Mol. Cryst. Liq. Cryst., 10, p. 1 (1970).
14 J. W. Doane, R, S. Parker, B. Cvilk, D. L. Johnson, and D. L. Fishel, "Possible Second-Order
Nematic-Smectic-A Phase Transition," Phys. Rev. Lett., 28, p. 1694 (1972).
15 P. H. Keyes, H. T. Weston, and W. B. Daniels, "Tricritical Behavior in a liqUid-Crystal System,"
Phys. Rev. Lett., 31, p. 628 (1973).
16 W. L. McMillan, "Measurement of Smectic-A Phase Order-Parameter Fluctuations near a Second
Order Smectic-A Nematic Phase Transition," Phys. Rev., 7A, p. 1419 (1973).
17 B. Cabane and W. G. Clark, "Orientational Order in the Vicinity of a Second Order Smectic-A to
Nematic Phase Transition," Solid Stale Comm., 13, p. 129 (1973).
18 L. Chueng, R. B. Meyer, and H. Gruler, "Measurement of Nematic Elastic Constants near a Sec-
ond Order Nematic-Smectic-A Phase Change," Phys. Rev. Lett., 31, p. 349 (1973).
19 M. Delaye, R. Ribotta, and G. Durand, ··Rayleigh Scattering at a Second Order Nematic to
Smectic-A Phase Transition," Phys. Rev. Lett., 31, p. 443 (1973).
20 S. Torza and P. E. Cladis, "Volumetric Study of the Nematic-Smectic-A Transition of N-p-cyano-
benzylidene-p-octyloxyaniline," Phys. Rev. Lett., 32, p. 1406 (1974).
21 W. J. Lin, P. H. Keyes, and W. B. Daniels, ··High Pressure Studies of Liquid Crystal Phase
Transitions in CBOOA,'· Phys. Lett., A49, p. 453 (1974).
22 P. G. de Gennes, "An Analogy Between Superconductors and Smectics-A," Solid State Comm.
10, p. 753 (1972).
23 P. G. de Gennes, "Some Remarks on the Polymorphism of Smectics," Mol. Cryst. Liq. Cryst., 21,
p. 49 (1973).
24 B. I. Halperin and T. C. Lubensky, "On the Analogy Between Smectic-A Liquid Crystals and Su-
perconductors," Solid State Comm., 14, p. 997 (1974).
Introduction to the Elastic Continuum Theory of
Liquid Crystals

Ping Sheng

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

In the preceding chapters of this book we have seen that in terms of


molecular theories1 ,2 one can calculate arid successfully explain various
properties of meosphase transitions. However, there exists a class of
liquid-crystal phenomena involving the response of bulk liquid-crys-
tal samples to external disturbances, with respect to which the use-
fulness of a molecular theory is not immediately obvious. These phe-
nomena are usually distinguished by two characteristics: (1) the ener-
gy involved, per molecule, in producing these effects is small com-
pared to the strength of intermolecular interaction; and (2) the char-
acteristic distances involved in these phenomena are large compared
to molecular dimensions. In describing these large-scale phenomena,
it is more convenient to regard the liquid crystal as a continuous me-
dium with a set of elastic constants than to treat it on a molecular
basis. Based on this viewpoint, Zocher,3 Oseen,4 and Frank5 devel-
103
104 CHAPTER 8

oped a phenomenological continuum theory of liquid crystals that is


very successful in explaining various magnetic (electric) field-induced
effects. It is the purpose of the present chapter to develop this elastic
continuum theory for nematic and cholesteric liquid crystals and to
discuss and illustrate its use. In this paper, the derivation of the fun-
damental equation of the elastic continuum theory is followed by the
application of the theory to four effects: (1) the twisted nematic cell,
(2) the magnetic (electric) coherence length, (3) the Freedericksz
transition, and (4) the magnetic (electric) field-induced cholesteric-
nematic transition.

2. The Fundamental Equation of the Continuum Theory of Liquid


Crystals

In earlier chapters we have seen that liquid crystals are characterized


by an orientational order of their constituent rod-like molecules.1,2,6,7
In nematic liquid crystals this orientational order has uniaxial (cylin-
drical) symmetry, the axis of uniaxial symmetry being parallel to a
unit vector ft, called the director. Let us now consider a very small
spatial region inside a macroscopic sample of nematic liquid crystal
that contains a sufficiently large number of molecules so that the
long-range orientational order is well defined within that region.
Such a spatial region can be characterized by a director pointing
along the of local axis of uniaxial orientational symmetry. Let us
imagine the division of the macroscopic sample into such small spa-
tial regions. In each of the regions, we define an orientational direc-
tor. In this manner the macroscopic sample of nematic liquid crystal
can be characterized by a local director at every spatial "point,"
where we use the term point loosely to mean a small region of space
as defined above. Obviously, this characterization of orientational
order by a director field, ft Cf), is not limited to nematic liquid crys-
tals, which we have used as an example in the above discussion. In
fact, the director-field characterization can be applied equally well to
cholesteric liquid crystals, since, locally, cholesterics also possess uni-
axial symmetry in their orientational order. However, for conve-
nience, we will continue to use nematic liquid crystals as our refer-
ence in the following discussion. The generalization of the theory will
be made at appropriate places to permit application to cholesteric
liquid crystals. In what follows, we first consider the free energy asso-
ciated with a distortion in the director field. Next, we examine the
free energy associated with the interaction of liquid crystals with ex-
ternal fields. The combination of all the free-energy contributions
then yields the fundamental equation, which is an expression for the
ELASTIC CONTINUUM THEORY 105

total free energy of a sample of nematic or cholesteric liquid crystal in


an external field.
The starting point for the development of continuum theory is the
consideration of the equilibrium state. In nematic liquid crystals,
parallel alignment of all the local directors represents the equilibrium
state, or the state of minimum free energy. However, when we pert-
urb the system by pinning the surface directors to the walls of the
container, applying an external field, or introducing thermal fluctua-
tions, the local directors will no longer be spatially invariant. A quan-
titative formulation of the above statements is that the quantities
dn aldx{3, (where x is the spatial variable and the subscripts ex,fJ =
1,2,3 denote the components along the three orthogonal axes of the
Cartesian coordinate system) are zero for the equilibrium state but
are nonzero for some, or all, values of ex and fJ when the system is dis-
torted. In other words, we can think of dn aldx{3 as the distortion pa-
rameters, and the equilibrium state is given by the uniformly aligned
state for which dn aldx{3 = 0 everywhere. Since the distorted state
represents a state with higher free energy than the equilibrium state,
we can write the free-energy density of the distorted state as

{(distorted) = fo(equilibrium) + ;1{, [1]

where f and fo are the free-energy densities of the distorted and equi-
librium states, respectively; N (>0) is a function of the na and the
dn aldx{3 that vanishes when all dn aldx{3 = O. Since, in general, dnal
dX{3 « (molecular dimension)-l for the phenomena of interest, we
can expand;1f as a power series in the na and the dn aldx{3 and retain
only the lowest-order non vanishing terms of the series. Let ~ D (sub-
script D stands for distortion) denote such an approximation to ;1f.
~ D must satisfy several requirements. First, since we are expanding
in powers of the dn aIdx {3 around dn aldx {3 = 0, which is a free-energy
minimum, the lowest-order nonvanishing terms must be quadratic in
the dn aldx{3 (i.e. proportional to terms of the form (dn aldx{3)' (dn-/
dXIJ». Second, since the "head" and the "tail" of a nematic director
represent the same physical state, ~D must be even in the na' Third,
~ D must be a scalar quantity. In addition, we will discard terms of
the form V . iter) where it(r) is any arbitrary vector field, since they
represent surface contributions to the distortion free-energy density
and are assumed to be small (by Gauss's Theorem, Iv· it(r)dV =
I d u • it(r) , where d u is a surface element with unit vector perpendic-
ular to the surface). With the above constraints, it can be shown 5 that
~D contains only three linearly independent terms. They are (1)
106 CHAPTER 8

[V. n(r))2, (2) [nm . V X n(r) ]2, and (3) [n(r) X V X nm)2. It is ob-
vious that all three terms satisfy the requirements stated above. In-
terested readers are referred to Refs. [5] and [8] for proofs that these
three terms are indeed unique. In Fig. 1 we show the physical distor-

\
I
I ----- \
\
\
\

I I
\ I I I I I
\ I 1 I I I
\ I 1 [ I I
\ I 1 1 I I
\ I I \ \ \
\II \ \ \

(01 SPLAY (blTWIST (c 1BEND

Fig. 1-Three types of distortion In a director field: (a) gives \7. fI(;) -,t. 0, (b) gives
flm· \7 X flm
-,t. 0, and (c) gives X \7 X flm flm
-,t. 0. Each director is
shown with double arrows in order to indicate that the "head" and the "tall"
directions represent exactly the same physical state In a nematic sample.

tions of the director field associated with the three terms. The first
term is called "splay," the second term "twist," and the third term
"bend." 5'D can now be written as

1 -- ----
5' 0 = 2IKll[V.n(r)]2 + K 22 [n(r)·V X n(r»)2

+ K33C;ii~ X V X ;;:(7»)21. [2J

where the constants K 11, K 22, K 33 are, respectively, the splay, twist,
and bend elastic constants and are named collectively as the Frank
elastic constants. The factor %is included so that the K's may agree
with their historical definitions. Since 5' D must be positive in order to
give stability for the uniformly aligned state, all the K's must be posi-
tive. As for their values, we note that the theoretical determination of
the K's from molecular parameters represents a task of linking the
continuum theory to the microscopic theories of liquid crystals and is
beyond the scope of the present paper. However, from dimensional
analysis we can get an order-of-magnitude estimate of what the
values of the K's should be. Since the K's ·are in units of energy/
length, we must look for the characteristic energy and length in the
problem. The only energy in the problem is the intermolecular inter-
ELASTIC CONTINUUM THEORY 107

action energy, which is estimated* to be "'0.01 eV, and the only suit-
able length is the separation between two molecules, which is ",10 A.
Therefore, K ~ 10-7 dyne, in order-of-magnitude agreement with
measured values 8 of 10-7_10-6 dyne. The K's are also temperature
dependent. In fact, it can be shown 9 that the temperature depen-
dence is of the form K '" <P 2 (cos 9»2, where < > denotes averaging
over that small volume of the sample that is characterized by a local
director fi Cf), P 2 is the Legendre polynomial of second order, and 9 is
the angle between any molecule (inside the volume where the average
is taken) and the local director fi(r). < P 2 (cos 9) > is just the local
order parameter measured with fiCf) as the axis of symmetry. The de-
pendence of the K's on the square of the local order parameter is
plausible if one thinks of the K's as the macroscopic analog of the an-
isotropic intermolecular interaction constants, the difference being
that, in place of molecules, we have small volumes of the sample with
well-defined long-range orientational order. In such an analogy
<P2(COS 9» plays L:1e role of a (temperature-dependent) dipole
st rength of the molecules.
Suppose now the sample of nematic liquid crystal is placed under
the influence of a magnetic or an electric field. Because the liquid
crystal molecules are generally diamagnetic, electrically polarizable,
and anisotropic in their magnetic and electric properties, the applica-
tion of a field usually contributes an amount of free-energy density
which is opposite in sign to that of the distortion free-energy density
(because fields help align molecules). Let us first discuss the magnet-
ic field contribution. Consider again a small region of the sample
characterized by a local director fiCf). The diamagnetic susceptibility
per unit volume in such a small volume is usually anisotropic. Let XII
denote the susceptibility per unit volume parallel to fi(r) and Xl.. de-
note the susceptibility per unit volume perpendicular to fiCf). The
difference, Llx == XII - Xl.., is a measure of the local anisotropy. As
shown in Ref. [6], LlX is equal to N < P 2 (cos 9) > (n - f.L), where N
is the number of molecules per unit volume, < P 2 (cos 9) > is the ori-
entational order within the small region under consideration, and
fll(f.L) is the diamagnetic susceptibility of a single rod-like molecule

* The intermolecular interaction energy responsible for the nematic


ordering can be estimated from the latent heat of isotropic-nematic
phase transition f1E ~ 300 cal/mol ~0.01 eV/molecule (see, for ex-
ample, G. H. Brown, J. W. Doane, and D. D. Neff, A Review of the
Structure and Physical Properties of Liquid Crystals, p. 43, CRC
Press, Cleveland, Ohio, 1971).
108 CHAPTER 8

parallel (perpendicular) to its long axis. In the following we will as-


sume that XII> X.L (i.e., positive anisotropy) and that the value of
<P2(COS 9) > is uniform throughout the volume of the sample,
implying that ~X, K u , K22, and K33 have no spatial dependence. In

A(t, ,THELOCAL AXIS OF


~ UNIAXIAL SYMMETRY
/l1?- I/.
A II "Iq
/Iv '?f./
'0 'ROD-LIKE MOLECULES

Fig. 2-OrIenlation of the local axis of uniaxial symmetry wHh respect to the external
magnetic field direction.

Fig. 2 we show the relative directions of magnetic field il and local


director n(r). The induced diamagnetic moments per unit volume
parallel and perpendicular to n(f) are, respectively,

M" = HX,cosO,
M .L = HX.L sinO.

The work done by the field per unit volume is

Wmagnetlc = LH(-M/sinO -M/cosO)dH'


H2
= - Z(X.L + ~XCOS20).

Discarding the spatially invariant term, -H2X.L/2, we obtain the


magnetic field contribution to the free-energy density

[3J

Using similar arguments as above, we obtain the electric-field contri-


bution to the free-energy density

[4]

where ~f = fll - f.L is the difference between the local dielectric con-
ELASTIC CONTINUUM THEORY 109

stants in directions parallel and perpendicular to the local director.


Values of ~x and ~E typically range from 10- 7_10- 6 cgs units for ~x
and 0.1-1 for ~E.
At this point we make a slight generalization so that the theory can
be applied to cholesteric liquid crystals as well. The basic difference
between cholesteric and nematic liquid crystals lies in the fact that
the equilibrium state of cholesterics is characterized by a nonvanish-
ing twist in the director field. If we denote the cholesteric helical axis
as the x -axis, the equilibrium state is characterized by

nx = 0,
ny = co'?l..:rrx/Ao),
n, = sine 7rX/Ao),

where AO is the pitch of the helix. Using this representation of the di-
rector field, the twist term, [fi(r) . Y' X n (r) )2, is calculated to be ('TrI
AO)2. This suggests that the twist part of free-energy density for the
cholesteric liquid crystals should be expanded around Ifi(r) • 'il X
fi(r)1 = 'Tr/Ao, where II is the absolute value sign. In fact, it can be
shown 5 that the appropriate form is Un(r) • Y' X fi(r)1 - 'Tr/Ao)2.
We are now in a position to combine all the free-energy density
terms to give a total free-energy density 5' of the system under exter-
nal fields:

= HKll[Y'.nC-;)]2 + K {1;i(;)·Y'
2 X ;;(r>I- ~oT
+ K33[~(-;) X Y' X ;;(;»)2 - ~X[H·~(--;)]2 - 4~~f[e:;;~)]2}. [5]

The total free energy of the sample is given by

F = [6]
volumt-
of the sample

Eqs. [5] and [6] are the fundamental equations of the elastic contin-
uum theory of nematic and cholesteric liquid crystals (for nematics AO
in Eq. [5] is set equal to <x». In the following section we use the funda-
mental equations to solve four examples as illustrations for their ap-
plications.
110 CHAPTER 8

3. Applications of the Elastic Continuum Theory

The basic principle involved in the application of the fundamental


equations to the solution of actual problems is that the equilibrium
state of the director field is always given by that director configura-
tion that minimizes the free energy of the system with specified
boundary conditions.
Before getting into actual calculations we first simplify Eq. [5] by
setting K 11 = K 22 = K 33 = K. This greatly facilitates the mathe-
matics but does not affect the qualitative behavior of the results. Ne-
glecting the term 7r/AO for the moment, we have

5' = ~{[V"n(--;)J2 + [n(;)-V' X ;;(;)J2 + [n(;) x V'. X n(r»)2

AX - - - 1 At - - - }
- K[Hn(r)J 2 - 47r K [E-n(r)]2

= K{[V"nC--;)]2 + [V' X nC-;)] 2 - AXC:iI'nC-;)J2


2 K
- 417r ~[E.nC-;:)]2}. [7J

"" z t
t
""
/
t
l
t
x
/ t

Fig. 3-The geometry of a 90°-twisted nematic cell.

3.1 Twisted Nematic Cell

In Fig. 3 we show a planar cell containing nematic liquid crystal, two


of whose bounding walls are rubbed or otherwise treated so that the
directors near the walls are pinned in the directions shown. Since the
ELASTIC CONTINUUM THEORY 111

directors at the two walls are perpendicular to each other, the local
nematic directors must undergo a 90° twist in passing from one wall
to the other. The question is how this twist is distributed across the
cell, i.e., should the distribution be uniform or nonuniform? To an-
swer this question, we must calculate the free energy of an arbitrary
twist pattern with the specified boundary conditions. The correct
twist pattern is then given by that director configuration that mini-
mizes the free energy of the system. Assuming all the directors lie in
the y-z plane, we write

nx = 0,
ny = sinB(x),
n z = cosO( x ).

From this representation of the director field, it is easily calculated


that V' • n(f) = 0 and V' X nCr) = (dO(x)/dx)[sinO] + COSOk], where we
use iJ,k to denote the unit vectors in the x,y,z directions, respective-
ly. Using Eqs. [6] and [7], we obtain

F =
K
fd3r"2[V' X n(r»)2
-

!i fd3r[dO(x)]2
2 dx
KA
2 f dx[d~~)J, [8J
thickness
of the cell

where A is the area of the cell in the y-z plane. To minimize F, we re-
call that if one wants to minimize the value of an integral

I - -J ( b
dy(x) x \-J
G y(x), (lX, rX
a

by varying the functional form of y(x), the optimal function y(x)


must satisfy the equation

[9]

which is called the Euler-Lagrange equation. In our present case 0


corresponds to y, and G = [dO(x)jdx]2; the application of Eq. [9]
112 CHAPTER 8

yields d 2 0(x)/dx 2 = 0, or dO(x)/dx = C, where C is an integration


constant that can be determined by the boundary condition C X (cell
thickness) = ±7r/2 (+7r/2 is indistinguishable from -7r/2 for nemat-
ics). This is the result we are looking for. It tells us that the twist will
be uniformly distributed across the cell. However, because dO/dx can
be either + or -, the twist can be either left handed or right handed.
In an actual 90°-twisted nematic cell, both senses of the twist are
usually present, a fact that is indicated by the existence of visible
disinclination lines separating regions of opposite senses of the twist.

3.2 Magnetic Coherence Length

The geometry of this problem is shown in Fig. 4. A semi-infinite sam-


ple of nematic liquid crystal with positive anisotropy is bound on one

I
vLx
t I "
Ii'
"'"
.,..,
t
t
I
I
.I'
.I' ..
"'"

...,., -
_H
t I .I' "'"
t I .I'
t I .I'
t I ,.I'
.,.,
)(=0 /9_ _
L~

Fig. 4-Distortion of the director field when the molecules are pinned to the wall per-
pendicular to the external magnetic field direction. The nematic molecules
are assumed to be diamagnetic with posHlve anisotropy.

side by a wall that is treated so that the directors near the wall are
pinned along the z- direction. A magnetic field is applied along the x-
axis, so that far away from the wall the directors would lie along the
field direction. There is a transition region near the wall where the di-
rectors gradually change from one direction to the other. The prob-
lem is to find the characteristic length of that transition region. From
ELASTIC CONTINUUM THEORY 113

Fig. 4 we have

nx = cos8(x),
ny = 0,
n z = sinO(x).

Straightforward calculation gives

Substitution into Eq. [7] yields

-F = -Kf= dx ([d(}(X)]2,lX
- - - - [H cosO(x)J2 ) [10]
A 2 dx K '
o

where A is the area of the bounding wall. Application of the Euler-


Lagrange equation to Eq. [10] gives an equation for the O(x) that
minimizes the free energy of the system:

K d 20(x) .
H 2,lX d";"2 - SInO cosO = o. [l1J

The combination K/(H2,lX) has the dimension of (length)2. We de-


fine

[12]

as the characteristic distance of the problem. Eq. [11] can now be re-
written as

sinOcosO dO = d 20 dO =!~ (dO)2


~M2 dx dx 2dx 2 dx dx
sinOdsinO = !~ (d8)2
~M2 dx 2dx dx
or
[13J
114 CHAPTER 8

Integration of Eq. [13] yields

(~~y = etOY + c.
The constant of integration is fixed by the condition that as x - 00,

0- 0 and dO/dx - O. Therefore C = 0, and

dO sinO
dx = ± ~M'

Choosing the - sign for x > 0 and integrating once more, we have

or

O(x) = 2 arctan[expl-x/~Ml]' where x ~ O. [14]

Eq. [14] is plotted in Fig. 5. As we can see, ~M is the characteristic


length scale below which the magnetic field does not have much in-

_ 90
(/)

t:l BO
ffiw 70
060
~ 50
-;; 40
;;; 30
20
10
°0L------L----~2----~~3==~~±4======5~

Fig. 5- Tilt angle (deviation from the field direction) of the director field plotted as a
function of distance from the wall. The distance is measured in units of mag-
netic coherence length, which is the characteristic length of magnetic phe-
nomena in nematic liquid crystals.

fluence on the relative orientations of the directors. Another way of


saying the same thing is that ~M defines the scale of magnetic phe-
nomena. ~M is usually called the "magnetic coherence length." For K
'" 10- 6 dyne, LlX '" 10- 7 cgs units, and H '" 1OkOe, ~M is about 3 /lm.
ELASTIC CONTINUUM THEORY 115

Suppose now that in place of the magnetic field an electric field is


applied. If impurity conduction and other dynamical effects are ne-
glected, the problem is qualitatively the same, and a quantity h can
be obtained which is the exact analog of ~M. Substitution of t:u/47r for
Llx and E for H in Eq. [12] gives

1/47rK [15]
~E == E Llt'

Setting ~M = ~E, we can compare the relat~ve effectiveness of magnet-


ic and electric fields in orienting the nematic directors. The relation
is

E = /4:~X H. [16J

Taking H = 1 Oe, Llx ~ 10- 7 cgs unit, and Llt ~ 0.1, we have

!47rLlX 1 statvolt
E = ~H ~ ylO-5H ~ 300 em = 1 V/em.

Therefore, one oersted of magnetic field is equivalent to the order of


one volt/cm of electric field in terms of effectiveness in orienting the
nematic liquid crystals.

3.3 Freedericksz Transition

Consider a planar cell of nematic liquid crystal with directors on both


surfaces anchored perpendicular to the walls as shown in Fig. 6. It
was first observed by Freedericksz lO in 1927 that such a cell would
undergo an abrupt change in its optical properties when the strength
of an external magnetic field, applied normal to the director (z- direc-
tion in Fig. 6), exceeded a well-defined threshold. (In the original ex-
periment one wall of the cell was concave in shape so as to give some
variation in the cell thickness.) Freedericksz further noted that the
strength of the magnetic field at threshold was inversely proportional
to the cell thickness d. 11 This Freedericksz transition is now a well-
studied phenomenon, and has found applications in liquid-crystal
display devices. The transition is essentially due to the magnetic
alignment of the bulk sample directors at sufficiently high field
strength. However, both the abruptness of its onset and the relation-
ship between the threshold field strength and the cell thickness are of
theoretical and practical interest. Here we will use the continuum
116 CHAPTER 8

theory to calculate the various properties of the Freedericksz transi-


tion. From Fig. 6 we get

nx cosO(x),
n) 0,
n, sinO(x ).

z
iH>HF
vLx
X- l ~

X- t ~

'/.. 'I ~
;/..

X ,.
~ ~

X- t ~

z tI ~
L\~_
x=-~ x=O x=+t

Fig. 6-Local nematic directors In a Freederlcksz cell when H > HF • Dotted lines Indi-
cate the equivalent tilt configuration of the directors. TIH angle (J Is defined
as shown.

From these expressions one obtains

[Y" n&)]2 + --
[Y' X n(rl]2 =
[dO(X)]2
(IX ,

and

Substitution into Eq. [7] and application of the Euler-Lagrange


equation results in

2 d 20 .
~M dx 2 + smOcosO = O. [17]

Using the same manipulations as those for Eq. [11], we get


ELASTIC CONTINUUM THEORY 117

(ddO)2
x
= C
-
sin 20
~M2·
[18]

The constant of integration C is obtained by noting that, from the


symmetry of the problem, dO /dx = 0 at x = O. Defining O(x = 0) as OM,
we get C = sin20M/~M2, and

[19]

where the + sign corresponds to the solution in region x < 0 and the
- sign corresponds to the solution in region x > O. Since the solution
is symmetric about x = 0, we will choose the + sign in the following
calculations. Integration of Eq. [19] yields

or
+ x)sinO M f
/j(x)

..l(lf:. dO' [20]


~M 2
o I [
1 - s.inO'
smOM
J2'
where -d/2 ~ x ~ o. The solution of Eq. [20] will proceed in two
steps. First OM will be determined as a function of ~M (or of H, since
~M = ~K/AX/H). Then this OM(H) can be substituted back into Eq.
[20] for the solution of O(x.) as a function of H.
From the definition of OM we get

d . 0 [21]
2~M sm M

This equation can be solved graphically as in Fig. 7 by plotting, as a


function of sinOM, the left- and right-hand sides on the same graph
and locating the points of intersection. By expanding the integral on
the right-hand side of Eq. [21], denoted here as L(sinOM), for small
values of OM, we get the slope

dL(SinO M )\ 11"
d sinOM 8M _ 0 = "2.
Therefore, for d/(2~M) < 11"/2 the only solution of Eq. [21] is OM = O.
118 CHAPTER 8

However, when d/(2~M) > 7r/2, a second solution with 8M ~ 0 is ob-


tained that gives lower free energy than the 8M = 0 solution. The crit-
ical magnetic field Hp for the transition is found by equating d/(2~M)

4.0

35

25

20

15

10

05

0~~--4~~6~~8--~IO--·
Sin 8M

Fig. 7-Graphical solution of Eq. [21].

and 7r/2. Substitution of vK/ D.x/H for ~M gives

[22]

which agrees with Freedericksz's observation that the threshold field


strength varies inversely with the thickness of the sample. The form
of Eq. [22] can be understood by a simple plausibility argument. In
section 3.2 we have seen that the magnetic coherence length ~M can
be thought of as that length below which the magnetic field does not
have much influence on the relative orientations of the directors. By
applying this interpretation of ~M to our present example it is clear
that only when ~M < d /2 would it be possible for the magnetic field to
have significant influence on the orientations of the directors. There-
fore, we would estimate
ELASTIC CONTINUUM THEORY 119

If)
w 90
w 80
a::
'"0
w 70
60
~
50
:E
40
'" 30
20
\0
0
2 3 4
H
HF

Fig. 8-Tilt angle of the directors at the center of Freedericksz cell, 8M, plotted as a
function of reduced field HIHF• For HIHF slightly greater than 1, 8M behaves
as -(HIHF - 1)1/2.

which differs with the exact result only by a factor of 7r/2. In Fig. 8,
OM is plotted as a function of H/HF. For H "" H F, Eq. [21] can be ex-
panded around 8M = 0 to give 8M 0: (H - HF)1/2.
Having obtained OM(H), we can now determine 8(x) as a function
of H. By writing d/2~M = 7rH /2HF, Eq. [20] is put in the form

7r H (
"2 HF 1 +d
2X) .
sm8M=
f0 /1 _ dO'(SinO')2
o(X)

[23]

sinOM

The right-hand side can be numerically integrated on computer, and


the results 8(x) are plotted in Fig. 9 for three different values of H.

8 8
t H=2HF t H =4HF
90· 90·
80· 80·
70· 70·
60· 60·
50· 50·
40· 40·
30· 30·
20· 20"
\0· \0"1

-\ o +\
o· -\ o +\
o· -\ o +\

h h .h.
d d d

Fig. 9-TIH angle of the directors plotted as a function of position In a Freederlcksz


cell for three different magnetic field strengths.
120 CHAPTER 8

The Freedericksz transition can be induced by an electric field as


well as by a magnetic field. The threshold electric field in that case is
given by

[24]
or

[25J

For K ,....., 10-6 dyne, Llx ,....., 10- 7 cgs unit, Llf ,....., 0.1, the critical magnet-
ic field HF is""'" 10 Oe for d ,....., 1 cm, and the critical voltage is ,....., 10 V,
independent of cell thickness.
To conclude the discussion of the Freedericksz transition, we note
that for H > HF (V > VF) there are two equivalent tilt configura-
tions of the directors, denoted by the solid and the dotted lines in Fig.
6. In practice, for H > HF (or V > VF) both tilt configurations are
usually present, and regions of different tilt patterns are separated by
visible disinclination lines.

END VIEW

Fig. 10-Two views of the local directors In a cholesteric liquid crystal. In order to In-
duce the cholesteric-nematic tranSition, a magnetic field H Is applied per-
pendicular to the cholesteric helical axis. The angle (J used In the calculation
Is defined as shown.

3.4 Field-Induced ChOlesteric-Nematic Transition

Consider a sample of cholesteric liquid crystal placed in a magnetic


field H, with the field direction perpendicular to the cholesteric heli-
cal axis as shown in Fig. 10. From Eqs. [5], and [6], the free energy of
the system over one period (or pitch) of the helix, A, can be written as
ELASTIC CONTINUUM THEORY 121

where A is the area of the sample in the y-z plane, assumed to be a


constant, >'0 is the pitch of the cholesteric helix at H = 0, 0 is the
angle between a local director and the y- axis as defined in Fig. 9,and
x = 0 is defined by any point at which 0 = 0 (or 11"). With

nx = 0,
ny = cosO(X).
n, sinO( x),

we have

dO(x)
----cIX

and

[27J

Here, we have to remember to take the absolute value of dO/dx. Ap-


plication of the Euler-Lagrange equation yields

~ 2_
d 20 + sinOcosO = 0
M 2dx '

which, as seen previously, can be put in the form

dO)2
( dx =
1(1
~i k 2 -
.)
sm 28 , [28J

where k 2 is an integration constant. At H = 0, it follows from Eq. [27]


that (dO/dx) equals a constant, (11"/>'0). Therefore, k must behave as
",H for H -- 0 in order to cancel the H2 from 1/~M2 in Eq. [28]. Writ-
ing Eq. [28] in the form

[28aJ

we note that for finite values of H, dO/dx is no longer a constant.


122 CHAPTER 8

Plotting the z-component of the local director, n z = sinO(x), as a


function of position along the helical axis (x- axis) reveals that the si-
nusoidal pattern for nz(x) at H = 0 becomes distorted at finite values
of H as shown in Fig. 11. The distortion makes nz(x) more square-
wave-like and lengthens the pitch of the helix. Both of these effects
can be understood on the basis that alignment along the magnetic

I"zl H=O

.~ . .5 1.0 1.5

H =0.9He A= 1.212 AO

~. 10 15

Fig. 11-Component of the cholesteric local director along the external field direction
( z) for two different magnetic field strengths. Note that at finite field strength
the helical pitch is lengthened and the sinusoidal shape of the curve at H =
o is distorted, becoming more square-wave-like.

field direction lowers the energy of the system. The ± signs for dO/dx
indicate the two possible senses of the helical twist. Since they are
equivalent, we choose the + sign in the following calculation. From
the expression for dO/dx we can get an expression for the pitch A:

A = f
A

dx So " dO-
o
dx
dO
= L" VI - ~M.k
0 k 2sin 20
dO

[29J

At this point, it becomes necessary to know k 2 as a function of H. To


do that, we must substitute Eq. [28a] back into Eq. [27] and minimize
the average free-energy density by varying k 2 , Let us rewrite Eq. [27]
as
ELASTIC CONTINUUM THEORY 123

[30]

where g is a dimensionless average free-energy density and qo =


7r/>"o. Substitution of Eq. [28a] for dx/dO and expansion of the terms
in the integrand gives

Detailed steps leading from Eq. [30] to Eq. [31] are given in the Ap-
pendix. Differentiation of g with respect to k 2 yields

[32J

The desired equation for determining k as a function of H is ob-


tained by setting dg/dk 2 = 0:

"2
~~f deV! - k 2 sin 28. [33]
o

Define

where E J and E 2 are the complete elliptic integrals of the first kind
and the second kind, respectively. Eqs. [29] and [33] can be put in the
form

and
124 CHAPTER 8

2
-k E.'(k),
7r -

Combining the two equations yields

[34J

E 1 (k) diverges at k = 1. Therefore, A/Ao diverges at a field given by


Eq. [33]:

~M7r = ~ E.,(l) = ~
Ao 7r - 7r'

which defines a critical magnetic field

[35J

If one takes K 22 ,...., 10- 6 dyne, Llx ,...., 10- 6 cgs units, Ao ,...., 10- 4 cm, He
is ,....,50 kOe. A similar threshold can be obtained if the magnetic field
is replaced by an electric field:

[36J

Eqs. [29], [33], [34], and [35] were first obtained by de Gennes. 12 In
terms of He, Eq. [33] can be put in the form

He [33aJ
- k = E.'(k).
H -

For H > He this equation has no solution. When H < He, the values
of k ranges from 0 to 1 as plotted in Fig. 12. In Fig. 13 we show a plot
of A/Ao vs. H/He. At H = He the pitch diverges and the cholesteric
phase transforms into the nematic phase. Experimentally this curve
is well verified. 13 ,14
Finally, it should be noted that if the field is initially applied paral-
lel to the helical axis, the cholesteric helix would usually rotate at H
< He so as to make the field perpendicular to the helical axis. There-
fore, the geometry shown in Fig. 10 is always the situation seen exper-
imentally just before the field strength reaches the cholesteric-nema-
tic transition threshold. 14
ELASTIC CONTINUUM THEORY 125

Fig. 12-Solutlon of Eq. [33a].

4. Concluding Remarks

The above discussion of the continuum theory of liquid crystals is by


no means complete. There exist many more effects that can be de-
scribed by the continuum theory, either in its present or modified
form. In view of the diverse applications of the theory, the selection

2.6

24

22
).
I;,
2.0

18

1.6

14

12

1.0
.88 .90 .92 .94 .96 .98 1.00
H
He

Fig. 13-Ratlo of the helical pnch In flnne field to the pHch In zero field, >..0\'0. ploHed
as a function of reduced field H/H". The divergence at H/Hc = 1 Is logarnh-
mlc In nature.
126 CHAPTER 8

of the four examples discussed in this chapter is based on the consider-


ation that they all have practical relevance to liquid-crystal display
devices. It is hoped that their description by the continuum theory
can, on the one hand, demonstrate the power and the flavor of the
theory and, on the other hand, complement the discussion of the de-
vice physics aspects of these effects in other papers of this series.

Appendix

In this appendix we show the steps leading from Eq. [30] to Eq. [31].
From Eqs. [28a] and [30] we have

g=-
1
A
i"
0
dO
VI -
k~M
k 2sin 20
{ 1- 2Vl - Psin 2 0
kqO~M
+ 1- k 2sin 2 0 _ Sin 2 0}
k qo2 ~M2
2 qo2~M2

= k~M dO r" dO _ 27r + 1


A Jo VI - k 2sin 20 Aqo k~Mq02

X 1"d8f1=k2sin20 - q02:~M2iAdX sin 28. [37J

By writing sin 20 = (liP) - ~M2(d8Idx)2, the fourth term in Eq. [37]


can be simplified as

-~
. 211
1 LAd XSlnu
qo I\<;M 0

-
_ _ 1
2 2t 2
k qo <;M
+ \ t1 2
k I\<;Mqo L" 0
dO V1 - . 211u
k2sm [38J

Therefore,

g [39]

References

1 P. J. Wojtowicz, "Introduction to the Molecular Theory of Nematic Liquid Crystals," Chapter 3.


2 P. J. Wojtowicz, "Generalized Mean Field Theory of Nematic Liquid Crystals," Chapter 4.
3 H. Zocher, "The Effect of a Magnetic Field on the Nematic State," Trans. Faraday Soc., 29, p.
945 (1933).
ELASTIC CONTINUUM THEORY 127

4 C. W. Oseen, "The Theory of Liquid Crystals," Trans. Faraday Soc., 29, p. 883 (1933).
sF. C. Frank, "On the Theory of Liquid Crystals," Faraday Soc. Disc., 25, p. 19 (1958).
6 E. G. Priestley, "Nematic Order: The Long Range Orientational Distribution Function," Chapter 6.
7 P. Sheng, "Hard Rod Model of the Nematic-Isotropic Phase Transition, Chapter 5.

8 P. G. de Gennes, Lecture Notes on Liquid Crystal Physics, Part I (1970).


9 A. Saupe, "Temperaturabhiinglgkeit und Grosse der Deformationskonstanten nematischer Fliissig-
keiten," Z. Naturforsch., 15a, 810 (1960).
10 V. Freedericksz and A. Replewa, "Theoretisches und Experimentelles zur Frage nach der Natur
der Anisotropen Flussigkeiten," Z. Physik, 42, p. 532 (1927).
11 V. Freedericksz and V. Zolina, "Forces Causing the Orientation of an Anisotropic Liquid," Trans.
Faraday Soc., 29, p. 919 (1933).
12 P. G. de Gennes, "Calcul de Is Distorsion d'une Structure Chlosteric par un Champ Magnetique,"
Solid State Comm., 6, p. 163 (1968).
13 G. Durand, l. Leger, F. Rondelez, and M. Veyssie, "Magnetically Induced Cholesteric-ta-Nematlc
Phase Transition in' Liquid Crystals," Phys. Rev. Lett., 22, p. 227 (1969).
14 R. B. Meyer, "Distortion of a Cholesteric Structure by a Magnetic Field," Appl. Phys. Lett., 14, p.
208 (1969).
Electrohydrodynamic Instabilities in Nematic
Liquid Crystals

Dietrich Meyerhofer

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

The best-known electrohydrodynamic instabilities in liquid crystals


are the Williams domains. 1 They are observed when an electric field
is applied to a thin layer of a nematic liquid crystal having negative
dielectric anisotropy and sufficient electrical conductivity. They
manifest themselves as a set of parallel straight lines separated by a
constant distance that is approximately equal to the cell thickness.
They appear above a well-defined threshold voltage and exist in their
original form over only a small voltage range. At higher voltages, the
pattern becomes more complicated and leads to heavily scattering
turbulence (dynamic scattering). 2
The structure of the domain instability is well-understood qualita-
tively. 3 The liquid flows in cylindrical motion at right angles to the
129
130 CHAPTER 9

domain walls, which represent the vortices of the motion. The hydro-
dynamic motion is not visible directly, but becomes manifest because
of the anisotropy in index of refraction. The pattern is only visible for
light polarized perpendicular to the domain walls, which means that
the director (parallel to the optic axis) is located in the plane perpen-
dicular to the walls. The flow and alignment pattern in this plane are
sketched in Fig. 1, following Penz. 3

'-------:o!<-- x

Fig. 1-Cross section through a liquid crystal cell at right angles to the domain lines.
The solid lines Indicate the director orientation n, the dashed lines represent
the flow Y.

It should be noted here that hydrodynamic instabilities are also


possible in isotropic liquids. Examples are the interaction of thermal
gradients and gravitational forces (Benard's problem 4 ) and electric-
field-driven flow under space charge limited current conditions. 5 The
latter is the likely cause of flow that has been observed in liquid crys-
tals above the nematic-isotropic transition 6 (N-L point). In contrast
to this, the instabilities discussed here depend on the anisotropic na-
ture of the material parameters.
The nature of the Williams domain instability was explained by
Helfrich who calculated the threshold voltage of domain formation
for the case of dc voltage and one-dimensional geometry.7 The calcu-
lations were extended to ac fields by Dubois-Violette 8 and to two-
dimensional geometry by Penz and FordY They show good qualita-
tive agreement with measured thresholds.
In this chapter we will extend the threshold calculations to show how
the instability arises and to combine both frequency dependence and
two-dimensional features. We will compare the results with measured
values of threshold voltage and domain spacing as function of ac fre-
quency.lO Good quantitative agreement will be demonstrated.
ELECTROHYDRODYNAMIC INSTABILITIES 131

2. Nature of the Instability and the Balance of Forces

We now calculate the threshold of domain formation following the


concepts of Helfrich. 7 We consider a thin planar cell with electrodes
on the two surfaces. We assume that the dielectric anisotropy is nega-
tive (d~ = ~II - ~1- < 0) and that the liquid crystal is aligned uniformly
parallel to the surface by suitable surface treatment. The geometry is
that of Fig. 1; the director lies in the x-z plane and is parallel to the
x- axis at the surfaces. There is no variation in the y- direction and all
variations in the x- direction are periodic with period A. In this geom-
etry, the electric field alone does not distort the liquid and any insta-
bility must be of hydrodynamic nature.
To calculate the threshold, we assume a small fluctuation of the di-
rector 8(x,z) and calculate whether the fluctuation grows or decays in
time. In the absence of applied fields, there is only an elastic torque
(r elast), which tends to restore the uniform alignment. An applied
electric field in the z-direction produces two kinds of forces. First,
there is a purely dielectric restoring torque (r diel) due to d~ < 0. Sec-
ond, the anisotropy in the conductivity produces a charge separation
(similar to the Hall effect) through which the applied field exerts a
force on the mass of the liquid causing it to flow. The nonuniformity
of the flow (shear) produces a torque on the director (r vise) that is in
such a direction as to increase the fluctuation. The magnitude of r diel
and r vise both increase as the square of the applied field so that, if
Ir vise! is larger than Ir died, there will exist a field at which I(r vise +
r diel)1 is larger than Ir elasd and the alignment will become
unstable.
In calculating the torque, we make the assumption that 8 « 7r/2.
This will lead to linearized equations and all torques will be propor-
tional to 8. This permits an accurate calculation of the threshold, but
does not allow the determination of the distortion above threshold for
which higher order terms in 8 are required.

3. Dielectric Response

Apply an external field EO = V /d. Thus, the field in the sample is


given by E = (Ex, 0, EO + Ez') where Ex, Ez' «Eo. Because \7 X E =
-aB/at = 0,

The director n lies in the (x, z) plane at. angle 8 to the x axis, so n =
132 CHAPTER 9

(cosO, 0, sinO) "" (1,0,0). The dielectric and conductivity tensors (to, IT)
are both uniaxial and parallel to n, and are given by

[2]

with an identical relationship for IT. If we keep only first order terms
in 0, Ex, E z ', we obtain

Dx = + t:.tOE z =
tllEx tllEx + t:.tOEO [3]
D z = t:.t8E x + t1E z = tlEO + tlE/,
jx = allEx + t:.aOEo [4J
jz = a lEo + alE/.

The free charge is given by

[5]

and from the charge continuity equation

[6]

Eqs. [1], [5], and [6] define Ex, E/, and p in terms of Eo and dO/dx.

4. Hydrodynamic Effects

The hydrodynamic equations describe the relationships between the


applied forces and the fluid velocities. As given by Leslie,11 in linear
approximation, they are

aa
+ "-')
I

F, L..ax = 0 r7]
J J

for the isothermal, steady-state case. Here, F is the applied body


force

[8]

and a' is the viscous stress tensor (a'ij is as defined by Penz 9 ). For an
ELECTROHYDRODYNAMIC INSTABILITIES 133

incompressible fluid, there is the additional requirement

v· v = 0, [9]

where v is the fluid velocity.


The viscous tensor as obtained from thermodynamic consider-
ations and symmetry properties is defined in terms of velocity gradi-
ents and director orientations as

a,/ = ~pO,)
.
+ alLnknIAkln,n)
kl
+ azn)N, + a 3n,N)

+ a 4 A,) + a 5Ln j nkA k, + aSLn,nkAkr [10]


k k

where

1 (au, au))
Aij = Z ax + ax, '
[11]
1
N = -z[V Xv] X n,

p is the hydrostatic pressure, and the ai's -are the viscosity coeffi-
cients. The velocity gradients are assumed small, so that the only
non-zero terms are

and

(Jx,Z' = z(-aa
1
+ a4 + as) ax +
au z 1
Z!a J + [12]
1 au, 1
= "2(-a 2 + a4 + ( ax +
5) Z! a 2 + a4

Inserting Eqs. [8J and [12J into Eq. [7], we obtain


134 CHAPTER 9

Eqs. [9], [13], and [14] can be solved for v x , vz , and p in terms of p
which was previously defined in terms of d8/dx.

5. The Boundary Value Problem In the Conduction Regime

The boundary conditions of the two-dimensional problem are that 8,


Ex, v x, and V z all must be zero at the electrodes (z = ±d/2). Penz and
Ford 9 have solved the various equations for these boundary condi-
tions in the dc case. The solution had to be obtained numerically,
which obscures the physical processes.
We can obtain a much simpler solution by relaxing one boundary
condition and letting Vx be finite at the electrode. While this can not
be the case in practice, due to viscous forces, we will show below that
the approximation is a good one. Penz 3 demonstrated experimentally
that the distortion of the director above the Williams domain thresh-
old may be described as

[15J

where q z = 7r/d, qx = 27r/A. This is the pattern plotted in Fig. 1. By


inspection, and by making use of Eqs. [1] and [9], we can write the
simplest trial functions

Ex = Ex' sin qxx cos qzz


E/ = SEx' cos qxx sin qzz
[16J
U, Su/ sin q.rX sin qzz
uz = u/ cos qxx cos qzz

where S = qz/qx = A/2d. The corresponding flow pattern is indicated


in Fig. 1 (the figure anticipates vz' being negative).
ELECTROHYDRODYNAMIC INSTABILITIES 135

Inserting the values of Ex and Ez' in Eqs. [5] and [6] and solving
for p, one obtains

Let the applied field be sinusoidal,

[18J

and assume that () and v are independent of time. This defines the
conduction regime of Dubois-Violette et al 8 in which Williams do-
mains can exist. Thus Eq. [17] is solved to obtain

[19]

where

in the notation of Dubois- Violette. T is the equivalent dielectric relax-


ation time.
Inserting Eqs. [18] and [19] into Eq. [5], one obtains

Ex' = [20]

the desired relationship between Ex and ().


Next, we insert Eq. [16] into Eqs. [13] and [14] and eliminate p to
obtain

ap
q}hl! + (cx! + 71! + 712)8 Z + 71z 84 Iuz' sin q,x cos q z 2 = - E oax'
[21]

where 711 = %( -CX2 + CX4 + CX5), 112 = %(CX3 + CX4 + CX6) are the Helfrich
viscosity parameters (Penz and Ford 9 use the same notation, but in-
terchange 711 and 712). The left-hand side of Eq. [21] is time indepen-
dent, so the right-hand side can be averaged over time (after inserting
Eq. [19]) to obtain
136 CHAPTER 9

1
-rUH E o2(q}£r sin q,x cos qzz)l + W 2T2"

Therefore,

[22]

6. The Torque Balance Equation

We can now calculate the various torques acting on the director. Be-
cause of the two-dimensional geometry, only the y- component of the
torque is non-zero. The dielectric and elastic torques are obtained
most easily by taking the functional derivative of the free energy with
respect to 0. 8

[231

The elastic free energy is given by Frank 12 as

Felast = ~[kll(V'"n? + kZ2 (n"V' x nf + k33(n x V' X n)2]

= kll (an x + an z )2 + k33 (an x _ an z)2 [24]


2 ax az 2 az ax"
Inserting the values for n, and keeping only linear terms,

a20 .+ k azO)
( k llaz2
- J3ax2
q}(k33 + S2k ll )8o sin qxx cos qzZ" [25]

This is a restoring torque (rio> 0), i.e., it tends to decrease 00 ,


The dielectric free energy is

[26]

Making use of Eq. [2] and the exact value of n = (cosO, 0, sinO), one
obtains

[27]
ELECTROHYDRODYNAMIC INSTABILITIES 137

Inserting Eqs. [15] and [20] into Eq. [27] and averaging over time, one
obtains

for Ilf < 0, r/B > 0, and this is also a restoring torque as predicted.
The viscous torque that the flow exerts on the director is given
by 9,13

r'\l~C ..\ = (Jx/ - (JZ)./

1 au, 1 au x
2(a" - a:, + an - a5~ + 2(a;1 - a 2 + a6 - ( 5)Tz
at" au
-KIa; + K2az' [29J
= q,v/(K i + S2K 2) sin q,x cos q,z I

Eoz(KI + S2K z) Ta H 0 .
1]1 + (a l + 1]1 + S2 +
1]2)
'-'4
1]2'-'
1 + W 2T 2 0 sm qr x cos q,z,

where Eq. [22J has been used. K 1 and K 2 are the combinations of vis-
cosity parameters Helfrich 7 calls the shear-torque coefficients. If the
Parodi relationship 13 (a6 - a5 = a2 + (3) is used, they become

[30J

Since K 1 is larger than K 2 and positive, r vis c/O < 0 and this can be
seen to be the driving torque.
All three torques have the same spatial dependence and are pro-
portional to B. This confirms that our set of trial functions, Eqs. [16],
are consistent. The threshold for domain formation may now be cal-
culated by setting the sum of the torques equal to zero. After some
rearrangement,
138 CHAPTER 9

where

Eq. [31] has been cast in this form to allow comparison with previous
work. If w = 0, Eq. [31] of Penz and Ford 9 is obtained «(\'1 is very small
°
or zero); if w = and S = 0, Eq. [5.2] of Helfrich;7 and if S = 0, Eq.

e w', o"/~(/ ~_____


I 63 / '~_ _ _ _ _ _ _ _
C)
~
o
7 I 717SS3~
~
«
11.
(/)

Z
6 S994~

~:
«
a:~
o
.2
~
Z
5 10 15 20 25
A PPLI ED VOLTAGE (V RMS )

Fig. 2-Plot of Eq. (32) for MBBA (Table 1): S = A/2 d versus V. The parameter Is the
reduced frequency w'. The dashed curve is the domain spacing given by
Penz and Ford. 9 Instability occurs at the point on the curve where the volt-
age Is a minimum.

[111,9] of Dubois-Violette 8 (except for a different definition of the vis-


cosity factor). We note, however, that Penz and Ford's Eq. [31] has a
different meaning, because it simply relates Eo and the various
values of S that form the more complicated solutions to the exact
boundary value problem (qz is not defined a priori). In the present
case, the form of the solution has been given in Eqs. [15] and [16] and
qz has been defined as 7f"/d. Therefore, we can rewrite Eq. [31]
ELECTROHYDRODYNAMIC INSTABILITIES 139

which relates the applied voltage V to the domain spacing A. The


threshold voltage is determined by minimizing V with respect to S.
Eq. [32] demonstrates that the relationship is independent of cell
thickness, which also appears to be the case for the exact result. 9

Table I-Material Parameters of MBBA9

kIl = 6.10 X 10- 12 newton


k33 = 7.25 X 10- 12 newton

<II = 4.72
'.L = 5.25
1/1 = 103.5 X 10- 3 kg/m/sec

1/2 = 23.8 X 10- 3 kg/m/sec

a<2 = -77.5 X 10- 3 kg/m/sec


<>3 = -1.2 X 10- 3 kg/m/sec
ull/U.L = 1.5

7. Numerical Results and Comparison with Experiment

Eq. [32] is plotted in Fig. 2 for various values of the normalized fre-
quency w' = w/w c = WT c , Tc = fll/lTll. The numerical values used are the
same as those used by Penz and Ford9 for MBBA (Table 1). For com-
parison, Fig. 2 also shows the complete solution of these authors (w =
0). As can be seen, our approximation leads to a somewhat lower
threshold, because we have neglected the viscosity force of the walls
on the liquid. Note that the threshold for the one-dimensional case (S
= 0, qx = 7r/d in Eq. [31]), 2.57 V, is considerably in error. From the
curves of Fig. 2, the values of threshold voltage and the corresponding
domain spacing are obtained as a function of voltage. They are plot-
ted in Fig. 3 as a function of frequency. Also shown are experimental
results of MBBA taken from the paper of Meyerhofer and Sussman lO
for a 10- ~m-thick cell of MBBA with the alignment parallel to the
surface. Since the conductivity had not been measured independent-
140 CHAPTER 9

ly, the experimental data were normalized to agree with the calculat-
ed threshold at the highest frequency. The agreement is very satisfac-
tory and shows that the inconsistency discussed in Ref. [10], that re-
sulted when the actual values of A are inserted in the Orsay equation,

50 10
X
~
Z
40 0
«
Ul
a.
0 (/)
l
a: 0
Z
> 30 4
:: 05:2
0<1 I
"0
> 20 o N
0 0
...J 0 W
0 N
I J
(/) 10
~x v th «
w :2
II: II:
I 0
f- 0 Z
05 10
w/=w~
0"11

Fig. 3-Threshold and domain spacing as function of w'. The lines are calculated from
the curves of Fig. 2, and the pOints are the experimental results. 10 The only
adjustment of the data was to determine Tc = ~II/O"II by fitting the measured
threshold to the calculated one at one frequency, because the value of 0"11
was not available.

is due to the one-dimensional nature of that theory, and it disappears


when the two-dimensional calculation is performed. The good agree-
ment is further demonstrated in Fig. 4, where the threshold voltage is
plotted versus domain size (Fig. 2 of Ref. [10]) and the calculated
curve has no adjustable parameters.

8. Range of Applicability

The calculations we have performed apply to the "conduction re-


gime" because of the assumption that 8 is constant in time. The range
of this regime has been discussed in detail by Dubois-Violette et alB
and shown experimentally by Meyerhofer and Sussman.l° The low-
frequency limit varies as d- 2; below this frequency both 8 and pare
ELECTROHYDRODYNAMIC INSTABILITIES 141

time dependent. The upper frequency limit, near We, is independent


of thickness; above this frequency the dielectric regime applies, where
p is constant and 8 varies.
The conduction regime as described by Figs. 2 to 4 occurs in suit-
ably doped nematic liquid crystals of negative dielectric anisotropy.
This can be seen qualitatively by studying the denominator of Eq.
[32]. The first (viscous) term must be larger than the second (dielec-

°
tric) term for the denominator to be positive and instability to occur.
This restricts negative values of Llf to the range -2 to for materials

/ Or-----------------------------,

05

0.2

5 /0 20 50

THRESHOLD Vth

Fig. 4- The data of Fig. 3 plotted as domain spacing versus threshold with frequency
as parameter. The crosses are the experimental results. 1o There are no ad-
Justable parameters.

similar to MBBA and PAA (K tini = 0.75, UI!/u 1- = 1.5 > fll/f1-). For
such materials, the threshold increases strongly with frequency, lead-
ing to a cutoff of the conduction regime near WT e = 1.
Eq. [32] shows that instability can also take place for Llf > 0, as dis-
cussed by Dubois-Violette8 and Penz.14 In that case, the threshold
voltage becomes independent of frequency at high frequency, but this
is not a conduction regime because V th is then only dependent on k
and f. Furthermore, this threshold is higher than that of the dielectric
Freedericksz transition I5 for the case of the entire sample deforming
uniformly (8 = 0, Vth = 1rvk l1 !Lld, and so will not occur. In fact,
142 CHAPTER 9

Williams domains will occur only over the range of Ll~ from 0 to ap-
proximately 0.5. These considerations are in agreement with experi-
mental observations. 16

References

1 R. Williams, "Domains in Liquid Crystals," J. Chem. Phys., 39, p. 384 (1963).


2 G. H. Heilmeier, L. A. Zanoni, and L. A. Barton, "Dynamic Scattering: A New Electrooptic Effect in
Certain Classes of Nematic Liquid Crystals," Proc. IEEE, 56, p. 1162 (1968).
3 P. A. Penz, "Voltage-Induced Velocity and Optical Focusing in Liquid Crystals," Phys. Rev. Lett.,
24, p. 1405 (1970); "Order Parameter Distribution for the Electrohydrodynamic Mode of a Nematic
Liquid Crystal," Mol. Cryst. Liq. Cryst., 15, p. 141 (1971).
• S. Chandrasekhar, Hydrodynamic and Hydromagnetic Stability, Clarendon Press, Oxford, England
(1961).
5 N. Filici, "Phenomenes hydro et Aerodynamiques dans la Conduction des dielectriques Fluids,"
Rev. Gen. Elect., 76, p. 717 (1969).
6 H. Gruler and G. Meier, "Correlation between Electrical Properties and Optical Behaviour of Nema-
tic Liquid Crystals," Mol. Cryst. Liq. Cryst., 12, p. 289 (1971).
7 W. Helfrich, "Conduction-Induced Alignment of Nematic Liquid Crystals: Basic Model and Stability
Considerations," J. Chem. Phys., 51, p. 4092 (1969). .
8 E. Dubois-Violette, P. G. deGennes, and O. Parodi, "Hydrodynamic Instabilities of Nematic Liquid
Crystals under AC Electric Fields," J. Phys. (Paris), 32, p. 305 (1971).
9 P. A. Penz and G. W. Ford, "Electromagnetic Hydrodynamics of Liquid Crystals," Phys. Rev., A6,
p. 414 (1972).
10 D. Meyerhofer and A. Sussman, "Electrohydrodynamic Instabilities in Nematic Liquid Crystals in
Low-Frequency Fields," Appl. Phys. Lett., 20, p. 337 (1972).
11 F. M. Leslie, "Some Constitutive Equations for Anisotropic Fluids," Quart. J. Mech. Appl. Math.,
19, p. 357 (1966); and "Some Constitutive Equations for Liquid Crystals," Arch. Ration. Mech. Anal-
ysis, 26, p. 265 (1968).
12 F. C. Frank, "On the Theory of Liquid Crystals," Discuss. Faraday Soc., 25, p. 19 (1958).
13 O. Parodi, "Stress Tensor for a Nematic Liquid Crystal," J. Phys. (Paris), 31, p. 581 (1970).
14 P. A. Penz, "Electrohydrodynamic Solutions for Nematic Liquid Crystals with Postivie Dielectric
Anisotrophy,"Mol. Cryst. Liq. Cryst., 23, p. 1 (1973).
15 V. Freedericksz and W. Zwetkoff, "Uber die Einwirkung des Elektrischen Feldes auf Anisotrope
Flussigkeiten. II. Orientierung der Flussigkeit im Elektrischen Felde," Acta Physiocochim. USSR, 3, p.
895 (1935).
16 H. Gruler and G. Meier, "Electric Field-Induced Deformations in Oriented Liquid Crystals of the
Nematic Type," Mol. Cryst. Liq. Cryst., 16, p. 299 (1972); W. H. deJeu, C. J. Gerritsma, and T. W.
Lathouwers, "Instabilities in Electric Fields of Nematic Liquid Crystals with Positive Dielectric Anisot-
ropy: Domains, Loop Domains, and Reorientation," Chem. Phys. Lett., 14, p. 503 (1972).
The Landau-de Gennes Theory of Liquid
Crystal Phase Transitions

Ping Sheng and E. B. Priestley

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

A physical system in which phase transition(s) can occur is usually


characterized by one or more long range order parameters (order pa-
rameter for short). For example, in nematic liquid crystals the order
parameter is the quantity S =
(P2(COS 8» as defined in previous
chapters;1-3 in ferromagnets the order parameter is the magnetiza-
tion in a single domain; and in liquid-gas systems the order parame-
ter is the density difference between the liquid and gas phases. In
each of the above cases the state of the system, at any fixed tempera-
ture, can be described by an equilibrium value of the order parameter
and fluctuations about that value. A phase transition can be accom-
panied by either a continuous or a discontinuous change in the equi-
librium value of the order parameter when the system transforms
from one phase to the other. (For simplicity we will consider temper-
ature as the only thermodynamic variable in this paper; the pressure
depedence of the various phenomena will be neglected).
From the above discussion it is apparent that an essential element
in the theory of phase transitions is the determination, at every tem-
143
144 CHAPTER 10

perature, of the equilibrium value and fluctuation amplitude of the


order parameter. In principle this can be accomplished if one can cal-
culate the free energy F[ti, T] of the system when it is in a state char-
°
acterized by a value of the order parameter and a temperature T.
F[ fJ, T] is sometimes called the Landau free energy function to distin-
guish it from the total free energy of the system. The equilibrium
value of the order parameter is that value fJeq(T) which minimizes
F[o, T]. The relative probability for the system to fluctuate to a state
ch~racterized by fJis proportional to expl-~(F[fJ,T] - F[fJeq(T),T])I,
where ~ = 1/hBT, hB being the Boltzmann constant. So far we have
neglected spatially nonuniform fluctuations of the system, i.e., fluctu-
ations that produce spatial variation of the order parameter. To in-
clude such fluctuations in the theory, one usually first calculates a
free energy density at each temperature as a function of the order pa-
rameter and its spatial derivatives. The free energy for an arbitrary
spatial variation of the order parameter, 0 (r), where r denotes spatial
position, is then obtained from the free energy density by a volume
integration. Thus, in the general case, the theoretical description of a
phase transition is equivalent to the determination of the free energy
density as a function of the order parameter, its spatial derivatives,
and the temperature. However, a rigorous determination of the free
energy density function is extremely difficult since it is almost equiv-
alent to the complete solution of the phase transition problem.
In 1937 Landau4 made an elegant and far-reaching speculation
about the functional dependence of the free-energy density on the
order parameter and its spatial derivatives near a second-order phase
transition point. Briefly, Landau speculated that near a second-order
transition the free-energy density function can be expanded as a
power series in the order parameter and its spatial derivatives, with
temperature dependent coefficients. Landau further argued that, suf-
ficiently close to the transition, only the leading terms of the series
are important, so that the expansion of the free-energy density func-
tion becomes a simple low-order polynomial. Despite certain short-
comings,5 Landau's theory of phase transitions has proven to be as
good as mean field theory in providing a semi-quantitative descrip-
tion of the specific heat, the order parameter, and the entropy in the
vicinity of a second-order phase transition. Moreover, the Landau
theory is mathematically simpler than mean field theory, and the in-
clusion of spatial variations of the order parameter gives it a new di-
mension not found in mean field theory. Of course, the Landau theo-
ry is useful only in a limited temperature range close to the transition
point, and it contains more phenomenological parameters than does
LANDAU-deGENNES THEORY 145

mean field theory. In these respects it is somewhat less satisfying


than mean field theory.
Although originally intended as a theory of second-order phase
transitions, the Landau theory can easily be generalized to include
first-order phase transitions. 6 de Gennes 7 was the first to successfully
apply Landau's theory to the first-order liquid-crystal phase transi-
tions. It is the purpose of the present chapter to develop this Landau-
de Gennes theory of liquid-crystal phase transitions and to discuss
and illustrate its use. In the following sections, the derivation and
discussion of the basic equations will be followed by application of
the theory to the calculation of thermodynamic properties and fluc-
tuation phenomena of liquid-crystal phase transitions, and by a de-
scription of some of the theory's more novel predictions and their ex-
perimental verifications.

2. Derivation of the Fundamental Equations of the Landau-de


Gennes Theory

2.1 The Partition Function

We begin by considering a macroscopic system whose equilibrium


state is characterized by a spatially invariant, dimensionless, scalar
order parameter 0'. Any disturbance in the system, such as thermal
fluctuations, produces spatial variations of the order parameter.
However, for low-energy (long wavelength) fluctuations, the spatial
variations occur on a scale much larger than the molecular dimension.
Therefore, if we limit our consideration only to the equilibrium state
and low-energy fluctuations about the equilibrium state, we can
imagine dividing the system into, say, M small, cubic, spatial regions
each of volume Ll V. Each of these regions contains a sufficient num-
ber of molecules so that long-range order is well defined inside the re-
gion, yet is small enough compared to the wavelength of low-energy
fluctuations that spatial variation of the order parameter within the
region is negligible. The partition function for one such region, say re-
gion a, is
[1]

where Ei(a) is the energy of region a (excluding its interaction with


the rest of the system) when it is in state i (a state is defined in quan-
tum systems by the eigenstate of the system and in classical systems
by a set of numbers giving the spatial positions and momenta of all
146 CHAPTER 10

the particles). In general, the summation extends over all possible


states of region a; however, we can make a simplifying approximation
since we are only interested in low-energy fluctuations, which pro-
duce negligible spatial variation of the order parameter in region a.
Thus out of all the possible states i of region a, we only sum over
those states p which can be characterized by a single value of the
order parameter. The resulting approximation to.} (a,T) will be de-
noted by Z(a,T). We now divide the possible values of the order pa-
rameter into small intervals of width ~u(a) and group together all
those states having values of the order parameter in the same inter-
val. By denoting the number of states in the interval centered at u(a)
as N[u(a)], it follows that

n( a, T) ~ Z( a, T) = Lp exp[-f3Ep( a)]
N [a( ad] l
= L
a(o<)
{
~
J,,1
ex p(-f3 E Aa(a)])r, [2]

where Ej[u(a)] is the energy of one of the states having an order pa-
rameter value within the interval centered around u(a). Note that
N[u(a)] depends on the width of the interval ~u(a) and, though re-
gion a is small, it still contains a sufficiently large number of mole-
cules that N[u(a)] must be very large in any finite interval ~u(a).1t is
therefore meaningful to define a density of states p[ u(a)] such that

( )] .
11m N[ a( a)]
p [a a = ( ) . [3]
aa(o<)-O ~a a

By reducing ~u(a) to a differential, du(a), and by assuming that the


energy of any state p in region a can be expressed as a continuous
function, E[u(a)], of the order parameter, Eq. [2] can be accurately
replaced by

Z( a, T) = f da( a)p[ a( a)] exp{-f3E[ a( a)]}, [4]

where the integral extends over all possible values of the order pa-
rameter.
We now define a Landau free energy function J[u(a),T] of region
a, when it is characterized by the value u(a), as

j[a(a), T] :: E[a(a)] - kBT In p[a(a)]. [5]


lANDAU-deGENNES THEORY 147

In terms of J[ u( a), T], Eq. [4] can be expressed as

Z( a, T) =j da( a) exp{-{3j[ a( a), Tn, [6)

and the total free energy of region a, F(a,T), is given by the usual
relation

F( a, T) = -kB T In Z( a, T) . [7)

Let us now consider the interaction energy between two neigh-


boring regions, say regions a and a + 1. When the order parameter
characterizing a physical system deviates from spatial uniformity,
there is always a restoring force tending to bring the system back into
spatial uniformity. It is therefore plausible to treat the interaction
between two neighboring regions as being elastic. That is, the interac-
tion energy I between regions a and a + 1, when they are character-
ized by the order parameter values u(a) and u(a + 1), is, to a first ap-
proximation, a function of the difference between u(a) and u(a + 1).
I has the property

I[a(a) - a(a + 1), T] =

o for a( a) a( a + 1)
[8)
positive and
increases with
Ia( a) - a( a + 1) I for a( a) "* a( a + 1),

where I I denotes the absolute value. The combined Landau free en-
ergy of regions a and a + 1, when they are characterized by u(a) and
u(a + 1), is then given by f[u(a),T] + J[u(a + l),T] + I[u(a) - u(a
+ l),T]. The complete partition function for the two regions, includ-
ing the interaction, can be written as

Z(a, a + 1, T) = jda(a)jda(a + 1) exp{ -(3(j[a(a) , T] +


j[a(a + 1), T] + I[a(a) - a(a + 1), T])}. [9)

It should be noted that if I = 0, Eq. [9] gives the expected result


F(a,a + 1,T) = -kBT In Z(a,a + 1,T) = F(a,T) + F(a + 1,T) for
two noninteracting regions. For the purpose of simplifying the count-
ing, we will define the quantities r(a,T), r(a + 1,T) such that
148 CHAPTER 10

1
r(a + 1, T) = rea, T) = 21[a(a) - a(a + 1), T). [10]

In other words, the interaction energy between two neighboring re-


gions is split equally and counted twice: once as belonging to region a
and once as belonging to region a + 1. Eq. [9] can now be put in the
form

Z(a, a + 1, T) = fda(a)jda(a + 1) exp {-{3(i[a(a) , T) +


j[a(a + 1), T) + rca, T) + r(a + 1, T))}. [9a]

We can easily generalize Eq. [9a] to obtain the partition function


Z(T) of the whole system, where each region interacts elastically with
its neighbors:

Z(T) = fda!1)Jda(2) . . Jda(AI) ex p [ -(3 ~ (j[a(a), T] + r(a, T))]

= fD{a(1).a(2), . . . a(l\ll)} ex p [-{3 ~ (j[a(a),T] + r(a,T))J, [11]

where r(a,T) is equal to half the interaction energy of region a with


its neighbors and is understood to depend on the difference in order
parameter values between region a and its neighboring regions.la(l),
a(2), ... , a(M)1 denotes a set of M numbers the value of each of
which is bounded by the maximum and minimum values of the order
parameter, and the integral over D!a(l), a(2), ... a(M)1 means inte-
gration over all possible sets of M numbers with the above constraint.
In Fig. 1 we illustrate schematically the equivalence of integration
over each a(a) individually and integration over all possible sets la(l),
a(2), ... , a(M) I.
It is convenient at this point to change notation somewhat. We will
henceforth label each region by the spatial coordinate r", of its center.
Since the volume of each region is ~ V, we can define a free energy
density,

j[ a( a), T) [12]
DoV
and an interaction energy density,

[13]
LANDAU-deGENNES THEORY 149

Since individual regions are small compared to the entire system, it is


the usual practice to regard each region as a spatial "point" in the
system. Thus u(ra) ..... u(r) and f[u(ra),T] ..... f[u(r),T], where u(r)

~ OF
REGION 2

~mox ------,---------------------,

{~(1).~(2)}

~ mi n -- - -- -I-----------------------i

~mox ~ OF
REGION I

Fig. 1-Schematic illustration of the equivalence of integrating over du(1) du(2) and
summing over all possible sets of numbers lu(1). u(2)1. A set of numbers lu(l).
u(2)1 correspunds to a point inside the square as shown; summing over all pos-
sible sets assures that the area of the entire square is included. Generalization
to any arbitrary number of variables is straightforward.

and f[ u(r), T] are defined at any arbitrary spatial "point" r. Recall


that, to a first approximation, r(a,T) depends only on differences in
the order parameter values between region a and its neighboring re-
gions, so that 'Y(rmT) can be replaced* by 'Y[t7u(r),T]. It therefore
follows that

L
M

0!=1
~
{f[a(QI) , T] + r(QI, Tn L
M

0!=1
{f[ a(rO!)' T] + y(rO!, T)}~ v

----- S volUll1e
tPr{f[a(r), T] + y[va(r), Tn, [14]

of the
sample

and each set of numbers lu(l), u(2) ... u (M)l is uniquely replaced by
• Theoretically the interaction term should depend on all orders of spatial derivatives of the order pa-
rameter. However. later development will show that one needs only retain the first order spatial de-
rivatives of uU) to obtain tlle Landau expression. Therefore. for simplicity. the function 'Y will be as-
sumed to depend only on V' u( ~.
150 CHAPTER 10

a spatial function u(f). Eq. [11] can now be rewritten as

Z(T) = f Da(r) exp{ -(3 Id 3r<f[a(y) , T] + y[Va(r) , T])}


[15]

where we have used Feynman's path integral notation8 Du(f) to de-


note integration over all possible functions u(f). Eq. [15] is the cen-
tral result of this subsection. Some of the implications of Eq. [15] are
discussed in Appendix A.

2.2 The Landau Expansion

Calculation of the physical properties of a spatially uniform, macro-


scopic system requires information about f[ u, T] only in the immedi-
ate vicinity of its minimum at temperature T, due to the sharpness of
the peak in the function expl-;3V{[u,T] I (See Appendix A). This
implies that for a physical system having a second-order phase transi-
ton at T = Te, with u = 0 for T > Te and u ~ 0 for T < Te (such as
the ferromagnetic phase transition), many features of the transition
can be deduced if {[ u, T] is known only in the neighborhood of u = 0,
T = Te. Landau speculated that the first few derivatives of the func-
tion f[ u, T] with respect to u exist, and that they have finite values
when evaluated at u = 0, T = Te. For a second-order phase transition
the value of u varies continuously and can be arbitrarily small near T
= Te; therefore, the Landau assumption enables one to express f[ u, T]
near T = Te as 4

f[a, T] = fo[T] + A(T)a + 21 A (T)a 2 +


1 B(T)a 3 + ~ C(T)a 4 + ... [16]

Various properties of the expansion coefficients A, A, Band C can be


obtained from quite general considerations.
For definiteness we will examine the expansion in relation to two
physical systems: (1) the CuZn (;3-brass) binary alloy and (2) a ferro-
magnet such as iron. It is well known9 that CuZn has a second-order,
order-disorder transition at Te = 742°K. The crystal structure at
OOK can be described by two interpenetrating simple cubic lattices
each with No sites. Let us suppose that lattice 1 is occupied by Cu
atoms and lattice 2 by Zn atoms. As the temperature is raised above
LANDAU-deGENNES THEORY 151

OaK, some Zn atoms will be found on lattice 1 and some Cu atoms on


lattice 2 but, so long as the temperature is less than T e, Ncu(1)/No >
%, where we have denoted the number of copper atoms on lattice 1 by
N Cu (1). For temperatures in excess of T e, complete randomization
occurs and Ncu(I)/No = Nz n (1)/No = %. A suitable order parameter
for the system can be defined as

[17]

which is zero for T ~ Te and in the range 0 to 1 for T < Te. It is also
obvious that the state characterized by -u = [Ncu(2) - Nz n (2)]lNo
is physically equivalent to the state characterized by +u because the
difference between the two can be ascribed to the interchange of la-
belings for lattices 1 and 2. It follows therefore that the free-energy
density for -u must be equal to that for +u and hence A(T) = B(T)
= O. In fact, all the terms having odd powers of u must necessarily
have vanishing coefficients. This is a general result not limited to the
CuZn system alone. Consider the ferromagnet, iron. In this case the
order parameter is given by a vector m defined as

[18]

where M is the magnetization vector and Mo is the magnitude of the


saturation magnetization at OaK. The equilibrium state is character-
ized by m = 0 for temperatures above the Curie point and by 0 < Iml
~ 1 for temperatures below the Curie point. In exact analogy with the
case of a scalar order parameter, the free energy density can be ex-
panded in terms of m. However, since the free energy density is a sca-
lar quantity, the Landau expansion of {[m,T] about m = 0 can only
contain scalar combinations of m. It is therefore obvious that A(T) =
O. B(T) must also vanish since it is impossible to construct a scalar
from three vectors. As before, we see that the coefficients of the odd
order terms in the expansion vanish. The above arguments can be
similarly applied to other examples of second-order transitions. lO •n
In fact, for a second-order phase transition, one can always define an
order parameter such that its equilibrium value is zero for T ~ Te
and nonzero for T < Te and such that A(T) = B(T) = 0 in the Land-
au expansion of {[ u, T].
For a second order phase transition, Eq. [16] thus takes the form

f[a, T] ~ fo[T] + ~A(T)a2 + iC(T)a4 + .... [19]


152 CHAPTER 10

At this point Landau4 assumed further that near Te the coefficient


C(T) is a slowly varying function of T compared to A(T) and there-
fore can be replaced by a constant C > 0 in the temperature range of
interest. Since the order parameter is zero at equilibrium in the high
temperature phase, it is necessary that u = 0 be a minimum of I[ u, T]
for T > Te and thus A(T) must be positive. However, in the low tem-
perature phase nonzero values of u must correspond to the minimum
in t[u,T]. This minimum occurs at u = ±[-A(T)/C(T»)1/2 provided
A(T) is negative. Therefore A(T) must be positive for T > Te and
negative for T < T e, which implies that A(T = Te) must be zero and,
in the neighborhood of T e, A(T) can be approximated by

[20]

where a is a positive constant. We can now rewrite Eq. [19] as

a,e > o.
[21]

Eq. [21] is illustrated schematically in Fig. 2.


So far we have considered only spatially uniform systems. General-
ization of Eq. [21] to include spatial variations of the order parameter
involves: (1) replacing u by u(f), and (2) including the contribution to
the free-energy density due to the interaction term I'[~ u(f), T]. Fol-
lowing Landau,4 we expand 1'[~u(f),T] in a power series in ~u(f) and
retain only the leading terms. Again, the expansion can contain only
scalar combinations of ~ u(f) and, since I'[~ u(f), T] = 0 when ~ u(f) =
0, we obtain*

[22]

In order that the spatially uniform state be the state of lowest free
energy, D(T) must be positive. Furthermore, near the critical tem-
perature, D(T) can be approximated by a constant D. Combining the
leading terms in the expansions of t[u(f),T] and 1'[~u(f),T], we ob-
tain

• TheoreticaJly, expansi9n of the interaction term can contain second order scalar combinations
such as 0"(,)V' 2 0"(,) and V' 2 0"(,). However, V' 2 0"(,) integrated over volume can be converted into a sur-
face integral and call be neglected since the surface contribution is assumed to be small. The vol-
ume integral of 0"(i')V' 2 0"(,) is equivalent to the integral of IV' O"(ilj2. Therefore, even in the most gen-
eral case, the leading term in the expansion is proportional to IV' 0"(,) F.
LANDAU-deGENNES THEORY 153

f[o{r),T] + Y[Va(r),T] ~ ;JL

1 ) 2~ 1 4~ 1 ~ ~ 2
== fo[T] + "2 a(T - Tc a (r) + 4Ca (r) + "2D['Va(r)] , [23]

where g; L is usually referred to as the Landau free energy density and


a, C, and D are all positive constants.

I
-.6 -.4 -.2 -.01 .2 .6
-.02

Fig. 2-lIIustration of the behavior of Eq. [21]. The Landau free energy density is sym-
metric about (J = 0; for T ~ Tc the disordered state «(J = 0) is the stable state
while for T < Tc the ordered state «(J ~ 0) is the state of lowest free energy.

2.3 Generalization of the Landau Expansion to Liquid Crystals

We must first identify a suitable order parameter for describing liq-


uid crystal phase transitions. In previous chapters, we have seen that
for nematic and cholesteric liquid crystals, the molecular ordering at
every spatial "point" f (where the term "point" has the meaning de-
fined in Section (2.1» is characterized by a director f2.(f) pointing
along the local axis of uniaxial symmetry, and by a quantity S(f) giv-
ing the local orientational order ofthe rod-like molecules. S(f) is de-
fined by

S(r) = « P2 (cos e) »;, [24]

where () is the angle between the long axis of any molecule in the
154 CHAPTER 10

small region of space associated with "point" r and the local director
ti(r), P2 is the second Legendre polynomial, and« »-;- denotes spa-
tial averaging over the configurations of the molecules at "point" r
(<< »-;- should not be confused with thermal averaging < >, which
gives the equilibrium value of a quantity. 8(r) should therefore be
distinguished from its equilibrium value <8(r», which is spatially
invariant and temperature·-dependent). In Chapter 6 it was shown
that in a coordinate system where ti(r) coincides with the external z-
axis (3-axis), the local order parameter of a nematic or cholesteric liq-
uid crystal is given by*
1
2
o o
1
S(r) o 2
o [25]

001

where Q(r) denotes that the quantity Q(r) is a tensor. For arbitrary
orientation of ti(r) relative to the external coordinate system, the
components of Q (r) can be expressed as

[26]

where i,j = 1, 2, 3 denote the components along the three orthogonal


axes of the Cartesian coordinate system and Oij = 1 for i = j and zero
otherwise. Comparing the order parameters for a ferromagnet and a
liquid crystal, we see that 8(r) corresponds to the magnitude of m,
and the matrix in Eq. [25] corresponds to the unit vector pointing
along the direction of magnetization; this matrix can be thought of as
a "unit tensor." To see that the order parameter for a liquid crystal
cannot be a vector, we observe that ti(r) and -ti(r) correspond to
physically equivalent states and Q(r) must therefore be proportional
to an even order combination of ti(r). The two lowest even-order
combinations of a unit vector are a scalar and a second rank tensor
and, since the liquid-crystal order cannot be completely described by
a scalar, we are left with the tensor as our only choice.
Let us now consider the Landau free-energy density expression for
liquid crystals. All isotropic-nematic (cholesteric) phase transitions

• Our definition of O(i) differs from that of de Gennes by a factor of 2.


LANDAU-deGENNES THEORY 155

are first order, and exhibit a discontinuous jump at T = Tc in the


equilibrium value of Q(f) from Q(f) = 0 in the isotropic phase to
some finite value in the low temperature phase. Consequently, we ex-
pect the Landau expansion about Q (f) = 0 to provide a better de-
scription of phenomena such as fluctuations in the high-temperature
isotropic phase than it does in the low-temperature phase, since the
expansion would not be accurate for the values of the order parame-
ter in the low-temperature phase. We will first develop the Landau-
de Gennes expansion about the isotropic phase of a nematic liquid
crystal and later generalize the expression to include the effects of ex-
ternal fields and the special symmetry of cholesteric liquid crystals.
Since the expansion can contain only scalar combinations of Q (f) and
its spatial derivatives, the term linear in Q(f) again vanishes because
3
the scalar L Qii(f) (the trace of Q(f» is identically zero. The
i=1
coefficient of the term linear in the spatial derivative of Q(f) must
also be zero because there is no way of forming a scalar quantity from
the derivative. However, unlike the situation described for ferromag-
netism and other second order phase transitions, the term cubic in
Q(f) does not have to vanish because it is possible to construct a sca-
lar from three tensors, and also because Q(f) and -Q(f) correspond
to physically different states, as illustrated in Fig. 3. Thus the free

1\
n

-- .
-
::::-

00)
S=.5 S=-.5

.... _.L
2.
O0) .... _.1
2.
0=.5 ( 0-2 0
00. 0=-.5 ( 0-1 0
00.
Fig. 3-Schematic illustration of the physical difference in molecular ordering for S=
0.5 and S = -0.5. The short lines represent projections of the molecular long
axes on the plane of the paper. Thus a molecule whose long axis is normal to
the paper is represented by a dot.
156 CHAPTER 10

energy density is no longer required to be symmetric about Q(f) = O.


Recalling that the trace of a matrix or of a product of matrices is al-
ways a scalar, we may write

fJ L =fo[T] + ~a(T - Tc*)QijCr)QjjCr) +

~BQij(y)Qjk(y)Qklr) + i C t[Qilr)Qj/r)]2 +

i C 2Qij(r)Q jk(r)QkZ(r)Q liCr) +

1 ~ ~ 1 (~) (~)
"2Lt8iQjk(r)8iQjk(r) + "2L28iQij r 8kQkj r , [27]

where Te* is a temperature slightly below T e, a > 0, B, Cl, C 2, L 1, L2


are constants; i, j, k, l = 1,2,3 denote the components along the three
orthogonal axes of the coordinate system; ai = a/aXi is the partial de-
rivative with respect to spatial coordinate Xi; and summation over re-
peated indices is implied. Two points regarding Eq. [27] are worthy of
special attention. First, because the order parameter is a tensor, ff L
contains two fourth-order terms and two spatial derivative terms, in
contrast with Eq. [23] where the order parameter is a scalar. Second,
a new phenomenological parameter Te* has been defined as the tem-
perature at which the curvature of ffL at Q(f) = 0 changes sign. Since
Te* < T e, we can find a temperature range, Te* < T < T e, in which
Q (f) = 0 is not the position of the absolute minimum of ff L (since
otherwise Q(f) = 0 would be the equilibrium state for T < T e, which
is a contradiction) and yet the curvature of ff L at Q (f) = 0 is positive.
In other words, Q(f) = 0 is a relative minimum of ffL for Te* < T <
Te. Since the existence of a relative minimum of ff L is a necessary
condition for supercooling, it follows that Te* can be interpreted
physically as that temperature below which supercooling becomes
impossible.
Eq. [27] can be put into a more physically interpretable form by
substituting Eq. [26] for Qij(f) and noting that

~ ~
n i ( r) 8 jn i ( r) = "21 8 j
[2~
nt (r) + n2 2~ 2~) 1 []
(r) + n3 (r) ="2 8 j 1 = 0

and

[n (r) • v]n(r) = ~ \7[ nCr) • ner)) - nCr) x [\7 x n(r)]


= - nCr) x [\7 x n er)) .
LANDAU-deGENNES THEORY 157

We obtain

5'L = fo[T] + 43 a(T - Tc*)S (r) +


2 ~
41 B ,Y(r)
<'.1 ~
+

~6 cster) + %L1[VSCr)]2 + ~ L1S2(r)(oin/r))(oin/r)) +


~L2[VS(r)F + ~L2[~(r). vSCr)F + ~L2S2Cr)[vo nCr)]2 +
%L 2S(r)[V. nCr)][lt(y) VSCr)] + 0

%L 2S(Y)[n(r) X (V X n(r))]. VS(y) +

~L2S2(r)[ncr) X (V X nCr))]2, [28]

where C - C1 + (C 2 /2). Eq. [28] can be further reduced by noting


that

(oin/r))(oin/r)) = [V onCr)]2 + [~rr)o (\7 X nCr))]2 +


[nCr) x (v X nCr)) F - v. [nCr)(v. nCr)) + nCr) X vx n(r)],
[29]

where the last term represents the surface contribution to the free-
energy density and therefore can be neglected (since the volume inte-
gral of~. VCr), where VCr) is an arbitrary vector field, can be convert-
ed to a surface integral by Gauss's Theorem). Substitution of Eq. [29]
into Eq. [28] yields

5=L =fo[T] + %a(T - Tc*)S2(r) + iBS3(r) +

196CS4Cr) + %(L1 + ~L2)[\7SCY)]2 + ~L2[n(11o \7S(;')]2 +

~S2(r){(L1 + 4L2) [V.n(r)]2 + L 1[n(r). vx n(r)]2 +

(L1 + 4L2 ) [nCr) x v X nCr)]2} + %L 2S(r)[V. nC;')] x

[n(r) VS(r)] +
0 %L 2S(r)[n(r) X v X nCr)] VSCr).
0 [30]
158 CHAPTER 10

There are four types of terms in Eq. (30). The first four terms concern
only the value of the orientational order S(r). The next two terms ac-
count for spatial variation of S(r). Next there is a term concerned
with the spatial variation of n(r); we have expressed this term in the
faminar form of splay, twist, and bend distortions 12 of the director
field n(r). It should be noted that to second order in the Landau ex-
pansion there are only two independent elastic constants, L1 and L 2 ,
whereas in the nematic phase there are known to be three indepen-
dent elastic constants.1 3 The last two terms in Eq. (30) represent the
interaction between spatial variations of S(r) and spatial variations
of n(r). Clearly, the mathematics can be quite complicated if S(r)
and n(r) are allowed to vary simultaneously.
Let us now examine the expansion coefficients B, C, L1. and L2 in
more detail. We will consider first the spatially uniform state which
can be described by the first four terms of Eq. (30). Following Land-
au, we choose C > 0, which requires that B < 0 if the equilibrium
value of S is to be positive in the low-temperature phase. We illus-
trate the behavior of g: L as a function of S for a spatially uniform sys-
tem in Fig. 4. Next we consider a state for which S(r) is constant but
n(r) is allowed to vary from point to point. Since the spatially uni-
form state must be stable against any distortion, it is clear from Eq.
(30) that

[31]
and

[32]

If now we fix n(r) and let S(r) vary, the same reasoning leads to the
ineq uali ties

1
L1 + 6 L2 > 0 for nCr) 1 VS(r) [33]

and
2
L1 + 3 L2 > 0 for nCr) II VS(r) . [34]

In order to clarify the physical meaning of L1 and L 2 , imagine a dis-


turbance of the equilibrium state Q(r) = 0 of the high temperature
LANDAU-deGENNES THEORY 159

phase such that S(f = O),e 0 along a certain director n. If L1 = L2 =


0, we see from Eq. [30] that the disturbance would be a delta function
at f = 0 since the rest of the system, including the immediate region

.2
aTe*
0\.1..-1152-0253 +035 4
Te* . .

T-1.033Te*oTc
.010
ToTc'"

.008

-.4

-.002

-.004

-.006

-.008

Fig. 4-lIIustration of the asymmetry in the Landau free energy density when a non-zero
third-order term is included. For T> Te. S = 0 is the absolute minimum. As T
approaches Te from above a second minimum appears and for T = Te the
physical states corresponding to the two minima have the same free energy
density and are separated by a barrier of height h. For Te· < T < Te the S ,e
o minimum represents the stable state and the S = 0 minimum represents the
metastable (supercooled) state. For temperatures below Te· the metastable
state becomes unstable.

surrounding f = 0, would still prefer to be in its lowest free energy


state, S(f) = o. However, as soon as L1, L2 ,e 0, the delta function be-
comes energetically unfavorable because of the divergence of the spa-
tial derivatives, and any disturbance at f = 0 must decay in a contin-
uous manner to the equilibrium value S(f) = 0 over some region of
160 CHAPTER 10

space surrounding f = 0. The characteristic decay length is deter-


mined by the competition between the first four terms of Eq. [30],
which tend to minimize the decay length so as to reduce their free-
energy contribution, and the spatial derivative terms, which tend to
maximize the decay length so as to reduce 1 ~ 8(f)l. The decay of a dis-
turbance of the sort described above (ncf) fixed, 8(f = 0) ."r, along
n) is illustrated schematically in Fig. 5. Note that for L2 ."r, 0, the
°
L 2.0
.......... . , .
.: ..... ', ',',':............
....... : :: ::: ,'::.
. ............... ,..
. ...............
.....................
......... ..............
.......................
................ ...........
:i iii! iii;::\ l·!>/iW
.:;:::::::':~}
{{iii
I:.'.: ;;::::::
............... .. .. .. .. ... .
.........................
,
.........................
.......................
............... - .......
- ....... ..... ...... ................
....... .......... .
..............
............ ..
( 0) (b) (e)

Fig. 5-lIlustration of the spatial anistropy in the decay of a disturbance in the local order
for different values of L2 . Regions of order are represented by double-headed
arrows whose length is proportional to the magnitude of the local order and
whose direction is parallel to the local director. The dots represent the sur-
rounding isotropic field. In (a), L2 > 0 and the disturbance (fluctuation) relaxes
faster in the direction perpendicular to the director, whereas in (c), L2 < 0 and
the relaxation is faster in the direction parallel to the director. In (b), L2 = 0 and
the decay pattern is isotropic.

decay length of the disturbance parallel to n is different from that


perpendicular to n. Specifically, spatial variation of 8(f) in a direc-
tion parallel to n contributes a free energy density

whereas spatial variation of 8 (f) perpendicular to n contributes a


free energy density

Therefore, if L2 > 0, the elastic constant governing the spatial varia-


LANDAU-deGENNESTHEORY 161

tion of S(r) in the direction parallel to n is larger than that in the di-
rection perpendicular to n. Indeed, the elastic constant along a direc-
tion which makes an arbitrary angle I/; with n is proportional to

The decay pattern of a disturbance for L2 > 0 is shown in Fig. 5a,


while that for L2 < 0 is shown in Fig. 5c.
Now, a word about the magnitude of the expansion coefficients a,
B, C, L 1 , and L 2 . Taking the intermolecular interaction energy (~
0.01 eV) as the characteristic energy,12 the intermolecular separation
(~ 10 A) as the characteristic length, and Tc ~ 300 0 K as the charac-
teristic temperature of nematic and cholesteric phases, we can esti-
mate the magnitudes of these coefficients using dimensional analysis.
The coefficient a has units of energy/(volume·oK). Combining the
characteristic values in the appropriate way leads to a ~ 0.005 J/
(cm 3 OK), compared to the measured value 14 of 0.042 J/(cm 3 OK) for
MBBA. Band C have units of energy/volume which gives IBI, ICI ~
1.6 J/cm 3 , in order of magnitude agreement with measured values 14
~ 0.5 J/cm 3 • Finally, we obtain IL11, IL21 ~ 10- 7 dyne which is in rea-
sonable agreement with the experimental value 15 of 10-6 dyne.
In the remainder of this section we will generalize Eq. [30] to in-
clude the effects of external magnetic and electric fields and also to
take account of the finite pitch in the equilibrium state of cholesteric
liquid crystals.
Suppose that a static magnetic field fI is applied to the system.
Since nematic and cholesteric liquid crystals are normally diamagnet-
ic, the energy density induced by the magnetic field is given by

[3~]

where Xij are the components of the susceptibility tensor in units of


susceptibility per unit volume. It was shown in Chapter 6 that

[36]

where I is the unit matrix, (~X)max = N(n - rJJ(N being the num-
ber of molecules per unit volume, and riM1-) the susceptibility of a
single rod-like molecule parallel (perpendicular) to its long axis), and
X is the susceptibility per unit volume when Q(r) = O. Using Eq. [13]
162 CHAPTER 10

of Chapter 6, we see that X = N (2t.L + tll)/3. Thus, Eq. [35] can be


rewritten as

Since the last term in Eq. [37] is independent of the order parameter
and its gradients, it can be absorbed into the term fo[T] of fh. There-
fore, the magnetic field contribution to the free energy density is

1 ~
= - 3 (t..X)maxHiQi/r)H j
=- i(t..X)max S(r){3[Ho n(r)]2 - H2}. [38]

Using similar arguments, we obtain the electric field contribution to


the free-energy density

where (~~)max is the maximum possible value of (~II - ~.L)' ~II(~.L)


being the dielectric constant parallel (perpendicular) to the local di-
rector.
It must be stressed that, unlike our expansion of ~h, Eq. [30],
which is valid only for vanishingly small values of the order parame-
ter, Eqs. [38] and [39] are valid for arbitrary values of the order pa-
rameter. In fact, the magnetic and electric energy density expressions
of the elastic continuum theory (Eqs. [3] and [4] of Chapter 8) are
special cases of Eqs. [38] and [39]. For example, consider the magnet-
ic energy density of the low temperature, anisotropic phase of a sys-
tem in which only spatial variation of n(r) is important. S(r) in Eq.
[38] can then be replaced by its equilibrium value <S>, where < >
denotes thermal averaging. The term proportional to <S>H2 can be
neglected because of its spatial invariance, and one obtains Eq. [3] of
Chapter 8 directly,

where ~x = (~X)max <S>.


The expression for the free-energy density can also be generalized
to include the special symmetry properties of cholesteric liquid crys-
LANDAU-deGENNESTHEORY 163

tals. Cholesteric order is indistinguishable from nematic order on a


microscopic scale (see Chapters 1 and 11). However, when we exam-
ine the spatial variation of the local order, we find that cholesteric
liquid crystals always exhibit helical ordering on a macroscopic scale
(compared to molecular dimensions). Moreover, for any given cho-
lesteric liquid crystal, the helical ordering has a definite handedness.
In other words, cholesteric liquid crystals are not symmetric under
the operation of spatial inversion, since a helix is always either left or
right handed and the handedness changes upon inversion. Since the
two cholesteric states related by spatial inversion are physically dif-
ferent,16 the free-energy density is no longer required to be invariant
under spatial inversion. This means that the free-energy density can
contain pseudoscalar terms as well as the usual scalar terms. de
Gennes found that the pseudoscalar term required by cholesteric
order is given by7
[40]

where the + and - signs refer to the two senses of helical pitch; qo =
7r/AO (> 0), A.o being the pitch of the cholesteric phase as T
-+ Tc; and

~ijkis the Levi-Cevita antisymmetric tensor of the third rank, which


has the property that

1 ijk = 123, 231, 312


Eijk ={ -1 ijk = 213, 321, 132
o otherwise

Eq. [40] can be reduced to the form

[41]

where the - sign refers to a left-handed helix (n(fH~ X n(f)] > 0)


and the + sign refers to a right-handed helix (n(r).[~ X ft(f)] < 0).
Eq. [41] can be combined with the twist term of Eq. [30]

to yield
(9/4)L 1S2(rHln(r) 0 V X nCr) I - qoF - (9/4)L 1S2(Y) qo 2,

where I I is the absolute value sign.


164 CHAPTER 10

The total Landau-de Gennes free-energy density for nematic or


cholesteric liquid crystals in an external field is given by

5" = 5"L + 5" M + 5"E + 5"c = fo [ T] +


%a(T - Tc * - 3L~q02) 5 2(11 + i B53 Cr) + 1~ Cs4Cr) +

%(L1 + i L2) [V5(;')]2 + ~ L2[~(;) 0 V5(y)]2 +

i 52 (r){ (L1 + ~ L2 )[vo n(;) F + L 1[ I ~(;)o ~ X ~(;) l-qoF +


( L1 + ~ L2 )[~Cy) X VX n(r) F} +
~ L 25(r) {[V ~(-;.) ][~(;)
0 0 V5(r)] + ~ [~(r) X VX n(;)] oV5(;)}-

i(~X)max5(r){3[Ho n(r)]2 - H2} -

1 ~ { ~ ~ ]2 2 [42]
247T (~E)max5(r) 3[Eo n(r) - E },
A

where qo = 0 for nematic liquid crystals. It should be noted that for


cholesteric liquid crystals Tc* is replaced by Tc** = Tc* + (3L 1qo2/a )
in Eq. [42]; we can interpret Tc** as the temperature below which su-
percooling of cholesterics is impossible. Eq. [42] can also be expressed
in terms of Qij(r),

1 ~ ~
5" = fo[T] + 2 a(T - Tc*)Qij(r)Qji(r) +
1 ~ ~ ~ 1 ~ ~. 2
3 BQij (r)Qjk(r)Qki(r) + 4 C [Qij.(r)Qji(r)] +
1 ~ ~ 1 ~ ~
Z-L18iQjk(r)8iQjk(r) + 2L28iQij(r)8kQk/r) -
1 ~ 1 ~
3(~X)maxHiQij(r)Hj - 127T (~E)maxEiQi/r)Ej

± 2QoL1EiikQjlC;)8kQi,(r). [43]
LANDAU-deGENNESTHEORY 165

In terms of 5= the partition function for the system is

Z( T) = f DQ(r) exp{ -(3f d 3 r ~[Q(r), aiQ(r), T]}, [44]

where DQ(f) denotes integration over all possible tensor fields Q(f),
and the dependence of 5= on Q (f), the spatial derivatives of Q (f), and
T is explicitly displayed. Eqs. [42] through [44] are the basic equa-
tions of the Landau-de Gennes theory of liquid-crystal phase transi-
tions.

3. Thermodynamic Properties of liquid Crystal Phase Transitions

In this section we use Landau-de Gennes theory to calculate the


thermodynamic properties of liquid-crystal phase transitions in
terms of the phenomenological parameters a, B, C, and Tc*. In par-
ticular, we will derive expressions for T e, the transition temperature;
<S>c, the equilibrium order parameter value in the low temperature
phase at the transition; and Lls, the transition entropy per molecule.
We will also calculate the temperature dependence of <S> (the equi-
librium value of S) for T close to Tc; and h, the height of the free-
energy barrier between <S> = 0 and <S> = <S>e at T = Te.
To calculate the thermodynamic properties listed above, it is suffi-
cient to consider a spatially uniform system in which the order pa-
rameter value is spatially invariant. This means that the spatial de-
rivative terms in the Landau free-energy density, which are impor-
tant for the calculation of fluctuation phenomena as shown in the
next section, can be neglected for the present purpose. Therefore,
from Eq. [42] we get

where we have set H = E = qo = O. The partition function is then


given by

Z(T) = fdS exp{-{3V~[S, Tn. [46]

From Z (T) one can calculate the free energy density of the system
5=°(T):

lim F(T) _ 1. -k B T In Z( T)
v-oo V
1m
V_oo
V . [47]
166 CHAPTER 10

Thermodynamic properties of the system can be derived directly


from ffO(T}.
In Appendix A it is shown that, if Z(T} is in the form of Eq. [46],
then ffO(T} is equal to the value of the absolute minimum of the
Landau free energy density ff at temperature T. Moreover, the (tem-
perature-dependent) equilibrium value of the order parameter, <S>,
is equal to the value of S which minimizes ff at each T. Therefore, in
order to obtain ffO(T}, we differentiate ff with respect to S and set the
result equal to zero for S = <S>:

0. [48]

Eq. [48] has three solutions:

(S) = 0,

(S) 1B / 1 B2 + 2 a (T * + B2 - T) [49]
6 C ± /324 CZ 3 C c 27aC .

The above solutions for <S> must be substituted into Eq. [45] to de-
termine which one gives the lowest value of ff. Straightforward calcu-
lation yields

g:o(T) =fo[T] + %a(T - Tc*HS)2 + iB(S)3 + 1~C(S)4,


[50]
where
B2
0, for T 2:: Tc * + 27aC
(S) 1 B + /1 B2 + 2 a (T * + ~_ T)
6 C /324 c 2 3C c 27aC '
forT~T*+~ [51]
c 27aC

Eq. [51] shows that the system transforms from the high-temperature
phase, <S> = 0, to the low-temperature phase, <S> ,t. 0, at a tem-
perature
LANDAU-deGENNES THEORY 167

[52]

The order parameter value of the low temperature phase at T = T c,


<S>c, can be obtained directly from Eq. [51]:

( S)
c
= _ ~6 B
C
+ ~
18
l!tl
C
= _.~9 B
C ' B < O. [53]

The entropy per molecule, s, can be obtained from ~O(T) by differen-


tiation:

S = _ ! ( a~o( T))
N aT
_
- So -
3a ( )2
4N S , [54]

where So = -(a!o[T]/aT)/N, and the term containing the tempera-


ture derivative of <S> is zero, because a~O(T)/a<S> O. Substitu- ==
tion of <S>, Eq. [51], into Eq. [54] gives the entropy per molecule of
the high temperature phase as So and the transition entropy per mol-
ecule ~s as

[55]

Eqs. [52], [53] and [55] are usually used as the means by which the
phenomenological parameters a, B, and C for any particular liquid
crystal are determined from the experimental values of <S>c, ~s,
and(Tc - Tc*). (The value of Tc* can be obtained by light scattering
experiments measuring order parameter fluctuations. See Section 5.)
In Table 1 we give the values of a, B, C and (Tc - Tc*) for MBBA.
In Fig. 4 we note that at T = Tc there is a free energy barrier be-
tween <S> = 0 and <S> = <S>c. It is interesting to calculate the
height h of this barrier in terms of the phenomenological parameters.
The peak of the barrier occurs at S = Sh for which

(a~/aS)S=Sh = 0 and (a2~/aS2)S=Sh < O.


~~ ~~

Simultaneous solution of these equations requires that Sh = -B/9C,


168 CHAPTER 10

which, when substituted into Eq. [45], gives

[56]

The value of h for MBBA is given in Table 1. Eq. [56] illustrates the

Table I-Values of the Various Phenomenological Parameters for MBBA'*'


Parameters Values Units
0.042 J/cm 3 oK
0.64 J/cm 3
0.35 J/cm 3
1 oK
6.1 X 10- 7 dyne

3.35 x 10-4 J/cm 3


'*' Data obtained from refs [14] and [15].

fact that the order of the phase transition in the Landau-de Gennes
theory is determined by the value of B. If B is nonzero, then h > 0,

°
and we have a first order phase transition with <S>e, Lls nonzero and
Te > Te*. For B = 0, h = and the transition is second order with
<S>e = Lls = 0, Te = Te*, and <S> '" (Te - T)1/2 for T < Te. For
liquid-crystal phase transitions the values of B are nonzero and nega-
tive (see Section 2). However, its value is small compared to that for
other first-order phase transitions, as evidenced by the relatively
small Lls, and that is why liquid-crystal phase transitions are some-
times described as "nearly second order."

4. Fluctuation Phenomena

In this section we use the Landau-de Gennes theory to discuss ther-


mal fluctuations in the isotropic phase of liquid crystals. For a physi-
cal system in thermal equilibrium, the instantaneous value of the
order parameter will almost always be equal or close to its mean value
(or equivalently, the equilibrium value). However, deviations from
the mean value of the order parameter do occur, and the problem is
to calculate the magnitude and the statistical distribution of these
deviations, or fluctuations. We distinguish between two types of fluc-
tuations: (1) homophase fluctuations, which occur within the range of
stability of a single phase and are completely described by the rms
deviation of the order parameter from its equilibrium value, and (2)
LANDAU-deGENNESTHEORY 169

heterophase fluctuations, which occur between two phases and are re-
lated to the problems of metastability and supercoolingP The quali-
tative difference between homo phase and heterophase fluctuations is
apparent in Fig. 4. Near T = Tc there is a free energy barrier of
°
height h (see Section 3), between the minima at S = and S = <S>c.
All fluctuations occurring around S = 0, and to the left of the barrier
are called homophase fluctuations. Fluctuations that carry the sys-
tem from one minimum, over the barrier, to the other minimum are
called heterophase fluctuations since the two minima correspond to
different phases.

4.1 Homophase Fluctuations in the Isotropic Phase

Since homophase fluctuations in the isotropic phase involve only


states close to Q(f) = 0, we can neglect the cubic and quartic terms in
the expansion of :fL. If we also set H = E = 0, Eq. [43] reduces to

1 ~ ~
g: = fo[T] + "2 a(T - Tc*)Qij(r)Qji(r) +
1 ~ ~ 1 ~ ~
"2LloiQjk(r)OjQjk(r) + "2L2oiQjj(r)okQkj(r)
± 2q OLi EijkQ H(r) 0kQ jzCr) • [57]

As shown in Section 2.3, Q (f) can be expressed in terms of S (f) and


n(f). If we were to make this substitution in Eq. [57], the resulting
expression would contain several interaction terms coupling S(f),
n(f), and their spatial derivatives. Thus, in general, the problem of
calculating the fluctuation spectra of S(f) and n(f) can be quite com-
plicated because of this coupling. In the latter part of this section we
will show one way to handle the coupling between S(f) and n(f), but
first we want to consider a simple system that illustrates both the
physics of homophase fluctuation phenomena and the mathematics
involved in manipulating the formalism developed in previous sec-
tions.

Simple Illustrative Example of Homophase Fluctuations

The system we wish to consider is one in which we can neglect the


spatial variation of n(f) and treat only the fluctuations of S(f). Set-
ting all spatial derivatives of n(f) to zero and limiting our discussion
for the moment to nematic liquid crystals (qo = 0), Eq. [57] becomes
170 CHAPTER 10

5" = fo[T] + ~a(T - Tc*)S2(y) +

~ (L1 + ~L2) [VS(;)]2 + ~L2[n.VSC;·)F, [58]

where n no longer depends upon r. The problem before us is to calcu-


late the mean square deviation of S(r) from its equilibrium value
S(r) = o. As a first step, we note that the partition function for the
system is given by

Z( T) = f DS( r) exp[ -(3 I volume d 3r ;t] .


of the
[59]
sample

The difficulty in using Eq. [59] for actual calculations lies in the func-
tional integral f DS (r) and we therefore digress briefly to consider
functional integration.

Functional Integration

It is customary to proceed by Fourier analyzing S(r) into its Fourier


components

[60]
-
q

where q is a particular wavevector, and S(q) is usually a complex


number representing the amplitude of the wave, exp(iq-r). S(q) can
be obtained from S(r) by the inverse transformation

[61]

Since S(r) is real, Eq. [61] implies that

[62]

where S*(q) is the complex conjugate of S(q). Since there is a one-


to-one correspondence between any function S(r) and its set of Fou-
rier coefficients IS(q)l, the functional integral is equivalent to inte-
grating over all possible sets of Fourier coefficients IS(q)l. This in
LANDAU-deGENNESTHEORY 171

turn is equivalent to a multi-dimensional integral in which each Fou-


rier coefficient S (q) is integrated over its possible values, with the
constraint S*(q) = S( -q) which implies that only half the S(q)'s in
any set IS (q)\ can vary independently. We will label the set of all in-
dependent S (q)'s by IS (q) 1'. IS (q) I' can be specified by the condition
that if S(q) is a member of IS(q)l', then S(-q) is not a member of
IS(q)l'.
The functional integral can then be written as

f DS(r) == J[ SCr) ;S(q) JIJ' {50 21T dcp S«1) f\


Seq) \ d\ Seq) \} ,
q [63]

where II' denotes the product only over those q's for which none is
the negative of any other, CPS(q) is the phase angle of the complex.
number S(q), IS(q)1 = [S*(q)S(q)]1/2 is the magnitude of S(q), and
J[S(r); S(q)] is the Jacobian of the transformation. In order to evalu-
ate the Jacobian we make use of the fact (Section 1) that DS(r) can
M
be written as II dS(rO/), where S(rO/) is the value of S in region Ol.
0/=1
Since each of the M regions has volume Ll V, we have from Eqs. [60]
and [61]
M
L S(QI3) ei~.;c<, [60a]
13=1

and

[61a]

It is clear from Eq. [60a] and the definition of a Jacobian that J[S(r);
S(q)] is simply the determinant of the M X M matrix whose Olfj - el-'
ement is exp(iqfjorO/). It is shown in Appendix B that this determinant
is equal to (VI Ll V)M/2. Substituting this factor into Eq. [63] yields

where we have used the fact that the product II' contains M 12 fac-
172 CHAPTER 10

tors. We can now proceed with our evaluation of the partition func-
tion, Eq. [59], for our idealized example system.
We first integrate Eq. [58] term by term since we will need the vol-
ume integral of ff in evaluating Z(T). The volume integrals of 8 2 (;'),
[~8(r)]2, and [n.~8(r)]2 can be expressed in terms of 8(ij) and ij in
the form

1d rS2Cr)
v
3

Jv
d 3r[VS(r) F V~q21 Seq) 12
q
2v~'q2ISCq)12,
q
[64b]

and

V~ (no q)21 S(q) 12


q

2V~ '(no q)21 S(q) 1 2, [64e]


q

where V is the volume of the sample and L:' denotes summing over
q
the ij's for which none is the negative of any other. In arriving at the
above expressions we have used the identity

V6(q + q') = fd 3r exp[i(q + q') 0 r]. [65]

From Eqs. [58], [64a], [64b], and [64c] we obtain

Iv d 3 ; ff = V!o[T] +

2v~'[~a(T - Tc*) + ~ (L1 + ~L2)q2 + %L 2(n oq)2] I S("q)12


= V!o[T] + ~aV(T - Tc*)~'[l + ~2(T,</Jq)q2]IS(q)12 [66]
q
where
~ o( </Jq) ( T c *) 1 /2
(T - Tc *) 1/2 [67]
with

[68]
LANDAU-deGENNESTHEORY 173

and V;q denotes the angle between nand q. The parameter ~(T,V;q)
has units of length and is referred to 'as a "correlation length" for rea-
sons that will become apparent later. If we substitute typical
values 14 ,15 for L 1 , L 2 , a, and Tc* in Eq. [68] we find ~o ~ lOA.. For sim-
plicity we will write ~(T,V;q) and ~o(V;q) as ~ and ~o, respectively. Sub-
stituting Eq. [66] into Eq. [59], we have for the partition function,

Z( T) = exp{ -(3 Vio [Tn x

fDSCr)I]'exP{-%(3Va(T - Tc*)(l + ~2q2)ls(q)12}, [69]


, q

which, according to Eq. [63], can be written

Z(T) = exp {-(3Vio[T]}I]' {:~ x


q

fdIS(Q)12exP[-%(3Va(T- Tc*Hl + ~2q2)IS(Q)12J}, [70]

where the angular integral has been explicitly evaluated. If we label


the factors in the product in Eq. [70] by Z(q, T), Eq. [70] reduces to

Z(T) = exp{-(3Vio[T]}II'Z(Q,T). [71]


Ii

The total free-energy density of the system is calculated from Eq.


[71] in Appendix B. For the moment we note that Eq. [71] can be
used to calculate the thermal averages of S(q) and IS(q)i2. Since the
partition function is the product of wavevector-dependent factors
Z(q,T), the thermal average of any quantity X[S(q)] associated with
wavevector q is given by

(x[S(q)]) =

[72]

where, in place of the maximum possible value of IS(q)1 2 , the upper


174 CHAPTER 10

limit of the dIS(q)1 2 integral has been extended to 00. This is possible
because the factor V in the exponent makes the integrand sharply
peaked at IS(q)1 2 = 0, and the error introduced by extending the limit
of integration is therefore negligible. We can now calculate the ther-
mal average of S(q) and IS(q)1 2 using Eq. [72]. Expressing S(q) in the
form IS(q)1 exp(iIPs(q) and substituting it for X[S(q)] in Eq. [72] we
find

(Seq) = 0,

(S(7]) S(q') = (S(7]) )(S(q') 0, q *- -q' [73]

Similar calculation yields

( I Seq) 12) - ~ k BT
- 3 Va(T - Tc *)(1 + ~2q2)

2 kB(T fTc *)
(£ * - 1)
[74]
= 3 Va [ + ~o2q2 ] '

where the last expression is obtained by the substitution of Eq. [67]


for ~. <I S ((j)l2> is the square of the amplitude of fluctuation in S (r)
with wavevector q, and Eq. [74] is simply an expression of the equip-
artition theorem. It is evident from Eq. [66] that in thermal equilibri-
um each of the independent modes in the set IS(q)1' contributes a
term

to the Landau free energy. Since S(q) is a complex number with two
degrees of freedom (real and imaginary parts), the equipartition the-
orem requires that

which is identical to Eq. [74]. The equipartition theorem therefore of-


fers a simple method for calculating the thermal averages of the
lANDAU-deGENNESTHEORY 175

square of the amplitudes for independent modes. * Three points con-


cerning Eq. [74] warrant further discussion.
First, it appears that all fluctuations vanish in the thermodynamic
limit V -- 00. However, this is not the case. Although the fluctuation
amplitude for each mode decreases as V increases, the density of
modes increases such that the sum of all fluctuation amplitudes re-
mains constant. This is simply demonstrated by calculating the aver-
age fluctuation amplitude in real space:

where we have llsed Eq. [60]. Making the replacement

~
~ --+ (27T)3
V I 3~
d q,
q

and substituting Eq. [74] for <IS(q)12>, we get

t
[ '>0
q
max
_ j~
Tc*
- 1 tan- 1(
jT
~oqmax )~ '
- -1
T*c
[76]

where qmax == 2'1l-j(smallest length for which the theory is valid) ~

° If, instead of considering only the independent modes, all possible modes were counted, then in
thermal equilibrium the contribution of each mode to the landau free energy would be (%) Va( T -
TeO) (t + ~2r1) <1S(Q)12>. However, each S(Q) now has only one degree of freedom due to the re-
ality restriction SO(Q) = S(-Q) and therefore can only have %keTof energy. The resulting equation

is identical to Eq. (75). Hence, the same expression for <1S(Q)12> is obtained independent of how
the modes are counted.
176 CHAPTER 10

27r/(fl. V)1/3, fl. V being the volume of an elemental region as defined in


Section 2. Eq. [76] expresses the physically reasonable result that the
average fluctuation amplitude in real space is independent of V.
The second point we wish to stress concerns the dependence of
<IS(q)1 2> on q and T, shown in Fig. 6. For a given temperature T>

A
N

1~
en
-V
01
>-'" m

r<>1C\J 10

2r-~ _ __

a !

.1 .2 3 .4 .5 6 .7 8 9 1.0 1.1 1.2 1.3 1.4


qe,

Fig. 6-The square of the fluctuation amplitude plotted as a function of wavevector for
different temperatures T ~ Tc·. Note that for fixed T the fluctuation amplitude
decreases monotonically with increasing q and that for fixed q it increases as T
decreases toward Tc •.

Tc*, <IS(q)i2> decreases with increasing q, and, for fixed q,


<IS(q)1 2> increases as T -- Tc* from above. We can qualitatively
understand this behavior from the following mechanical analogy.
Consider the string of coupled pendula shown in Fig. 7. Let the gravi-
tational field play the role of the term %a(T - Tc*)S2(r) in ~L, and
the coupling springs play the role of the terms proportional to the
spatial derivatives of S(r). It is obvious that there are two kinds of os-
LANDAU-deGENNES THEORY 177

cillations; one is the oscillation in the gravitational field and the other
is the normal mode oscillation of a string of mass points connected by
springs. If q denotes the wavenumber associated with the normal

--

Fig. 7-Mechanical analog illustrating the qualitative features of Fig. 6. The springs play
the role of the [~S(r)J2 terms and the gravitational field plays the role of the (T
- Tc • )S2(r) term in the Landau free energy density function.

mode oscillation of the elastic linear chain, then for q = 0 the springs
are not distorted and we have uniform in-phase oscillation of all the
pendula in the gravitational field. For q ~ 0 each mass point experi-
ences the force from the springs (caused by the relative motion of
neighboring mass points) as well as the force due to the gravitational
field. The square of the amplitude for the mode with wavevector q is
analogous to IS(q)1 2 • Suppose now that the system of pendula is in
thermal equilibrium at temperature T with each mode having energy
kBT/2. For the q = 0 mode the entire kBT/2 of energy goes into the
oscillation in the gravitational field. However, for a q ~ 0 mode part
of the kBT/2 of energy is spent in distorting the springs, the remain-
der going into oscillation in the gravitational field. Thus the ampli-
tude of oscillation for any q ~ 0 mode is smaller than that for the q =
o mode. Since the energy stored in the springs increases with q, it fol-
lows that, for a fixed temperature, the amplitude of oscillation (and
by analogy <IS(q)1 2 » should decrease with increasing q. What hap-
pens as T -+ Tc *? This corresponds in our analogy to weakening the
gravitational field, with the result that, for a fixed energy input and
fixed q, the amplitude of oscillation increases. This increase in fluctu-
ation amplitude as T decreases toward Tc* is an indication of the
growing instability in the high-temperature phase. In fact, at T = Tc*
the fluctuation amplitude <IS(q)1 2 > diverges for q = O. Tc* is thus
the temperature below which the isotropic phase is absolutely unsta-
ble.
178 CHAPTER 10

The third point concerns the physical significance of ~. Consider


the correlation function <S(rdS(r2». It is obvious that if L1 = L2 =
0, a disturbance at r1 would have no effect on any other spatial point
r2 ~ r1. The relative phase of fluctuations at two different points
would be completely random. Therefore, it is expected that

The interesting case arises when Ll, L2 ~ 0. Then a thermal fluctua-


tion at r1can cause a certain amount of in-phase fluctuation at r2 ~
r1 and as a result <S(rdS(r2» should be nonzero. In general, we can
write

~ ~ (S(q) Seq') exp {i(q. r1 + q,. r2)}. [77]


11 q'

<S(q)S(q'» can have the following values:

~ ~, =
~

_ {(S(q)(S(q') 0, q' *- -q
(S(q)S(q) - (I
S(q) 2),
1
q' = -q
[78]

Therefore,

~ (I S(q) 12) exp {iq. (r1 - r2)}


q

In converting the sum over all q vectors to an integral, we have made


the approximation of replacing qmax, the upper limit of the integral
over dq, by 00, valid for qmaxlr1 - r21 » 1. From Eq. [79] it can easily
be checked that if Ll, L2 -+ 0, then ~ -- 0, and <S(rdS(r2» = for
r1 ~ r2. For Ll, L2 ~ 0, Eq. [79] states that a thermal disturbance at
°
r1 decays exponentially to the equilibrium condition of its surround-
lANDAU-deGENNES THEORY 179

ings in a characteristic distance ~. Since ~ is the distance over which


fluctuations occur in phase, it is called the "correlation length." ~ div-
erges as T -+ Tc*, and for materials with Tc sufficiently close to Tc*
(such as liquid crystals) this behavior of ~ gives rise to various observ-
able phenomena, such as the increase in the light scattering cross sec-
tion and in the Cotton-Mouton coefficient when Tc is approached
(See Sections 5 and 6).

The General Case of Homophase Fluctuations

We will now investigate in more detail what happens when the re-
striction placed on n(r) in the previous subsection is relaxed. It was
pointed out earlier that in such a situation the fluctuations of S(r)
cannot be calculated independently from the fluctuations of fi (r), due
to interaction terms between them in Eq. [57]. The problem of cou-
pling between different fluctuation modes can be avoided if one con-
siders the thermal fluctuation amplitudes of the tensor order parame-
ter components, Qij(r) , rather than the fluctuation amplitudes of
S(r) and fi(r).
Since Q(r) is a symmetric, traceless tensor, only five of its nine
components are independent, viz. Qll(r) - Q22(r), Q33(r), Q12(r),
Q13(r), and Q23(r). As before we Fourier analyze each independent
component,

~ Q i/q) exp {iq r} 0


[80]
q

with Qij(9) defined as

The volume integral of~, (Eq. [57]), can then be expressed as

[82]

For a wavevector q pointing along the 3-axis (z-axis), ~(q) has the
form
180 CHAPTER 10

g:(q) = a(T - T c *)[%(1 + ~112q2)IQ33(q)12 +

i(1 + ~12q2)IQ11(q) - q2iq)12 + (1 + ~12q2)IQ1lq)12 +

(1 + ~/q2)( 1Qdq) 12 + 1Q23(q) 12)] ±

2qqoL1i [Qdq)(Q11*(q) - Q22 *(q)) +

Q12*(q)(Q22(q) - Ql1(q)) + ~3(q)Q13*(q) - Qdq)Q23*(q)] ,


[83]

where ~11(td is the value of ~ when 1/;q is zero (7r/2) and 6 is the value
of ~ when L2 = 0. It is clear from Eq. [83] that for qo = 0, i.e., for
nematics, the last term of Eq. [83] vanishes and Q33(q), Q11 (q) - Q22
(q), Q12(q), Q13(q), Q23(q) are independent quantities. Since there
are no interaction terms containing both q and q' (q ~ q') in Eq. [82],
we are free to choose the spatial axes for each q. Thus, Eqs. [82] and
[83] represent a general solution, valid for any q, provided we remem-
ber that the 3-axis is parallel to q. The thermal average of the square
of the fluctuation amplitude for each independent mode can be calcu-
lated in exactly the same manner as was that for S(r) in Section 4.1.1.
As before, the results can be obtained directly from the equipartition
theorem. Therefore, we have, for qo = 0,7

[84a]

[84b]

[84c]

and

2Va(T - Tc *)(1 + ~J.2q2)·


[84d]

When qo ~ °
the situation is quite different, and some of the
°
modes which are independent of each other for qo = become mixed.
LANDAU-deGENNES THEORY 181

However, if we again take the 3-axis parallel to q, Eq. [83] can be di-
agonalized to give

It can be seen that the independent modes for cholesterics are


different from those for nematics. If we denote the quantity
[Qu(q) - Q22(q) ± (8qoq 6 2i/(1 + 6 2q2))Q12(q)] by Qa(q) and
[Q13(q) ± (2qoq 6 2i/(l + ~-L 2q2))Q23(q)] by Qb(q), then the thermal
averages of the fluctuation amplitudes of the independent modes are7

\ ~ \2 2 kBT
(Q33(q) ) - 3 Va(T - Tc *)(1 + ~112q2) , [86a]

1 ~ 12 - 2kBT
(Qa(q) ) - Va(T - Tc*)(l + ~12q~ , [86b]

1
~ 12 - k B T(l + ~12q2)
(Qdq) ) - 2Va(T - Tc*)[(l + ~12q~2 - 16qo2q2~14J'
[86c]

(I Qb(q) 12) = 2Va(T _ :cB*r(l + ~J.2q~ [86d]


and
182 CHAPTER 10

The thermal averages <IQll(q)i2>, <IQ22(q)1 2 >, and <IQ13(q)1 2>


can also be calculated in terms of the thermal averages of the inde-
pendent modes given in Eq. [86]. Remembering that Qll(q) + Q22(q)
+ Q33(q) = 0, we obtain

12) = 12)
I~
<IQ22(q) <IQll(q)

(I "2 [~
i 8q oq ~ 1 2i ~ ~
Qa(q) =F (1 + ~12q2) Q12(q) - Q33 (q)
]
/

-"41 [<I Qa~q + ~12q~2 <I Q12(q)


'--)1 2) + (164qo2q2~14 ~ 12 ) + <1Q33 (q)
~ /2 ) ]

= <IQ 12 (q)12) + i<IQ33(q) 1


2), [86f]

<I Q2iq) 12) = <I Q1lq) 12)


= (I Q/q) =F (12~q t~~2) Q2iq) 1 ?
~ 12
= <I Qb(q) 4qo2q2~14
) + (1 + ~.l2q2)2
<I Q23 (~12)
q) . [86g]

In deriving Eqs. [86f] and [86g] we have used the fact that the ther-
mal averages of the cross terms between independent modes, such as
<Qa(q)Q12*(q», all vanish.
Due to the fact that the order parameter Q is related to macroscop-
ic observable quantities such as the susceptibility tensor X (Eq. [32])
and the dielectric tensorf, fluctuations in the components of Q are
directly manifested as fluctuations in E and in X and are therefore ex-
perimentally measurable. In Section 5 we will show how the fluctua-
tion amplitude, <IQij(q)1 2 >, can be used to calculate cross section of
light scattering by fluctuations in the isotropic phase of liquid crys-
tals.

4.2 Heterophase Fluctuations

In this section we will investigate fluctuations that result in the for-


mation of small spatial regions of the low temperature, liquid-crystal
phase in the isotropic phase, as the temperature T approaches Tc
from above. These so-called "heterophase" fluctuations produce a
LANDAU-deGENNES THEORY 183

°
sudden change in the value of the order parameter, in a local spatial
region, from S = to a value appropriate to the liquid crystal phase
at temperature T. Referring to Fig. 4, such fluctuations carry local re-
gions from the free energy minimum at S = 0, over the barrier to the
free energy minimum near S = <S>e. The small regions of liquid
crystal phase so produced are sometimes called "embryos"; their sta-
bility and statistical distribution for T", Te are the subject of the fol-
lowing discussion.
Let 5i denote the value of the Landau free energy density of the
spatially uniform isotropic phase and g;a denote the analogous quan-
tity for the anisotropic, liquid crystal phase. ffi and 5 a correspond to
the values of the two minima in Fig. 4 for T close to T e , and their dif-

negative for T < Te. For the moment let us set Ll = L2 = in the
Landau free energy expression, which implies perfectly sharp
°
ference, Llg; = ffa - ffi' is positive for T > T e, zero for T = T e, and

boundaries between the embryos and their surroundings; that is,


every point in the system sits in one or the other of the two minima.
In this case, the probability of heterophase fluctuations can be easily
calculated. If we let P a denote the probability for a spatial region of
volume Vern to spontaneously fluctuate from the isotropic phase to
the anisotropic phase and Pi be the probability that the same spatial
volume would remain in the isotropic phase, then in thermal equilib-
rIum

[90]

Eq. [90] is correct, independent of the barrier height h between the


two minima of the Landau free energy density, so long as we neglect
the energy contributions due to the gradient terms. In this case the
barrier height influences only the rate at which the thermal equilibri-
um distribution is reached, but not the distribution itself. According
to Eq. [90] one should expect to observe large heterophase fluctua-
°
tions near any first order phase transition since Ll5 "'" for T close to
Te. This, of course, is not the case and the reason lies precisely in our
neglect of the gradient terms in the Landau free energy density ex-
pansion.
If we let L1, L2 ~ 0, it is clear that formation of an embryo with an
infinitely sharp boundary is energetically unfavorable. Therefore,
there must necessarily be a boundary layer within which the value of
the order parameter varies continuously from that inside the embryo
to that of the surrounding isotropic fluid. The tendency of the gradi-
184 CHAPTER 10

ent terms to maximize the spatial extent of the boundary layer is op-
posed by the accompanying increase in energy associated with values
of the order parameter between the two minima, and the boundary
layer that obtains represents a compromise between these two forces.
In reality, then, we see that the barrier height plays a vital role not
only in the rate of approaching the thermal distribution of embryos
but also in the final distribution itself. That is, in addition to the free
energy ~5= Vem, creation of an embryo also requires a certain amount
of surface energy f.,t per unit areaP Later in this section we will esti-
mate the magnitude of f.,t for liquid crystals, but for the present we
will calculate the total work <I> required to produce a spherical embryo
of radius p.
From the above discussion we have

[91]

so that the probability for thermally generating an embryo of radius p


is ~ exp[-~<I>(p)l. Fig. 8 shows the behavior of <I>(p) for three different
cases: ~5= > 0, ~5= = 0, and ~~ < 0. For ~~ ~ 0, T ~ T e , <I>(p) is a
monotonically increasing function of p, and embryos of any radius,
once generated, inevitably shrink and disappear. It is also clear that if
the value of the surface energy f.,t is large, then even at T = T e , where
~g: = 0, the probability for the occurrence of heterophase fluctua-
tions is small. Since the magnitude of the surface energy f.,t is directly
related to the barrier height between the two free energy minima
(higher barrier ..... larger value of f.,t), it follows that in physical sys-
tems where the barrier is high, such as in the liquid-solid phase tran-
sitions, fluctuations are usually small and unobservable.

°
The interesting case arises when ~5= < 0, i.e., for T < Te. In this
case S = is a relative minimum, and it is possible to have a super-
cooled metastable state. In Fig. 8 it is shown that for ~g: < 0, <I>(p) has
a maximum at

p* [92]

due to the competition between the volume free energy, which is neg-
ative, and the surface energy, which is positive. For embryos of radius
p < p* the surface energy dominates, and the embryos tend to shrink.
However, for those embryos whose radius is greater than p* the vol-
umes free energy dominates, and the embryos tend to grow until the
LANDAU-deGENNESTHEORY 185

whole sample transforms into the low temperature phase. It follows


that if one could avoid producing embryos with radius p ~ p*, then
supercooling of the sample would occur.

ARBITRARY
UNITS

~_---'I_ _ _ l!. j<O.T<Tc


I
I
o t--.......= - - - - - - - - - + I:--_ _ _~~----~p(ARBITRARY
p* UNITS)

Fig. a-Plot of the work required to generate an embryo as a function of its radius p at
different temperatures. For T ~ Tc the work required is an increasing function
of p so that an embryo of any radius. once generated, must ultimately shrink
and disappear. For T < Te. embryos whose radius is greater than a critical ra-
dius p' can grow indefinitely until the entire sample has transformed to the or-
dered phase. Note that if the formation of embryos of radius p > p' can be
prevented. the sample can be supercooled.

Since the phenomenon of supercooling depends critically on the


absence of embryos with radius p > p*, it is of interest to estimate the
relative probability A of thermally generating an embryo with radius
p* by heterophase fluctuations. We have, for T near T e ,

A '= Probability of generating an embryo with radius p*


Probability of the system remaining in isotropic state

[93]

Since Ilff changes sign at T = T e, we approximate it by the form


186 CHAPTER 10

-
~~ = canst. (T - Tc)

for T close to Te. The constant must be equal to -Nfl.s, the transi-
tion entropy per unit volume, since, by definition, -afl.ff/aT\ T=Tc =
N fl.s. Thus,

[94]

which can be substituted into Eq. [93] to give 18

[95]

From Eq. [95] it is evident that if the dimensionless ratio I.N


(kBTe3N2(fl.s)2) is large, A is small and supercooling should be easily
observable.
We can estimate the value of 1.1. within the framework of Landau-de
Gennes theory. For this purpose we will use Eq. [42] with fI, E, qo,
and L2 set equal to zero and we will neglect any spatial variation in
nCr). Eq. [42] then becomes

3 ~
~ = fo[T] ,+- "4a(T - Tc*)S2(r)

9 ~ 3 ~ ~ 2
16 cst(r) + "4 L1 [V'S(r)] . [96]

We will consider a semi-infinite planar sample in which SCr) varies


only as a function of z with boundary values S( - <Xl) = <S>e and
S( <Xl) = O. The origin z = 0 is defined arbitrarily by the condition
S(z = 0) = <S>e/2. From arguments presented previously we expect
to find a transition layer in which the value of S varies continuously
from <S>e to O. The free energy per unit area of the transition layer
is given by

F{~Z)} = Jroro dz{a' - forT]}

= Jroro dz {~a( T - Tc *) 5 2(z) ; {B5"(Z) + 1~ Cs4(z) + ~ Ll ( d~~Z) ) 2}


[97]
LANDAU-deGENNESTHEORY 187

where the dependence of F on 8(z) is shown explicitly. The variation


of 8(z) with z is obtained by minimizing the free energy F. The re-
sulting 8(z) must satisfy the Euler-Lagrange equation

d~(Z) [%a(T - T e*)S2(Z) + i Bs3 (z) + 1~ Csi(z) ]


_ ~ L d 2S(z)
- 2 1 dz 2 [98]

Multiplying both sides of Eq. [98] by d8(z)/dz and integrating once


with respect to z gives

const. + %a( T - Te *) S2(Z) + i BS3(z) + 1~ CsI(z)

= ~L (dS(Z))2 [99]
4 1 dz

The constant in Eq. [99] must be zero since 8( ex» = 0 and [d8(z)/
dzl z ==]= O. Therefore, using the fact that z = 0 is defined as the
point at which 8 = <8>c/2, we have

aT*)1I2 z
- (~
3L 1
Z
or C[S(z)] - C[(S)e/2], [100]
~1
where

G(S) ~ In {-h[ (TIT,*) - 1] + (B/~T,*)S + (9C/4aT,*)S'

J3[(TITc*) - 1] _ (BlaTe *) }
[101]
S 2/3 [( TIT e *) - 1] .
188 CHAPTER 10

For T = T e, Eq. [100] can be simplified to the form

5(z) = (5)c [lOOa]


exp (~~) + 1

with the aid of Eqs. [52] and [53]. Equations [100], [100a], and [101]
give the spatial variation of 5 as a function of z. Using Eq. [100a] and
<S>e = 0.33, we obtain S(z) as plotted in Fig. 9 for T = Te. The fact

------------<q --------------
03

02

0.1

-2 -I o 2
ZI!,

Fig. 9-Variation of the order parameter S as a function of distance in the boundary


layer separating the ordered phase, S =
< S> e, from the disordered phase, S
= 0, at T = Te.

that S(z) decays most rapidly for 5 ~ 0.2 is understandable because


5 = 0.17 is the position of the barrier peak and therefore such values
of 5 should occupy the smallest portion of the boundary layer in
order to minimize the total free energy.
We can also calculate the surface energy7 per unit area f.l using Eq.
[99]. Substituting Eq. [99] into Eq. [97], we get

~L
2 1
f 00

-00
dz d5(z) (d5(Z))
dz dz

[102]

Using Eq. [99] for dS(z)/dz (with the minus sign for z > 0), Eq. [102]
becomes
LANDAU-deGENNESTHEORY 189

(3L aT *)1/2
J.1. = 1 2C x

f <s )c
O dS'S' 3
[( T
Tc* - 1
) B
+ ----y;; S' +
a c
~ S/2 ]
4aT *
c
1/2

[103]

Setting T = T e , L1 = 6 X 10- 7 dyne, and using the values of a, B, C,


T e, <S>e, and Te* for MBBA, we find Jl ~ 0.02 erg/cm2 • This is in
good agreement with the measured value 19 of Jl ~ 0.023 erg/cm2 •
Using this value of Jl, Lls ~ 0.16 kB, and N ~ 2.3 X 1021 cm- 3 , we ob-
tain the dimensionless ratio Jl3/(kBTe3N2(Lls)2) ~ 0.7 X 10- 6 for
MBBA. Putting this value into Eq. [95] yields the result that A is of
order e- 1 (<f>(p*) = kTe) when T ~ Te - 10 K, which means that
MBBA can at most be supercooled to about 10 K below Te. t Recalling
from Section 2.3 that Te* was interpreted as that temperature below
which supercooling is impossible, we conclude from the above calcu-
lation that for MBBA, Te* ~ Te - 10 K. This is in very good agree-
ment with experimental value 14 of Te - Te* ~ 10 K.

5. Observation of Fluctuations Using Light Scattering

Fluctuations in the order parameter are reflected in various physical


properties of a liquid crystal material. In this section we will focus on
the elastic (Rayleigh) scattering of light by such fluctuations in the
isotropic phase of nematic and cholesteric materials near Te.
We begin by expressing the dielectric constant at an arbitrary
point r in terms of the order parameter

[104]

where E is the dielectric constant of the isotropic liquid and 1 is the


unit tensor of the second rank. Since the polarizability a(r) [E(r) - ==
1]1411", we can write

[105]

which, upon Fourier transformation, becomes

t At this temperature the value of p' calculated from Eqs. [92] and [94] is about 80 A.
190 CHAPTER 10

a(q) [106]

Consider an incident light wave (Fig. 10) of the form

[107]

with frequency wand wave vector Ikinl = wlc. ein in Eq. [107] is a unit

~O"t ~
r,

Fig. 10-Graphic illustration of the vector identity. Eq. [110].

polarization vector pointing along the electric field direction of the


incident light. This wave will induce a dipole moment per unit vol-
ume PCr) in the sample

At a point r1 far from the scattering volume the radiation from a vol-
ume d 3r around r due to the induced oscillating dipole moment P(r)
is 20

where c is the speed of light and k out is a vector of magnitude wlc


pointing in the direction of (r1 - 1-). Letting eout be the unit vector
pointing along the direction of dE out , we see from Fig. 10 that

[(P(;') x k out) X kout ]


[110]
Ik out 12
LANDAU-deGENNESTHEORY 191

Substitution of Eqs. [108] and [110] into Eq. [109] yields

w2Eo
d Eout = eout~R [ . (~ ~
exp z k out o r1
A

c 0
exp[iqo r]d 3 r, [111]

where Ro = Irl - rl and q = k out - kin. Integrating over the volume


of the sample with Ro treated as a constant (rl is very far from the
sample), we obtain an expression for the total scattered electric field
at rl

Eout wt) ](e out • a(q) ° ein),


[112]

and for the differential Rayleigh scattering cross section dUR

[113]

where dO is the differential solid angle in the direction of k out . Sub-


stituting Eq. [106] into Eq. [113], we find the light scattering power,
or Rayleigh ratio, to be

R = l dUR [114]
- V dn

for q ~ O. In the isotropic phase of a nematic or cholesteric liquid


crystal the scattering results from thermal fluctuations, and the ex-
perimentally measured Rayleigh ratio is therefore given by

R [115]

In Fig. 11 we show a scattering geometry in which the wavevector


of the incoming light makes an angle v with that of the outgoing light
and €out is perpendicular to €in. If €l, €2, and €3 are unit vectors along
the X,y, and z axes, respectively, with q I z, then
192 CHAPTER 10

and

[116]

It follows that

[117]

Since Q23(q) and Q12(q) are independent, the cross terms in Eq.
[117] average to zero, leaving

q). ein 12> =


<1eAout • Q (~A COS
2"2
U 1 ~ 12
< Q32(q) ) + . 2 U <1Q12(q)
sm"2 ~ 12)

[118]

Substituting Eq. [118] into Eq. [115] and using the results of Eq. [86]
we have, finally, that the Rayleigh ratio for the scattering geometry of
Fig. 11 is

where Iql = (2wlc) sin v12.


For nematic materials qo = 0 and Eq. [119] shows that the scatter-
ing intensity decreases with increasing q. For cholesteric materials,
on the other hand, qo ~ 0 and the second term of Eq. [119] can be
peaked for a value of q such that

(4q ~ - 1) 1/2
q = 0 1 > 0, [120]
~1

provided q06 > %. Physically this means that if the correlation


length 6 is of order "olIO then fluctuations in cholesteric order will
produce incomplete helices with -1f1O of a pitch, which couple strong-
ly to the incident and scattered light waves when the transfer wave-
LANDAU-deGENNES THEORY 193

vector q satisfies Eq. [120]. As T-Tc* from above, 6 increases and


the peak in the second term of Eq. [119] div~rges at Qo6 = %which
corresponds to the temperature T = Tc* + 4qo2Lda (Recall that 6 =
[Lda(T - T c*)]1/2).

LIGHT LIGHT
ISOURC~ ~TOR
~ a in
1\
e out

------~~-----1I~~--~~--------~x

LIQUID
CRYSTAL
® SAMPLE
Y
Fig. 11-lIIustration of the geometry used in the light scattering experiment discussed in
q
Section 5. Note that is chosen to be along the z-direction to conform with
the convention used in the text.

The Rayleigh ratio has been measured in the isotropic phase of


both nematic 14,21 and cholesteric22 liquid crystals. Agreement be-
tween these measurements and the predictions of the Landau-de
Gennes theory is excellent in both cases.

6. Magnetic Birefringence and the Para nematic Susceptibility

Consider the effect of applying a static or slowly varying magnetic


field to a sample of nematic liquid crystal in its isotropic phase. As-
suming that n II fl, we can write the magnetic contribution to the free
energy density as (see Eq. [38])

[12l}

The equilibrium state in the presence of the magnetic field will have
finite order since s= M is negative; the average value of this order S (H)
can be calculated straightforwardly. We have
194 CHAPTER 10

JDs(;)[~fd3rS(;) ] exp {-f3fd3r (3'L - (A~max S(;)H 2)}


JDS(;') exp {-f3f d3;Y (3'L - (Art u S(~H2)} [122]

where < >H denotes the thermal average with the magnetic field- H
present.
8(H) depends on H quadratically and, furthermore, since 8(H) is
expected to be small and 8(H = 0) = 0, we can make the following ex-
pansion

S(H) = 1]H 2 + .
where

1] = as(H)
2
I
a(H) H2=0

is the paranematic susceptibility. Substituting for S(H) from Eq.


[122] and evaluating the resulting expression in the limit H2 -+ 0, we
find

1] = kBIT (~~max V{<I Seq = 0) 1 2)H2=0 - <S(q = 0)2H2=0}

= (~:kma,;V <IS(q = 0)1 2), [123]


B

In the approximation B = C = 0, we can substitute Eq. [74] into Eq.


[123] and get

[124]

Taking (LlX)max !:>!. 10-6 cgs units, a !:>!. 4.2 X 105 erg/oK cm3 , T = T e,
and Te - Te* = 10 K, we find 1/ ~ 5 X 10-13 cm3/erg. For a field H ~
10 KOe this results in an equilibrium order 8(10 KOe) !:>!. 5 X 10-5 •
The paranematic susceptibility is directly related to the phenome-
non of magnetic birefringence. Any anisotropic property is propor-
LANDAU-deGENNES THEORY 195

tional to the induced order. Specifically we can write

where Mo is a constant and tllkd is the long wavelength (q = 0) di-


electric constant parallel (perpendicular) to fl. However, Llt = (tIl 1/2
- t-L 1/2)(tI1 1/2 + t-L 1/2) ~ 2nLln, n being the refractive index, and
using our previous result for S(H), Eq. [125] becomes

[126]

Rearranging Eq. [126] gives

Lln _ Mo(LlX)max:
[127]
H2 9an(T - Tc *)
_ MoTl
- 2n

The quantity Lln/H2 is called the Cotton-Mouton coefficient which is


seen to diverge at T = Tc* and to fall off as (T - T c*)-1 above Tc*.
Measurements of magnetic birefringence have been made for
MBBA,14 and the inverse Cotton-Mouton coefficient was plotted as a
function of temperature. The experimental behavior is in complete
agreement with Eq. [127]. Analysis of the data yields Tc* ~ Tc -
1 0 K.

Acknowledgment

We express our appreciation to Dr. R. Cohen for helpful discussions


and to Dr. P. J. Wojtowicz for assistance in some numerical calcula-
tions and careful reading of the manuscript.

Appendix A

In this appendix we illustrate some implications of Eq. [15]. Consider


a hypothetical system in which the function 'Y [t7 u(f), T] is given by

[128]

The interaction energy defined by Eq. [128] reflects a rigid coupling


196 CHAPTER 10

between neighboring spatial regions. Substituting Eq. [128] into Eq.


[15], we see that the only form of u(f) that can contribute to the inte-
gral is u(f) = constant, since any spatial variation of u(f) would make
the integrand vanish due to the particular form of I'[~ u(f), T]. The
delta function thus simplifies the functional integral f Du(r), which
can be written as f du(l) ... f du(M), (see Section 1) to a one-dimen-
sional integral f du. Eq. [15] therefore becomes

Z(T) = fda exp {-/3V![a, Tn. [129]

At any particular temperature, f[ u, T] has a minimum. If uo(T) is the


value of u which minimizes t[ u, T] at temperature T, then the inte-
grand in Eq. [129] has a peak at u = uo(T). In the spirit of Landau's
approach, let us approximate Z(T), Eq. [129], by using a Taylor se-
ries approximation for t[ u, T] around its minimum:

1
![a,T] ~!o[T] + 2A(T)[a - a o(T)]2, A(T) > 0 [130]

where forT] = t[ uo(T), T]. Then


Z(T) ~ eX P {-/3V!o[Tnj da exp{-/3VA(T)[a - a o(T)]2/2}.
[131]

Note that the integrand in Eq. [131] is a Gaussian centered at uo(T)


with half width [,BVA(T)]-1/2. As V -- 00, the width of the peak ap-
proaches zero and the approximation, Eq. [131]' becomes exact. It
also follows that in the limit V -- 00 the thermal distribution func-
tion, expj-,BVf[u,T]l/Z(T), approaches a delta function peaked at
uo(T). This observation is the basis of the well-known result that in
the thermodynamic limit the thermal average of u, denoted by <u>,
is equal to the value of uo(T) that minimizes the Landau free energy
function f[ u, T], that is, the equilibrium value.
For a sufficiently large volume V the integral in Eq. [131] can be
evaluated to give

21T \112
Z(T) ~ ( /3VA(T)) exp{-/3 V!o[T]}. [132]
LANDAU-deGENNESTHEORY 197

From Eq. [132] the free energy density in the thermodynamic limit V
--+ 00 is evaluated to be

.
~:~
F( T) _
---v- - forT] .
+ ~:~ 2V
kB T [ 21T
In (3VA(T)
J_
- forT]. [133]

Eq. [133] states the result that in a spatially uniform system, where
the partition function can be expressed by Eq. [129], the free energy
density at temperature T is equal to the value of the minimum of the
Landau free energy density function I[ u, T] at that temperature.
The magnitude of the fluctuations in this hypothetical system can
also be evaluated. If the thermal average of the quantity u is defined
by

[134]

then the root mean square fluctuation of the order parameter is

[135]

Carrying out the calculation with the approximation of Eq. [130] and
V large, we obtain

[136]

which vanishes in the limit of V --+ 00 unless, of course, A(T) --+ O.


This result, which is the same as the corresponding result of the mean
field theory, is purely an artifact of having required the whole system
to fluctuate in phase by the specific choice of "Y[t7 u(f), T] in Eq. [128].
To see the effect of specifying a different "Y[t7 u(f), T], let us set

[137]

and calculate the free energy density. In this case Eq. [15] becomes

Z( T) = f J)(J(;') exp {-(3(f d 3;f[ a(-;') , T])}

= fda(l) . . . fda(M) exp {-(3~V ~f[a(O'), T]}


= [ f da exp {-{3~ Vf[a, Tn JM, [138]
198 CHAPTER 10

where ~ V is the elementary volume as defined before, M = V / ~ V,


and V is the volume of the whole system. Using the same approxima-
tion for f[ u, T] as that given by Eq. [130], we get

27T ) 11
Z(T) ~ ( f3~VA(T) 2 exp{-f3Vfo[T]}. [139]

The free energy density is then

1I. mF(T)
V_oo
--
V = JOf' [T] + 11'm k BT M [1
v_ oo 2V
27i
n f3~VA(T)
]

[140]

As V --+ ro, we have M --+ ro such that V / M = ~ V = constant. There-


fore,

. F(T) ~[ 27Tk B T ]
hm V = fo[T] + 2 ~V In ~VA(T) . [141]
V-oo

Eq. [141] differs from Eq. [133] by an additional term which can be
attributed to spatial fluctuations of the order parameter. In contrast
to the previous case where the whole system fluctuates in phase, fluc-
tuations in a system where 1'[\7 u(r), T] = 0 are completely uncorrelat-
ed from one spatial region to the next.
In real systems, coupling between a small part of the sample with
the rest is neither zero nor rigid but can be described as elastic. This
implies that for physical systems the order parameter fluctuations at
two different spatial points are partially correlated, the degree of cor-
relation being a decreasing function of the separation between the
two points. These expectations are made explicit by calculations in
Section 4.

Appendix B

In this appendix we wish to obtain the Jacobian factor for Eq. [63]
and thence to evaluate the free energy density given by Eq. [70].
From Eq. [60a] it is clear that the Jacobian is given by
lIexp(iqjj-r"JII, the determinant of the M X M matrix with elements
exp(iqjj-r a) in column a and row (j. The inverse for the matrix for
which the a{j - element is [exp(iqwra)] can be obtained from Eqs.
[60a] and [61a];
LANDAU-deGENNESTHEORY 199

= ll. V L L exp{-i"%. r,,} exp{-i"% •rex}S(r,,) [142]


v 13 "
which implies

{~
y = O!
[143]
=
y '* O!

Therefore, the inverse matrix has element

-ll. V e -iq13orex
V

in row a and column {j. From matrix algebra and Eq. [143] we get

[144]

Therefore, apart from a phase factor, which we set equal to zero, the
Jacobian is given by (VI d V)MI2.
From Eq. [70] the integral is easily evaluated to give

[145]

where the upper limit of integration for dIS(q)i2 is set equal to 00.

The total free energy of the system is given by

F(T) = -kBT In Z(7j = Vfo[T] -

k T"\;"'/I [21T kBT


B L::- n ""3 aVa(T - Tc *)(1 + eq2)j=
l Vfo T
[] -
q

kBT
2
V
(21T) 3
jqmax
0
2 [21T
41Tq dq In ""3 aVa(T -
kBT
Tc *)(1 + ~2q'l)
]
[146]

where the summation over the independent set of q vectors is extend-


200 CHAPTER 10

ed to every q with the final sum multiplied by %. The integral can be


evaluated to give a free-energy density expression

where we have set Ll V =


(27r)3/q3 max . Eq. [147] can be easilYl"educed
to the two limits, Eqs. [133] and [141], discussed in Appendix A. On
the one hand, the rigid coupling limit can be obtained by reducing
the degrees of freedom of the system from M independently varying
components to a single independently varying component. This can
be accomplished by letting qmax -'0 in Eq. [147], which is equivalent
to requiring Ll V -. V and M -. 1. The zero coupling limit, on the
other hand, can simply be obtained by letting ~ -. 0 in Eq. [147].

References:

1 P. J. Wojtowicz, "Introduction to the Molecular Theory of Nematic Liquid Crystals," Chapter 3.


2 P. Sheng, "Hard Rod Model of the Nematic-Isotropic Phase Transition," Chapter 5.
3 E. B. Priestley, "Nematic Order: The Long Range Orientational Distribution Function," Chapter 6.
4 L. D. Landau, "On the Theory of P!lase Transitions, Part I and Part II," Collected Papers of L. D.
Landau, Edited by D. ter Haar, Gordon and Breach, Science Publishers, N. Y., 2nd Edition, p. 193-
216 (1967).
5 The values of the various critical exponents predicted by the Landau theory has been shown to dis-
agree with experimental findings. See H. E. Stanley, Introduction to Phase Transitions and Critical
Phenomena, Oxford University Press, N. Y. (1971).
6 C. Kittel, Introduction to Solid State Physics, 4th Ed., John Wiley and Sons, N. Y., p. 477.
7 P. G. de Gennes, "Short Range Order Effect In the Isotropic Phase of Nematics and Cholesterics,"
Mol. Cryst. Liq. Cryst., 12, p. 193 (1971).
8 R. P. Feynman and A. R. Hibbs, Ouantum Mechanics and Path Integrals, McGraw-Hili Book Co., N.
Y. (1965).
9 L. D. Landau and E. M. lifshitz, Statistical Physics, 2nd Edit. Addison-Wesley, Reading, Mass. p.
425 (1969).
10 J. A. Gonzalo, "Critical Behavior of Ferroelectric Triglyclne Sulfate," Phys. Rev., 144, p. 662
(1966).
11 V. L. Ginzburg, "On a Macroscopic Theory of Superconductivity for All Temperatures," Soviet
Phys. Doklady, 1, p. 541 (1956-57).
12 P. Sheng, "Introduction to the Elastic Continuum Theory of Liquid Crystals," Chapter 8.
13 F. C. Frank, "On the Theory of Liquid Crystal," Faraday Soc. Disc., 25, p. 19 (1958).
14 T. W. Stinson and J. D. Litster, "Pretransitional Phenomena in the Isotropic Phase of a Nematic
Liquid Crystal," Phys. Rev. Lett., 25, p. 503 (1970).
15 T. W. Stinson and J. D. Litster, "Correlation Range of Fluctuations of Short-Range Order in the
IsotropiC Phase of a Liquid Crystal," Phys. Rev. Lett., 30, p. 688 (1973).
16 The cholesteric liquid crystal state consists of either all lett-handed helices or all right-handed heli-
ces. The physical difference between these two states is manifested in their optical activities. See,
for example, H. de Vries, "Rotatory Power and Other Optical Properties of Certain Liquid Crystals,"
Acta. Cryst., 4, p. 219 (1951).
LANDAU-deGENNESTHEORY 201

17 J. Frenkel, Kinetic Theory of Liquids, Oxford University Press, London, Chapter VII (1946).
18 Ref. 9, p. 473.
19 D. Langevin and M. A. Bouchiat, "Molecular Order and Surface Tension for the Nematic-Isotropic
Interface of MBBA, Deduced from Light Reflectivity and Ught Scattering Measurements," Mol. Cryst.
Lfq. Cryst., 22, p. 317 (1973).
20 L. D. Landau and E. M. Ufshitz, The Classical Theory of Fields, Addison-Wesley, Reading, Mass.,
~. 200 (1962).
1 T. W. Stinson, J. D. Litster and N. A. Clark, "Static and Dynamic Behavior Near the Order-Disor-
der Transition of Nematic Uquid Crystals," J. Phys. (Paris), Suppl. 33, CI-69 (1972).
22 C. C. Yang, "Ught Scattering Study of the Dynamical Behavior of Ordering Just Above the Phase
Transition to a Cholesteric Uquid Crystal," Phys. Rev. Lett., 28, p. 955 (1972).
Introduction to the Optical Properties of
Cholesteric and Chiral Nematic Liquid Crystals

E. B. Priestley

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

The results derived here, and the related discussion, apply both to
cholesteric and to chiral nematic liquid crystals; however, in the in-
terest of brevity, we refer specifically only to cholesteric liquid crys-
tals.
The helical arrangement of the molecules in a cholesteric phase has
been described in an earlier chapter. 1 On a sufficiently microscopic
scale one cannot distinguish between cholesteric and nematic or-
dering. However, as we consider larger and larger volumes of the two
types of material, a difference in the molecular ordering begins to be-
203
204 CHAPTER 11

come apparent; we observe that the cholesteric director n follows a


helix

[lJ

as shown in Fig. 1, whereas this secondary, helical structure is absent


in the nematic phase. In general, both the direction of the helix axis z
in space and the magnitude of the constant <p are arbitrary. It is evi-
dent from Fig. 1 that the structure of a cholesteric liquid crystal is pe-

~----
'"
~
A
----= )
----~
- ---
-~~---

---- ---
---

"
~ ,,~"" """~"
",,::::::"" ""~
'-.....: ""::::"" ':'-..~
"""" """'' ' " I
1r
qo
I
I

Fig. 1-A schematic representation of the helical arrangement of the constituent mol-
ecules In a chOlesteric liquid crystal.

riodic with a spatial period

[2J

In a right-handed coordinate system, Eq. [1] describes a right-handed


helix for positive q 0 and a left-handed helix for negative q o. Thus
the sign of q 0 determines the sense of the helix and its magnitude de-
termines the spatial periodicity.
Several unique optical properties arise from this spatially periodic,
helical structure of cholesteric liquid crystals.
(1) Bragg reflection of light beams is observed. For light incident
parallel to the helix axis z, only the lowest-order reflection is al-
OPTICAL PROPERTIES 205

lowed; for oblique incidence the higher-order reflections also be-


come allowed.
(2) The Bragg reflected light is circularly polarized if the incident
wave propagates parallel to the z- axis and elliptically polarized
for oblique incidence.
(3) Only the component of optical polarization for which the instan-
taneous spatial electric field pattern matches the spiraling cho-
lesteric director is strongly reflected. The other component is
transmitted with no significant reflection loss.
(4) Very strong rotatory power is observed. Rotations of tens of revo-
lutions per millimeter are typical, compared to the fraction of a
revolution per millimeter characteristic of isotropic, optically ac-
tive liquids.
A detailed treatment of the optical properties of cholesteric liquid
crystals for obliquely incident light beams involves extensive numeri-
cal calculations 2 and is outside the scope of this chapter. However,
much of the physics underlying the observed optical behavior of these
spiral structures can be understood by considering the more restrict-
ed case in which the wave vector of the incident light is everywhere
normal to the local director, i.e., kliZ. The development presented
below parallels closely that of de Vries;3 however, the approach is
somewhat different. 4

2. Maxwell's Equations

As noted above we treat only waves propagating along the helix axis
z; D and E are therefore confined to the xy- plane and are related by
a two-dimensional, second-rank dielectric tensor t. In addition to this
restriction, we neglect (1) the weak intrinsic optical activity of the
constituent molecules which persists even in the isotropic phase, (2)
energy dissipation by absorption, and (3) magnetic permeability (~ =
1). Finally, we assume all waves to be of the form

ReI {(z)exp( -iwt)1 [3]

(Re = real part) so that \7 2 --- a2 jaz 2 and a2 jat 2 --- -w 2 . With these
assumptions, Maxwell's equations reduce to

[4]

One could at this point proceed to solve Eq. [4] with 'tsuitably ex-
pressed as a periodic function of z in the fixed laboratory frame of
206 CHAPTER 11

reference. Then, by Floquet's theorem,6 we know that there exist so-


lutions to Eq. [4] such that

E(z + L) = K'E(z) [5]

where K is a constant that may be complex. However, it is somewhat


simpler to solve Eq. [4] if we first transform it to a coordinate system
that rotates with the cholesteric helix. In the rotating frame 7 has a
simple diagonal form for all values of z. In making this transforma-
tion we will utilize the Pauli matrices and we therefore digress briefly
to consider the properties of these matrices.
Pauli Matrices

a1 = (01 01) ,a (0i -i)0' a (10-10)


2 = 3 =

are known as the Pauli matrices. Their principal properties, which


can be deduced from their explicit form, are summarized by

[6]

where j, k and I can independently take on the values 1, 2 or 3. ~jkl in


Eq. [6] is the Levi-Cevita anti symmetric symbol,6 which behaves as
follows:
I if j kl = 123, 231, 312
tiki =
{ -1 if jkl = 213,321,132
o otherwise.
aQ is the 2 X 2 unit matrix. Inspection of Eq. [6] reveals that the Pauli
matrices anticommute, i.e., aj Uk + Uk Uj = 0, j 7"'- k.
In transforming Eq. [4] to the rotating frame of reference we make
use of the exponential operator, exp( -i Uj (J). Its properties can best
be seen from a series expansion of the exponential, viz,

[7]

Regrouping the terms in this expansion, and bearing in mind the


properties ofthe Pauli matrices summarized in Eq. [6], we obtain
OPTICAL PROPERTIES 207

(P ()4
exp(-iu,() = Uo ( 1 - 2! + 4! - + ... )
()3 ()5 )
- iu, ( () - 3! + Sf - + ... , [8J

where the two power series in parentheses will be recognized as ex-


pansions of cosO and sinO. Thus

[9J

The reader can check that for j = 2, Eq. [9] becomes

. (CoS() -Sin()) [10J


exp( -lU 2() = . () () ,
sm cos

which is the rotation matrix for a vector in a plane. The utility of Eq.
[10] in our present problem results from the ease and compactness
with which the coordinate transformation can be made using the ex-
ponential operator notation. Notice that the z- dependence is con-
tained in the relationship

() = 27rZ / P, [l1J

where P = 2L is the pitch of the cholesteric structure.


As mentioned previously, 1 can be written in diagonal form in the
rotating coordinate system. Letting fll and f-L represent the dielectric
constant parallel and perpendicular to the local director ft, respec-
tively, it follows that

which, in terms of Uo and (13, is simply

[12J

fO= (fll + f-L)/2 is the mean dielectric constant, and fl = (fll - f-L)/2 is
a measure of the dielectric anisotropy, in the xy plane.
The wave equation in the rotating frame of reference is (see Ap-
pendix A)
208 CHAPTER 11

(~;') - e~(T2)(~:') + [(~y(f~O + fl(T3) - e~~O)J E' = 0


[13]

where the x and y axes of the rotating coordinate system have been
fixed parallel and perpendicular to ft, respectively. We try as a solu-
tion to Eq. [13]

[14]

where
[15]

is a (complex) two-component vector that describes the state of po-


larization of the wave in the rotating frame, Eo is a real constant that
gives the amplitude of the wave, and X is the wavelength in vacuo of
the wave. The product m fOl/2 plays the role of the refractive index of
the cholesteric material; however it is not strictly correct to think of it
as such. As we shall see later, the m values are complicated functions
of the pitch and dielectric anisotropy of the medium and of the wave-
length of the electromagnetic wave. fO in Eq. [12] is an average (opti-
cal frequency) dielectric constant defined above. Eq. [13] then reduc-
es to

Notice that 47r 2c 2/w 2 is just the square of the vacuum wavelength. X.
Defining a = fdfo and X' = X/fOl/2P simplifies Eq. [16] to

[17]

which can be written explicitly as

-2imX'
1 - (X')2 - m2
[18]

The two simultaneous equations represented by Eq. [18] have a non-


trivial solution only if
OPTICAL PROPERTIES 209

1 - (A')2 - m 2 + a -2imA'
[19]
1
2imA' 1 - (A')2 - m 2 -
0'1 = O.

The resulting fourth-order equation in m


[20]

has solutions

[21]

It will be useful to have more explicit expressions for the roots


given by Eq. [21] in two limiting cases. In both limits we will expand
Eq. [21] and keep only the lowest-order terms.
(a) The 4 (A')2/O' 2 « 1 Limit
In this limit the pitch is large compared to the wavelength of the
light. The square root in Eq. [21] can be expanded to give

m" = 1 + (A')2 ± (a + 2(:7 + ... ) [22]

whence
(A')2(O' - 2)
ml = -m3 = (1 - 0')1/2 + 20'(1 _ 0')1/2 + [23]

and
. + )112 + (A''f(o' + 2) + [24J
m2 = -m4 = (1 a 20'(1 + 0')1/2

The positive and negative roots are associated with waves traveling
through the cholesteric medium in the positive and negative z direc-
tions. respectively.
(b) The 4 (A')2/a 2 » 1 Limit
In this case the pitch is small compared to the wavelength of the light
and the appropriate expansion of Eq. [21] is

m2 = 1 + (A,)2 ± (2A' + ~, + ... ), [25]

from which it follows that


0'2
[26]
8A'(1 - A') + ...
210 CHAPTER 11

and

m2 = -m 4 = 1 + A' + SA /(1 + A') + [27J

Again the plus and minus signs correspond to waves traveling in op-
posite directions along the helix axis.
The results we have obtained above are identical to those given by
de Vries. 3
Now that we have explicit expressions for the roots m, it is simple
to determine the polarization vectors it of the corresponding modes.
We begin by rewriting Eq. [18] schematically as

(a + ex'
-ib )(I-Ll) = 0, [28J
ib a - ex' I-Lz

where a = 1-(A/)2_m2 and b = 2m A'. Expanding Eq. [2S], we have

0, [29J
and
0. [30J

Eqs. [29] and [30] require

[31J

which, when combined with the normalization condition,

leads to

[32J

[33J

The relative phase of I-Ll and I-L2 is fixed by Eq. [29], viz

[34J

Substituting for a and b in Eqs. [32] through [34] yields the following
OPTICAL PROPERTIES 211

expressions which, together with the appropriate m values, deter-


mine the polarization vectors A:

m2 - a
1 - (A')2 -
J.11J.1t = 2[1 - CA? - m 2J ' [35J

1 - (A'? - m 2 + a
J.12J.12 *= 2[1 - (A')2 - m2J ' [36J

J.12 =
J.11
_i(l - (A'?2mA-, m+ a).
2
[37J

For the case 4(A')2/a 2 « 1, we have

[38J

[39J

These vectors describe linearly polarized light. Thus, the modes are
essentially linearly polarized near the origin of the A' axis in Fig. 2. In
the other extreme, 4(A')2/a 2 » 1, the eigenvectors are

~ ~(-~) [4OJ

~ ~C). [ 41J

which describe circularly polarized light. Thus the modes are essen-
tially circularly polarized for large values of A'.

3. Discussion

Eq. [21] has been evaluated numerically with a arbitrarily set equal
to 0.1. The results for m 1 and m 2 are plotted in Fig. 2. The qualita-
tive features of Fig. 2 are apparent from the approximate expressions
for the roots, Eqs. [23], [24], [26], and [27]. For example, in the limit
of very small A', Eqs. [23] and [24] reduce to

These equations also show that the initial dependence on A' is qua-
212 CHAPTER 11

dratic. For large values of A', on the other hand, Eqs. [26] and [27] be-
come linear in A', as observed in Fig. 2.

m1 = 1 - A' and m2 = 1 + A'.

20

15

10 -~--~-- -~-- J-~-

05 -

o /1
m,
-05

-10

Fig. 2-Solutlons given by Eq. [21], with a = 0.1 as a function of reduced wavelength
X' for the two waves propagating In the positive z direction. In the shaded
region the m1 wave Is strongly reflected.

We have seen by Eqs. [38] and [39] that the normal waves are lin-
early polarized in the local frame of reference for A' - O. As the local
frame rotates (8 = 27rz/P), so do the polarization vectors of the nor-
mal waves. This is the "waveguide" regime discussed by de Gennes 7
and is also the regime in which twisted nematic field-effect devices 8
operate (values typical of such devices are P "" 50 J,Lm, €Ol/2 "" 1.65 and
A"" 0.5 J,Lm, leading to a value of A' "" 0.006).
For values of A' near unity, i.e., within the shaded region of Fig. 2,
m 1 and m3 are imaginary and the corresponding waves are nonpro-
pagating. It is apparent from Eq. [21] that this reflection band ex-
tends over the range of A' .

The m 2 and m4 waves are unaffected and are observed to propagate


freely for all values of A'. Considering now only the waves traveling in
OPTICAL PROPERTIES 213

the plus z- direction, we see from Eq. [41] that the m 2 wave is left cir-
cularly polarized and that its instantaneous electric field pattern is of
opposite sense to the (right-handed) cholesteric helix (see Appendix
B). The strongly reflected m 1 wave, on the other hand, is right circu-
larly polarized and has an instantaneous electric field pattern that is
superposable on the cholesteric helix. Thus, for X' ,..., 1, a right-handed
cholesteric liquid crystal reflects right circularly polarized light and
transmits left circularly polarized light. The reverse is true for a left-
handed cholesteric material.
For larger values of X', the normal waves are nearly circularly po-
larized (see Fig. 3), have opposite signs of rotation, and propagate

3.0r--,----,----r--.--,-7"7I=--r----r--r--~-__,

2.5

2.0

CIRCULAR POLARIZATION

Fig. 3-The eillptlcities 8, = I~,(m, )/~2(m,)1 and e2 = -1~2(m2 )/~,(m2 ~ as a


function of reduced wavelength X' for the two waves propagating In the posi-
tive z direction.

with different phase velocities. A superposition of two waves of oppo-


site circular polarization can be thought of as a linearly polarized
wave whose plane of polarization lies along the bisector of the instan-
taneous angle between the two rotating vectors. Since the two circu-
larly polarized waves travel at different velocities in the cholesteric
medium, there is a net rotation of the plane of polarization when a
linearly polarized wave is passed through a slab of the cholesteric ma-
terial. We can estimate the magnitude of this rotation in a straight-
forward manner using the results derived above.
214 CHAPTER 11

Again we consider only the m 1 and m 2 waves and we restrict our-


selves to the 4(>"')2/a 2 » 1 limit. We have seen that in this limit the
m 1 wave has the form

in the rotating coordinate system, with components

[43J

and
[44J

Thus, by Eq. [52], we can write (E l)x in the laboratory frame of ref-
erence as

[45J

which, upon substitution of Eqs. [43] and [44], reduces immediately


to

(E 1 )x = -v'Eo2 [21rZ
cos -p- -
(wt -
21rmltoIlZZ)]
-~A~- [46J

From Eq. [46] we see that the phase angle fh of the ml wave is

[47J

in the laboratory frame of reference. Similarly we find for the m 2


wave

O2 =
21rZ(
P 1 mz) + wt.
- A' [48J

The angular position of the resultant linear polarization vector is de-


termined by
OPTICAL PROPERTIES 215

[49J
which is the bisector of the instantaneous angle between the two ro-
tating electric vectors. Combining Eqs. [47], [48] and [49], and substi-
tuting the results of Eqs. [26] and [27], we find
dl/; 211" a"
dz = -P 8(,\')2[1 - (A')2J [50J
for the rotation per unit length. Taking a = 0.1, P = 0.5 ~m, EOI/2 =
1.65 and A = 1 ~m, d tJ;/dz is calculated to be approximately 36 revo-
lutions cm- I .
Finally, we plot in Fig. 3 the ellipticities el = 1~I(md/~2(mdl and
e2 = -1~2(m2)/M(m2)1 of the waves corresponding to ml and m2 as
a function of reduced wavelength A'. These have been defined so that
linear polarization is associated with a value of zero; one could equal-
ly well have chosen to plot the reciprocal relations in which case lin-
ear polarization would correspond to infinite ellipticity. With either
definition, the ellipticity of circularly polarized waves is unity. It is
apparent that the ellipticity of the m I wave is anomalous in the re-
gion of the reflection band whereas the m 2 wave is "well behaved"
for all values of A'.

4. Conclusion

The optical properties of cholesteric liquid crystals have been exam-


ined in the restricted case of waves propagating parallel to the helix
axis. Solutions to the wave equation were obtained in order to deter-
mine the polarization states of the normal waves in the medium. It
was found that the waves corresponding to two of the four solutions
are strongly reflected when the wavelength becomes comparable to
the helix pitch. Closer examination revealed that the instantaneous
electric field pattern of these waves is superposable on the cholesteric
helix. That is, the instantaneous electric field pattern of the reflected
waves is a right-handed helix that, when the wavelength is correct,
matches exactly the right-handed cholesteric helix. An expression
was derived for the optical rotatory power (rotation per unit length),
that correctly accounts for the sign and magnitude of observed opti-
cal rotations in cholesteric liquid crystals.

Acknowledgment

I am indebted to P. Sheng for helpful discussions.


216 CHAPTER 11

Appendix A

In this appendix we transform the wave equation to a coordinate sys-


tem that rotates with the cholesteric helix described by Eq. [1]. Let
q 0 be pos~tive so the resulting helix is right handed. Then, an arbi-
trary electric field vector E with components Ex and Ey in the labo-
ratory frame of reference has components Ex' and Ey'

E/ = Excos8 + E,sin8 }
[51J
Ev' = - Exsin8 + Evcos8

in the z = P 8/27r plane of the rotating coordinate system. Thus, by


Eq. [9], we can write

E = exp(-ia)J)E'}.
and o = exp( -ia"lJ)D' [52J

Also, in the rotating frame of reference, we know that

[53J

where flocal has the si~ple diagonal form given by Eq. [12]. Hence the
wave equation

[54J

in the laboratory frame of reference becomes

~~Z~' - (47r~(J~)~~' + [(~y(tlJ(J" + tj(Jj) - 4~~IJJEI = 0

[55J

in the rotating frame of reference.

Appendix B

Here we consider the spatial and temporal variation of the electric


vector for a circularly polarized light wave. We begin with a plane
wave propagating in the positive z direction

[56J
OPTICAL PROPERTIES 217

where Re = real part, Ais a vector describing the polarization state of


the wave, and Eo is the amplitude of the wave. Eq. [56] can be rewrit-
ten

E(t,z) sin[w(t -~)JJ}.


[57]

toO
t= !!
w

t=L
2w

(0)
E(t,O)

(b)
~(O,z)

Fig. 4-Behavlor of the electric vector for a right circularly polarized light wave: (a)
rotation of the electric vector as a function of time t at a fixed position on
the z axis; (b) posHlon of the electric vector as a function of z at a fixed In-
stant In time.

Choosing

~ = Vz(-~). [58J

it is apparent that

[59]
218 CHAPTER 11

and
[60]

Consider first the temporal development of the wave at some fixed


value of z, say z = 0 for convenience. Examination of Eqs. [59] and
[60] shows that at t = 0, the electric vector lies along the x axis while
at a later time t = 'Ir/2w it is along the negative y axis lmd at a still
later time t = 'Ir/w it is along the negative x axis, and so on. Thus, as
viewed from a point on the positive z axis, looking back toward the
origin, the electric vector in the z = 0 plane rotates in a clockwise
sense from left to right, and by convention this is called a right circu-
larly polarized wave (see Fig. 4a).
Next we consider the spatial distribution of the electric vector
along the z axis at one instant of time, e.g., at t = O. We see from Eqs.
[59] and [60] that at z = 0 the electric vector points along the x axis,
as before. As we move along the positive z axis we find the electric
vector rotates first to the positive y direction at z = 'lrC /2w and then
to the negative x direction at z = 'lrC /w, etc. Thus the instantaneous
electric field pattern traces out a right-handed spiral (see Fig. 4b). All
directions are reversed for left circularly polarized light, which is de-
scribed by the polarization vector

~
J.l
1(1)i .
=.y'2 [61]

These properties of circularly polarized light are important in under-


standing the optical properties of cholesteric liquid crystals.

References

, E. B. Priestley, "Liquid Crystal Mesophases," Chapter 1.


2 See. for example, D. W. Berreman and T. J. Scheffer, "Reflection and Transmission by Single-
Domain Cholesteric Liquid Crystal Films: Theory and Verification," Mol. Cryst. and Liquid Cryst., 11,
p. 395 (1970); R. Dreher and G. Meier, "Optical Properties of Cholesteric Liquid Crystals," Phys.
Rev., A8, p. 1616 (1973).
3 This problem has been studied in detail by H. De Vries, "Rotatory Power and Other Optical Proper-
ties of Certain Liquid Crystals," Acta Cryst., 4, p. 219 (1951).
4 The approach presented in this chapter follows that contained in some unpublished lecture notes of
P. S. Pershan on the optical properties of cholesteric liquid crystals.
5 See, for example, A. Messiah, Quantum Mechanics, Vol. II, pp. 544-549, John Wiley & Sons, Inc.
(1966).
6 J. Mathews and R. L. Walker, Mathematical Methods of Physics, W. A. Benjamin Inc., N. Y. (1965).
7 P. G. de Gennes, The Physics of Liquid Crystals, p. 228, Oxford University Press, Oxford (1974).

8L. A. Goodman, "LiqUid-Crystal Displays-Electro-optic Effects and Addressing Tech-


niques," Chapter 14.
Liquid-Crystal Displays-
Packaging and Surface Treatments

L. A. Goodman

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

For the proper operation of liquid-crystal display devices, appropri-


ate cell construction and surface treatment methods are necessary. In
this chapter, first, some of the general packaging techniques that are
being used today are outlined. Second, the common conductive coat-
ings are listed and standard preparation methods given. Third, the
diverse surface treatment methods are summarized and, finally, some
relationships between cell construction and device operation are dis-
cussed.

2. Packaging

The standard sandwich cell configuration for liquid-crystal displays


is schematically illustrated in Fig. 1. Normally, low cost soda-lime
soft glass is used. However, more expensive borosilicate and fused sil-
ica substrates can also be utilized. The spacing between top and bot-
tom glass plates varies from 5 to 50 /lm with nominal values being in
the 10 to 20 /lm range. The spacer composition is restricted by possi-
219
220 CHAPTER 12

ble chemical reactions with the liquid crystal. Glass frits and relative-
ly inert organic materials such as Teflon and Mylar can be used as
the spacer materials.
The sealing materials are limited both by the need for compatibil-
ity with the liquid crystal and by the need for a fairly hermetic seal

GLASS
- - - PLATE
r-~~~~~~~~~~
CONDUCTIVE
~~iz$$~¢;2¢ ' - - - + - - COATING
" ' ,"> :. . ~' . ; .':»-:- " . SEALING
MATERIAL
SPACER

L10UID
CRYSTAL

Fig. 1-Slde view of schematic representation of a liquid-crystal cell.

since both moisture and oxygen can react with many mesomorphic
systems in a deleterious fashion. Other criteria for selecting sealing
materials are thermal expansion match with the glass plates, bonding
strength, sealing temperature, and manufacturing cost. Glass frits,
solder glasses, and polymeric materials are suitable sealing materials.

3. Electrodes

At least one of the conductive coatings in a liquid-crystal display


must be transparent. The most common transparent conductive coat-
ing material is a mixture of indium oxide and tin oxide. This material
can be prepared by several different techniques which include (1) rf
sputtering of indium-tin oxide powder targets,l (2) dc sputtering of
highly conducting indium-tin oxide powder targets,2 (3) dc reactive
sputtering of an indium-tin alloy,3 (4) thermal evaporation,4 and (5)
high-temperature (400°C) bake-out of spun-on solutions. 5 The sheet
resistance of as-deposited indium-tin oxide coatings can vary from 2
ohms/square up to many kilohms per square. 2 After bake-out in air at
temperatures in excess of 500°C, the resistance increases by a factor
of 3 to 10. 1 ,2 The sheet resistance for liquid-crystal displays is nomi-
nally in the 100 to 500 ohms/square range. The optical transmission of
these films is excellent, and some typical results are given in Fig. 2.
Another compound commonly used for conductive coatings is anti-
mony-doped tin oxide, which is usually deposited by a pyrolitic spray
L1QUID·CRYSTAL DISPLAYS-PACKAGING 221

process at temperatures between 600 and 700°C. The chemical dura-


bility, abrasion resistance, electrical conductivity, and typical trans-
mission are all very adequate. Surface resistances as low as 40 ohms/
square can be achieved with a transmission in excess of 75-80%.

.

'-
,.-'"';..tc:.,-'-' - " _ .
'". / .
-,.:.:::-:.:.==:.:--=::.:.....-. - ' -
--
z
0 60
(/)
(/)
14.0./0
40
~ 28.0./0
(/)
Z 20 320.0./0
«
a:
f- 0
4000 4800 5600 6400 7200 8000
WAVELENGTH ( A)
Fig. 2- Transmission versus wavelength for In203-Sn02 films deposited in argon to dif-
ferent thicknesses. 1

The chemically deposited tin oxide suffers from two important dis-
advantages as compared to the vacuum-deposited indium-tin oxide.
At the high temperatures necessary for the spray deposition process,
soft glass loses its flatness. The different approaches previously listed
for creating the indium-tin oxide films can all be performed at suffi-
ciently low temperature that the soft glass substrates do not warp. In
liquid-crystal displays, where a relatively narrow spacing between op-
posite electrodes is required, the loss in flatness caused by the spray
process is a distinct drawback.
Antimony-doped tin oxide can also be deposited by sputtering, but
the sheet resistance is not as low as with sputtered indium-tin oxide.
In addition, for the tin oxide concentrations normally used, the in-
dium-tin oxide films can be readily etched in hydrochloric acid. 3 ,6
Antimony-doped tin oxide films are not readily soluble in acids or
bases, but they can be etched by a procedure using zinc dust and hy-
drochloric acid. 7
Reflective electrodes are readily obtained by the evaporation of
metals-for example, aluminum or chromium. Other coatings such as
layers of insulating dielectrics are also used to obtain specular surfac-
es. These can be deposited by both electron-beam and thermal-evap-
oration methods. 4
222 CHAPTER 12

4. Surface Orientation

A unit vector called the director denotes the average orientation of


the long molecular axes in any local region of the fluid. For the pur-
pose of maximizing the contrast ratio of the display, it is desirable
that the orientation of the director be the same throughout the fluid
whether the applied voltage is on or off. Two important examples of
maximum ordering are shown in Fig. 3. For perpendicular (or homeo-

TRANSPARENT
CONOUCTIVE
GLASS COATING GLASS

GLASS aLASS

(A) (8)

Fig. 3-Side view (A) of homeotropic and (8) of homogeneous orientations.

tropic) alignment, all of the long molecular axes are perpendicular to


the cell walls. Looking down through the cell, the fluid appears to be
isotropic, and this state is optically clear. When the mesomorphic me-
dium possesses uniform parallel (or homogeneous) orientation, the
director is parallel to the cell walls over dimensions of a millimeter or
more and points in only one direction. As observed in a top view, the
fluid has the optical properties of a uniaxial single crystal with the di-
rector describing the main axis (see Fig. 4A). The "head" and the
"tail" of the director can be interchanged without changing the ob-
servable fluid properties.
There are two other examples of fluid orientation that are closely
related to uniform parallel alignment. In the first case, the director
also lies parallel to the cell surfaces. However, the director orienta-
tion in the plane parallel to the cell walls is not uniform; rather, it
changes randomly over dimensions on the order of micrometers. This
orientation is known as random parallel alignment (see Fig. 4B).
The second related example is the planar or Grandjean texture of
the cholesteric mesophase. In the planar state, the main helix axis is
perpendicular to the electrode surfaces of the cell. Consequently, the
director is always oriented parallel to the surface with the orientation
LIQUID-CRYSTAL DISPLAYS-PACKAGING 223

of the director varying in helical fashion with linear distance along


the helix axis (see Figure 4C).
Four states of bulk orientation have been described so far. Because
of the long-range ordering forces that operate in liquid crystals, the
preceding bulk orientations can be produced by the proper treatment

(AI

(el

(el

Fig. 4-Top view (A) of homogeneous orientation and (8) of random parallel alignment
(the arrows represent the directors) (C) Side view of cholesteric planar
texture. Length of director arrows Illustrates the amount of twist of each
layer.

of the surface region between the liquid crystal and the cell walls.
The mechanisms of surface alignment have been poorly understood,
but recent investigations seem to be leading toward a better compre-
hension of them.
In his work on optical textures in mesomorphic materials, Friedel
presented an interesting discussion of both the uniform parallel and
perpendicular states. s However, the techniques he describes for
achieving these two states are rather cumbersome. Since that time, a
number of investigators have described different methods for ob-
taining perpendicular alignment. These have included chemical etch-
ing,9,10 coating with lecithin,ll and physical adsorption of organic sur-
factant additives such as the polyamide resin Versamid 12 or impuri-
224 CHAPTER 12

ties in the fluid 13 that are the thermal decomposition products of the
mesomorphic material.
Petrie et aP4 indicated that additives of certain surface-active mol-
ecules such as tertiary amines, quaternary ammonium salts, and pyri-
dinium salts produced perpendicular alignment. Specifically, Haller
and Huggins 15 reported that the quaternary ammonium salt, hexade-
cyltrimethylammoniumbromide (HT AB), caused perpendicular
alignment.
Zocher 16 stated that uniform parallel alignment could be obtained
if the cell surfaces were first unidirectionally rubbed. Various materi-
als such as paper, tissue, and cotton wool were used to achieve uni-
form parallel alignment. Chatelain 17 hypothesized that the parallel
orientation resulted from the forces generated by the presence of an
adsorbed layer of fatty contaminants on the cell surface, and the di-
rectionality was achieved by the unidirectional rubbing However, he
could not eliminate the possibility that mechanical deformation of
surface might have induced the observed alignment.
Several recent articles support the theory that physicochemical
forces, e.g., van der Waals, hydrogen bonding, and dipolar forces, are
dominant. Proust et aps have obtained both uniform parallel and
perpendicular alignment by depositing monolayers of hexadecyltri-
methylammoniumbromide (HTAB) from aqueous solution onto glass
slides prior to the insertion of the liquid crystal between the slides.
They stated that perpendicular alignment was obtained when the
bromide compound was densely packed in the monolayer; conse-
quently, the molecules were oriented normal to the glass slide. Uni-
form parallel alignment was observed when the monolayer was less
densely packed and the molecules were oriented parallel to the sur-
face. They achieved the directionality necessary for uniform parallel
orientation by withdrawing the glass slides from the bromide-con-
taining solution in one direction.
Kahn 19 has also demonstrated the importance of physicochemical
forces. He has prepared highly stable surface aligning conditions by
the utilization of silane coupling agents that were chemically bonded
to the glass surface. The preparation of the organosilane layer was af-
fected by the nature of the metal oxide surface, the degree of surface
hydration, and the pH of the solutions used for deposition. The al-
koxysilane monomers of the general type RSiX 3 were found to be
quite useful for aligning liquid crystals. R is an organofunctional or-
ienting group and X designates a hydrolyzable group attached to the
silicon. Schematic illustrations of the cured coatings and the chemical
formulas for the silane agents are given in Fig. 5. The DMOAP coat-
ing resulted in perpendicular alignment of the liquid crystal; parallel
LlQUID·CRYSTAL DISPLAYS-PACKAGING 225

alignment of the mesomorphic molecules was produced by the cured


MAP layer.
Berreman 20 ,21 calculated the difference in elastic strain energy that
occurs when a nematic fluid lies parallel to grooves in a surface rather
than perpendicular to them. As a consequence of his calculations and
experimental observations made by himself and Dryer,n Berreman
has suggested that the anisotropy in elastic strain energy of the liquid
crystal is sufficient to induce the alignment of the fluid director par-
allel to the grooves in the surface. Furthermore, he concluded that

Fig. 5-Chemlcal formulae for two silane coupling agents and their geometric relation-
ship to a substrate when cured. 19

the elastic energy considerations explain the tendency of some nema-


tic molecules to align perpendicular to a surface that is rough in two
dimensions even if physicochemical forces would normally cause the
molecules to lie parallel to a flat surface of the same solid.
Creagh and Kmetz 22 ,23 have recently suggested an explanation of
surface orientation forces that is a synthesis of both the physico-
chemical and geometric factor hypotheses. Using chemically cleaned
tin oxide coated glass substrates, they investigated a number of dif-
ferent surface aligning conditions. Thorough cleaning with chromic
acid was required because they found that, in the absence of the
cleaning, trace remnants of carbon on the glass surface caused nemat-
ic fluids such as MBBA and PAA (p-azoxyanisole) to align with the
long axes of the molecules parallel to the surface. With the appropri-
ate cleaning, these materials aligned in the perpendicular state.
The data presented in Fig. 6 summarizes the results of their tests
on surface alignment. Cells were prepared with different aligning
226 CHAPTER 12

Surfactant

Grooves None Lecithin Carbon

No
~ndom ti1tIWith EJ
-1 Random II
-1 -1
Yes
!Tilt along grOOIes with EJ
II
Homogeneous
I
Fig. 6-Matrlx 0' alignment resuHs 'or various grooving and surfactant condHlons. 23

agents and either with or without grooves. The grooves were obtained
by rubbing the surfaces with a diamond paste. Fig. 7 is a typical elec-
tron micrograph of a rubbed surface. The tabulated results show that
the evaporated carbon layer gave parallel orientation with the
grooves providing the unidirectionality necessary for uniform parallel
alignment.
With both lecithin-coated and chemically cleaned surfaces, rub-

Fig. 7-Scannlng electron micrograph of a substrate grooved with 1-#lm diamond


paste and then cleaned. 23
LIQUID-CRYSTAL DISPLAYS-PACKAGING 227

bing did not result in uniform parallel alignment. In both situations,


perpendicular orientation occurred for no applied voltage. When a
voltage was applied to the fluid, either one of two configurations ap-
peared. With no grooving, the nematic fluid adopted the random par-
allel condition, whereas when the surfaces were rubbed with the dia-
mond paste, the nematic director adopted a uniform orientation in
the direction of rubbing.
As a result of their experimental observations, Creagh and Kmetz
claimed that the determination of whether a liquid crystal adopts the
perpendicular or parallel orientation can be made on the basis of the
relative surface energy of the substrate and the surface tension of the
fluid. They asserted that when the surface energy of the substrate is
low compared to that of the liquid crystal the fluid does not wet the
surface, and intermolecular forces within the fluid produce the per-
pendicular texture. However, if the relative surface energy of the
solid is high, the fluid wets the substrate, and the long axes of the
molecules align parallel to the surface. They also quoted the results of
Proust et aP8 in support of their theory. Apparently, the surface en-
ergy is lower when the hexadecyltrimethylammoniumbromide mole-
cules are densely packed on the surface than when they are diffusely
packed.
Their observation that the nematic director was normal to thor-
oughly cleaned surfaces is somewhat at variance with the above con-
clusions, since one would expect the solid surface to possess a high
surface free energy, but they have given an explanation. The appar-
ent discrepancy was presumably caused by the presence of a thin
layer of water on the surface. They cited the results of Shafrin and
Zisman,24 namely, that at normal relative humidity, the critical sur-
face tension of glass at 20°C is about 30 dynes/cm for non hydrophilic
liquids. This surface water can only be removed by prolonged heat-
ing. In further support of their theory, Creagh and Kmetz were able
to achieve random parallel alignment on flame-fired platinum sub-
strates. Fig. 8 is a summary of their results on the effect of surface en-
ergy on orientation.
In the model of Creagh and Kmetz, the main effect of grooving ap-
pears to have been the provision of a preferred direction for the mole-
cules in accord with calculations made by Berreman 20 .21 and Grab-
meier et a1. 25.26 The physicochemical forces determined whether the
molecules were parallel or perpendicular to the substrate surface.
They concluded that Chatelain's hypothesis was correct-that the
conventional rubbing technique provided uniform parallel alignment
by grooving an organic surface layer produced by the rubbing medi-
um, be it cloth or paper.
228 CHAPTER 12

Kahn, Taylor, and Schonhorn 27 have expanded upon the ideas dis-
cussed by Creagh and Kmetz. 22 ,23 In essence, Kahn et al agreed with
the concepts propounded by Creagh and Kmetz with regard to the
relative importance of the physicochemical and elastic forces in the

MBBA
Wet Glass C
Lecithin

y
10 dyne,lcm

Fig. 8-Allgnment of MBBA on substrates with different surface energies. 23

fluid. However, they did not describe the physicochemical interaction


between the liquid and the solid in terms of the liquid surface ten-
sion, 'YL, and the surface tension of the solid, 'Ys. Rather they charac-
terized the solid by its critical surface tension, 'Yc, and the liquid by
its surface tension, 'YL. The critical surface tension of a solid is an em-
pirically determined quantity first proposed by Zisman and asso-
ciates 28 to classify low-energy solid surfaces.
As explained by Adamson,29 'Yc is not a fundamental property of
the solid surface alone, but also depends upon the nature of fluid in
contact with the solid. Apparently, it is only a fixed quantity for a
given homologous series of organic liquids on the solid. 'Y c values can
vary somewhat from one homologous series of liquids to another.
Kahn et al,27 also noted that the 'Yc of a solid, in particular one of
high surface energy, can be drastically lowered by the presence of a
layer of liquid molecules that has been preferentially adsorbed at the
solid-liquid interface. If the molecules in the monolayer have both
polar and nonpolar ends, the polar ends will often attach themselves
to the high-energy surface, with the nonpolar portion facing into the
liquid. As far as the bulk liquid is concerned, the nonpolar groups
present a lower surface energy to the fluid than the free solid surface,
and consequently the effective 'Y c is lowered. The adsorbed mono-
layer can consist of intentionally added impurities in the fluid 12-14 or
unintentional impurities. In addition, for certain impurity-free liq-
uids, the molecular structure may be such that the fluid cannot
spread on its own monolayer. These are known as "autophobic" liq-
uids. 28
Proust and Ter-Minassian-Saraga30 have measured the contact
LIQUID-CRYSTAL DISPLAYS-PACKAGING 229

angle 4> and calculated the work of adhesion, W A = 'YL (1 + cos 4», of
MBBA to glass surfaces coated with a monolayer of HTAB. As pre-
viously described,18 the MBBA had either uniform parallel or per-
pendicular orientation depending upon the surface density of the
HTAB. They have found that the contact angle associated with par-
allel alignment was 32° instead of 0°. Also, in spite of the fact that
the alignment changed sharply from parallel to perpendicular due to
a small increase in the surface density of HT AB, the contact angle
only increased slightly-to 37°. The work of adhesion gradually de-
creased during the transition from parallel to perpendicular orienta-
tion. The general criterion enunciated by Creagh and Kmetz 22 .23 that
parallel alignment is caused by stronger solid-liquid forces than those
present with perpendicular alignment was verified, but parallel align-
ment was not associated with wetting of the solid by the surface.
Haller 31 has measured the contact angle at the solid-liquid inter-
face for three different liquid crystals in contact with coated solid
surfaces and has also observed the alignment properties of the vari-
ous liquid crystals in sandwich cells with the specially treated surfac-
es. The surface energy of the glass plates was varied by treatment
with either an organic monolayer or multi-molecular layer. Of the
three liquid crystals, only MBBA exhibited perpendicular orientation
on glass surfaces coated with octadecyltrichlorosilane, HT AB, or bar-
ium stearate. All three liquid crystals, MBBA, BECS (4-n-butyl-4-
ethoxy-2-chlorostilbene) and LiCristal IV (isomeric mixture of 4-
methoxy-4'-butylazoxybenzenes) had surface tensions far in excess of
the 'Yc's of the coatings that were measured with unnamed isotropic
test liquids. It should be noted that the three liquid crystals had dif-
ferent central linkages, and it is possible that this variation in chemi-
cal structure accounted for the diverse results. Haller concluded that
the proposed correlation between wetting and alignment properties
was too general, and specific details of the interaction forces were
necessary for a precise prediction of the alignment properties.
A technique for producing uniform parallel alignment that does
not involve organic coatings has been described by Janning. 32 The
surface preparation consisted of evaporating materials such as gold,
aluminum, platinum, or silicon monoxide onto the substrate at an
angle of 85° to the substrate normal. The films were 100 A or less
thick. Cells containing MBBA and using glass plates coated in this
manner exhibited uniform parallel alignment.
Recently Guyon, Pieranski, and Boix 33 investigated the depen-
dence of liquid-crystal orientation on the angle, 0, between the evapo-
ration direction and the normal to the substrate. Their principal re-
sults were:
230 CHAPTER 12

(1) 0 < 8 < 45°. No preferred direction for the alignment of MBBA
was observed.
(2) 45° < 8 < 80°. Uniform parallel alignment was obtained with the
preferred direction being perpendicular to the plane containing
the direction of evaporation and the normal to the substrate.
(3) 80° < 8 < 90 0. The fluid director possessed a preferred direction
of orientation in the plane of evaporation; however, the direction
did not lie parallel to the substrate, but instead it was tilted out
of the substrate plane at an angle between 20° and 30. °
The authors hypothesized that the evaporated coatings were de-
posited with a sawtooth surface profile whose shape depended on the
oblique angle of incidence. For 45° < 8 < 80, ° it was felt that the long
axes of the fluid lay parallel to the long axes of the sawtooth grooves.
They also argued for 80° < 8 < 90° that the liquid crystal director
was pointed into the teeth of sawtooth. Most of their data was ob-
tained with SiO, but they stated that similar data was achieved with
C and Au.
Complete microscopic confirmation of their model has not yet been
obtained, although using high magnification with an electron micro-
scope, Dixon, Brody, and Hester 34 have observed fine structure in 85°
evaporated SiO x films with the structure oriented in the plane of
evaporation.
Guyon et aP3 did not state the exact nature of the liquid-crystal
orientation with films evaporated at 8 = 0°. Meyerhofer 35 has mea-
sured random parallel alignment for 0 = 0° evaporations. A reason-
able conclusion that can be drawn from the data is that, when in con-
tact with mesomorphic fluids, freshly evaporated SiO x films produce
parallel alignment in liquid crystals through physicochemical forces.
At the present time, it is still not clear whether the directionality in-
duced in the liquid crystal by obliquely evaporated SiO x film is
caused by the sawtooth model or- some other anisotropic property of
the SiO x .

5. Influence of Packaging on Surface Orientation

In the section on packaging we have indicated that some liquid crys-


tals are adversely affected by coming in contact with moisture; conse-
quently, for these materials a good water-tight seal is necessary. The
moisture can increase the fluid conductivity, produce a lowering of
the mesomorphic-isotropic transition, or can participate in the pro-
duction of impurities in the fluid that are adsorbed onto the surface
LIQUID-CRYSTAL DISPLAYS-PACKAGING 231

and that can modify the alignment of the liquid crystal through a
change of the forces at the solid-liquid interface.
Not only may the choice of sealing technique be important for con-
trollable liquid-crystal orientation, but the effect of the interaction at
the glass-liquid crystal interface must be considered in the selection
of the type of glass to be used in the liquid crystal cells. W or kers at E.
Merck Co.36 stated that it was easier to achieve perpendicular align-
ment on borosilicate glass than soda-lime glass when both types of
glass pieces had been heated above lOO°C. They claimed that the al-
kali impurities in the soda-lime glass caused a disorientation of the
liquid crystal in some unknown manner. They were able to prevent
the alkali-induced misorientation by rinsing the he~ted plates in
chromosulfuric acid or water. They suggested that over coating the
glass surface with inorganic thin films such as MgF 2 and Si0 2 should
alleviate the condition. Also, they found that the addition of certain
surfactants such as lecithin or tetraalkyl ammonium salts to the liq-
uid crystal produces good perpendicular alignment even between two
soft glass plates that had been heated above lOO°C.
We have found 37 that even with as much as 0.1% of HTAB added
to MBBA, misorientation gradually occurred on solid glass surfaces
that had been heated to 400°-500°C. When the cells were first made,
good perpendicular alignment was achieved. However, over a period
of time of up to several months at 25-30°C, or within a few hours at
85°C, the fluid slowly lost its perpendicular orientation. Cells made
with fused silica plates did not exhibit this degradation. The data ob-
tained from an ion-scattering analysis of the surfaces of the glass
plates fired at high temperature showed a strong excess of cations, in
particular, alkali ions. These results strongly support the original
suggestion of alkali-induced misalignment.

6. Summary

A short outline of the standard conductive coatings and packaging


methods has been presented.
Most of the chapter has been devoted to a discussion of the different
surface treatments used to obtain controllable molecular alignment
in liquid-crystal cells. The methods reviewed for obtaining parallel or
perpendicular alignment were the use of surfactant additives, organic
coatings on the solid surface, rubbing of the solid surface, and evapo-
ration of inorganic materials. The data in the literature strongly
suggests that the strength of the physicochemical forces at the liquid-
crystal-solid interface is the most important element in determining
232 CHAPTER 12

whether parallel or perpendicular orientation will occur. However,


the details of the interaction are sufficiently complicated that it is
very difficult at the present time to make totally valid predictions
about orientation for a specific solid surface and a particular liquid
crystal. The directionality necessary for uniform parallel alignment
can be provided by physical grooving of the solid surface, unidirec-
tional withdrawal of the glass substrates from a coating solution, or
by using an oblique angle of incidence during the evaporation of inor-
ganic materials onto the glass surfaces.
Finally, two examples have been presented which show the rela-
tionship between packaging techniques and liquid-crystal molecular
orientation.

References
1 J. L. Vossen, "RF Sputtered Transparent Conductors II: The System In203-Sn02," RCA Rev., 32,
p. 289, June 1971.
2 D. B. Fraser and H. D. Cook, "Highly Conductive Transparent Films of Sputtered In2.xSnx-
Oa.y," J. Electrochem. Soc., 119, p. 1368 (1972).
3 F. H. Gillery, "Transparent Conductive Coatings of Indium Oxide," Information Display, 9, p. 17
(1972).
• R. Clary, Optical Coating Lab., Inc., Santa Rosa, California; private communication.
S Emulsltone Solution No. 673, Emulsltone Co., Millburn, N.J.
6 Bulletin on Indium Oxide Conductive Coatings, Optical Coating Laboratory, Inc., Santa Rosa, Calif.
7 Bulletin entitled "Nesa and Nesatron Glass," PPG Industries, Industrial Glass Products, Pittsburgh,
Penna.
6 G. Friedel, "The Mesomorphic States of Matter," Ann. Physique, 18, p. 273 (1922).
9 H. Zocher, Z. Phys. Chern., 132,285 (1928).
10M. F. Schiekel and K. Fahrenschon, "Deformation of Nematic liquid Crystals with Vertical Orien-
tation In Electrical Fields," Appl. Phys. Lett., 19, p. 391 (1971).
11 J. F. Dryer, "Epitaxy of Nematic liquid Crystals," p. 1113 In Liquid Crystals 3, G. H. Brown and M.
M. Labes, eds., Gordon and Breach, London (1973).
12 W. Haas, J. Adams, and J. Flannery, "New Electro-Optic Effect in a Room-Temperature Nematic
liquid Crystal," Phys. Rev. Lett., 25, p. 1326 (1970).
13 T. Uchida, H. Watanabe, and M. Wada, "Molecular Arrangement of Nematic Liquid Crystals," Jap.
J. Appl. Phys., 11, p. 1559, (1972).
14 S. E. Petrie, H. K. Bucher, R. T. Klingbiel, and P. I. Rose, "Aspects of Physical Properties and Ap-
plications of Liquid Crystals," Organic Chemical Bulletin 45, No.2 (1973), Eastman Kodak Co.,
Rochester, N.Y.
151. Haller and H. A. Huggins, Additive for Liouid Crystal Material, U.S. Patent 3,656,834, April 18,
1972.
16 H. Zocher and K. Coper, Z. Phys. Chem., 132, p. 195 (1928).
17 P. Chatelain, Bull. Soc. Franc. Miner. Christ. 66, p. 105 (1943).
18 J. E. Proust, L. Ter-Minassian-Saraga, and E. Guyon, "Orientation of a Nematic Liquid Crystal
By Suitable Boundary Conditions," Sol. Sf. Commun., 11, p. 1227 (1972).
19 F. J. Kahn, "Orientation of Liquid Crystals by Surface Coupling Agents," App. Phys. Lett., 22, p.
386 (1973).
20 D. W. Berreman, "Solid Surface Shapes and the Alignment of an Adjacent Nematic Liquid Crys-
tal," Phys. Rev. Lett., 28, p. 1683 (1972).
21 D. W. Berrernan, "Alignment of Liquid Crystals by Grooved Surfaces," Mol. Cryst. and Liq. Cryst.,
23, p; 215 (1974).
22 L. T. Creagh and A. R. Kmetz, "Performance Advantages of Liquid Crystal Displays with Surfac-
tant-Produced Homogeneous Alignment," Digest of 1972 Soc. for Information Display International
Symp., San Francisco, Calif., p. 90.
LIQUID-CRYSTAL DISPLAYS-PACKAGING 233

23 L. T. Creagh and A. R. Kmetz, "Mechanism of Surface Alignment in Nematic Liquid Crystals,"


Mol. Cryst. and Uq. Cryst., 24, p. 59 (1973).
24 E. G. Shafrin and W. A. Zisman, "Effect of Adsorbed Water on the Spreading of Organic Liquids
on Soda-Lime Glass," J. Arner. Ceramic Soc., 50, p. 478 (1967).
25 J. G. Grabmeier, W. F. Greubel, H. H. Kruger, and U. W. Wolff, "Homogeneous Orientation of
Liquid Crystal Layers," 4th International Liquid Crystal Conf., Kent, Ohio, Aug. 1972, Paper No.
103.
26 U. W. Wolff, W. F. Greubel, and H. H. Kruger, "The Homogeneous Alignment of Liquid Crystal
Layers," Mol. Cryst. and Uq. Cryst., 23, p. 187 (1973).
27 F. J. Kahn, G. N. Taylor, and H. Schonhorn, "Surface-Produced Alignment of Liquid Crystals,"
Proc. IEEE. 61, p. 823 (1973).
28 W. A. Zisman, "Relatiofl of the Equilibrium Contact Angle to Liquid and Solid Constitution," Adv.
Chem. Ser., 43, p. 1 (1964).
29 A. W. Adamson, Physical Chemistry of Solids. 2nd ed. New York: Interscience, 1967, Chap. VII.
30 J. E. Proust and L. Ter-Minassian-Saraga, "Notes des Membres et Correspondants et Notes Pre-
sentees ou Transmisses Par Leurs Soins," C. R. Acad. Sci.• 276C, p. 1731 (1973).
3'1. Haller, "Alignment and Wetting Properties of Nematic Liquids," Appl. Phys. Lett., 24, p. 349
(1974).
32 J. L. Janning, "Thin-Film Surface Orientation for Liquid Crystals," Appl. Phys. Lett., 21, p. 173
(1972).
33 E. Guyon. P. Pieranski, and M. Boix. "On Different Boundary Conditions of Nematic Films Deposit-
ed on Obliquely Evaporated Plates," Letters in Appl. and Eng. Science. 1, p. 19 (1973).
3 4 G. D. Dixon, T. P. Brody, and W. A. Hester, "Alignment Mechanism in Twisted Nematic Layers,"
Appl. Phys. Lett.• 24, p. 47 (1974).
35 D. Meyerhofer, Private Communication.
36 Current Information on Liquid Crystals No.4 (1973), E. Merck Co., Darmstadt, Federal Republic of
Germany.
37 L. Goodman and F. DiGeronimo, "Nematic Liquid Crystal Misalignment Induced by Excess Alkali
Impurities in Soft Glass," 5th International Liquid Crystal Conf., Stockholm. Sweden. June 1974.
Pressure Effects in Sealed Liquid-Crystal Cells

Richard WIlliams

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

Liquid crystal cells are hermetically sealed glass containers complete-


ly filled with liquid. Two plane-parallel plates are sealed all around
the edges to a frit glass spacer. The cell is then filled with liquid
through two holes and sealed off with plugs of fusible metal. This
construction gives rise to some internal pressure effects, because the
thermal expansion coefficient of the liquid is about 100 times that of
the glass. If the cell is filled and sealed off at room temperature, the
liquid will exert a pressure at all higher temperatures. At lower tem-
peratures it will be under tension. The pressure will deform the cell,
making the walls bow out. This makes the volume enclosed by the
cell a little larger and reduces the pressure but does not eliminate it
completely. Some pressure or tension will always remain for tempera-
tures different from the filling temperature. Repeated expansion and
contraction may lead to loss of hermeticity or other cell failure. In
what follows, the magnitude of the effect is calculated and the impor-
tant factors are analyzed.
235
236 CHAPTER 13

Fig. 1 shows the effect schematically and gives the notation used
for the cell dimensions. The thickness of the layer of the liquid crys-
tal is z, the length and width of the cell are b and a, and t is the thick-
ness of the glass plates.

T-T
o
lezzzzzzzzzzmzli ~ ~

(A)

.1
c: ________ -1- _____ :J }
,- a, b

max

(8)

Fig. 1-(A) Expansion and contraction effects as the filled cell Is maintained at a tem-
perature T higher or lower than the filling temperature To. (B) Deformation
of one of the plates: a, b are the width and length and t Is the thickness of
the glass. The maximum displacement wmax Is at the center of the plate.

If the temperature is raised after filling, the liquid expands, and


the cell walls bow out to accommodate the change in volume. We can
neglect the thermal expansion of the glass and the changes of the liq-
uid volume due to changes in pressure, since both these effects are
small compared to the volume change of the liquid due to thermal ex-
pansion. The cell walls are held at the edges and the pressure of the
liquid exerts a uniform force per unit area over the surface.
Consider the case where the cell is filled and sealed at temperature
To and later warmed to temperature T. As the walls bow out, the
edges stay fixed. The maximum displacement, wmax , of the plate
from its unstressed position will be at the center (Fig. 1). We need to
determine the volume change .lv due to this deformation of the cell
walls. The volume Vo of the original undeformed cell is abz. Simple
geometric considerations show that when W max « a the increase in
volume due to the bowing out of the walls is

.lu = ab wmox [lJ


2

.lv is also equal to the increase in volume of the liquid due to thermal
PRESSURE EFFECTS IN SEALED CELLS 237

expansion. Using the thermal expansion coefficient Cl'L of the liquid,


we can express this as
[2J

Since Eqs. [1] and [2] must be equal,

[:3]

To get the pressure p required to give a deformation W max is a


standard problem in the strength of materials. It involves the cell di-
mensions and the mechanical properties of glass. For our particular
case of a plate of uniform thickness held at the edges, l the solution is

[4]

where E is Young's modulus and, for typical glasses, has the value 6
X 1011 dynes/cm 2• C is a tabulated function that depends on the ratio
alb, i.e., the ratio of cell width to cell length. t is the thickness of the
glass. From Eqs. [3] and [4] we get

[5]

For MBBA (N-(p'-methoxybenzylidene}-p-n butylaniline), Cl'L has


the value of 0.85 X 10- 3 over the range of interest. 2 (The volume
change at the nematic-isotropic transition is small compared to the
thermal expansion and will be neglected.)

2. Effect of Temperature Change

The magnitude of the pressure developed by thermal expansion is


shown in Fig. 2 for a cell 1 cm wide, 2 cm long, made of glass 1.0 mm
thick and filled with a layer of MBBA 12.5 ~m thick (% mil). For
these dimensions the value of C is 0.11.
Three cases of filling and sealing are shown, corresponding to three
different temperatures To: 0 0 , 25 0 , and 50 0 e. The pressure change
l1p may be either positive or negative, depending on whether the am-
bient temperature is greater than or less than To. The final cell pres-
sure difference may amount to about 1 atmosphere for an operating
range of 100 0 e. The pressure inside the cell will fluctuate contin-
uously as the ambient temperature changes. This will be a continuous
test of the hermetic sealing plugs. Eq. [5] shows the factors in cell de-
sign that lead to high pressures. These are the thickness of the liquid
238 CHAPTE R 13

CELL DIMENSIONS I X2 em
GLASS THICKNESS' 1.0 mm
I X 106 LIQUID LAYER 12.5,. THICK 15

-IN
0.5XI06 7.5
-...
·ii
~ E
...
...
" u
<I
0
<I 75 100

Fig. 2-Pressure changes /1p that result when a cell Is filled at one temperature, To.
and put In an ambient at another temperature, T.

layer, the thickness of the glass plates, and the overall cell size. The
effect will be most serious for small cells, such as watch displays, and
it can be alleviated by using thinner glass and thinner layers of liquid
crystal.

3. Effect of Glass Thickness

Fig. 3 shows the effect of the thickness of the glass plates for a given
temperature change with other conditions fixed. This is the t 3 depen-

IXl em CELL
3 X10' T-To • 100' e 45
LIQUID LAYER 12.5,. THICK

....
30 -~
...
<I
...
<I
~
I XIO' 15

O'--_O=::::;;.L...-_ _.L-_ _....L.-~O


o 0.5 1.0 1.5
, Imml

Fig. 3-EHect of the thickness t of the glass plates on the pressure change /1Pl00,
caused by heating the cell 100 0 above the filling temperature.
PRESSURE EFFECTS IN SEALED CELLS 239

dence of Eq. [5] and emphasizes the merits of using thin glass, insofar
as this is compatible with other cell requirements. In general, the
pressure effects are less serious for larger cells and would be negligi-
ble for cells larger than 2 inches on a side.

4. The Case of a Rigid Container

In very small cells or in cells made of thick glass, there may be so lit-
tle deformation of the cell walls that the liquid behaves as if it were in
an ideally rigid container. Very high pressures develop and these can
be estimated from readily available thermodynamic data. The quan-
tity required is (ap/ aT)v, the pressure increase per degree of temper-
ature rise when the volume is held constant. From general thermody-
namic arguments this can be related to aL and the compressibility fh
of the liquid.

[6J

For most liquids the magnitude of aL is around 1 X 10- 3 deg- 1 and


the magnitude of fh is around 1 X 10-4 atm -1. The pressure rises by
about 10 atmospheres per degree of temperature rise. This would be a
severe design limitation. About the only way it might arise in practice
would be if the lateral dimensions of the cell were small; if, for exam-
ple the individual elements of a numeric display were made as sepa-
rate closed-off cells.
An interesting possibility, especially important in the rigid con-
tainer case, arises when the cell temperature is lowered after filling
and sealing. The resulting negative pressure tends to make any dis-
solved gas come out of solution in the form of a bubble-a gas "embo-
lism." This is very similar to what happens when a diver gets the
bends. Dissolved gas, equilibrated in a liquid at a high pressure,
comes out of solution when the pressure is lowered, with disastrous
consequences. The same remedy may be useful in both cases-the
use of a helium atmosphere. The solubility of helium in organic liq-
uids is about one-tenth that of air. By handling and storing liquid-
crystal materials under an atmosphere of helium, the gas available to
produce an embolism would be reduced by a factor of ten.
In summary, the liquid-crystal cell, though ideally a closed system,
is one that reacts significantly to changes in external conditions, and
this is an important consideration in applications.
240 CHAPTER 13

Acknowledgment

I am indebted to L. A. Goodman, D. Meyerhofer, E. B. Priestley, and


P. J. Wojtowicz for valuable discussions ofthis problem.

References

1 J. P. Den Hartog, Advanced Strength of Materials, pp. 132, McGraw-Hili Book Co., New York
(1952).
2 M. J. Press and A. S. Arrott, "Expansion Coefficient of Methoxybenzylidene Butylaniline through
the Liquid-Crystal Phase Transition," Phys. Rev., AS, p. 1459, Sept. 1973.
Liquid-Crystal Displays-Electro-Optic Effects
and Addressing Techniques

L. A. Goodman

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

Many of the physical properties of mesomorphic materials, such as


birefringence, optical activity, viscosity, and thermal conductivity are
sensitive to relatively weak external stimuli. Electric fields, magnetic
fields, heat energy, and acoustical energy can all be used to induce
optical effects. At the present time, most of the display-related re-
search is centered on the application of electro-optic effects because
of the relative ease and efficiency of excitation with an applied volt-
age as compared with other means of stimulation. Liquid-crystal elec-
tro-optic effects are important because they do not require the emis-
sion of light; instead they modify the passage of light through the liq-
uid crystal either by light scattering, modulation of optical density, or
color changes. The salient properties are low-voltage operation, very
low power dissipation, size and format flexibility, and washout immu-
nity in high-brightness ambients.
This chapter is divided into three major sections. The first describes
241
242 CHAPTER 14

the various liquid-crystal electro-optic phenomena; the second dis-


cusses important display-related parameters; and the third describes
tlie operation of liquid-crystal devices in matrix-addressed and
beam-scanned modes of operation.

2. Electro-optic Phenomena

Liquid-crystal electro-optic phenomena can be divided into two cate-


gories-those caused only by dielectric forces and those induced by
the combination of dielectric and conduction forces. The two conduc-
tion-induced phenomena discussed later are dynamic scattering and
the storage effect. Four of the dielectric phenomena, or field effects
as they are sometimes known, are discussed first: (1) induced bire-
fringence, (2) twisted nematic effect, (3) guest-host interaction, and
(4) cholesteric-nematic transition.
In all of the present theories about the excitation of nematic or
cholesteric liquids by an electric field, the mesomorphic material is
treated as a continuous elastic anisotropic medium. The Oseen 1-
Frank 2 elastic theory is used to describe the interaction between the
applied field and the fluid. The application of an electric field causes
the liquid crystal to deform. For a material with a positive dielectric
anisotropy, .::l€ = €II - €-L > 0, the director aligns in the direction of
the field; if the dielectric anisotropy is negative, the director tends to
align perpendicular to the applied field. The elastic forces attempt to
restore the field-driven fluid to the initial orientation, which is deter-
mined by the surface alignment. The interplay between dielectric and
elastic torques leads to the occurrence of the threshold voltage or
field. In another chapter in this book,3 the interaction between the liq-
uid crystal and an applied field is discussed in detail. The interested
reader can refer to that chapter and to papers published elsewhere 4- 6
for the calculations of the field-liquid-crystal interaction.

2.1 Field-Induced Birefringence

The first electro-optic effect is field-induced birefringence or defor-


mation of aligned phases. 7- 10 Schematic representations of the fluid
with zero applied volts and for a voltage exceeding the threshold volt-
age are presented in Fig. 1. With no applied voltage, the nematic liq-
uid is in the perpendicular state. In the discussion of induced bire-
fringence and the other effects considered, the assumption is made
that the surface orientation of the molecules remains constant even
when the field is applied, while the voltage-induced deformation in-
creases toward the center of the cell. When the applied voltage ex-
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 243

ceeds the threshold voltage, the liquid crystal distorts if it has nega-
tive dielectric anisotropy. A maximum director rotation of 90° is pos-
sible. The threshold voltage is given in m.k.s. units by

IK;;
V TH = 7rV~, [lJ
fo~f

where K 33 is the bend elastic constant.n· 12


The perpendicular texture is optically isotropic to light propagat-
ing perpendicular to the cell walls. Consequently, with crossed polar-
izer and analyzer, no light is transmitted through the analyzer. Dur-

• ANALYZER

~ 'oi;:'8~"h~""
o
VUIQIIll?l???nm
CONDUCTIVE
COATING
{ GLASS ~

1
. . . POLARIZER

h.

Fig. 1-Schematlc Illustration of the Induced birefringence effect wHh and wHhout an
applied voHage. Thin arrows represent director orientation.

ing fluid deformation, the liquid crystal becomes birefringent to the


transmitted light, and part of the light passes through the analyzer.
The intensity of the emerging light is expressed by 9

I = I p sin 2 2,1.
'f'
sin 2 ~2 [2J

where 0 = [2 7rd ~n(V) fA], I p is the light transmitted through two


parallel polarizers, ¢ is the angle between the input-light optical vec-
tor and the projection of the director on the plane parallel to the cell
walls, d is the cell thickness, ~n(V) is the voltage-induced change in
birefringence, and A is the wavelength of the light.
The transmitted light intensity is maximum when ¢ = 45°. Nor-
mally, the angle ¢ is not well-defined because of the cylindrical sym-
metry that results when a perpendicularly aligned fluid is deformed
by the electric field (see Fig. 2b). However, as described in the surface
244 CHAPTER 14

investigations of Creagh and Kmetz,13 a preferential direction can be


established in a plane parallel to the cell walls by grooving or rubbing
the substrates. Samples prepared in this manner deform with a well-
defined direction for the fluid director.1 4 This preferred direction can
be set at a 45 0 anglE\ to the crossed polarizer and analyzer.

(8 )

Fig. 2-Top views of perpendicularly oriented cells. (A) V = 0 and the long axes of
the molecules are perpendicular to the electrodes. (B) The fluid Is partially
deformed with no prescribed direction.

The maximum value of f1n(V) is the index of refraction anisotro-


py, f1n = nil - n -L. Typically, f1n is about 0.2 to 0.3 This anisotropy
is so large that, for monochromatic radiation, the transmitted light
intensity undergoes many maxima and minima as the voltage in-
creases above threshold. With white light, variable colors can be ob-
served as a function of voltage.
To first order, the frequency response of induced birefringence is
constant in amplitude from low frequency to the molecular disper-
sion frequency in the dielectric constant where the dielectric anisot-
ro{)¥ changes. 9 This property is typical of all the field effects.
For a nematic material with positive dielectric anisotropy, induced
birefringence can also be observed. However, the liquid crystal must
be in the uniform parallel orientation at zero voltS. 15 Above the
threshold voltage, the director aligns itself parallel to the applied
field. With crossed polarizer and analyzer, the voltage dependence of
the light intensity is reversed from that described previously for a
fluid of negative dielectric anisotropy.6,1l
For materials with positive dielectric anisotropy, the threshold
voltage can be as low as 1.0 volt, whereas devices using negative di-
electric anisotropy fluids typically possess threshold voltages in the
4-6 volt range. Since the elastic constants are relatively independent
of material, the difference in threshold voltages is ascribed to the
much larger magnitude of anisotropy normally found in fluids with
positive anisotropy as compared to that occurring in materials with
negative anisotropy.
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 245

2.2 Twisted Nematic Effect

The twisted nematic field effect is probably the most important of


the field effects because of its combined properties of very low volt-
age threshold, low resistive power dissipation, and relatively wide
viewing angle in the reflective mode.
The typical cell structure used in the twisted nematic device is
shown in Fig. 3. 16 The molecules in each surface layer of the liquid
crystal are uniformly aligned in one direction, but with a twist angle
of 90° between the preferred direction for the two surfaces. With no

• ANALYZER

~ e???l~~:':~??I?11~
- - TRANSPillRENT
CO DUCTIVE
r--RPZ2??Z2?1?IZI?ii.i??Zi?lzZ"ZZ???~??r...,., COAT I NG
{ GLASS ~
. . . . POLARIZER

1 h.

(A) ( 9)

Fig. 3-Slde view of twisted nematic effect for (A) V = 0 and (B) V> VTH• The thin
arrows represent the orientation of the nematic molecules.

applied voltage, the bulk fluid distorts so as to provide a gradual


rotation of the molecular alignment from one cell wall to the other.3
With a nematic fluid of positive dielectric anisotropy, voltages ex-
ceeding the threshold voltage cause the nematic director to become
untwisted and to tend to align parallel to the applied field. The
threshold voltage is

[3]

where K 11 is the splay elastic constant, K 2 2 is the twist elastic con-


stant, and the twist angle !po is equal to 7r/2. Eq. [3] is the corrected
version of the expression derived by Schadt and Helfrich for the di-
electric analog to the magnetic case originally solved by LeslieP The
theoretical dependence of the threshold voltage on the anisotropy has
been verified. IS Also, several mixtures with large dielectric anisotropy
have been reported with threshold voltages less than 1.0 volt. I9- 2I
The optical properties of the twisted nematic field effect are par-
246 CHAPTER 14

ticularly interesting. Linearly polarized light propagating perpendic-


ular to the cell is rotated by approximately 90° as it passes through
the fluid when there is no applied voltage. 22 ,23 Maximum light trans-
mission is obtained for the zero-field case by orienting the crossed po-
larizer and analyzer with the polarizer optic axis parallel to one of the
preferred surface alignment directions in the cell. The transmitted
light decreases when the applied voltage exceeds the threshold volt-
age, and the fluid starts to align in the perpendicular state. Fig. 4 pre-
sents the data obtained by Schadt and Helfrich 16 with parallel polar-
izer and analyzer, which can be used instead of crossed polarizer and
analyzer. Extinction is obtained with no applied voltage, while light
transmission occurs for voltages exceeding the threshold voltage.
Hence, depending on the orientation of the polarizer and analyzer, ei-
ther a black-on-white or white-on-black display can be obtained.
The formula presented in Eq. 3 indicates the threshold voltage at
which the director starts to reorient. Gerritsma, DeJeu, and Van Zan-
ten 24 have measured the magnetic threshold by both capacitive and
optical techniques and found that the capacitive threshold is lower
than the optical one. Van Doorn 25 has shown that this difference is to
be expected, since the fluid starts to reorient by the tilting of the di-
rector toward the applied magnetic field before the twist has appre-
ciably changed. Consequently the capacitive threshold, which occurs
when the director starts to tilt toward the applied field, is lower than
the optical threshold, which occurs when the twist becomes suffi-
ciently nonuniform that the optical vector of the light does not "fol-
low" the twist. A similar difference has been observed in twisted
nematic devices excited with electric fields. 26 ,27 Berreman's28 expla-
nation of the static characteristics of electric-field-excited devices is
similar to that of Van Doorn. 25
The data presented in Fig. 4 and the optical results given by Ger-
ritsma et al 24 were obtained with light normally incident upon the
cllll and coaxial with the detector. Recently, the angle and voltage de-
pendence of the light transmission characteristics of twisted nematic
devices prepared by the rubbing technique has been measured by
several investigators. Kobayashi and Takeuchj29 have obtained the
data presented in Fig. 5. The transmitted light as a function of volt-
age is not symmetric with the viewing angle. The curves also demon-
strate that the apparent optical threshold is a function of viewing
angle. Both of these phenomena have been explained in terms of the
relative ability of the incident light to "follow" the twist for different
angles of incidence. 3o As a result, the true threshold is the capacitive
threshold, and it is always lower than the angle-dependent optical
threshold. Nonuniformities in cell appearance caused by reverse
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 247

'"'C
~IO
"
z
o
iii
II)

i
~ 5
<I
a::
f-

VOLTAGE

Fig. 4-Curve (a) shows rotation angle of linearly polarized light versus voltage for a
nematic liquid crystal at room temperature and 1 kHz; curve (b) is transmis-
sion versus voltage with parallel polarizers (Ref. [16]).

tilt 3 ,31 and reverse twist 32 ,33 have been studied and can be eliminated
by the addition of a cholesteric to the nematic and by proper surface
preparation. 33 ,34
With field-induced birefringence, variable colors are transmitted
when the voltage exceeds the critical value. These angle-dependent
color effects are unavoidable and constitute one of the main disad-
vantages of the induced birefringence effect. In the twisted nematic

_~\9_
L'9ht~Detector

Cell
z
o
B'l
!
.'"
100
z
OV
>- 10V
...
<> 4V

.
N
:::;
,. 50
'"oz
20V

-60 -30 30 60

ANGLE 8 (DEG.I

Fig. 5-Light transmission of a twisted nematic cell with crossed polarizers as a func-
tion of turning angle for various values of applied field al 5 kHz. Cell thick-
ness is 30 JJ.m (Ref. [29]).
248 CHAPTER 14

phenomenon, relatively little birefringence occurs except at applied


voltages just above the threshold voltage. For a twisted nematic de-
vice operating either with zero volts or with a voltage more than two
or three times the threshold value, where most of the molecules are
aligned perpendicular to the walls, the principal color phenomena are
minor effects in the fieldless state associated with the inhomogenous
thickness of the fluid. 35
In reflective applications, the small angle visibility of twisted
nematic devices about the normal to the cell can be much better than
with scattering displays. For both scattering and nonscattering de-
vices, a specularly reflecting mirror would normally be used behind
the cell to obtain a bright, legible display. However, not only is the
desired display information seen by the observer, but quite often un-
wanted specular glare can also be detected.
With nonscattering displays, such as the twisted nematic device,
this glare can be prevented by the use of a diffuse reflector in place of
the mirror reflector. However, there is a loss of brightness due to the
depolarization of radiation reflected from the diffuse surface. The
compromise is acceptable because the glare-free viewability is usually
considered more important than high brightness.

PLEOCHROIC
DYE MOLECU LES
\A
ASS

POLARIZER TRA SPARE T


CO OUClOR

Ass L
Fig. 6-Schematic diagrams of electronic color switching phenomenon.

2.3 Guest-Host Effect

A third phenomenon depending solely upon dielectric forces is the


guest-host or electronic color-switching interaction:l6 ,:l7 in which
"guest" pleochroic dyes are incorporated within nematic "host" ma-
terials. The dyes have different absorption coefficients parallel and
perpendicular to their optical axes. As illustrated in Fig. 6, the dye
molecules can be oriented by the liquid crystal. With zero field, the
liquid crystal is in the uniform parallel orientation and the dye mole-
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 249

cules are aligned with long axes parallel to the optical vector of the
linearly polarized light. In this configuration, the dye molecules have
absorption bands in the visible. Above the threshold voltage, the
nematic fluid of positive dielectric anisotropy tends to align parallel
to the field. This is the condition for low dye 'absorption. Consequent-
ly, a color variation can be observed between the two states. Only one
polarizer need be used with this effect. With optimum dye concentra-
tions, optical density changes as large as 1.5 have been measured and
a threshold voltage of approximately 2.0 volts has been observed. 37

2.4 Cholesteric-to-Nematic Transition

The electric-field-induced cholesteric-to-nematic phase transition


was observed by Wysocki et a1. 38 The magnetic analog had been pre-
viously measured by Sackmann et al,39 and the theoretical magnetic
and electric field dependence has been calculated by deGennes 40
and Meyer. 41
The phase transition, which is illustrated in Fig. 7, only occurs in
an electric field with a cholesteric fluid of positive dielectric anisotro-

GLASS
L. CHOLESTERIC M ESOPH A SE
(SCATTERING TEXTURE I

..
-= - ~~ . /
/' ~Z! ~~~ .:. !. /
~!
"eo & ~
.~~ ~

./
-
~
/
~ ,. ~~
Z~
- ... .,. ..
:--~
i-
! .;
~
/' /
TRANSPARENT
CONOUCTIVE
COAT I NG

Fig. 7-Side view representation of the electric-field-induced cholesteric-to-nematic


phase change.

py. The cholesteric planes are approximately perpendicular to the


cell walls with zero applied field. The helical axes have random orien-
tation and this state is strongly scattering in appearance. 42 ,43 As the
electric field approaches the critical value, the helices begin to un-
wind and dilate. Above the threshold field, all the molecules, except
for surface layers, are aligned parallel to the electric field. This latter
condition is the perpendicular or homeotropic texture. 38 When the
field is lowered below threshold, the scattering texture returns. The
theoretical calculations performed by deGennes:l ,4o result in the fol-
250 CHAPTER 14

lowing expression for the threshold field.

E
TH
= -7r
Po
2 (K- -22 )112
.6.flo '
[4J

where Po is the undeformed helix pitch.


Meyer 44 performed optical measurements of the magnetically"in-
duced pitch dilation. By mixing nematic and cholesteric materials,
Durand et al 45 were able to vary the pitch and to verify that the
threshold magnetic field was inversely proportional to the unde-
formed helix pitch.
In the original experiment of Wysocki et aI, the threshold field was
about 105 V/cm. 38 Heilmeier and Goldmacher 46 reduced the thresh-
old field to 2 X 10 4 V/cm by using a mixture of cholesteric and high
positive dielectric anisotropy nematic materials. More recently, the
threshold field has been reduced to 5 X 103 V/cm by use of the posi-
tive biphenyl compounds. 20,47
The model presented so far is accurate, but incomplete. In actuali-
ty, the scattering texture with the helical axes parallel to the cell
walls is not a stable state without an applied field. Rather, it is a met-
astable state that has a lifetime of from minutes to months, depend-
ing upon the surface alignment, fluid thickness, and pitch. 48 The sta-
ble state is the planar or Grandjean texture. Kahn 43 made the initial
measurements of the various steps in the texture changes. Additional
experiments with both electric and magnetic fields have added more
detail to the model.
With no applied electric field, the liquid crystal exists in the planar
texture, in which it remains until the voltage exceeds the critical
field, EH, which is proportional to 1/ VPOL. 49-51 At E H, a square
grid pattern of periodic distortions of the helical axes occurs. 42 ,52-54
In these perturbations, the direction of the helix axis periodically
varies between being perpendicular and somewhat non perpendicular
to the cell walls. As the field increases further, the distortions grow
until the helix planes become perpendicular to the walls. 42 .43 With
further increase in voltage, the fluid becomes perpendicularly aligned
as described previously. When the voltage is shut off, the fluid re-
turns to the metastable scattering state. The transient decay from the
perpendicular texture to the scattering state is retarded by the pres-
ence of a bias voltage that is below the threshold voltage for the field-
induced transition. 55 ,56 For no bias level, the decay to the scattering
texture proceeds rapidly with the natural relaxation time. As the bias
level approaches the threshold voltage, the relaxation slows consider-
ably.
LIQUID-CRYSTAL DISPlAYS-ElECTRO-OPTIC EFFECTS 251

Greubel 57 has recently analyzed the cholesteric-to-nematic transi-


tion for an applied voltage. With perpendicular alignment of the liq-
uid crystal molecules at the cell surfaces, he found that the field tran-
sition for the cholesteric-to-nematic transition was higher than the
opposite nematic-to-cholesteric transition. With one mixture, the
ratio of field thresholds was approximately 2.5. This work points out
the iinportance of the proper surface orientation and cell cleanliness
in achieving the best bistability.
White and Taylor58 have described a new device that combines
both the guest-host effect and the cholesteric-to-nematic transition.
The pleochroic dye is added to a cholesteric material and the cell
transmission changes when the cholesteric undergoes the cholesteric-
to-nematic transition. Because of the rotational symmetry of the long
axes of the dye molecules about the cholesteric helical axes, contrast
ratios of greater than 4 to 1 were obtained without the use of a polar-
izer.
2.5 Dynamic Scattering

In nematic materials with negative dielectric anisotropy and electri-


cal resistivity less than 1-2 X 10 10 ohm-cm, conduction-induced fluid
flow occurs during the application of an applied voltage. The wide-
angle forward-scattering phenomenon known as dynamic. scatter-
ing 59 ,60 is the most important manifestation of the turbulence that

( AI (81 (C I

Fig. a-Side view of the various steps In the formation of dynamic scattering for uni-
=
form parallel orientation. (A) V 0, (8) V = Vw, and (e) V> Vw.

accompanies the electrohydrodynamic flow. The light scattering


arises from micron-sized birefringent regions in the turbulent fluid. 61
The various steps in the production of dynamic scattering are por-
trayed schematically in Fig. 8 for initial uniform parallel alignment.
Below voltage V w, no change in orientation occurs. At V w, the fluid
becomes unstable and it deforms into the periodic structure shown in
Fig. 8. 62 Two similar mechanisms are responsible for the creation of
the instability, one for ac voltages and a second for dc voltages. Both
252 CHAPTER 14

interactions require the presence of space charge in the fluid, but dif-
fer in the means by which the space charge is generated.
In the ac case, space-charge separation perpendicular to the ap-
plied field is caused by the anisotropy in conductivity.62,63 The ap-
plied field produces a force upon the liquid crystal because of the
space charge. This stimulus drags the fluid toward the walls. The
cells walls impose boundary conditions that necessitate vortical flow
of the fluid. The fluid shear torque aligns the director in the direction
of the fluid flow, while the dielectric and elastic forces oppose the
fluid deformation. At the threshold voltage, the fluid becomes unsta-
ble and the periodic distortion takes place.
When the cell is observed through a microscope, the periodic defor-
mation appears as a series of alternating dark and light domains,
known as Williams' domains that run perpendicular to the original
homogeneous alignment. The wavelength of the periodic deformation
is determined by the thickness of the cell. 64 The birefringence of the
nematic fluid and the periodicity of the instability combine to form
periodic cylindrical lenses in the fluid. 65 The domain lines are the re-
sult of light focused by the periodic array of lenses. This periodic lens
array also acts as a transmission phase grating. Consequently, co-
linear diffraction spots can be observed when a laser beam propa-
gates through the liquid crystal lens array.61,66,67
The presence of fluid flow in the seemingly static periodic domains
has been observed by the motion of dust particles in the fluid. 65
Below the threshold voltage, the dust particles are stationary. At
threshold, they move in an oscillatory pattern closely related to the
domain spacing. The velocity of the particles increases with increas-
ing voltage and with decreasing cell thickness.
Neglecting the frequency response of the electrohydrodynamic
flow, Helfrich62 calculated the threshold voltage for the domain in-
stability. A slightly rewritten form 68 of his expression is

V o2
VW 2

[5J

where 0' 1- and 0'11 are the perpendicular and parallel components of
the conductivity, and 170 and 'Yare viscosity coefficients. Eq. [5] is in
good agreement with the experimental data for p- azoxyanisole. The
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 253

Orsay Liquid Crystal Group 68-70 solved the electrohydrodynamic


problem for a variable frequency, sinusoidal voltage source. The fluid
instability occurs at the frequency-dependent threshold voltage

V o2(1 + (2rr{)2T2)
[6J
V W2 = f2 - (1 + (2rrfJ"T2)
where { is the frequency, T = fllfo/<T is a dielectric relaxation frequen-
cy, and Vo and r are the same terms as in Eq. 5. A cutoff frequency

results from Eq. [6]. The cutoff frequency is directly proportional to


the conductivity.
The theoretical analysis predicts the existence of two regimes with
different frequency dependences. For applied frequencies, f < fe, the
space charge in the fluid oscillates at the same frequency as the driv-
ing signal. This is the region of cellular fluid flow described in the
preceding discussion. Because this region exists for frequencies less
than the dielectric relaxation ftequency, it is called the "conduction
regime." The thickness-independent portion of the solid voltage-fre-
quency curve in Fig. 9 is experimental verification of Eq. 6.
When the applied frequency is greater than {e, the space charge
does not oscillate. The fluid interacts with the applied field to result
in the "dielectric regime." The threshold field is proportional to {1/2
(see Fig. 9). Contrary to the low-frequency situation, the field is
thickness independent. Periodic deformations, known as "chevrons,"
result from the field-fluid interaction. The spatial frequency of the
chevron striations is a monotonically increasing function of the drive
frequency.
In the calculation of the threshold voltage, Helfrich assumed that
the spatial periodicity of the fluid deformation was proportional to
the thickness of the cell. Recently, Penz and Ford 72 ,73 have solved the
boundary-value problem associated with the electrohydrodynamic
flow process. They have reproduced Helfrich's results and have also
shown several other possible solutions that may account for the
higher-order instabilities that cause turbulent fluid flow. Meyerhof-
er 74 has analytically solved the two-dimensional problem by making
one simplifying assumption. He has been able to obtain good agree-
ment between the experimental results and the calculated frequency
dependence of the domain spacing and the threshold voltage.
254 CHAPTER 14

Experimentally, it is observed that, when the applied voltage sur-


passes the threshold voltage by an increasing amount, the rotational
velocity of the fluid increases. 75 The fluid gradually becomes more
turbulent until the applied voltage exceeds twice the threshold volt-
age. The intense wide-angle forward scattering accompanying the
strong turbulence is the dynamic scattering region. 59 •61 Dynamic scat-
tering only happens below the critical frequency and above the
threshold voltage (see Fig. 9). As the voltage increases even further
(beyond twice the threshold voltage), the scattering likewise increases
and the fluid becomes even more turbulent 61 with virtually all traces
of the underlying domain structure disappearing. The increasing tur-
bulence as a function of voltage causes the scattering intensity and
transient kinetics to be angle dependent. 76- 78

100
THICKNESS
(mils)
,
I • 4
DYNAMIC SCATTERING I
I v 2
fc
I
10 - -....
---11
_ _.........-..... o I
FIELD INDUCED REALIGNMENT
____ v+_o __ c __ o_ +2~ _0 _ _ -.J/l-_-+- 4 i
• !
4
+ !
8

10 100 IK 10K
FREQUENCY (Hz)

Fig. 9-Varlous threshold phenomena for nematic fluids with negative dielectric an-
Isotropy and perpendicular alignment. The dashed horizontal line is the
threshold voltage for Induced birefringence. The curved solid line describes
the frequency dependence of the threshold voltage for domains. The sloped
dashed lines are the threshold plots for chevron formation. The material Is
MBBA at 25°C (Ref. [86]).

Though dynamic .scattering usually results when the applied volt-


age exceeds the domain voltage, this need not be true. Dynamic scat-
tering only seems to occur if the fluid is thick enough (~6 /-lm) and
has sufficiently low resistivity (less than 1-2 X 1010 ohm-em) and for
negative dielectric anisotropy. Domains have been observed without
dynamic scattering when any of the preceding three conditions have
been violated. 79 ,80 With the high resistivity and thin cells, the spatial
frequency of the domains is voltage-dependent. 81 At the present time,
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 255

no theory exists that defines the exact relationships between the vari-
ous degrees of fluid instability.
The domain instability and dynamic scattering are also observed
for both low-frequency ac and for dc applied voltage. The volume
space charge, necessary for hydrodynamic motion, is not produced by
the conductivity anisotropy, but by injection of charge from the elec-
trodes. 70,82-84 Meyerhofer and Sussman85 have measured the voltage-
frequency plot for the formation of the domains from very-low-fre-
quency ac to the cutoff frequency fc. Below a certain frequency,
which they relate to the transit time for ions, they have found that
the domain threshold voltage decreases from the ac value toward the
dc value. Also, the domain spacing changes below the inverse transit-
time frequency. This transit-time frequency is of the order of 5 to 10
Hz. At present, the injection mechanism is unknown. It has been hy-
pothesized that the double-layer space charge present at the elec-
trode-liquid interface is responsible for the fluid flow. 86 Vortical fluid
flow 87 and laser diffraction patterns67 have also been observed with
dc applied voltage.
Most of the discussion presented above for uniform parallel align-
ment is still valid with zero-field perpendicular orientation. However,
there is one change. The voltage sequence for the production of dy-
namic scattering has an extra step. For most of the materials used at
present, the voltage threshold for induced birefringence is lower than
the threshold for domain formation (see Fig. 9), although there are
exceptions.!l As a function of increasing voltage, the fluid progresses
from the undeformed state to the induced birefringence texture, then
to the presence of domains, and finally to the occurrence of dynamic
scattering.
Initial perpendicular alignment provides for greater circular sym-
metry in the scattering distribution. As suggested in Fig. 2, the director
for the deformed fluid has only medium range order, approximately
50 J,tm or less. The projections of the directors in a plane parallel to
the fluid are randomly oriented. Because of the circular symmetry
about the axis perpendicular to the cell, the laser diffraction pattern
consists of a set of circular rings instead of colinear spotS. 6l ,66 The
same symmetry is observed for the angular dependence of the dy-
namic scattering 76 (see Fig. 10).

2.6 Storage Mode

Optical storage effects in mixtures of nematic and cholesteric materi-


als with negative dielectric anisotropy were first observed by Heil-
meier and Goldmacher.88 They reported the following sequence of
256 CHAPTER 14

events (see Fig. 11). Initially, with no applied voltage, the sample was
in a relatively clear state. The application of a dc or low-frequency ac
voltage of sufficient magnitude induced the intense scattering known
as dynamic scattering. When the voltage was removed, the dynamic
scattering disappeared, but a quasi-permanent, forward scattering

(0) (b)

(e1

Fig. 10-Diffractlon patterns for a perpendicularly oriented cell: (a) V = 10 Vpp, very
light diffuse scattering; (b) V= 13 Vpp, diffraction rings due to domain for-
mation; (c) V = 16 Vpp, strong diffuse scattering characteristic of dynamic
scattering. The black square In each picture is a piece of tape placed on the
screen.

state remained. The reported decay time was on the order of hours at
elevated temperature. The scattering texture could be returned to the
clear state by the application of an audio frequency signal (greater
than 500-1000 Hz). Since the original reports, other investigators
have also observed the same effect.89-91 The off-state has been identi-
fied as the Grandjean texture. 89 Due to imperfections in the Grand-
jean planes, the off-state is slightly scattering.
Rondelez, Gerritsma, and Arnould 54 have reported the presence of
two-dimensional deformations at the threshold voltage for scattering.
Electrohydrodynamic instabilities were first predicted for negative
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 257

cholesteric materials by Helfrich. 5o Hurault51 has combined Hel-


frich's theory with the time dependent formalism used by Dubois-
Violette, deGennes, and Parodi. 68 His calculations predicted a volt-
age-frequency relationship similar to that observed for pure nematics
of negative dielectric anisotropy (see Fig. 9). Experimental verifica-

CHOLE:STE:RrC ....TERr .. l
«PLANAR TEXTURE )

/
OFF STATE " WR ITE: ' MODE: / STORA GE MODE

.10.. ,.. ,-1 . ...


1··II'·, . ....1l'4
' 1" " 'lf .. ,,"

"'....1 """"'"
.,.. ,,......
•,· .... ·'1.1'l1'li
... I I ~I I I . , ....

.. ERASE · MODE

Fig. 11-Schematic illustration of the effect of low- and high-frequency signals on cho-
lesteric fluids with negative dielectric anisotropy.

tion of both "conduction" and "dielectric" regimes has been estab-


lished. 54 ,87 The domain periodicity is proportional to (PoL) 1/2 in
agreement with theory, where Po is the zero field value of the pitch
and L is the cell thickness. The threshold voltage in the conduction
regime is not thickness independent, but is proportional to (LIP o) 1/2.
The exact means by which the grid structure becomes distorted so
as to form the strongly scattering state is not known. The scattering
texture is approximately the same as that found in the cholesteric-
nematic transition,48.6o,89 even th-ough the mechanisms for producing
the scattering states are probably quite different. As explained, in
cholesteric-nematic mixtures of positive dielectric anisotropy, the di-
electric forces are sufficient to tilt the fluid into the scattering state.
With the materials of negative dielectric anisotropy that are used in
the storage effect, the strongly scattering state must arise from the
strong fluid deformations associated with dynamic scattering. The
light scattered from the storage state is relatively independent of the
258 CHAPTER 14

concentration of cholesteric material, the direction of the incident


light, and the cell thickness. 9o
The restoration of the planar state by the applied field is purely a
dielectric interaction. It can only take place when the signal frequen-
cy is greater than cutoff frequency fe, which is proportional to the
conductivity of the liquid crystal.

2.7 Transient Response

The theoretical expressions describing the transient response of the


fluid deformations are of the same general form for the different ef-
fects. A characteristic response time for the director reorientation is
given approximately by91

[7J

where 1/ is the proper fluid viscosity, E is the applied field, K is the


appropriate elastic constant, and q is the wave-vector of the distur-
bance. For the phenomena occurring in pure nematics, the wave-vec-
tor is approximately 7r/L where L is the cell thickness. Consequently,
the rise time and decay time should be of the forms

[8J
and

[9J

Experimental observations for induced birefringence,14,92 twisted


nematic,91 and dynamic scattering60 ,78,84 are in reasonable agreement
with the theory. For a 12-/lm-thick fluid, 10 V dc, and 20°C, rise
times of the order of 10 msec have been observed with the twisted
nematic effect and 200 to 300 msec for dynamic scattering. Decay
times are approximately 400 msec for the twisted device and 100
msec for dynamic scattering. For both dynamic scattering and the
field effects, the presence of high voltages causes second-order effects
to occur and the decay time becomes voltage dependent. 93 ,94
The response times for both field- and conduction-induced phe-
nomena can be changed by the presence of a second voltage source
whose frequency is above some critical cutoff frequency, while the
main source has a frequency less the critical frequency. For dynamic
scattering, the critical frequency is fe, which is inversely proportional
to the dielectric relaxation time of the fluid. For an applied frequency
f > fe, the conduction torques do not affect the fluid and, through
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 259

the negative dielectric anisotropy, the dielectric torque causes a re-


turn of the fluid to its nonscattering state. Consequently, the decay
time can be significantly shortened as compared to the case when no
high-frequency source excites the ce11. 95 ,96
For field-effect materials, the critical cutoff frequency occurs only
in materials that have positive dielectric anisotropy at low frequen-
cies. Above the critical frequency, the dielectric anisotropy is nega-
tive. Several materials have been developed with a critical frequency
as low as a few kHz at 25°C.97,98 As with dynamic scattering, the ap-
plication of a high-frequency source can produce a decay time much
shorter than the natural decay time. 98-101 The decay time is inversely
proportional to the square of the amplitude of the high-frequency
voltage. 98 ,100
In the cholesteric-nematic phase change with LIP o > 1, the wave-
vector is given by 7rIP o not 7r1L. The experimental rise and decay
times are consistent with the theory for the field-induced phase tran-
sition. 91
The texture change from the highly scattering cholesteric state to
the Grandjean texture is described by different kinetic relationships.
The erasure time of the scattering state for the storage effect is pro-
portional to v-m where V is the audio frequency signal and m varies
between 1 and 3 depending upon the materia1. 88 ,102 The natural
decay time from the scattering to planar texture is approximately ex-
ponentially dependent on the LIP 0 rati065 and the inverse of the
sample temperature. 102 The inverse temperature dependence
suggests that the decay time of the storage state is also directly pro-
portional to the viscosity, which is exponentially dependent on the
inverse temperature. 86

3. Display-Related Parameters

3.1 Display Life

The determination of the operating life is fraught with complications


because of the difficulty of defining the conditions that describe the
end of useful operation. Subjective evaluation of the steady-state
cosmetic appearance and quantitative examination of the variations
in response time, nematic-isotropic temperature, and power dissipa-
tion are necessary. The changes in cosmetic appearance and response
times are usually manifestations of misalignment of the fluid at the
liquid-solid interface. The misalignment may be caused either by the
application of voltage or by chemical interaction between the fluid
and the substrate surface. Time-dependent variations of the current
260 CHAPTER 14

may also arise from chemical interaction between the fluid and the
cell walls. Some of the commonly used liquid-crystal materials are de-
leteriously affected by the presence of moisture and UV radiation.
Proper cell packaging is then necessary to minimize these two un-
wanted agents.
Sussman 103 has examined the dc electrochemical failure mecha-
nism in the dynamic scattering material p- methoxybenzylidene-p'-
aminophenyl acetate (APAPA). Using the loss of 50% of the cell scat-
tering area as his criterion for the end of life, he showed that the
amount of charge passed through the cell determined the operating
life. The failure mode was traced to the production of an insulating
film at the anode. The utilization of ac drive signals, instead of dc,
greatly diminishes the likelihood of failure being caused by electro-
chemical effects. Consequently, commercial dynamic scattering dis-
plays are driven by ac signals. AC operating life is cited as being
greater than 5-10,000 hours. 76 ,104 Equally long operation is to be ex-
pected from field-effect displays operating with ac excitation.

3.2 Temperature Dependence

Until the resurgence in liquid-crystal research in the 1960's, most of


the mesomorphic materials were solid at room temperature. Today,
there are many liquid-crystal systems that exhibit the mesophase
over a wide temperature range around 20°C.
The temperature variation of the fluid properties is important for
display applications. The rise and decay times are directly propor-
tional to the fluid viscosity as shown in Eqs. [8] and [9]. Since the vis-
cosity is approximately exponentially dependent on the inverse tem-
perature, the response times are strongly temperature dependent. 78 ,86
At low temperatures, even when the material is still mesomorphic,
the viscosity may be so high as to preclude the operation of the liq-
uid-crystal display because of the sluggish transient characteristics.
The threshold voltage for the field effects should be mildly temper-
ature sensitive due to the relatively weak temperature variation of
the elastic constants and the dielectric anisotropy. Experimental re-
sults on the induced birefringence phenomena confirm this statement
so long as the operating temperature is less than 95% of the nematic-
isotropic temperature. 92 ,105 Measurements show that the threshold
voltage and contrast ratio in dynamic scattering devices are almost
completely insensitive to temperature throughout the nematic
range. 78
The conductivity in nematics is governed by ionic equilibrium 106
and is inversely proportional to the viscosity and square root of the
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 261

dissociation constant. Both the viscosity and equilibrium constant


are exponentially dependent on temperature as is the conductivity.
The critical cutoff frequency for both conduction-induced and field
effect phenomena is exponentially dependent on temperature. In the
former case, the temperature dependence occurs because the cutoff
frequency is proportional to the conductivity, whereas for field effect
materials with positive dielectric anisotropy the exponential tem-
perature dependence is a property of the molecular relaxation process
which is responsible for the cutoff frequency.
The variation with temperature of the different properties of the
materials must be taken into account during device design so that the
driving signal can always induce the electro-optic effect over the de-
sired temperature range with acceptable speed, wide viewing angle
contrast ratio, and low power consumption.

4. Addressing Techniques

The presentation of visual information by a display requires a meth-


od or methods for exciting multiple positions in the display medium.
The process of transmitting signal information throughout the dis-
play and exciting the different positions in the display medium is
known as addressing. Two general approaches to addressing are use-
ful with liquid crystals. One method involves beam steering, which
includes electron-beam addressing (as performed in a cathode ray
tube) and light-beam scanning. The alternative approach, which is
discussed first, is that of matrix addressing or multiplexing. The two
words are equivalent, but, for historical reasons, matrix addressing is
used when referring to displays with a large number of elements, and
multiplexing is reserved for displays with a relatively small number
of elements.

4.1 Matrix Addressing

One of the strong motivating factors in liquid-crystal research is the


possibility of constructing two-dimensional displays that dissipate
little power. The third dimension, that of the glass-liquid-crystal-
glass sandwich, is of the order of 1fs inch and is small compared to the
other two dimensions. An example of a multi-element liquid-crystal
display is the seven-segment, five-digit dynamic scattering display
shown in Fig. 12. Each segment of each digit could be individually ad-
dressed by a driving signal whose frequency components are lower
than the critical cutoff frequency, but this approach is wasteful of
both driving circuitry and interconnections between the display and
262 CHAPTER 14

DIG IT BACKPLANE LEADS

//
r-l, rJL,r,rol,, r-l, ,.R_,
I
I
I,
I I, •
I
'I
I I
I
I

: :: :: : 1 ::
J I

f
I I I I I
LI ....... __ J' I ____ JI ,I.. _ ___ ..I 'L "
,
____ J
"
L. ___ J, ,

SEG MEN T LEADS

Fig. 12-Top view of a seven-segment five-digit display with electrode leads.

the circuitry. A much more economical approach is one in which the


display is rearranged inta an X-Y matrix as indicated in Fig. 13. One
segment from the first digit is connected to its counterpart on each of
the other digits. This interconnection scheme is repeated for the
other six segments of the first digit and their counterparts. The seven
row lines in the matrix represent the seven coordinated segment
leads, while the five column lines are the same as the five digit back-
plane leads. Consequently, the five-digit seven-segment display with
35 leads can be electrically viewed as a five-by-seven matrix with only
12 leads. In general, a display with M X N resolution elements can be
treated as a matrix array in which only M + N leads are required.

~2 j. ~ +\ ';2 + ';2
L ---.l
- 'f2 -'f2 • ~2 • ~2

~2 r. ~
L --.J
+ 'f2 + '92 +~2

-V'2 - 'f2 -~2 - ~2

-~2 -\ ~2 -V~

2 - V'2 - V'2 - V'2 -V'2


SEGMENTS

2 - V'2 - V'2 - V'2 - V'2

o o o o
INPUT VOLTAGE
DIGITS

Fig. 13-Representatlon of the seven-segment five-digit display with leads rearranged


In matrix fashion. Half-voltage selection pulses are applied.
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 263

...a::-
a
... UI
t-!::
t-z
c(:>
u .
UlIII
a::
t-c(
:J:_
C!I
:::i

VOLTAGE (VOLTS)

Fig. 14-Typical scattered light versus voltage curve for dynamic scattering.

For the proper operation of a matrix array, it is necessary to excite


only the desired element in the matrix and no other. As an example,
let us discuss the 5 X 7 array. Only the first and third segments of the
second digit should be scattering light with all the other segments
being nonscattering. For this situation to occur, the voltages ± V /2
must be less (in an absolute magnitude sense) than the threshold
voltage required to initiate light scattering. The light scattered versus
voltage transfer function for a typical dynamic scattering cell is given
in Fig. 14. Since the contrast ratio at 2Vr can be in excess of 20:1, it
would appear that dynamic scattering displays can be matrix-ad-
dressed without any difficulty.
So far, the matrix array has been treated in a purely static fashion.
In actuality, the segment data ent.ers in parallel from the drive cir-
cuitry and each digit is selected sequentially in time. After all of the
digits have been addressed, the cycle is repeated. The temporal de-
pendence of an arbitrary segment is presented in Fig. 15. The seg-

VOLTS t-
i
.V :J
...aa::
III
t-
o r-~+;--~--~~----~
!cu
UI
-~2 T
FRAME
TIME
TIME

(0) (b)

Fig. 15-(a) Applied voltage versus time for an arbitrary segment and (b) time depen-
dence of light scattering for the same segment.
264 CHAPTER 14

ment is only excited for a time TIN where T is the frame time of the
data (typically 30 to 60 Hz) and N is the number of digits in the dis-
play. The scattered light as a function of time is modified from the
voltage waveform by the finite response times of the phenomenon
(see Fig. 15B).
Several devices and circuit parameters must be properly controlled
to maximize the contrast ratio when operating in a scanning mode.
One important consideration is that the rise time should be as short
as possible. In practice, fast rise times are only achieved by using
voltages that are far greater than twice the threshold voltage, and
therefore elements that should be off turn on. Consequently, the con-
tradictory requirements of short rise times and sufficient half-select
capability restrict the number of digits that can be addressed in a
multiplexing mode.
In order to improve the applied-voltage discrimination ratio for
matrix arrays, the so-called one-third-voltage selection method
shown in Fig. 16 can be used. This method clearly offers the advan-
tage over the half-select method of applying more voltage to the "on"
elements without exceeding the threshold voltage for the elements
that are supposed to be nonscattering.
Second, when the decay time is longer than the frame time T, the
cell integrates the successive series of input signals that occur every
frame. So long as the integration property is obeyed, it has been
found that both field-effect 107 •108 and dynamic-scattering 109 devices
respond to the driving signal in a root-mean-square fashion. As a con-
sequence of the rms behavior, the contrast ratio for a scanned multi-

"l3 -"l3 "l3 vl3 "l3 -~

'"'" \
-~3 +\ V'3 -'+3 -~3

~ +"l3 +"l3 f+vl +~ +"l3 +rv-13


o
> - -v '+~ -1"vi3 -v -y
...::> "l3 ~3 ~3 ~3 ~3

11- V'3 -~3 +~3 -V'3 -V'3 +'3


z
-~3 -V'3 + '+3 - V'3 -'+3 -~3
SEGMENTS
-'l
~3 ~3 + V'3 V'3 - ~3 -~'3

o -2"l3
I
0 0 o
INPUT VOLTAGES

DIGITS

Fig. 16-Applied voltages for the one-third selection method.


LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 265

plexed display can be no higher than the contrast ratio obtained with
a continuous-drive waveform whose rms amplitude is the same as the
rms content of the scanned waveform. Kmetz 107 has calculated the
ratio of the rms value for an on element to that for an off element in a
V: V /3 scheme to be

[lOJ

where N is the number of digits being scanned.


Because of the rms behavior, the contrast ratio is reduced at a fixed
viewing angle. In addition the angular dependence of the contrast
ratio, which is most noticeable when the applied voltage is only some-
what greater than the threshold voltage during continuous drive con-
ditions, becomes more important. For example, in Fig. 5, the level of
transmission at 4 V is very asymmetric in its angular dependence,
whereas, at 10 V, it is not. In a multiplexed mode, as the number of
digits increases, the rms behavior dictates that the angular depen-
dence of the transmitted light becomes closer to that for the 4-V
curve than for the 10-V curve. Consequently, the number of digits
that can be multiplexed with good contrast is not only a function of
the driving signal amplitudes, but also is dependent upon the viewing
angle.
Another condition for the proper implementation of the present
multiplexing technique is that enough of the power in the driving
signal be at frequencies much lower than the cutoff frequency of the
electro-optic effect being utilized. Frequency components just under
the cutoff frequency are not as effective as lower-frequency compo-
nents at producing the excited state, as can be seen for dynamic scat-
tering in Fig. 9. As explained previously, signals whose frequency con-
tent is higher than the cutoff frequency cause a return to the unexcited
state.
Alt and Pleshko 109 have extended Kmetz's analysis of rms-re-
sponding liquid-crystal devices. They have shown that the voltage-
drive configuration that optimizes the number of scanned digits is
not the V: V /3 scheme, but one in which the ratio of peak voltage to
bias voltage is greater than 3:1. They also explicitly demonstrate that
the more nonlinear the transmitted light versus voltage curve, the
greater the multiplexing capability for a given set of drive-signal am-
plitudes.
Varying degrees of success have been reported for matrix ad-
dressing liquid-crystal displays. For displays using dynamic scatter-
ing, anywhere from 3 or 4 digits to 7 digits have been reported as the
266 CHAPTER 14

maximum addressing capability.1°7,l1o Twisted nematic devices pos-


sess approximately the same matrix addressing limitations as dynam-
ic scattering. 107 Hareng, Assouline, and Leiba III have reported ma-
trix addressing a 50 X 50 element array, while Schiekel and Fahren-
schon 112 have successfully operated a 100 X 100 array. Both of these
induced birefringence devices are two-color displays with rise times
on the order of 1 second.1 08,l1I Due to the nature of the electro-optic
process, the display appearance is a sensitive function of viewing
angle, fluid thickness, voltage, and temperature. The small field of
view probably limits matrix-addressed birefringence devices to pro-
jection display applications. Takata et aPlO have reported the opera-
tion of a 260 X 260 liquid-crystal display using the storage-mode ef-
fect. Because of the slow rise times at the voltage levels appropriate
to matrix addressing, 10 to 20 seconds are required to adddress the
whole display. The long decay times associated with the storage effect
permit the maintenance of the displayed information for hours or
more and the rms behavior does not occur.
The cholesteric-to-nematic phase transition effects have also been
utilized in matrix-addressed displays.55,113 Up to 28 lines have been
scanned in the V: V /3 mode with a bias voltage of 35 Vrms and a con-
trast ratio of 15:1. The relatively large multiplexing capability is due
to the long decay time produced by the bias voltage. 56
The discussion until now has centered on the utilization of driving
signals whose frequency components are much lower than the cutoff
frequency. However, it is possible to implement useful multiplexing
schemes with drive signals whose frequency components are both
below and above the cutoff frequency. This approach will first be an-
alyzed for dynamic-scattering devices.
Imagine a cell containing a liquid capable of dynamic scattering.
Let us drive this cell with two sinusoidal voltage sources in series, one
with applied frequency 11 «Ie and the other with frequency 12 »fc.
The high-frequency signal retards the occurrence of dynamic scatter-
ing, and it can be shown that the low-frequency threshold voltage for
the formation of dynamic scattering is related to the high-frequency
signal by the following equation 114,115

ell]

where Vo is the threshold voltage derived by Helfrich62 for dc and


very low frequency ac signals, and 'Y is a parameter dependent upon
several material properties such as dielectric constant, shear torque,
viscosity, and conductivity. For the common liquid crystal MBBA, 'Y
= 0.5 at 32°C.115
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 267

+VL
12
'it.12- v.. r. ~
L --..J
VL - VH
12
VL -
~
"" \ VH

-VL/2 -~2- VH -~ Yt1


-VLAi VH
-~ VH

'"C!> ,-v..
~
+V~
2 I- ~ VL;
VH
V~ VH '\:-2 Yt1
0 L --..J
>
~
~
-V
L/2
-Il - VH
lt2
-V -Il
L/2 H -V~VH -V~ v..
11.
!!!:
~
Z
-YL
12
-V~2 - v.. -VL/2-
~ -Vlt. -ivH
2
-V~VH

'"::E
C!>
'"'"
-VL/2 -YL - VH -Il
L/2
-r".H -Vlt, -VH -VL -IvH
12 2 ~

-V~2 -V~2 -VH -VL/2-


~ -V~-ivH -V~
~

-VL
'2
DIGIT INPUT VOLTAGE

Fig. 17-Dual-frequency addressing of a matrix array.

The increase in threshold voltage for dynamic scattering induced


by the high frequency can be readily utilized in a matrix display as
shown in Fig. 17. VL/2 and V H are the zero-to-peak amplitudes of the
low- and high-frequency signals. With the simultaneous application
of the low- and high-frequency voltages to those elements that are
not supposed to be scattering, the low-frequency signal can be in-
creased in amplitude so that the light intensity from the "on" ele-
ments is greater than it would have been if a single low-frequency sig-
nal had been applied. With pure nematic materials, approximately 16
lines can be addressed with reasonable contrast.l 16 The addition of
cholesteric material to the nematic should lengthen the decay time
and thereby increase the upper limit on the number of lines that can
be matrix-addressed.
The two-frequency approach can also be used with field-effect ma-
terials. Bucher, Klingbiel, and Van Meter98 have shown that the low-
frequency threshold voltage for the twisted nematic effect is in-
creased by the superposition of a signal whose drive frequency is
greater than the critical frequency where the dielectric anisotropy be-
comes negative. They claim that

V
/.F
2 = V 02 + I!:J.fH
!:J.fLF
F IV
HF
2 [12J
268 CHAPTER 14

where V LF is the amplitude of the low-frequency signal, VHF is the


amplitude of the high-frequency signal, Vo is the threshold in the ab-
sence of the high-frequency signal, ~~HF is the dielectric anisotropy
at the high-frequency, and ~~LF is the same quantity at the low fre-
quency. The material they used has a crossover frequency of 2.5 kHz
at 25°C, whereas other materials published in the literature have
higher critical frequencies. 97 ,117
Because of the combined threshold-voltage and rise-time require-
ments, none of the approaches described here are capable of matrix
addressing a high-resolution, high-speed display. Lechner, Marlowe,
Nester, and Tults 96 have investigated the application of liquid-crystal
matrix displays to television and have concluded that a nonlinear
threshold or isolation device, such as a diode or transistor, must be
inserted in series with the liquid-crystal element at each matrix inter-
section to obtain the required speed and legibility for line-at-a-time
addressing.
A television-rate line-at-a-time display operates in a manner simi-
lar to the small 5 X 7 matrix described previously, with the main dif-
ferences being size and speed. The frame time is 30 msec and the line
time is 30 X 10- 3/500 sec, or 60 J.lsec. A small section of a much larger
matrix with diodes serving as the isolation devices is shown in Fig. 18.
Assuming that the liquid crystal is exhibiting dynamic scattering, the
response of a single diode-liquid-crystal combination is given in Fig.
19. Due to the inherent dielectric relaxation time, du, the addressing
voltage across the cell is stored for a time (1 to 10 msec) sufficient to
cause the excitation of the fluid deformation. The same basic descrip-
tion of the time response applies to all the electro-optic phenomena.
Of course, ~/ u varies from material to material.
The design of a color-television display panel that uses the twisted
nematic phenomena and polycrystalline thin-film transistors (TFT's)

Fig. 18-A small section of a IIquid-crystal-diode matrix display.


L1QUID·CRYSTAL D ISPLA YS-E LECTRO·OPTIC EFFECTS 269

as the isolation devices has been described by Fischer, Brody, and Es-
cottP8 In the TFT's, CdSe serves as the semiconductor and Ah03 as
the gate insulator. The display has been constructed and is 6 X 6
inches in size with 14,400 TFT's in a 120 X 120 array.1 19 Operation of
the entire panel has recently been demonstrated, but defects were
still present in the panel and the display was only black and white.1 2o
Good reliability is claimed for the display.

(0) EQUIVALENT CIRCUIT


FOR AN ELEMENT
IN THE MATRIX

:--- 33m sec.~


V(tl "
VJf60JLIOC fL
I

(b) ADDRESSING SIGNAL

~ (C) VOLTAGE ACROSS


CELL

""~ ~NTRAST
RESPONSE

Fig. 19-Tlme response of a single Intersection In the IIquld-crystal-diode array.

Lipton and Koda 121 have also recently presented results on a CdSe
TFT-liquid-crystal panel. They used relatively high-conductivity dy-
namic-scattering material; consequently, they added a capacitor in
parallel with each TFT and display element to obtain the required
electrical decay time. Brody1l8,120 and co-workers were able to utilize
the long dielectric relaxation time associated with the twisted nemat-
ic fluid and did not have to add the supplemental capacitance.
Liquid-crystal displays have also been constructed on a matrix of
single-crystal silicon MOS FET's.1 22 Pictorial-gray-scale images have
been created on a 1 inch display although line defects were present.
Though large arrays of TFT's should be much more economical than
silicon MOS FET's, TFT's have suffered in the past from stability
and reliability problems. More experimentation is necessary to prove
their capabilities as the threshold devices in a liquid-crystal panel.
270 CHAPTER 14

4.2 Beam Scanning

Images are produced on a cathode-ray tube by scanning a high-volt-


age electron beam across the surface of the cathodoluminescent ma-
terial. Each position on the phosphor is excited sequentially as the
beam is scanned by the deflection electron optics. This is an example
of element-at-a-time addressing. Beam scanning in an element-at-a-
time mode can be performed using either an electron beam or a light
beam. Both techniques have been implemented with liquid crystals.
Van Raalte 123 was the first to describe the results of an electron-
beam-scanned dynamic-scattering display. A schematic diagram of
his demountable cathode-ray tube and some typical images obtained
from the liquid-crystal display are presented in Fig. 20. The liquid

CONDUCTivE
COATING WIRE I GLASS MOSAIC

\""~.l Q UI O
CRYSTAL
ELECTRON
L AYER
GUN

REAL- TI ME DIS PLAY

Fig. 20-IIIustratlon of the electron-beam-scanned liquid-crystal display with some Im-


ages created In the display (Ref. [123]).

crystal was sandwiched between a tin-oxide-coated glass slide and a


mosaic feed-through plate constructed with fine wires inserted in
glass. A segmented mirror was evaporated on the wire pin mosaic
array that provided electrical contact between the electron beam and
the liquid crystal. With the electron beam scanning at video rates, a
maximum contrast ratio of 7.5 to 1 was obtained. Though very inter-
esting, the display suffered from two deficiencies; (1) the liquid crys-
LlQU ID-CRYSTAL 0 ISPLA YS-E LECTRO-DPTlC EFFECTS 271

tal was addressed by unipolar voltage pulses and (2) the pin mosaic
was extremely difficult to manufacture at the necessary resolution.
Gooch 124 and co-workers have presented a solution to the first prob-
lem. They have devised an electron-beam-scanned liquid-crystal dis-
play using bistable secondary-electron emission techniques combined
with a dielectric layer to drive the liquid crystal with an ac potential.
The economical construction of a hermetically sealed high-resolution
wire mosaic is still very difficult, although progress has been made. 125
The complexity of the feed-through mosaic is one of the motivating
factors in the development of the light-beam-addressing approaches.
In the most common variation of this second method, light illumi-
nates a thin-film photoconductor-liquid-crystal sandwich (see Fig.
21). Margerum, Nimoy, and Wong 126 presented the operation of a

TIN·
OXIDE ff~
GLASS ;1'

Fig. 21-Schematlc diagram of a Ilquld-crystal-photoconductor structure.

photoconductor-liquid-crystal device. Both dynamic-scattering and


storage effects were produced with light irradiating the photoconduc-
tor through the glass plate. Unilluminated, the resistance of the ZnS
layer was higher than that of the liquid crystal. With UV radiation of
sufficient intensity, the resistance of the photoconductor dropped
below that of the liquid crystal and the voltage activated the liquid
crystal. The ZnS was insensitive to the visible radiation that was used
to observe the image in the liquid crystal. A sensitivity of 0.1 mJ/cm 2
was obtained.
White and Feldman 127 improved the sensitivity of the display to
light by using evaporated selenium as the photoconductor. An
opaque, highly reflecting light barrier was placed between the photo-
conductor and the liquid crystal to prevent the excitation of the pho-
toconductor by the viewing light. Assouline, Hareng, and Leiba 128
used CdS as the photoconductor, and measured a sensitivity of 5 X
272 CHAPTER 14

10-6 J/cm 2. Similar results with CdS have been obtained by Jacob-
son et a1. 29 Haas, Adams, Dir, and Mitchell 130 obtained a sensitivity
of 2.5 X 10- 6 J /cm 2 using an unspecified photoconductor and storage-
effect liquid crystals.
All of the liquid-crystal-photoconductor structures discussed so
far were excited by a dc voltage source. Under dc operation, various
electrochemical life-degrading interactions have been observed at the
liquid-crystal-photoconductor interface. 130 Beard, Bleha, and.
Wong l31 have reported on a photoconductor-liquid-crystal valve for
projection applications that is driven by an ac voltage source (see Fig.
22). In addition, the display is operated in the reflection mode with a
multilayer combination of dielectric mirror and absorbing layer to
separate the CdS photoconductor from the projection light. With 200
lumens/cm 2 of white projection light irradiating the valve, no notice-
able interaction between the projection light and the photoconductor
has been observed. The light ·valve has been excited with the project-
ed image from a CRTl32 as well as a static slide image. At the present
time, the contrast ratio and speed are still below television standards.
The above structure with the induced birefringence mode has been
used as the light valve in a color projection displayJ33
The photoconductor-liquid-crystal sandwiches incorporate either
a dc driving source or a somewhat complicated multilayer structure.
Maydan, Melchior, and Kahn l34 have circumvented both of these as-
pects by utilizing the thermo-optic properties of nematic-cholesteric
mixtures reported by Soref. 135 He showed that nematic-cholesteric

GLASS 103'+--- -t- TRANSPARENT


CO NDUCTIVE ELECTRODE

PROJECTION
LIGHT
-
WR ITING LIGHT

--
AH TiREfLECTIO
COATI G

TRANSPAREHT PHOTOCONDUCTOR
COND UCTIVE
COU TEft ELECT RODE

Fig. 22-Slde view of an ac addressed IIquld-crystal-photoconductor light valve (Ref.


[131]).
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 273

GLASS
LEHS
!ISOTROPIC a
LASER SCATTERING
BEAll PHASES )

HOT SPOT

LlQUIO CRYSTAL
(CHOLESTERIC PHAS£ )--c:...----'i~~....J

ERASE
VOLTAGE

Fig. 23-A laser-beam-addressed thermo-optic liquid-crystal light valve (Ref. [134]).

mixtures can be converted from the clear planar texture to the highly
scattering state by heating the material from the mesophase into the
i~otropic phase and then letting it cool. In their laser-beam-scanned
display, Maydan and co-workers use the heat absorbed from the laser
beam by the In2-x Snx 03-y coatings to change the nematic-cholester-
ic storage material to the scattering state (see Fig. 23). The stored in-
formation is removed by exciting the liquid crystal with a high-fre-
quency erase signal.
An improvement of this device has been described in which the
nematic-cholesteric mixture is replaced by a smectic material. 136
Thermal writing induces the change of the smectic from the perpen-
dicular to a scattering texture. Unlike the nematic-cholesteric mate-
rials, selective erasure is possible with the smectic device. The ther-
mal writing is too slow for television-rate applications because of the
thermal inertia of the glass-liquid-crystal system. With a laser-beam
power of 20 m W, addressing speed is approximately 10 4 elements/sec
for the smectic device. In the projection mode, the resolution is 50
lines/mm at a contrast ratio of approximately 10:1.

5. Summary

The different electro-optic phenomena have been classified into


those that involve only dielectric forces and those that depend upon
the interaction of conduction and dielectric torques. The field-effect
phenomena possess several common properties. The resistivity of the
materials may be as high as chemically practical, i.e., p ~ 1011 ohm-
cm. For the induced birefringence, twisted nematic, and guest-host
color switching effects, the threshold voltages are less than 3 or 4
274 CHAPTER 14

volts, with an observed minimum of approximately one volt. The first


two properties imply a very low power dissipation of less than 1 /.l W /
cm 2• In ambient lighting, a diffuse reflector can be used with both the
guest-host and twisted nematic effects to obtain a high-contrast and
glare-free display.
Dynamic scattering and the storage effect are characterized by for-
ward scattering, turbulent fluid motion during the presence of a field,
and fluid resistivities that range between 108 and 1010 ohm-cm. Both
effects can only occur with the applied frequency less than the dielec-
tric relaxation frequency of the liquid crystal. The power dissipation
is between 0.1 and 1 mW/cm 2 at an operating voltage of 15 Vrms • Nei-
ther a polarizer nor an analyzer are required. The static scattering
can be stored in the nematic-cholesteric mixtures for anywhere from
seconds to months.
Electrical addressing of liquid-crystal displays is accomplished ei-
ther by X-Y matrix addressing or a scanning electron or light beam.
The electron-beam approach is analogous to a cathode ray tube with
the phosphor screen replaced by a liquid-crystal-pin wire mosaic
feedthrough. Similarly, an amplitude-modulated scanning light beam
can activate a photoconductor-liquid-crystal cell to provide the de-
sired spatial pattern in the liquid crystal.
Small and/or slow matrix displays have been fabricated using the
inherent threshold characteristics of liquid crystals. Both single-fre-
quency and a.ual-frequency schemes have been described. The latter
approach possesses greater multiplexing capability than the former,
but at the price of higher applied voltage and larger power dissipa-
tion. However, due to speed of response and contrast ratio limita-
tions, the construction of a 500 X 500 element liquid-crystal display
operating at video rates does not appear feasible without the addition
of a nonlinear device at each X-Y intersection. Thin-film transistor
arrays are being fabricated to serve as the nonlinear devices. Their
performance is being evaluated.

References

1 C. W. Oseen, "The Theory of Liquid Crystals," Trans. Faraday Soc., 2, p. 833 (1933).
2 F. C. Frank, "On the Theory of Liquid CrystalS," Disc. Faraday Soc., 25, p. 19 (1958).
3 P. Sheng, "Introduction to the Elastic Continuum Theory of Liquid Crystals, "Chapter 8.
4 P. deGennes, The Physics of Liquid Crystals, Oxford University Press, London (1974).
5 W. Helfrich, "Electric Alignment of Liquid Crystals," Mol. Cryst. and Liq. Cryst., 21, p. 187 (1973).
6 H. Gruler, T. J. Scheffer, and G. Meier, "Elastic Constants of Nematic Liquid Crystals I. Theory of
the Normal Deformation," Z. Naturforsch. A, 27a, p. 966 (1972).
7 M. F. Shiekel and K. Fahrenschon, "Deformation of Nematic Liquid Crystals with Vertical Orienta-
tion in Electrical Fields," Appl. Phys. Lett., 19, p. 393 (1971).
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 275

8 F. J. Kahn, "Electric-Field Induced Orientational Deformation of Nematic liquid Crystals: Tunable


Birefringence," Appl. Phys. Lett., 20, p. 199 (1972).
9 R. A. Soref and M. J. Rafuse, "Electrically Controlled Birefringence of Thin Nematic Films," J.
Appl. Phys., 43, p. 2029 (1972).
10 H. Mailer, K. L. Likins, T. R. Tayloer, and J. L. Fergason, "Effect of Ultrasound on a Nematic Liquid
Crystal," Appl. Phys. Lett., 18, p. 105 (1971).
11 H. Gruler and G. Meier, "Electric Field Induced Deformations in Oriented Liquid Crystals of the
Nematic Type," Mol. Cryst. and Liq. Cryst., 16, p. 299 (1972).
12 H. Deuling, "Deformation of Nematic Liquid Crystals in an Electric Field," Mol. Cryst. and Liq.
Cryst. 19, p. 123 (1972).
13 L. T. Creagh and A. R. Kmetz, "Mechanism of Surface Alignment in Nematic Liquid Crystals,"
Mol. Cryst. and Liq. Cryst., 24, p. 59 (1973).
14 J. Robert and G. Labrunie, "Transient Behavior of the Electrically Controlled Birefringence in a
Nematic Liquid Crystal," J. Appl. Phys., 44, p. 4689 (1973).
15 G. Heilmeier and J. Goldmacher, U.S. Patent No. 3,499,702 (1970).
16 M. Schadt and W. Helfrich, "Voltage Dependent Optical Activity of a Twisted Nematic Liquid Crys-
tal," Appl. Phys. Lett. 18, p. 127 (1971).
17 F. M. Leslie, "Distortion of Twisted Orientation Patterns in Liquid Crystals by Magnetic Fields,"
Mol. Cryst. and Liq. Cryst., 12, p. 57 (1970).
18 C. J. Alder and E. P. Raynes, "Room Temperature Nematic Liquid Crystal Mixtures with Positive
Dielectric Anisotropy," J. Phys., 06, p. L33 (1973).
19 A. Boller, H. Scherrer, and M. Schadt, "Low Electro-Optic Threshold in New liquid Crystals,"
Proc. IEEE, 60, p. 1002 (1972).
20 A. Ashford, J. Constant, J. Kirton, and E. P. Raynes, "Electro-Optic Performance of a New Room
Temperature Nematic Liquid Crystal," E/ec. Lett., 9, p. 118 (1973).
21 R. R. Reynolds, C. Maze and E. P. Oppenheim, "Design Considerations for Positive Dielectric
Nematic Mixtures Suitable for Display Applications," Abstracts of Fifth International liquid Crystal
Conf., Stockholm, Sweden, June 1974, p. 236.
22 C. H. Gooch and H. A. Tarry, "Optical Characteristics of Twisted Nematic Liquid Crystal Films,"
E/ec. Lett., 10, p. 2 (1974).
23 J. Robert and F. Gharadjedaghi, "Rotation du Plan de Polarisation de la Lumiere dans une Struc-
ture Nematique en Helice," C. R. Acad. Sc. Paris, 2788, p. 73 (1974).
24 C. J. Gerritsma, W. H. DeJeu and P. VanZanten, "Distortion of a Twisted Nematic Liquid Crystal
by a Magnetic Field," Phys. Lett., 36A, p. 389 (1971).
25 C. Z. VanDoorn, "On the Magnetic Threshold for the Alignment of a Twisted Nematic Crystal,"
Phys. Lett., 42A, p. 537 (1973).
26 A. I. Baise and M. M. Labes, "Effect of Dielectric Anisotropy on Twisted Nematics," Appl. Phys.
Lett., 24, p. 298 (1974).
27 D. Meyerhofer, "Electro-optic Properties of Twisted Field Effect Cells," Abstracts of Fifth Interna-
tional Liquid Crystal Conf., Stockholm, Sweden, June 1974, p. 220.
28 D. W. Berreman, "Optics in Smoothly Varying Anisotropic Planar Structures: Applicated to liquid
Crystal Twist Cells," J. Opt. Soc. Amer., 63, p. 1374 (1973).
29 S. Kobayashi and F. Takeuchi, "Mullicolor Field-Effect Display Devices with Twisted Nematic liq-
uid Crystals," Proc. of S.I.D., 14, p. 115 (1973).
30 C. Z. VanDoorn and J. L. A. M. Heldens, "Angular Dependent Optical Transmission of Twisted
Nematic Liquid Crystal Layers," Phys. Lett., 47A, p. 135 (1974).
31 F. Brochard, "Back flow Effects in Nematic Liquid Crystals," Mol. Cryst. and Liq. Cryst., 23, p. 51
(1973).
32 C. J. Gerritsma, J. A. Geurst and A. M. J. Spruijt, "Magnetic-Field-Induced Motion of Disclinations
in a Twisted Nematic Layer," Phys. Lett., 43A, p. 356 (1973).
33 E. P. Raynes, "Twisted Nematic Liquid Crystal Electro-Optic Devices with Areas of Reverse
Twist," E/ec. Lett., 9, p. 101 (1973).
34 E. P. Raynes, "Improved Contrast Uniformity in Twisted Nematic liquid Crystal Electro-Optic Dis-
play Devices," E/ec. Lett., 10, p. 141 (1974).
35 P. J. Wild, "Twisted Nematic Liquid Crystal Displays with Low Threshold Voltage," Comptes Ren-
dus des Joumees d'Electronique, EPFL, p. 102 (1973).
36 G. H. Heilmeier and L. A. Zanoni, "Guest-Host Interactions in Nematic Liquid Crystals-A New
Electro-Optic Effect," Appl. Phys. Lett., 13, p. 91 (1968).
37 G. H. Heilmeier, J. A. Castellano, and L. A. Zanoni, "Guest-Host Interactions in Nematic Liquid
Crystals," Mol. Cryst. and Liq. Cryst., 8, p. 293 (1969).
276 CHAPTER 14

38 J. J. Wysocki, J. Adams, and W. Haas, "Electric-Field-Induced Phase Change in Cholesteric Liquid


Crystals," Phys. Rev. Lett., 20, p. 1024 (1968).
39 E. Sackmann, S. Meiboom and L. C. Snyder, "On the Relation of Nematic to Cholesteric Meso-
phases," J. Am. Chern. Soc., 89, p. 5981 (1967).
40 P. G. deGennes, "Calcul de la Distortion D'une Structure Cholesterique Par un Champ Magneti-
que," Sol. St. Commun., 6, p. 163 (1968).
41 R. B. Meyer, "Effects of Electric and Magnetic Fields on the Structures of Cholesteric Liquid Crys-
tals," Appl. Phys. Lett., 12, p. 281 (1968).
42 F. Rondelez and J. P. Hulin, "Distortions of a Planar Cholesteric Structure Induced by a Magnetic
Field," Sol. St. Commun., 10, p. 1009 (1972).
43 F. J. Kahn, "Electric-Field-Induced Color Changes and Pitch Dilation in Cholesteric Liquid Crys-
tals," Phys. Rev. Lett., 24, p. 209 (1969).
44 R. B. Meyer, "Distortion of a Cholesteric Structure by a Magnetic Field," Appl. Phys. Lett., 14, p.
208 (1969).
45 G. Durand, L. Leger, F. Rondelez and M. Veyssie, "Magnetically Induced Cholesteric-to-Nematic
Phase Transition in Liquid Crystals," Phys. Rev. Lett., 22, p. 227 (1969).
46 G. H. Heilmeier and J. E. Goldmacher, "Electric-Field-Induced Cholesteric-Nematic Phase Change
in Liquid Crystals," J. Chern. Phys., 51, p. 1258 (1969).
47 G. W. Gray, K. J. Harrison and J. A. Nash, "New Family of Nematic Liquid Crystals for Displays,"
E/ec. Lett., 9, p. 130 (1973).
48 J. P. Hulin, "Parametric Study of the Optical Storage Effect in Mixed Liquid Crystal Systems,"
Appl. Phys. Lett., 21, p. 455 (1972).
49 W. Helfrich, "Deformation of Cholesteric Liquid Crystals with Low Threshold Voltage," Appl. Phys.
Lett., 17, p. 531 (1970).
50 W. Helfrich, "Electrohydrodynamic and Dielectric Instabilities of Cholesteric Liquid Crystals," J.
Chem. Phys. 55, p. 839 (1971).
51 J. Hurault, "Static Distortions of a Cholesteric Planar Structure Induced by Magnetic or A.C. Elec-
tric Fields," Fourth International Liquid Crystal Conf., Kent, Ohio, Aug. 1972.
52 C. J. Gerritsma and P. VanZanten, "Periodic Perturbations in the Cholesteric Plane Texture,"
Phys. Lett., 37A, p. 47 (1971).
53 T. J. Scheffer, "Electric and Magnetic Field Investigations of the Periodic Gridlike Deformation of a
Cholesteric Liquid Crystal," Phys. Rev. Lett. 28, p. 598 (1972).
54 F. Rondelez, H. Arnould and C. J. Gerritsma, "Electrohydrodynamic Effects in Cholesteric Liquid
Crystals Under AC Electric Fields," Phys. Rev. Lett., 28, p. 735 (1972).
55 J. J. Wysocki et al., "Cholesteric-Nematic Phase Transition Displays," Proc. SID, 13, p. 115
(1972).
56 T. Ohtsuka and M. Tsukamoto, "AC Electric-Field-Induced Cholesteric-Nematic Phase Transition
in Mixed Liquid Crystal Films," Jap. J. Appl. Phys., 12, p. 22 (1973).
57 w. F. Greubel, "Bistability Behavior of Texture in Cholesteric Liquid Crystals in an Electric
Field," Appf. Phys. Lett .. 25, p. 5 (1974).
58 D. L. White and G. N. Taylor, "A New Absorptive Mode Reflective Liquid Crystal Display Device,"
J. Appf. Phys .• 45, p. 4718 (1974).
59 G. H. Heilmeier, L. A. Zanoni and L. A. Barton, "Dynamic Scattering in Nematic Liquid Crystals,"
Appl. Phys. Lett., 13, p. 46 (1968).
60 G. H. Heilmeier, L. A. Zanoni and L. A. Barton, "Dynamic Scattering: A New Electro-Optic Effect
in Certain Classes of Nematic Liquid Crystals," Proc. IEEE, 56, p. 1162 (1968).
61 C. Deutsch and P. N. Keating, "Scattering of Coherent Light from Nematic Liquid Crystals in the
Dynamic Scattering Mode," J. Appl. Phys., 40, p. 4049 (1969).
62 W. Helfrich, "Conduction-Induced Alignment of Nematic Liquid Crystals: Basic Model and Stability
Considerations," J. Chern. Phys., 51, p. 4092 (1969).
63 E. F. Carr, "Ordering in Liquid Crystals Owing to Electric and Magnetic Fields," Advan. Chern.
Ser., 63, p. 76 (1967).
64 R. Williams, "Domains in Liquid Crystals," J. Chern. Phys., 39, p. 384 (1963).
65 P. A. Penz, "Voltage-Induced Vorticity and Optical Focusing in Liquid Crystals," Phys. Rev. Lett.,
24, p. 1405 (1970).
66 T. O. Carroll, "Liquid Crystal Diffraction Grating," J. Appl. Phys., 43, p. 767 (1972).
67 G. Assouline, A. Dmitrieff, M. Hareng, an.d E. Leiba, "Diffraction d'un Faisceau Laser par un Cris-
tal Liquide Nematique Soumis it un Champ Electrique," C. R. Acad. Sci. Paris, 271B, p. 857 (1970).
68 E. Dubois-Vilette, P. G. deGennes, and O. Parodi, "Hydrodynamic Instabilities of Nematic Liquid
Crystals Under AC Electric Fields," J. Physique, 32, p. 305 (1971).
LlQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 277

69 Orsay Liquid Crystal Group, "Hydrodynamic Instabilities in Nematic Liquids Under AC Electric
Fields," Phys. Rev. Lett., 25, p. 1642 (1970).
70 P. G. deGennes, "Electrohydrodynamic Effects in Nematics," Comments Sol. St. Phys., 3, p. 148
(1971).
71 R. A. Kashnow and H. S. Cole, "Electrohydrodynamic Instabilities in a High-Purity Nematic Liquid
Crystal," J. Appl. Phys., 42, p. 2134 (1971).
72 P. A. Penz and G. W. Ford, "Electrohydrodynamic Solutions for Nematic Liquid Crystals," Appl.
Phys. Lett., 20, p. 415 (1972).
73 P. A. Penz and G. W. Ford, "Electromagnetic Hydrodynamics of Liquid Crystals," Phys. Rev., 6A,
p. 414 (1972).
74 D. Meyerhofer, "Electro Hydrodynamic Instabilities in Nematic Liquid Crystals," Chapter 9.
75 T. O. Carroll, "Dependence of Conduction-Induced Alignment of Nematic Liquid Crystals Upon
Voltage Above Threshold," J. Appl. Phys., 43, p. 1342 (1972).
76 L. Goodman, "Light Scattering in Electric-Field Driven Nematic Liquid Crystals," Proc. SID, 13, p.
121 (1972).
77 L. Cosentino, "On the Transient Scattering of Light by Pulsed Liquid Crystal Cells:' IEEE Trans.
Electron Devices, ED-1, p. 1192 (1971).
78 L. Creagh, A. Kmetz and R. Reynolds, "Performance Characteristics of Nematic Liquid Crystal
Display Devices," IEEE Trans. Electron Devices, ED-18, p. 672 (1971).
79 W. F. Greubel and U. W. Wolff, "Electrically Controllable Domains in Nematic Liquid Crys-
tals," Appl. Phys. Lett., 19, p. 213 (1971).
80 W. H. DeJeu, C. J. Gerritsma, and A. M. VanBoxtel, "Electrohydrodynamic Instabilities in Nematic
Liquid CrystalS," Phys. Lett., 34A, p. 203 (1971).
81 L. K. Vistin, "Electrostructural Effect and Optical Properties of a Certain Class of Liquid Crystals
and Their Binary Mixtures," Sov. Phys. erystallogr., 15, p. 514 (1970).
82 N. Felici, "Phenomenes Hydro et Aerodynamiques dans la Conduction des Dielectrique Fluide,"
Rev. Gen. Elec., 78, p. 717 (1969).
83 Orsay Liquid Crystal Group, "AC and DC Regimes of the Electrohydrodynamic Instabilities in
Nematic Liquid Crystals," Mol. Cryst. andUq. eryst., 12, p. 251 (1971).
84H. Koelmans and A. M. VanBoxtel, "Electrohydrodynamic Flow in Nematic Liquid Crystals," Mol.
eryst. and Uq. Cryst., 12, p. 185 (1971).
85 D. Meyerhofer and A. Sussman, "The Electrohydrodynamic Threshold in Nematic Liquid Crystals
in Low Frequency Fields," Appl. Phys. Lett., 20, p. 337 (1972).
86 A. Sussman, "Electro-Optic Liquid Crystal Devices: Principles and Applications," IEEE Trans.
Parts, Hybrids and Packaging, PHP-8, p. 28 (1972).
87 G. Durand, M. Veyssie, F. Rondelez and L. Leger, "Effect Electrohydrodynamique dans un Crlstal
Liquide Nematique," C. R. Acad. Sc. Paris, 2708, p. 97 (1970).
88 G. H. Heilmeier and J. E. Goldmacher, "A New Electric Field Controlled Reflective Optical Storage
Effect in Mixed Liquid Crystal Systems," Proc. IEEE 57, p. 34 (1969).
89 G. Dir et al., "Cholesteric Liquid Crystal Texture Change Displays," Proc. SID, 13, p. 105 (1972).
90 D. Meyerhofer and E. F. Pasierb, "Light Scattering Characteristics in Liquid Crystal Storage Mate-
rials," Mol. Cryst. and Uq. Cryst., 20, p. 279 (1973).
91 E. Jakeman and E. P. Raynes, "Electro-Optic Response Times in Liquid Crystals," Phys. Lett.,
39A, p. 69 (1972).
92 J. Robert, G. Labrunie and J. Borel, "Static and Transient Electric Field Effect on Homeotropic
Thin Layers," Mol. eryst. and Uq. eryst., 23, p. 197 (1973).
93 A. Sussman, "Secondary Hydrodynamic Structure in Dynamic Scattering," Appl. Phys. Lett., 21,
p. 269 (1972).
94 C. J. Gerritsma, C. Z. VanDoorn and P. VanZanten, "Transient Effects in the Electrically Con-
trolled Light Transmission of a Twisted Nematic Layer," Phys. Lett., 48A, p. 263 (1974).
95 C. H. Gooch and H. A. Tarry, "Dynamic Scattering in the Homeotropic and Homogeneous
Textures of a Nematic Liquid Crystal," J. of Phys. D. Appl. Phys., 5, p. L25 (1972).
96 B. J. Lechner, F. Marlowe, E. Nester, and J. Tults, "Liquid Crystal Displays," Proc. IEEE, 59, p.
1566 (1971).
97 W. H. DeJeu, C. J. Gerristma, P. VanZanten, and W. J. A. Gossens, "Relaxation of the Dielectric
Constant Electrohydrodynamic Instabilities in tl Liquid Crystal," Phys. Lett., 39A, p. 355 (1972).
98 H. K. Bucher, R. T. Klingbiel, and J. P. VanMeter, "Frequency-Addressed Liquid Crystal Field Ef-
fect," Appl. Phys. Lett., 25, D. 186 (1974).
99 E. P. Raynes and I. A. Shanks, "Fast Switching Twisted Nematic Electro-Optical Shutter and
Color-Filter," Elec. Lett., 10, p. 114 (1974).
278 CHAPTER 14

100 T. S. Chang and E. E. Loebner, "Crossover Frequencies and Turn-Off Time Reduction Scheme
for Twisted Nematic Liquid Crystal Displays," Appl. Phys. Lett., 25, p. 1 (1974).
101 G. Baur, A. Stieb, and G. Meier, "Controlled Decay of Electrically Induced Deformations in Nema-
tic Liquid Crystals," Appl. Phys., 2, p. 349 (1973).
102 B. Kellenevich and A. Coche, "Relaxation of Light Scattering in Nematic-Cholesteric Mixtures,"
Mol. Cryst. and Liq. Cryst., 24, p. 113 (1973).
103 A. Sussman, "Dynamic Scattering Life in the Nematic Compound p-Methoxybenzylidene-p-
Amino Phenyl Acetate as Influenced by Current Density," Appl. Phys. Lett., 21, p. 126 (1972).
104 L. Pohl, R. Steinstrasser, and B. Hampel, "Performance of Nematic Phase V and VA in Liquid
Crystal Displays," Fourth Internat. Liq. Cryst. Conf., Kent, Ohio; Aug. 1972, Paper No. 144.
105 I. Haller, "Elastic Constants of the Nematic Liquid Crystalline Phase of p-Methoxybenzylidene-p-
n-Butylaniline (MBBA)," J. Chern. Phys., 57, p. 1400 (1972).
106 A. Sussman, "Electrochemistry in Nematic Liquid-Crystal Solvents," Chapter 17.
107 A. R. Kmetz, "Liquid Crystal Displays Prospects in Perspective," IEEE Trans. Elec. Dev., EO-20,
p. 954 (1973).
108 M. Hareng, G. Assouline and E. Leiba, "La Birefringence Electriquement Controlee dans les Cris-
taux Liquides Nematiques," Appl. Opt., 11 p. 2920 (1972).
109 P. M. All and P. Pleshko, "Scanning Limitations of Liquid Crystal Displays," IEEE Tran:;. Elec.
Dev., EO-21, p. 146 (1974).
110 H. Takata, O. Kogure, and K. Murase, "Matrix-Addressed Liquid Crystal Display," IEEE Trans.
Elec. Dev., EO-20, p. 990 (1973).
111 M. Hareng, G. Assouline, and E. Leiba, "Liquid Crystal Matrix Display by Electrically Controlled Bi-
refringence," Proc. IEEE, 60, p. 913 (1972).
112 M. F. Schiekel and K. Fahrenschon, "Multicolor Matrix Displays Based on Deformation of Verti-
cally Aligned Nematic Liquid Crystal Phases," Digest 1972 Soc. for Information Display International
Symp., San Francisco, Calif., p. 98.
113 T. Ohtsuka, M. Tsukamoto, and M. Tsuchiya, "Liquid Crystal Matrix Display," Jap. J. Appl. Phys.,
12, p. 371 (1973).
114 C. R. Stein and R. A. Kashnow, "A Two Frequency Coincidence Addressing Scheme for Nematic
Liquid Crystal Displays," Appl. Phys. Lett., 19, p. 343 (1971).
115 P. J. Wild and J. Nehring, "An Improved Matrix Addressed Liquid Crystal Display," Appl. Phys.
Lett., 19, p. 335 (1971).
116 C. R. Stein and R. A. Kashnow, "Recent Advances in Frequency Coincidence Matrix Addressing
of Liquid Crystal Displays," Digest 1972 Soc. for Information Display International Symp., San Fran-
cisco, Calif. p. 64.
117 M. Schadt, "Dielectric Properties of Some Nematic Liquid Crystals with Strong Positive Dielectric
Anisotropy," J. Chern. Phys., 56, p. 1494 (1972).
118 A. G. Fischer, T. P. Brody, and W. S. Escott, "Design of a Liquid Crystal Color TV Panel," IEEE
Conf. Record 1972 Conf. on Display Devices, New York, NY, p. 64.
119 T. P. Brody, J. Asars and G. D. Dixon, "A 6 X 6 Inch 20 Lines per Inch Liquid Crystal Display
Panel," IEEE Trans. Elec. Dev., EO-20, p. 995 (1973).
120 T. F. Brody, F. C. Luo, D. H. Va"ies, and E. W. Greeneich, "Operational Characteristics of a 6 X
6 Inch, TFT Matrix Array, Liquid Crystal Display," Digest 1974 Soc. for Information Display Interna-
tional Symp., San Diego, Calif., p. 166.
121 L. Lipton and N. Koda, "Liquid Crystal Matrix Display for Video Applications," Proc. SID, 14, p.
127 (1973).
122 M. Ernstoff, A. M. Leupp, M. J. Little and H. T. Peterson, "Liquid Crystal Pictorial Display," Tech-
nical Digest 1973 International Electron Devices Meeting, Washington, D.C., p. 548.
123 J. A. van Raalte, "Reflective Liquid Crystal Television Display," Proc. IEEE, 56, p. 2146 (1968).

124 C. H. Gooch et aI., "A Storage Cathode-Ray Tube with Liquid Crystal Display," J. Phys., 06, p.
1664 (1974).
125 C. Burrowes, "Electrical Fiber Plates-A New Tool For Storage and Display," IEEE Can'. Record
1970 IEEE Can'. on Display Devices, New York, NY, p. 126.
126 J. D. Margerum, J. Nimoy and S.-Y. Wong, "Reversible Ultraviolet Imaging with Liquid Crystals,"
Appl. Phys. Lett., 17, p. 51 (1970).
127 D. L. White and M. Feldman, "Liquid Crystal Light Valves," Elec. Lett., 6, p. 837 (1970).
128 G. Assouline, M~ Hareng, and E. Leiba, "Liquid Crystal and Photoconductor Image Converter,"
Proc. IEEE, 59, p. 1355 (1971).
LIQUID-CRYSTAL DISPLAYS-ELECTRO-OPTIC EFFECTS 279

12 9A. Jacobson et al., "Photoactivated Liquid Crystal Light Valve," Digest 1972 SID International
Symp., San Francisco, Calif., p. 70.
130 W. Haas, J. Adams, G. Dir and C. Mitchell, "Liquid Crystal Memory Panels," Proc. SID, 14, p.
121 (1973).
131 T. D. Beard, W. P. Bleha and S.-Y. Wong, "Alternating Current Liquid Crystal Light Valve," Appl.
Phys. Lett., 22, p. 90 (1973).
132 W. P. Bleha, J. Grinberg, and A. D. Jacobson, "AC Driven Photoactivated Liquid Crystal Light
Valve," Digest 1973 SID International Symp., New York, NY, p. 42.
133 J. Grinberg et al., "Photoactivated Liquid Crystal Light Valve for Color Symbology Display," Conf.
Record 1974 IEEE-SID Conf. on Display Devices and Systems, New York, p. 47.
134 D. Maydan, H. Melchior and F. Kahn, "Thermally Addressed Electrically Erased High-Resolution
Liquid Crystal," Appl. Phys. Lett., 21, p. 392 (1972).
135 R. A. Soref, "Thermo-Optic Effects in Nematic-Cholesteric Mixtures," J. Appl. Phys., 41, p.
3022 (1970).
136 F. J. Kahn, "IR-Laser-Addressed Thermo-Optic Smectic Liquid Crystal Storage Displays," Appl.
Phys. Lett., 22, p. 111 (1973).
Liquid-Crystal Optical Waveguides

D. J. Channin

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

In recent years a new optical technology has developed in the field of


light guiding in thin films and fibers. Liquid-crystal materials have
yet to contribute to this technology to the extent that they have to
display devices and optical beam processing; nevertheless, examples
of liquid-crystal waveguide modulators,1,2 switches,3 and deflectors 4
have been demonstrated. Furthermore, optical waveguiding holds po-
tential as a tool for investigating physical and chemical processes in
thin films, 5 including liquid crystals.
This chapter reviews the basic theory of optical wave guiding in pla-
nar structures, with emphasis on the general concepts of mode spec-
tra, energy distribution within the guide, phase matching, and scat-
tering. Experimental techniques appropriate to liquid crystals are
discussed, and some experimental results on attenuation measure-
ments and electro-optic effects are described. More detailed discus-
sions of optical waveguide theory and technology (often called inte-
grated optics) are available in recent reviews and books. 6- 8
281
282 CHAPTER 15

2. Guided Optical Waves

The simplest optical waveguide structure is an infinite planar slab of


perfectly transparent, isotropic, dielectric material with refractive
index ng bounded on both surfaces by similar material of index no <
ng (see Fig. 1). According to Snell's law, light rays in the slab will be
totally internally reflected if they have an incidence angle (J with a
surface satisfying

sinO> no. [1]


ng

Since both surfaces of the slab are parallel, the light will undergo an
identical reflection at the opposite boundary and be trapped within

RADIATION GUIDED
(LEAKY GUIDE) LIGHT
~------------~~~~~4-----N

Fig. 1-Total Internal reflection of light in slab waveguide. N = c/v is the effective re-
fractive Index for phase velOCity v.

the slab. The phase velocity v of the light as it travels within the slab
is given by
c
v = ng sinO'
[2]

where c is the speed of light in free space. From Eqs. [1] and [2] it is
apparent that the guided light may have phase velocities within the
range

[3J
LIQUID-CRYSTAL OPTICAL WAVEGUIDES 283

Propagation under such conditions is called wave guiding.


If () is too small to satisfy Eq. [1], some light is transmitted out of
the slab at each encounter with the surface. Nevertheless, sufficient
reflection may occur that the light in the slab can be characterized as
a "leaky" wave, subject to loss by radiation away from the slab. It is
convenient to deal with the effective refractive index N = c/v to
characterize both leaky and truly guided light.
It has been tacitly assumed that the thickness of the slab is much
greater than an optical wavelength, so that geometric optics is appli-
cable. Should the thickness become comparable to the wavelength,
destructive interference between the multiple reflections prevents
guided propagation except at specific incidence angles for which the
interference is constructive. Waveguiding then occurs only for a dis-
crete set of guided modes, with effective indices Nj determined by
the guide thickness d as well as the two refractive indices ng and no.
This situation, though retaining many features of the geometric op-
tics limit, is properly described by wave optics.
Suppose now that the slab and surroundings are replaced by a me-
dium that is translation ally invariant in the y-z plane, but has refrac-
tive index n(x) that varies with position in the x direction. The con-
ditions

1 ~ no ~ n(x) ~ ng [4a]
n(±ro) = no [4b]

are imposed on n(x).


Consider light propagating parallel to the z axis, translationally
invariant along the y axis, and having an electric field f y . Such light
satisfies the wave equation

o. [5]

Separating the variables by coordinates yields

f, = f(X) exp!iK zl [6J

[7]

In these equations Ko = 27r/Ao, where Ao is the free-space optical


wavelength. It is convenient to make again the identification of KjK o
= c/v = N, the effective refractive index of the guided modes.
284 CHAPTER 15

Equations formally identical to Eq. [7] arise in many physical


problems, such as the Schrodinger equation for a particle in a poten-
tial well. N is the eigenvalue for this equation and, in accordance
with Eqs. [4a] and [4b], will take on a discrete and finite spectrum No,
N 1, . . . Nj, ... N max within the range no < Nj < n g , and a continuous
spectrum in the range 1 < N < no. The former characterizes the
guided modes and the latter the so-called radiation modes. The two
kinds of modes together are associated with a complete orthogonal
set of eigenfunctions of Eq. [7]. The leaky modes are solutions of Eq.
[7] but are not part of the complete set of eigenfunctions. They have a
discrete spectrum within the range of N spanned by the continuous
radiation mode spectrum. The ranges spanned by the different modes
are shown in Fig. 2.

GUIDED
MODES
RADIATION
(LEAKY MODES)
~------------+-~~~~~----N

Fig. 2-Ranges of effective Indices for guided and radiation modes.

We return now to the particular case of the slab waveguide defined


by

{
ng,lx l < d/2
n(x) = no, Ixl > d/2 [8J

The guided modes have effective indices N j determined by the solu-


tions of the dispersion relation

ni - N/ .
N }2 _ 2') = 1,3, ...
tan(Kifi
2
vn 2 -
g
N
}
2) = {-
IN ~ _;°2 ,
2 -
no
2
[9J
j = 0,2, ...
ng }

The number of modes increases with the guide thickness d. The low-
est-order mode No is closest in effective index to n g , while the high-
est order mode has effective index N max closest to no.
The spatial distributions of the electric fields for these modes are
given by
LlQUID-CRYSTAL OPTICAL WAVEGUIDES 285

~
sin(xKoVn/ - N i),
j = 1,3, ... lxl < d/2
exp(iKoN,z)
cos(xKoVn/ - N I),
j = 0,2, ...
[10]

~
exp( -xKoV N, 2 no~),
X > d/2
exp(iKoN,z) 1/~::-:::------'::
exp(+xKov N/ - n o2 )
X < d/2

The exponentially decaying field outside the slab is called the eva-
nescent field. Since the electric field Ey lies in the waveguide plane
the modes are called transverse electric or TE. A corresponding set of
transverse magnetic or TM modes also exist. These satisfy a wave
equation for Hy similar to Eq. [7]. The dispersion relationship for
TM modes is
(n)2 n 2 - N 2
Ng 2- '2,j=1,3, ...
K d {- ---.Q

IN - n
nil, no
tan(-o-vn 2 - N 2) = .
2 g , (nn:
)2: 2- 2
ng~ N~2' j = 0;2, ...
[11]

Eq. [11] differs slightly from Eq. [9], indicating differing phase veloci-
ties for TE and TM waves despite the assumption of isotropic mate-
rials. The velocity differences come from the differing phase shifts on
reflection at the interfaces for light polarized parallel to (TE) or per-
pendicular to (TM) the reflecting surface. Should the waveguide be
composed of anisotropic material there will be an additional velocity
difference due to the variation of guide index ng with optical polar-
ization.
In practice, many optical waveguides are asymmetric, comprising
thin films with substrate on one side and air on the other. Others
have graded continuous ind'ex distributions produced by diffusion of
atoms into or out of a bulk material. For such waveguides the range of
effectiye indices spanned by the guided modes is
[12]
where (n,,) max is the greatest of the indices surrounding the guide.
The spacing of the modes and the light distribution within the guide
is determined by the specific refractive index profile, though some
features are common to all guides.
In general, the optical energy distribution is concentrated in the
286 CHAPTER 15

highest-index material for low-order modes, and spreads out into the
lower-index material as the mode order increases. Thus, the optical
field of a multimode waveguide will sample different regions of the
guide as different modes are excited. The techniques of optical wave-
guiding are therefore potentially useful in studying the structure of
thin films and layers.

3. Phase Matching and Coupling

The recent surge of activity in planar optical waveguides was initiat-


ed by the development of practical and efficient ways to couple light
between guided modes and beams in free space. 9 •lO These techniques
are based on the introduction of an additional element to the wave-

RADIATION
MODE
SCATTERING
(GRATING)

(a)

RADIATION
MODE\

(b)

Fig. 3-Wave vector diagrams for (a) grating coupler and (b) prism coupler.

guide to match the phase velocities of the guided modes to those of


the radiation modes.
Fig. 3 represents the extension of Figs. 1 and 2 to include the plane
perpendicular to the waveguide. The axes now represent wave vec-
LIQUID-CRYSTAL OPTICAL WAVEGUIDES 287

tors, the magnitudes of which are related to the effective indices by


multiplication by Ko. Fig. 3a shows the basis for a grating coupler. A
small periodic modulation of the guide thickness or refractive index
phase matches the guided mode with wave vector KaNj to radiation
in the form of a ray at angle <p to the guide surface. The periodic mod-
ulation has wavelength As such that

Ks = r211"
s
= Ko(N} - no cayp). [13J

The basis for a prism coupler is shown in Fig. 3b. A small part of
the waveguide is bounded on one side by material of index np > n g .
The condition
[14J

establishes coupling between the guided mode N j and a ray with inci-
dence angle <Pp inside the high-index material. In practice the high-
index material takes the form of a prism that refracts the internal ray
into free space at a convenient angle.
Both kinds of couplers allow selective excitation of particular
waveguide modes. TE or TM modes are determined by setting the
polarization of the incident beam parallel to or perpendicular to the
waveguide plane. The particular effective index N j is determined by
varying the coupling angle after having chosen a prism index np or
periodic modulation spacing As. In output coupling, a waveguide
with one or more modes excited will cause beams of light to be cou-
pled out at angles corresponding to the effective indices of the excited
modes.

4. Scattering

Inhomogeneities in the waveguide material will scatter the guided


light and thereby attenuate the waveguide modes. Such scattering
may put light jnto other guided modes or into radiation modes. For
waveguides with sharp boundaries such as the uniform slab, a distinc-
tion is made between scattering centers located at the material inter-
faces (surface scattering) and scattering centers distributed through-
out the waveguide material (bulk scattering). In both cases the atten-
uation is a function of Nj because the different spatial distributions
of light in the guide have differing overlap of the regions containing
scattering centers. 8 •11
For TE modes in a symmetric slab guide with lossless surround-
ings, the bulk loss ab is given by
288 CHAPTER 15

ngYN/ n02 YN/ - n0 2 + Kod(nl - n02)


ab = abo-"--....:.N-;-;-",----'- (n/ - no2 )(1 + KodYN/ - no2 ) ,
[15]

where abo is the bulk scattering coefficient of the waveguide materi-


al. If'the scattering itself is wavelength dependent, abo will also vary
with Nj. For example, if the attenuation is due to Rayleigh scatter-
ing, abo is proportional to (Nj) -4. Eq. [15] shows that when abo is
constant, ab is maximum at the lowest order modes (j = 0), and de-
creases rapidly at the highest modes, where most of the light is in the
evanescent field and out of the scattering material.
The surface loss as for TE modes of a symmetric slab waveguide is
given by

[16]

If the surface-scattering coefficient a so is constant, as is seen to ap-


proach zero for the lowest-order mode as Nj= 0 approaches n g . The
increase in loss for higher mode numbers comes from the increasing
optical field at the scattering interface. For the highest-order modes
the scattering goes again to zero, since the light is distributed over a
large volume in the evanescent field.
The dependence of bulk and surface scattering on N in multimode
waveguides is sketched in Fig. 4. The total scattering is the sum of
the two contributions, and this of course is what is determined when
the attenuation of the various modes is measured. In many cases it is
possible to resolve the total scattering into the bulk and surface com-
ponents. This is based on the observation that (1) the surface scatter-
ing is zero and the factor multiplying abo is unity for N = ng and (2)
the bulk scattering is nearly constant except near N = no, so that the
rate of increase in scattering for N < ng is proportional to the surface
scattering. The second condition fails if either abo or a so depends on
N. A more detailed discussion of waveguide scattering has been pub-
lished elsewhere. 12

5. Liquid-Crystal Waveguides

Optical waveguides have been made of nematic liquid-crystal layers.


There is no apparent reason why light could not be guided in other
mesophases as well, but to date this has not been reported, and the
subsequent discussion is restricted to nematics. Since the liquid-crys-
LIQUID-CRYSTAL OPTICAL WAVEGUIDES 289

tal layer must be confined on both surfaces and since the bounding
material can easily be identical on both sides, the symmetric slab
model is appropriate for such waveguides_ The high birefringence of
nematics makes it necessary that the materials be well aligned and
free from domains of differing orientation. The resulting uniaxial
layer is of course different from the isotropic materials of the previ-

BULK SCATTERING

~(
'---~-----'---N
no

SURFACE SCATTERING

VI
VI
o
...J

'---~n-o----~nQ--- N

Fig. 4-Dependence of scattering loss on attenuation for multlmode slab waveguides.

ous discussion. However, if propagation and polarization directions


are restricted to the principle axes, the theory of the isotropic case is
applicable if the appropriate value of ng is chosen for the particular
mode polarization (TE or TM) involved.
The technique used for coupling light into the liquid-crystal wave-
guide must be compatible with the need to confine the layer on both
surfaces. We have found it possible to combine the optical coupling
prism and one bounding surface into a unitized structure. Alterna-
tively, light may be coupled into a solid-film waveguide that termi-
nates at the liquid crystal, but this technique makes it difficult to se-
lectively excite specific modes in the liquid-crystal layer.
Fig. 5 shows some coupling arrangements we have used. In Fig. 5a
the prism with index np = 1.95 is bonded to low-index glass or fused
quartz of index no = 1.48-1.51. The base of the structure is optically
polished. The laser beam is focused just to the left of the prism cor-
290 CHAPTER 15

nero Leaky modes excited in the liquid-crystal film under the high-
index prism become true guided modes after passing into the film
bounded by low-index material. A second prism bonded on the other
end of the low-index glass may be used to couple the guided light out
again.

LASER

\\\
,\,\
\ -----.------~

(a)

LASER DETECTOR

\\
\ \ /~
\ \ ------~// / /
\

(b)

Fig. 5-Prlsm coupler for exciting (a) guided or (b) leaky modes in liquid-crystal (LC)
layers.

A leaky wave coupler is shown in Fig. 5b. Here light coupled into
guided modes of the liquid-crystal layer is directly coupled out again.
This structure is useful for measuring the mode structure of the liq-
uid-crystal film by measuring the discrete angles for which light rays
are coupled into and out of the particular waveguide modes.
The attenuation by scattering of light guided in the arrangement of
Fig. 5a is determined by observing the intensity of light scattered
from the guided modes into radiation. The streak of scattered light is
imaged on a slit mounted in front of a photomultiplier, and this as-
sembly is translated to record the decrement of scattered light inten-
sity as a function of distance from the input coupling point.
Fig. 6 shows attenuation data obtained at room temperature for
LlQUID·CRYSTAL OPTICAL WAVEGUIDES 291

TE and TM modes in guides of 6- to 12-Jlm thickness. The liquid-


crystal layers were MBBA aligned with lecithin to have optic axis
perpendicular to the bounding surface. For TM modes, ng = next =
1.75, while for TB modes, ng = nord = 1.54. Each data point repre-
sents a separate guided mode.

100r-~~-'-----'----'-----r----.-----,

E
~
m
~
z
o
!;i 50
::>
zw
I-
~
• 12fLm LAYER
o 6fLm LAYER
X = 63281.

o 1.50 1.60 1.70 1.80


EFFECTIVE INDEX. N

Fig. 6-Attenuatlon measured 'or nematic waveguides consisting 0' MBBA layers with
hemeotroplc alignment.

That the measured attenuation is due to scattering rather than ab-


sorption was established from the absence of significant loss when the
liquid crystal was heated above the nematic-isotropic phase transi-
tion. The waveguide layers were well aligned, homogeneous, and free
from disclinations. The scatteririg results from thermal fluctuations
in molecular order l3 over volumes comparable to optical wavelengths.
As described previously, the scattering of the lowest order modes is
equal to the bulk scattering coefficient of the waveguide materiaL
This was in the range 35-45 dB/cm for TE modes and 45-60 dB/cm
for TM modes. According to Eq. [15], the bulk scattering strength ab
is independent of waveguide thickness d in the limit Nj -+ n g . The
differences in attenuation must be attributed to differences in abo
due, perhaps, to different sample purities or to experimental inaccu-
racy. The measured attenuation in all samples measured, including
samples with molecular alignment parallel to the waveguide plane,
was much greater than that reported by Sheridan et aL 2 The very low
attenuations they report (less than 1 dB/cm) could not be reproduced
in any of our experiments.
The TM mode attenuation increases with increasing mode number.
The rate of increase with decreasing N j is seen to be independent of
292 CHAPTER 15

sample thickness. Surface scattering strength increases with mode


number, but is inversely proportional to the waveguide thickness (see
Eq. [16]). Since the thickness dependence is not observed, the mode
dependence of the TM scattering is attributed to wavelength depen-
dence in the bulk scattering. The solid curve was fitted to the TMo
attenuation measured for the 12-J,Lm guide, and then scaled by the
factor (N/NoJ -4 to represent the wavelength dependence of Ray-
leigh scattering. The fit to the data is obviously very good. Further
work is needed, however to understand in detail the scattering mech-
anisms. It may be particularly rewarding to try and resolve any sur-
face scattering effects resulting from various techniques of molecular
alignment.
Electro-optic effects were produced in nematic waveguides by
applying pulsed electric fields. In one series of experiments, interdigi-
tal electrodes were deposited on one of the bounding surfaces and
driven with voltage pulses. The light coupled out of the guide down-
stream from the electrodes was monitored with a photomultiplier and
the response viewed on an oscilloscope.

Fig. 7- Transmission of light in nematic waveguide is diminished in response to a


pulsed electric field (rectangular pulse) and recovers when field is off (drive
voHage of 1-msec duration and 50-volt amplitude).

Fig. 7 shows a voltage pulse (1 msec duration and 50 V amplitude)


and its effect on waveguide propagation. Transmission drops during
the pulse duration and recovers exponentially afterwards. The light
transmission may be completely turned off by this process. Response
LIQUID-CRYSTAL OPTICAL WAVEGUIDES 293

times T off and recovery times Ton are shown in Fig. 8. It is presumed
that the modulation mechanism is refraction of light out of the wave-
guide by the index modulation caused by molecular rotation in re-
sponse to the applied electric field.
As in the case of the attenuation, a more detailed investigation of
these phenomena is warranted. For example, the variation of the re-

;;; 10

§
on

!?
....
on
IX
.
10

PULS

Fig. a-Response times to application of field T off and recovery times after removal of
field Ton for waveguide transmission shown in Fig. 7.

covery time T on with pulse voltage is quite unexpected, and suggests


that complex patterns of molecular rotation may be excited by the in-
terdigital electrodes. Such effects are relevant to display-device tech-
nology, and optical-waveguide techniques appear to be an advanta-
geous way of studying them.
Dynamic scattering has also been used to modulate guided light. In
this experiment a liquid-crystal layer was established on the surface
of a passive thin-film waveguide. Electrically excited turbulance scat-
tered light in the evanescent field of the thin-film waveguide, which
extended into the liquid crystal. The extinction of waveguide light as
a function of ac and dc dynamic-scattering excitation voltages is seen
in Fig. 9. Very high voltages are necessary, since only a small part of
the optical field interacts with the liquid crystal.
294 CHAPTER 15

6. Conclusions

Optical-waveguide propagation in liquid-crystal layers has been


achieved, and basic electro-optic device phenomena have been dem-
onstrated. The usefulness of such devices has yet to be established,
particularly in view of the considerable progress being made with
other active waveguides, such as single-crystal electro-optic films.14
Liquid-crystal materials offer the advantages of high electro-optic
index changes and a variety of birefringence and scattering operation

1.0

z 8 AC
2
f- \-/~ ---
u DC
z /.
i= .6 /.
x
w /,
w /,
0 /,
'5 .4
<!) /,
w
~
I
~ /
.2

/
/
00 500 1000 1500
EXCITATION VOLTAGE

Fig. 9- Transmission reduction In thin-film waveguide when dynamic scattering Is ex-


cited In nematic overlay.

modes. They are handicapped at present by slow response and recov-


ery speeds, high scattering loss, and the need for confinement in
sealed cells. Possibly the most fruitful area of application may be for
side-illuminated display panels, should useful materials with low
scattering loss be developed.
Optical waveguiding may be most useful as a technique for studies
of liquid-crystal phenomena. Accurate measurements of the mode
spectrum can be used as a probe of transient changes in the refractive
index profile of the liquid-crystal layer. Light scattering into radia-
tion modes and between guided modes determines the spatial fre-
quency spectrum of the alignment inhomogeneities and fluctuations.
Scattering between closely spaced modes is a sensitive measure of
very-small-angle forward scattering. These same techniques may be
useful in studying physical phenomena in other liquid layers as well
as liquid crystals.
LIQUID-CRYSTAL OPTICAL WAVEGUIDES 295

References

1 D. J. Channin, "Optical Waveguide Modulation Using Nematic Liquid Crystal," Appl. Phys. Lett.
22, p. 365 (1973).
2 J. P. Sheridan, J. M. Schnur, and T. C. Giallorenzi, "Electro-Optic Switching in Low-Loss Liquid
Crystal Waveguides," Appl. Phys. Lett., 22, p. 561 (1973).
3 J. P. Sheridan, "Liquid Crystals in Integrated Optics," OSA Topical Meeting on Integrated Optics,
New Orleans, Jan., 1974.
• Chenming Hu, John R. Winnery, and Nabil M. Amer, "Optical Deflection in Thin-Film Nematic-Liq-
uid-Crystal Waveguides," IEEE J. Quan. Elect., QE-10, p. 218 (1974).
S H. A. Weakliem, D. J. Channin, and A. Bloom, "Determination of Refractive Index Changes in
Photosensitive Polymer Films by an Optical Technique," Applied Optics, 14,560 (1975).
6 P. K. Tien, "Light Waves in Thin Films and Integrated Optics," Appl. Opt., 10, p. 2395 (1971).
7 Dietrich Marcuse, ed., Integrated Optics, IEEE Press, New York (1973).
S Dietrich Marcuse, Light Transmission Optics, Van Nostrand Reinhold Co., New York (1972).
9 P. K. Tien, R. Ulrich, and R. J. Martin, "Modes of Propagating Light Waves in Thin Deposited Semi-
conductor Films," Appl. Phys. Lett., 14, p. 291 (1969).
10 M. L. Dakss, L. Kuhn, P. F. Heidrich, and B. A. Scott, "Grating Coupler for Efficient Excitation of
Optical Guided Waves in Thin Films," Appl. Phys. Lett., 16, p. 523 (1970).
11 J. Kane and H. Osterberg, "Optical Characteristics of Planar Guided Modes," J. Opt. Soc. Am.,
54, p. 347 (1964).
12 D. J. Channin, J. M. Hammer, and M. T. Duffy, "Scattering in ZnO-Sapphire Optical Waveguides,"
Applied Optics, 14,923 (1975).
13 P. G. deGennes, The Physics of Liquid Crystals, Oxford University Press, London (1974).
14 J. M. Hammer and W. Phillips, "Low-Loss Single-Mode Optical Waveguides and Efficient High-
Speed Modulators of LiNbxTa'-x03 on LiTa03," Applied Phys. Lett., 24, p. 545 (1974).
The Electro-Optic Transfer Function in Nematic Liquids*

Alan Sussman
RCA Solid State Division, Somerville, N.J. 08876

1. Introduction

When an electric field is applied across transparent plane-parallel


electrodes containing mesomorphic liquids, many complex phenomena
occur that depend on the optical, dielectric, and elastic properties of
the liquid, the geometry of the test situation, and the nature of the
electrical signal. 1 The electro-optic transfer function is a way of
specifying such optical changes. An ever increasing interest in liquid-

* Presented in part as an invited paper at the Optical Society of


America annual meeting, Rochester, N.Y. Oct. 1973.

297
298 CHAPTER 16

crystal electro-optic phenomena particularly, but not entirely, in the


field of display devices has caused a corresponding growth of the litera-
ture 2 ; hence, this chapter is limited to steady-state properties of nematic
liquids.

(J) FREQUENCY (Hz) DC 40 400 4K


0::
<I
..J
o
11.

-'"
X

.
g
III
<t
++--1>
z
2
(J)
(J)

::;:
(J)
Z
<I
0::
f-

ELECTRIC FIELD

Fig. 1-Field effect (lateral electrode geometry). The change in optical


orientation properties with frequency is a result of changing field
penetration. The dielectric relaxation frequency is 80 Hz. This
material of positive dielectric anisotropy does not show any tur-
bulent flow (R-A. Soref9 ).

It is convenient to divide nematic liquids into two classes depend-


ing on the dielectric anisotropy-positive for those in which the mole-
cular orientation described by a vector, called the director, tends to
align parallel to an electric field, and negative when that alignment
is perpendicular to the field. Those phenomena in which reorientation
is the major result are considered "field effects". This designation is
not arbitrary, since it also separates these dielectric effects from those
whose mechanism has additionally a hydrodynamic flow, Le., realign-
ment or disruption of the liquid by material transport. Sometimes,
material parameters completely specify electro-optical performance,
while in other cases, the two regimes are selectable by choosing the
frequency of excitation.
In materials of positive dielectric anisotropy, most electro-optic
phenomena are frequency independent field effects; hydrodynamic ef-
fects, occurring with certain boundary conditions result in stable
(laminar) flow,3 and no turbulent-flow reorientation is observed.4
ELECTRO-OPTIC TRANSFER FUNCTION 299

In the storage effect,12 on the other hand, a write mode is obtained


below the dielectric relaxation frequency. This is a hydrodynamically
produced turbulent flow that disturbs the (cholesteric) order. An erase
mode is produced at high frequencies, where the ordered structure is
restored by dielectric reorientation forces.
Depending on the geometry, dielectric reorientations are sometimes
accompanied by hydrodynamic transients, i.e., a backflow.!3 (The re-
orientation caused by hydrodynamic flow is well known.!4) Such back-
flow transients, under certain conditions, can give rise to turbulent
effects similar to dynamic scattering; the latter term should be re-
served specifically for the sequence: domains, domain instability, and
scattering, produced by conductance hydrodynamics.
Another important frequency is the inverse of the time it takes
for an ion to traverse the whole of the way across the electrode dis-
tance tiT = V2/ f1.d, where V is uhf' voltage, f1. the mobility, and d the
electrode spacing.!S Unlike the dielectric relaxation frequency, f (tran-
sit- 1 ) depends on voltage and cell thickness. It effectively separates ac
from dc effects whose hydrodynamic mechanisms are related to elec-
trode effects.!S Note that for high enough voltages and/or small enough
spacing, f(transit- 1 ) can exceed the electrical operating frequency.
This condition is particularly favorable for loss of conducting ions by
electrosorption; such effects have been noted, especially in nematics
with low (initial) conductivity.17
Differences in threshold behavior when other than sinusoidal excita-
tion is used in dynamic scattering!' have been observed.!9
Modifications to the threshold-voltage characteristic may also be
obtained by adding a high-frequency signal to the low-frequency
drive.'o This adds a dielectric orienting field that tends to suppress
hydrodynamic flow, thereby increasing the low-frequency voltage re-
quired to cause a given scattering as compared to that required with-
out the high frequency.

2. Geometric Considerations In Optical Measurements

Careful consideration must be given to using an appropriate geometry


when taking measurements, particularly if the results are to be used
to evaluate the visual performance of a display system.
In dielectric reorientation effects (such as voltage-controlled bire-
fringence;! twisted nematic structure effects," and guest-host effects"),
the optical changes are basically variations in either optical density or
wavelength of absorption. As such, the optical measurements are quite
straightforward. The plane-parallel-geometry structures may be con-
sidered neutral density filters, with angle-dependent transmission or
300 CHAPTER 16

wavelength characteristics. It is necessary to apply corrections for


refraction because of index mismatch," which is particularly important
for measurements on optical parameters where a liquid crystal is used
as a switch to couple into a waveguide," to modulate and switch in a
waveguide,26 or to control total reflectance at a prism/liquid-crystal
interface.27

Fig. 2-Diffraction by domains. A nematic orientation that is originally


planar usually results in a well-ordered domain structure that
behaves like a phase grating, and a diffraction pattern consisting
of a series of spots is obtained (top). When the orientation is
originally perpendicular, clusters of domains may give rise to
diffraction spots or rings depending on the relative diameter of
the probe (w) and the dimensions of the domain clusters.

When scattering or diffraction phenomena are studied, however, the


geometry becomes particularly important, both because of the rich
nature of the optical processes and because of the more varied ways in
which such processes are applied to display devices. In the case of
diffraction by domains,2' different results may be obtained depending
on the original orientation of the fluid and the relation between the
beam probe size and the domain cluster size.
When domains are observed in configurations where the surfaces
induce planar orientation,29 they exhibit the properties of a phase grat-
ing, and unpolarized laser illumination at normal incidence produces a
spot diffraction pattern"· (Fig. 2, upper part).
ELECTRO-OPTIC TRANSFER FUNCTION 301

If, however, the orientation is originally perpendicular, the first


electro-optic effect is the conversion to a planar orientation in the bulk,
followed at higher voltage by the domains, whose cylindrical axes may
not be correlated over large distances, and "clusters" result. The dif-
fraction pattern will be a ring if the probe diameter (w) is large
compared to the cluster size (Fig. 2, lower part).
To measure the angle dependence of scattering in a transmission-
type display, three general schemes may be employed. In Fig. 3 they
are illustrated schematically, and qualitative electro-optic transfer
curves are given for each.

'----v
o
I NA RROW BEAM, REAR

n EXTENDED SOURCE, REAR

ABSORBING
LIGHT
MATERIAL ~i ~ </> B 0

ill EXTEN DED SOURCE, FRONT

Fig. 3-Influence of geometrical parameters on scattering electro-optic


transfer function measurements, Three cases are illustrated, with
their corresponding electro-optic transfer functions (highly sche-
matic).

Type I-Narrow Beam, Rear

In type I, a laser beam is used as a coherent probe; such a configura-


tion has given a good deal of information on the ordering of the
nematic liquid,"' on domains,a. and on the structural modifications that
occur during dynamic scattering. aa Both the angle of incidence and the
detector angle can be varied; such data can be reported in compact
302 CHAPTER 16

graphical form. 34 This configuration is also applicable to optical data


processing 35 and holographic information-storage techniques. 36
With illumination normal to the plane of the structure (Fig. 3, top)
and no voltage applied (0 curve), only a small inherent scattering
may be measured off-axis.31 On-axis, of course, the unscattered beam
is virtually undiminished in intensity. With the application of voltage
V, there is finite scattering into off-axis angles (increase in signal),
but an attenuation of the on-axis brightness signal; a contrast ratio
as high as 10 4 has been measured for spatial filtering applications. 37

Type II-Extended Source, Rear

Type II has an extended-area source behind the cell, with the detector
in front. The liquid-crystal plane when excited reduces the intensity,
and may be thought of as a variable density screen, similar to type I
before the crossover, with the liquid crystal plane integrating the
light, which arrives at all angles. This is a rather uninteresting con-
figuration, although some knowledge of the angle dependence of the
scattering is needed when the liquid-crystal plane is used for a view-
ing screen,38 as a variable optical stop, or as an image plane for
projection.

Type III-Extended Source, Front

In type III the illumination is from the viewing side, with a light-ab-
sorbing screen behind. This configuration is not used frequently in
digital displays, because the scattering to the rear is not particularly
efficient.17 By illuminating the screen from the back, full use is made
of the scattering in the propagation direction, i.e., toward the viewer.
A unique configuration for display devices is to illuminate from the
rear through a screen consisting of blackened vanes adjusted so as to
give the maximum viewing angle with minimum total thickness of
the system. 39
A similar configuration, but with a mirror surface as the rear
electrode, has a wide application in watch and clock displays. Best
results are obtained when the viewer sees reflected in the mirror a
light-absorbing surface. The on-state is particularly bright because
both transmitted light and scattered light are reflected back toward
the viewer. A type III electro-optic transfer curve is thereby obtained.
If, however, the mirror reflects a bright surface, it is possible that
contrast reversal will occur, i.e., the type II configuration. For some
angles and brightnesses, a no-contrast result is possible. In the case
ELECTRO·OPTIC TRANSFER FUNCTION 303

of field effects, particularly the twisted structure, type III effects are
especially subtle when the angle of incidence is varied.


£>4>=5"

UJ
<!)

....<l £>4>=4"


.J
0
>
<!)
0
.J £>4>=3.-

£>4>=2,.

£>4>=.,
Vth

ITIllJJ]] v =0
Ife
LOG FREQUENCY

Fig. 4-Retardation as a function of voltage. The continuing reorientation


of the structure (left side) with increasing voltage results in an
ever increasing retardation, i.e., birefringence. The retardation
act> continues to increase unless another electro-optic regime ensues;
viz, below the dielectric relaxation frequency, domains, and above,
chevron distorions (M. Hareng, et al 21 ) .

3. Field Effects-Negative Dielectric Anisotropy

With an orientation that is originally perpendicular to the electrodes,


and material of negative dielectric anisotropy, an electric field interacts
with the fluid to cause a tilt in the optical axis, beginning at the center
of the cell, when a threshold voltage is reached. 21 This tilt results in
birefringence that continues to increase with increasing voltage, unless,
of course, the threshold voltage for another phenomenon is exceeded,
e.g., domains at low frequency or "chevrons" at high frequency'o (Fig.
4). As the field increases, the tilt angle continues to increase as does
the retardation, measured as a phase difference Do cp." To get uniform
retardation over large areas, the surface may be prepared to give a
slight bias, i.e., a uniform but small tilt off-normal."
304 CHAPTER 16

The electro-optic transfer function at a fixed angle between crossed


polarizers is shown in Fig. 5. By swinging between two voltages, it
is possible to modulate the color equivalent to a phase retardation at
one wavelength (first voltage) to produce a color equivalent to a dif-
ferent wavelength (second voltage). This is the basis for a matrix-

"

'"a:<t
..J
o
"-
I
-x" I
z /
o I
'"'" /
:;; /
'"z<t /
a:
I-
/
/
/
,/

VOLTAGE

Fig. 5-Electro-optic transfer as a function of voltage. At a fixed view-


ing angle, the transmission of monochromatic light between crossed
polarizers undergoes a series of maxima, which occur with in-
creasing voltage and represent phase retardations of 'IT, 2'IT, .•• etc
(M. Hareng, et al 21 ) . •

addressed color crossed-array display 43 where the angle dependence of


the color produced by the phase retardation (or transmission of light
between crossed polarizers when monochromatic light is used, as shown
in Fig. 6) has been avoided by using a projection system. Inter-
digitated eleetrodes may be used to create small areas of different re-
tardation, the overall area being angle independent. 44
The angle dependence can be used to make accurate measurements
of the electrode spacing in liquid crystal cells. 45
Advantage may be taken of the property of certain pleichroic dyes
in which the wavelength of absorption in the crystal is different along
different crystalline axes. If such a dye is dissolved in and follows the
alignment of the liquid crystal of the previous example'· with the
analyzer removed, the absorption in the voltage-off state would depend
on the extinction coefficient perpendicular to the dye crystal axis;
when the structure is changed by application of voltage, the absorption
ELECTRO..QPTIC TRANSFER FUNCTION 305

becomes converted to that related to the extinction coefficient parallel


to the crystal axis. If one of these absorptions is not in the visible, no
polarizer is required to see the color change, and the angle dependence
results in a change in optical density at the visible wavelength.
Discussion of the chevron regime, although also a field effect, is
described in the section on diffraction .

.~
0

'"
)(
iL
>
C
'"
)(

LL
.-<

'"<Ia:
..J
0
Q.

.'"
x
i
!2
'"'"
i
'"<IZ
a:
....

ANGLE ;

Fig. 6-Transmission as a function of viewing angle. At a constant voltage,


and for a fixed wavelength, the transmission between crossed
polarizers shows an angular dependence. At angles other than
normal, a correction for refraction by the glass is necessary (zero
voltage is illustrated).44

4. Field Effects-PosHlve Dielectric Anisotropy

Since materials of positive anisotropy turn their orientation parallel


to the direction of a field, a necessary requirement is an off-state of
uniform planar alignment. Then, a change analogous to that just dis-
cussed may be obtained by the application of a field; above a threshold
voltage, the parallel orientation (beginning at the center) eventually
becomes indistinguishable optically from one that is perpendicularly
aligned. Between crossed polarizers, the result is a lessening of the
retardation with increasing voltage, with an electro-optic transfer
function similar to that of Fig. 5. The change is maximum when the
306 CHAPTER 16

polarizer axes are normal to the planar axes. At any voltage, the
transmission is angle dependent.47 with a symmetry related to whether
or not the angle is varied in a plane containing the undisturbed liquid
orientation.
Incorporation of a pleiochroic dye and elimination of the analyzer
results in voltage-controlled optical absorption, the guest-host effect. 4s
The "twisted" nematic structure has unique optical properties: the
plane of linearly polarized light, of wavelength shorter than the dis-
tance over which the twist makes one revolution, is rotated following
the twist ..• In static patterns a 90 rotation is maximum, greater angles
0

being unstable. Placed between crossed polarizers, such a cell would


cause maximum transmission. The converse would be true if the
polarizers were initially parallel.

50%1---_-::::: - - - - - - - - - - - - - - - - - J - - - --

z
o
(/)
(/)

:;;
(/)
z
'a:f-"

V,
VOLTAGE

Fig. 7-Electro-optic transfer function of twisted structure. Above thres-


hold, increasing the voltage casuses the structure to lose the ability
to rotate the plane of polarized light; eventually, the optical be-
havior of the structure approaches that of a perpendicular home-
otropic nematic. The difference (e,), however, is still finite at three
times threshold. 52

If the material has a positive dielectric anisotropy, then an elec-


tric field can convert the twisted structure to one that eventually has
the optical properties of a perpendicularly oriented one." A threshold
voltage, as shown in Fig. 7, is present; only at sufficiently high voltages
is the difference between the still-twisted and the non-twisted structure
undetectable. As in the voltage-controlled birefringence, the tilt of the
ELECTRO-OPTIC TRANSFER FUNCTION 307

orientation continues with increasing voltage, being greatest at the


midpoint between the electrodes (Fig. 8, top). This causes birefringent
color changes, but an overshadowing optical effect occurs because of
the modifications to the twist. Initially, the twist may be considered
uniform throughout the cell thickness (Fig. 8, bottom). With increas-

90°

W
..J
'-'
Z
<[

e- v, _ _
'"~ v2 - -
e-
v3 - - - -
v4 - · -

05 10

FRACTION OF CELL THICKNESS

Fig. 8-Variation of tilt and twist angles throughout twisted structure as


a function of voltage. Top figure shows symmetric tilt occuring as
the voltage increases (as in Fig. 7). Bottom figure shows the
symmetric twist that, while uniform at voltages below threshold,
becomes distorted with increasing voltage. When most of the twist
occurs over a distance small compared to the wavelength of light,
the structure can no longer rotate the plane of polarization. 52

ing voltage, the twist becomes distorted. Above that value of voltage
corresponding to the optical threshold, the total distributed twist is
converted to distortions near the wall, so that the rotation conditions
cannot be met50 and the optical density of the structure changes. Even-
tually, increasing voltages result in almost total loss of rotation; for a
cell between crossed polarizers initially, the optical transmission would
increase from that of the parallel polarizers to one that approaches that
of the crossed polarizers. The coherence of light so rotated is not af-
fected greatly.51
Fig. 9 shows calculated curves· 2 for angle-dependent transmission
308 CHAPTER 16

GLASS 10° 20° 30° 40°


AIR 15 5° 31.8° 50.4° 818°

ANGLE FROM VERTICAL, IN PLANE PARALLEL TO ORIENTATION

Fig. 9-Electro-optic transfer coefficient along surface orientation for


twisted structure. With the source (S) in a plane perpendicular
to the electrodes and parallel to the optic axis of the nematic
orientation at that electrode, the optical transmission is a mono-
tonic function of angle and voltage (calculated curves).52

in which the light being measured emerges perpendicular to the elec-


trode face, while the source (S) is rotated in a plane that is perpendi-
cular to the electrode and contains the orientation of one of the
electrodes. Note that the results are relatively monotonic. If how-
ever, the source is in a plane rotated so as to be 45° to the orientation
direction in the electrode face, the results are complex (Fig. 10).*
Type I, II, and III measurements may be expected to yield such com-
plex results.
Type II is at present the basis for most digital applications. If a
specular back is used (behind the polarizer), the reflected brightness
is maximized, but at the expense of curtailed viewing angle; in Type
III configuration, with a scattering back plane acting as a (partially
depolarized) source of illumination, the optical results would be ex-
pected to be similar to those with Type II.

5. Hydrodynamic Effects-Diffraction by Domains

Liquid-crystal hydrodynamic phenomena are extremely complex, and


the mechanisms for their production are far from being completely
" The reason is again that the light in the probe beam now is resolved
into birefringent components. In these calculations, the index of the glass
was assumed to be equal to the higher index of the nematic liquid and the
necessary refraction for other than normal incidence corrected for. Note
that this results in a considerable difference between the geometrical
beam/glass angle and the actual beam/liquid angle.
ELECTRO-OPTIC TRANSFER FUNCTION 309

elucidated. The differences between positive and negative dielectric


anisotropy material are more distinct than in the field effects. The
negative materials are more likely to produce striking opto-hydrody-
namic changes. There are, however, indications that positive materials
may participate in similar although more subdued effects.53

~ -.........'J~
'\ s

r-------_" '---
~ ,
w
~ V3 \ <t --..- / : s

; \'<-----, /~",
j

~
Q
\. j'
1\\
~
~
\ / \
:=t----____ - - _ _ _ _,;y
V2 "----" \

v, \

GLASS 10' 20' 30' 40'


AIR 155' 31.8' 50.4' 81.8'

SOURCE ANGLE FROM VERTICAL, IN PLANE AT 45' TO ORIENTATION

Fig. lO-Electro-optic transfer coefficient at 45° to surface orientation for


twisted structure. When the source (S) is in a plane perpendi-
cular to the electrodes but not parallel to the optic axis of the
nematic in the electrode (director), the variation of transmission
between crossed polarizers as a function of angle and voltage is no
longer simple, because of anisotropic propagation within the liquid
(calculated curves) .52

The following is a typical series of events when a material of


negative dielectric anisotropy, originally in the planar orientation, is
subject to an electric field of a frequency between the dielectric relaxa-
tion frequency and the inverse transit-time frequency. Dielectrically,
the field should cause no change; above a threshold voltage, however, a
periodic pattern of light and dark lines appears. These domains;4 as-
sociated with vortical liquid flow;5 have a spacing proportional to the
electrode separation. The refractive properties are not unlike an array
of cylindrical lenses, with alternate convergent and divergent bands,
which are the domain boundaries, visible in ordinary light. This focal
length decreases with increasing voltage and represents the change in
curvature of each lens due to increased velocity of the fluid. (The
number ,density of domains need not change.)
The periodic pattern also exhibits the properties of a phase grating,
and a diffraction pattern may be observed (Fig. 2). When the voltage
is raised, the steady-state result is a higher domain periodicity, i.e.,
310 CHAPTER 16

new domains have made their appearance. There being no smooth way
for a new domain to decrease the periodicity, a transient diffraction
occurs, which relaxes into a new equilibrium steady-state pattern whose
intensity distribution may be correlated with the spatial frequency of
the domain."o Such new domains, by the nature of the symmetry of
the electrode configuration, might be expected to also have a sym-
metrical flow pattern. With further increase in voltage, the new
domains apparently cannot enter into an orderly array, and the
turbulent properties of· dynamic scat:tering begin to appear. Optically
speaking, the structure behaves like a deep phase screen, i.e., at each
position there is a phase retardation resulting in a path difference
which may be several wavelengths. 33
At frequencies below the inverse transit time, the electrode effects
may no longer be geometrically symmetric ;17 the domain periodicity is
nonuniform because a second diffraction pattern, harmonically un-
related to the original set, has made its appearance. I '
The stability of the hydrodynamic modes of cylindrical domains has
been investigated using a model with the appropriate boundary condi-
tions:· The results indicate a stability curve with two branches-one
on which the domain spacing remains constant with voltage, the other
on which the spacing varies inversely with the voltage (as just de-
scribed). These appear in Figure 11. Under certain conditions, which
may be depend on material parameters as well as electrode spacing,
dc excitation does not result in dynamic scattering, but only in a
continuing increase in domain periodicity.57 This result is shown in
Fig. 12. Here, the simplified experimental data show the intensity dis-
tribution of a diffraction ring. The original orientation was per-
pendicular to the electrodes, so that the domain clusters would be ex-
pected. In this case, an unusual domain structure was observed. 58 A
tree-like pattern, as shown in the upper inset of Fig. 11, allowed an
almost continuous variation in domain spacing as the voltage changed.
The new domains appeared as new branches and the growth was quite
smooth. When the voltage was increased, the diffraction pattern would
become diffuse for an instant, then relax into the sharp diffraction
rings, as shown in Fig. 12. The intensity profile of the highly reflected
domains was a sharp as that for less diffracted ones, indicating an
improvement in diffraction efficiency with increasing voltage.
At frequencies above the dielectric relaxation frequency, another
diffraction with a threshold makes its appearance (see Fig. 4). This
is the "chevron" regime,S> which is not hydrodynamic in origin but
depends on interaction between the anisotropic dielectric properties of
the fluid and the field. It represents a sinusoidal modulation of the
ELECTRO-OPTIC TRANSFER FUNCTION 311

idl

.u
C>
z

a.

.
<Il

~
::;;
0
I~
0

<Il
<Il
W
Z

'"
~
:J:
>--
..J
..J
W DOMAIN SPACING" CONSTANT
U

VOL TAG E

Fig. ll-Domain spacing as influenced by hydrodynamic regime. The pe-


riodicity of domains shows a voltage variation that depends on
hydrodynamic modes, represented by two branches of the stability
curves. 56 Insets: right, domains are cylindrical; left, a tree
structure allows a continuous variation of the spacing d. s,

--IOV
- - - 20V
- - - - 30V
- - - 40V
>- ............ 60V
>--
<Il
Z
W
>--
Z

DIFFRACTION ANGLE

Fig. 12-Voltage dependence of diffraction by domains. The intensity dis-


tribution of the diffraction rings (for a single wavelength) with
increasing dc voltage demonstrates a constant intensity profile
even at 60 volts where the diffraction angle is nearly 50°. No
dynamic scattering occurs.57
312 CHAPTER 16

director, with a threshold field proportional to the square root of the


frequency, and results in a phase-grating condition:· The spatial fre-
quency of these domains is much greater than that of the low-frequency
domains; the diffraction also gives rise to angle-dependent chromatic
effects. The model based on a symmetric periodic slit grating, which is
satisfactory for low-frequency domains, must be corrected for refrac-
tion within the liquid."· An added stabilizing field, at frequencies above
threshold for chevrons, causes an increase in period."!

L'l
Z

......'"
UJ

«
u
(f)
tcp
V

LC
~
flV

...,.
(f)
z
UJ
a
...J
«
u
...a.
o

VOLTAG E

Fig. 13-Photoconductor control of dynamic scattering. Electro-optic trans-


fer function is shown in the dark and under saturated photocon-
ductor conditions. For the liquid crystal to be below threshold
voltage in the dark, the value of the applied voltage must not
exceed the original threshold voltage (V,,) plus the voltage drop
across the dark photoconductor (/::, V) . With suitable choice of
photoconductor parameters, the light-saturated photoconductor al-
lows the liquid-crystal response to behave almost as if no photo-
conductor were present.

6. Dynamic Scattering

As domains become increasingly unstable with increasing voltage, the


phenomenon known as dynamic scattering 18 occurs. Fig. 13 shows a
a typical electro-optic transfer function. Note the lack of saturation of
scattering with voltage in the type I case. At higher voltages, changes
in flow patterns occur at definite thresholds, and result in changed op-
tical, electrical, and kinetic properties."' Another phenomenon, caused
by certain patterns of flow, particularly near the dielectric relaxation
ELECTRO-OPTIC TRANSFER FUNCTION 313

frequency, converts an originally perpendicularly oriented off-state to


one that is (temporarily) planar.·'
The angular dependence of the scattering has been observed as a
function of voltage,"4 and the results in Fig. 14 are typical. (The

....
I
<!l
..J
0
UJ
Q:
UJ

........
'"
u
<f)

<!l
0 60 V
..J

40 V

20 V

ov
0 40·
ANGLE ~

Fig. 14-Angular dependence of dynamic scattering as a function of volt-


age. Under narrow beam condition (Type I), the scattering of
the original beam near the optic axis results in a reduction in
transmission; off-axis, however, the scattering results in an in-
crease in light detected. (The crossovers do not necessarily
occur at the same angle.)

crossover, however, is not always found at the same angle for every
voltage.) The effect of increasing cell thickness is an increase in the
voltage at which equivalent scattering occurs; the threshold is un-
changed. The angular dependence is also relatively independent of
thickness." The scattering at a particular value of thickness, angle,
and voltage has a direct dependence on the optical anisotropy of the
nematic compound under study."' The scattering for any compound is
independent of the current density·· as long as the measurement is
made sufficiently far from the dielectric relaxation frequency.
When polarized light is used, the angular scattering dependence
is a function of the angle between the polarizer and the direction of
the (original) planar orientation."
314 CHAPTER 16

7. Photoconductor Control

The incorporation of a photoconductor layer in series 67 with the liquid


crystal layer allows direct conversion from optical input to optical out-
put; such devices hold promise of usefulness in the field of image and
data pl'ocessing,43 optical conversion:' and storage:" As of this review,
photoconductor control of dynamic scattering and storage effects only
have been reported. Optimum performance of such devices is achieved
when the resistance of the photoconductor is matched to that of the
liquid crystal, so that the voltage drop across the photoconductor, Ll V,
leaves the voltage across the liquid crystal below the threshold value
when the photoconductor is unilluminated; correspondingly, when the
photoconductor is illuminated, the voltage drop should be as low as
possible, allowing full scattering operation of the liquid crystal. Fig.
13 shows the electro-optic transfer functions for the dynamic scatter-
ing discussed above for the series photoconductor saturated and in the
dark. The optimum applied voJtage, as will be seen later, is V th + Ll V.
The relationship between the photoconductor and liquid resistance
in the dark and under illumination has been treated for the case of
pure resistance. There are reported cases in which the electro-optic
response depends on whether the illumination falling on the photo-
conductor reaches it from the liquid-crystal side or from the trans-
parent-conductor side. T·his rectification is not observed in all liquid-
crystal photonconductor systems!O However, the generality of the above
discussion is not affected.
To achieve sufficient control over the photoconductor, with either
a bulk or barrier photoconductor, requires a photon flux of 10 11 /cm 2 !'
This value is comparable to the flux required for other processes, such
as Xerographic or electroluminescent outputs. When the photoconduc-
tor is nonohmic, higher minimum exposures may be required.
Operation in the transmission mode allows direct insertion of the
image plane into an optical crystal. This, of course, requires the photo-
conductor to operate outside the spectral range of the projection source.
In reflectance modes, the photoconductor may be isolated from the
secondary illumination by means of an insulating and/or opaque dielec-
tric layer!2 Such a configuration requires ac operation, which happily
circumvents some of the electrochemical problems associated with dc. 73
Using the somewhat artificial assumption that equal increments of
illumination result in equal changes in Ll V, Fig. 15 was developed. The
liquid-crystal sc.attering is plotted against the illumination up to the
value at which the photoconductor "saturates", i.e., when increasing
light no longer significantly changes ~ V. If the values are properly
E LECTRO-OPTIC TRANSFE R FUNCTION 315

'";;;a:
w
l-
I-
<t
u
(f)

>-
I-
Ui
Z
W
0

.J
<t
U
;::
0-
0

0 5
ILLUMINATION

Fig. 15-Dynamic scattering range as a function of light intensity and


applied voltage. The operating voltage VI. + /:',. V may be optimized.
At lower than optimum voltages, the maximum scattering (S)
with a properly selected photoconductor, is reduced. At even lower
voltage, the scattering reaches a saturation value, but is not zero
even at low light levels (calculated curves).

chosen, the behavior of the liquid crystal should approach that observed
with no series photoconductor. At voltages above and below the opti-
mum, V th + AV, the scattering curve was estimated from Fig. 13. Note
that for voltages lower than optmium, maximum scattering cannot be
obtained; at the lowest voltages, the scattering is not zero even when
the photoconductor is unilluminated. The shape of the original scatter-
ing-voltage curve has a significant effect on the dynamic range. For
example, a steep electro-optic function, which is desirable for matrix
addressing or digital display, will reduce the halftone range in the
photoconductor combination, as might be expected.

References
1 Several review articles, emphasizing electro-optic properties of liquid crystals are
available: R. A. Soret, "Liquid Crystal Light Control Experiments," in The Physics of
Opto-Electronic Materials, ed. W. A. Albers, Jr., Plenum Press, New York (1971); A.
Sussman, "Electro-Optic Liquid Crystal Devices: Principles and Applications," IEEE
Trans. Parts, Hybrids, and Packaging, Vol. PHPB, p. 24 (1972); and A. Sussman, "Liquid
Crystals in Display Systems," in Liquid Crystalline Systems, ed. G. W. Gray and P. A.
Winsor, Ellis Horwood, London (in press).
2 For example in 1969, there were 70 papers listed in Physics Abstracts. In 1970, 250;
in 1971, 315; and in 1972, 390.
3 P. DeGennes, "Electrohydrodynamic Effects in Nematic Liquid Crystals I. DC Effects,"
in Comments on Solid State Physics, Vol. 3, p. 35 (1970); and P. A. Penz, Electrohydro-
dynamic Solutions for Nematic Liquid Crystals with Positive Dielectric Anisotropy,"
Mol. Cryst., Vol. 23, p. 1 (1973).
316 CHAPTER 16

4 W. H. DeJeu, C. J. Gerritsma, and Th. W. Lathouwers, "Instabilities in Electric Fields


of Nematic Liquid Crystals with Positive Dielectric Anisotropy: Domains, Loop Domains,
and Reorientation," Chem. Phys. Lett., Vol. 14, p. 503 (1972).
5 G. H. Heilmeier and W. Helfrich, "Orientational Oscillations in Nematic Liquid
Crystals," Appl. Phys. Lett., Vol. 16, p. 155 (1970).
6 H. S. Harned and B. B. Owen, The Physical Chemistry of Electrolytic Solutions, 2nd
ed., Reinhold Publishing Co., New York, 1950, Chap. 4.
7 A. Sussman, "Ionic Equilibrium and Ionic Conductance in the System Tetra·iso-pentyl
Ammonium Nitrate p-Azoxyanisole," Mol. Cryst. and Liq. Cryst., Vol. 14, p. 182 (1971).
8 D. Meyerhofer, A. Sussman, and R. Williams, "Electro-Optic and Hydrodynamic Prop-
erties of Nematic Liquid Films with Free Surfaces," J. Appl. Phys., Vol. 43, p. 3685
(1972).
9 R. A. Soref, "Transverse Field Effects in Nematic Liquid Crystals," Appl. Phys. Lett.,
Vol. 22, p. 165 (1973); and N. V. Madhusudana, P. P. Karat, and S. Chandrasekhar,
"Some Electrohydrodynamic Distortion Patterns in Nematic Liquid Crystals," Current
SCience, Vol. 42, p. 147 (1973).
10 W. Haas, J. Adams, and J. B. Flannery, "New Electro·Optic Effect in a Room·Tem·
perature Nematic Liquid Crystal," Phys. Rev. LeU., Vol. 25, p. 326 (1970).
11 W. Helfrich, "A Simple Method to Observe the Piezoelectricity of Liquid Crystals,"
Phys. Lett., Vol. 35A, p. 393 (1971).
12 G. H. Heilmeier and J. Goldmacher, "A New Electric Field Controlled Reflective
Optical Storage Effect in Mixed Liquid Crystal Systems," Proc. IEEE, Vol. 57, p. 34
(1969).
13 F. Brochard, "Backflow Effects in Nematic Liquid Crystals," Mol. Cryst. and Liq.
Cryst., Vol. 23, p. 51 (1973).
14 W. Helfrich, "Molecular Theory of Flow Alignment of Nematic Liquid Crystals," J.
Chem. Phys., Vol. 50, p. 100 (1969); and "Conduction·lnduced Alignment of Nematic
Liquid Crystals: Basic Model and Stability Considerations," J. Chem. Phys., Vol. 51,
p. 4092 (1969).
15 D. Meyerhofer and A. Sussman, "The Electrohydrodynamic Instabilities in Nematic
Liquid Crystals in Low-Frequency Fields," Appl. Phys. Lett., Vol. 20, p. 337 (1972).
16 N. Felici, "Phenomenes Hydro et Aerodynamiques dans la Conduction des Dielectric
Fluides," Rev. Gen. Elect., Vol. 78, p. 717 (1969); A. Sussman, "Contribution of the
Ionic Double Layer to the DC Hydrodynamic Instabilities in Nematic Liquids," Paper
presented at Fourth International Liquid Crystal Conf., Kent, Ohio, Aug. 1972; and R. J.
Turnbull, "Theory of Electrohydrodynamic Behaviour of Nematic Liquids in a Constant
Field," J. Phys. D: Appl. Phys., Vol. 6, p. 1745 (1973).
17 A. Derzhanski and A. G. Petrov, "Inverse Currents and Contact Behaviour of Some
Nematic Liquid Crystals," Phys. Lett., Vol. 36A, p. 307 (1971).
18 G. H. Heilmeier, L. A. Zanoni, and L. A. Barton, "Dynamic Scattering: A New Electro·
Optic Effect in Certain Classes of Nematic Liquid Crystals," Proc. IEEE, Vol. 56, p. 1162
(1968).
19 W. H. DeJeu, "Instabilities of Nematic Liquid Crystals in Pulsating Electric Fields,"
Phys. Lett., Vol. 37A, p. 365 (1971); and Orsay Liquid Crystal Group, "Transition Be·
tween Conduction and Dielectric Regimes of the Electrohydrodynamic Instabilities in a
Nematic Liquid Crystal," Phys. Lett., Vol. 39A, p. 181 (1972).
20 P. Wild and J. Nehring, "Turn-on Time Reduction and Contrast Enhancement in
Matrix-addressed Liquid Crystal Valves," Appl. Phys. Lett., Vol. 19, p. 335 (1971); and
C. Stein and R. Kashnow, "A Two-Frequency Coincidence Addressing Scheme for
Nematic-Liquid-Crystal Display," Appl. Phys. Lett., Vol. 19, p. 343 (1971).
21 M. Schiekel and K. Fahrenschon, "Deformation of Nematic Liquid Crystals with Verti-
cal Orientation in Electrical Fields," Appl. Phys. Lett., Vol. 19, p. 391 (1971); R. A.
Soref and M. J. Rafuse, "Electrically Controlled Birefringence of Thin Nematic Films,"
J. Appl. Phys., Vol. 43, p. 2029 (1972); F. J. Kahn, "Electric·field·iT1duced Orientational
Deformation of Nematic Liquid Crystals: Tunable Birefringence," Appl. Phys. Lett., Vol.
20, p. 199, (1972); and M. Hareng, E. Leiba, and G. Assouline, "Effet du Champ Elec-
trique sur la Birefringence de Cristaux Liquides Nematiques," Mol. Cryst. and Liq.
Cryst., Vol. 17, p. 361 (1972).
22 M. Schadt and W. Helfrich, "Voltage-Dependent Optical Activity of a Twisted Nematic
Liquid Crystal," Appl. Phys. Lett., Vol. 28, p. 127 (1971).
23 G. H. Heilmeier and L. A. Zanoni, "Guest·Host Interactions in Nematic Liquid Crys·
tals. A New Electro-Optic Effect," Appl. Phys. Lett., Vol. 13, p. 91 (1968).
24 I. Haller, H. A. Huggins, and M. J. Freiser, "On the Measurement of Indices of Re·
fraction of Nematic Liquids," Mol. Cryst. and Liq. Cryst., Vol. 16, p. 53 (1972).
ELECTRO-OPTIC TRANSFER FUNCTION 317

25 D. J. Channin, "Optical Waveguide Modulation Using Nematic Liquid Crystals," Appl.


Phys. Lett., Vol. 22, p. 365 (1973).
26 J. P. Sheridan, J. M. Schnur, and T. G. Giallorenzi, "Electro-Optic Switching in Low-
Loss Liquid Crystal Waveguides," Appl. Phys. LeU., Vol. 22, p. 560 (1973).
27 R. A. Kashnow and C. R. Stein, "Total-Reflection Liquid-Crystal Electro-Optic De-
vice," Appl. Optics, Vol. 12, p. 2309 (1973).
28 G. Assouline, A. Dmitrieff, M. Hareng, and E. Leiba, "Diffraction d'un Faisceau Laser
par un Cristal Liquide Nematique Souvris a un champ Electrique," C. R. Acad. Sci.,
Vol. B271, p. 857 (1970).
29 W. Helfrich, "Orientation Pattern of Domains in Nematic p-azoxyanisole," J. Chem.
Phys., Vol. 51, p. 2755 (1969).
30 T. O. Carroll, "Liquid-Crystal Diffraction Grating," J. Appl. Phys., Vol. 43, p. 767
(1972).
31 C. Deutsch and P. N. Keating, "Scattering of Coherent Light from Nematic Liquid
Crystals in the Dynamic Scattering Mode," J. Appl. Phys., Vol. 40, p. 4049 (1969); and
Orsay Liquid Crystal Group, "Viscosity Measurements by Quasi-Elastic Light Scattering
in p-azoxyanisole," Mol. Cryst. and Liq. Cryst., Vol. 13, p. 187 (1971).
32 P. A. Penz, "Order Parameter Distribution for the Electrohydrodynamic Mode of a
Nematic Liquid Crystal," Mol. Cryst. and Liq. Cryst., Vol. 15, p. 151 (1971).
33 E. Jakeman and P. N. Pusey, "Light Scattering from Electrohydrodynamic Turbulence
in Liquid Crystals," Phys. Lett., Vol. 44A, p. 456 (1973); and F. Scudieri, M. Bertolotti,
and R. Bartolino, "Light Scattered by a Liquid Crystal: A New Quasi-Themal Source,"
Applied Optics, Vol. 13, p. 181 (1974).
34 D. Meyerhofer and E. F. Pasierb, "Light Scattering Characteristics in Liquid Crystal
Storage Materials," Mol. Cryst. and Liq. Cryst., Vol. 20, p. 279 (1973).
35 R. B. MacAnally, "Liquid Crystal Displays for Matched Filtering," Appl. Phys. Lett.,
Vol. 18, p. 54 (1971).
36 G. W. Taylor and W. F. Kosonocky, "Ferroelectric Light Valves for Optical Memo-
ries," Ferroelectrics, Vol. 3, p. 81 (1972).
37 H. J. Caulfield and R. A. Soref, "Optical Contrast Enhancement in Liquid Crystal
Devices by Spatial Filtering," Appl. Phys. Lett., Vol. 18, p. 5 (1971).
38 E. Tomkins, "Liquid Crystal Viewing Screen," Opt. Soc. Am. Meeting, Tucson, Ariz.
(1971).
39 A. Sussman, "Illumination Scheme for Liquid Crystal Displays," U.S. Patent pending.
40 R. A. Kashnow and H. S. Cole, "Electrohydrodynamic Instabilities in a High-Purity
Nematic Liquid Crystal," J. Appl. Phys., Vol. 42, p. 2134 (1971).
41 M. Hareng, G. Assouline, and E. Leiba, "La Birefringence Electriquement Contr61ee
dans les Cristaux Liquides Nematiques," Appl. Optics, Vol. 11, p. 2920 (1972).
42 L. T. Creagh and A. R. Kmetz, "Performance Advantages of Liquid Crystal Displays
with Surfactant-produced Homogeneous Alignment," Soc. for Information Display, 1972
International Symp. Dig. Tech. Papers (Lewis Winner, New York), p. 90; and F. J. Kahn,
"Orientation of Liquid Crystals by Surface Coupling Agents," Appl. Phys. Lett., Vol. 22,
p. 386 (1973).
43 G. Assouline, M. Hareng, and E. Leiba, "Liquid Crystal and Photoconductor Image
Converter," Proc. IEEE, Vol. 59, p. 1355 (1971); and M. Hareng, G. Assouline, and E.
Leiba, "Affichage Bicolore a Cristal Liquide (Two Color LiqUid-Crystal Display),"
Electron. Lett., Vol. 7, p. 699 (1971).
44 T. Shimojo, K. Matsuda, and K. Kasano, "Singular Electro-Optical Characteristics of
Liquid Crystal Display with Interdigital Electrodes," S.I.D. International Symp. Digest,
1973 (p. 36).
45 R. A. Kashnow, "Thickness Measurements of Nematic Liquid Layers," Rev. Sci. Inst.,
Vol. 43, p. 1837 (1972).
46 J. A. Castellano and M. T. McCaffrey, "Liquid Crystals IV. Electro-Optic Effects in
p-alkoxybenzylidene-p'-aminoalkyphenones and Related Compounds," in Liquid Crystals
and Ordered Fluids, ed. J. F. Johnson and R. S. Porter, Plenum Press, New York,
p. 293 (1970).
47 U. Bonne and D. P. Cummings, "Properties and Limitations of Liquid Crystals for
Aircraft Displays," Contract #N00014-71-C-0262, ONR Task No. NR 215-173, Honeywell,
Inc., Oct. 1972, Chap. VII.
48 G. H. Heilmeier, J. A. Castellano, and L. A. Zanoni, "Guest-Hose Interactions in
Nematic Liquid Crystals," Mol. Cryst. and Liq. Cryst., Vol. 8, p. 293 (1969).
49 J. Dryer, "Liquid Crystal Optical Devices," Reported at Second International Liq.
Cryst. Conf., Kent, Ohio, 1968.
50 H. DeVries, "Rotary Power and Other Properties of Certain Liquid Crystals," Acta.
Cryst., Vol. 4, p. 219 (1951).
318 CHAPTER 16

51 C. B. Burckhardt, M. Schadt, and W. Helfrich, "Holographic Recording with an


Electro-Optic Liquid Crystal Cell," Appl. Optics, Vol. 10, p. 2196 (1971).
52 D. W. Berreman, "Optics in Smoothly Varying Anisotropic Planar Structures: Applica-
tion to Liquid-Crystal Twist Cells," J. Opt. Soc. Am., Vol. 63, p. 1374 (1973).
53 G. H. Heilmeier, "Some Cooperative Effects in Butyl p-Anisylidene-p-Amino Cinna-
mate," in Ordered Fluids and Liquid Crystals, Advances in Chemistry Series #63, p. 68,
American Chemical Society, Washington, D.C. (1967); A. Takase, S. Sakagami, and
M. Nakamizo, "Light Diffraction in a Nematic Liquid Crystal with Positive Dielectric
Anisotropy," Japan J. Appl. Phys., Vol. 12, p. 1255 (1973); W. H. DeJeu and C. J.
Gerritsma, "Electrohydrodynamic Instabilities in Some Nematic Azoxy Compounds with
Dielectric Anisotropies of Different Sign," J. Chem. Phys., Vol. 56, p. 4752 (1972); and
Ref. 4.
54 R. Williams, "Domains in Liquid Crystals," J. Chem. Phys., Vol. 39, p. 384 (1963).
"P. A. Penz, "Voltage-Induced Vorticity and Optical Focusing in Liquid Crystals,"
Phys. Rey. Lett., Vol. 24, p. 1405 (1970).
56 P. A. Penz and G. W. Ford, "Electromagnetic Hydrodynamics of Liquid Crystals,"
Phys. Rey., Vol. 6A, p. 414 (1972).
57 H. Greubel and U. Wolff, "Electrically Controllable Domains in Nematic Liquid Crys-
tals," Appl. Phys. Lett., Vol. 19, p. 213 (1971); and l. K. Vistin, "New Electrostructural
Phenomenon in Liquid Crystals of Nematic Type," SOY. Phys. Crys., Vol. 15, p. 514
(1970).
58 A. Sussman, unpublished results.
59 Orsay Liquid Crystal Group, "Hydrodynamic Instabilities in Nematic Liquids Under ac
ElEctric Fields," Phys. Rey. Lett., Vol. 25, p. 1642 (1970).
60 R. A. Kashnow and J. E. Bigelow, "Diffraction from a Liquid Crystal Phase Grating,"
Appl. Oplics, Vol. 12, p. 2302 (1973).
61 Y. Galerne, G. Durand, M. Veyssie, and V. Pontikis, "Electrohydodynamic Instability
in a Nematic Liquid Crystal: Effect of an Additional Stabilizing ac Electric Field on the
Spatial Period of 'Chevrons'," Phys. Lett., Vol. 38A, p. 449 (1972).
62 A. Sussman, "Secondary Hydrodynamic Structure i.1 Dynamic Scattering," Appl. Phys.
Lett., Vol. 21, p. 269 (1972).
63 J. Nehring and M. S. Petty, "The Formation of Threads in the Dynamic Scattering
Mode of Nematic Liquid CrystalS," Phys. Lett., Vol. 40A, p. 307 (1972).
64 L. Goodman, "Light Scattering in Electric-Field Driven Nematic Liquid Crystals," Soc.
for Information Display 1971 International Symp. Dig. Tech. Papers, (Lewis Winner,
New York), p. 124. See also References [31] and [47].
65 l. T. Creagh, "Nematic Liquid Crystal Materials for Displays," Proc. IEEE, Vol. 61,
814 (1973); and Eastman Liquid Crystal Products, Bulletin JJ-14 (1973).
66 G. Assouline and E. Leiba, "Cristaux Liquides," Rey. Tech. CSF, Vol. 1, p. 483 (1969).

67 J. D. Margerum, J. Nimov, and S.-Y. Wong, "Reversible Ultraviolet Imaging with


Liquid Crystals," Appl. Phys. Lett., Vol. 17, p. 51 (1970).
68 D. H. White and M. Feldman, "Liquid Crystal Light Valves," Electron Letters, Vol. 6,
p. 837 (1970).
69 J. D. Margerum, T. D. Beard, W. P. Bleha, Jr., and S.-Y. Wong, "Transparent Phase
Images in Photoactivated Liquid Crystals," Appl. Phys. Lett., Vol. 19, p. 216 (1971).
70 A. D. Jacobson, "Photo-Activated Liquid Crystal Valve," Soc. for Information Display,
1972 International Symp., Dig. Tech. Papers (Lewis Winner, New York), p. 70.
71 A. Rose, "The Role of Space-Charge-Limited Currents i.1 Photoconductivity-con-
trolled Devices," IEEE Trans. on Electron Dey., Vol. ED19, p. 430 (1972).
72 T. O. Beard, W. P. Bleha, and S.-Y. WOilg, "AC Liquid Crystal Light Valve," Appl.
Phys. Lett., Vol. 22, p. 90 (1973).
73 A. Sussman, "Dynamic Scattering Life in the Nematic Compound p-Methoxybenzy-
Iidene-p-amino phenyl acetate as Influenced by the Current Density," Appl. Phys. Lett.,
Vol. 21, p. 126 (1972).
Electrochemistry in Nematic Liquid-Crystal
Solvents

Alan Sussman

RCA Solid State Division, Somerville, N. J. 08876

1. Introduction

This chapter discusses the electrolytic-solution properties of low-di-


electric-constant nematic solvents. Dissolved substances, if electro-
lytes, can contribute only a fraction of their ions to the conductance
because of equilibrium between the free ions and ion pairs. If the so-
lute forms ions through intermediate charge-transfer reactions, addi-
tional equilibria must be considered. For nematics, the solvent fluidi-
ty is anisotropic, and the conductance depends on the direction of
current flow with respect to the orientation of the fluid. The variation
of the conductance with temperature is directly related to the varia-
tion with temperature of both the ionic equilibrium and the fluidity.
319
320 CHAPTER 17

A considerable background of both theoretical and experimental


work is available.
The properties of the interface between conductors and ordinary
electrolytic solutions are exceedingly complex; Jor low-dielectric-con-
stant solvents, details of the double layer are lacking, but dimensions
may be estimated from simple theory. In cells of the dimensions of
the usual liquid-crystal devices, many of the properties of the inter-
face can assume an increased significance as the usual dimensional
differences between the bulk and the interface become less distinct.
Many problems of charge transport are incompletely solved, but
through the use of carefully purified solvents, specially prepared elec-
trodes, and well-defined experimental conditions, it is possible to
separate the contributions of bulk processes, electrode processes, and
diffusion. Some kinetic studies of transport phenomena, operating
life, and a few electrochemical reactions are discussed.
The relationship between electrolytic and hydrodynamic solution
properties is still under intensive study and is not treated in this
paper. Many instances of specific electrolyte-low-dielectric solvent
interaction need to be investigated fully. The equivalent problems in
anisotropic solvents are not completely understood. This review is
presented with that thought in mind.

2. Equilibrium Properties of Bulk Solutions

The equivalent conductance of a solution is a convenient chemical


quantity. It is defined as the hypothetical conductance of one chemi-
cal equivalent of a dissolved substance; A = a N° p.,e, where a is the
fraction of the dissolved substance (solute) in the ionic form, and N°,
p." and e are, respectively, Avogadro's number, the ionic mobility, and
the elementary charge. The equivalent conductance is related to the
conductance (J = p.,e by the relation A = 1000 (Jle, where e is the so-
lute concentration in moles per liter. With solvents of dielectric con-
stant greater than 30, solutions of simple electrolytes generally may
be expectd to be fully ionized at all concentrations, i.e., a = 1. Upon
dilution of concentrated solutions in which the mobility of the ions is
reduced by interionic forces, the variation of A with concentration
follows the general limiting law: l

A = a(Ao - SVaC) [1J


where S is the limiting slope and a - 1. Ao is obtained by extrapola-
tion of Eq. [1] to "infinite" dilution; Ao is about 50% greater than A at
0.01 mil in a typical simple electrolyte.
ELECTROCHEMISTRY IN NEMATIC SOLVENTS 321

In solvents of low dielectric constant, on the other hand, the princi-


pal variation of A on dilution results from an increase in free ion con-
centration due to the dissociation of ion pairs:2

[2J

where A +, B- and A+B- are the positive, negative, and paired ions,
respectively. The pairs are the direct result of interionic attraction,
since the electrostatic force between ions is shielded less in the low
dielectric constant solvent than in the more usual solvents. There is
no charge transfer in the formation of ion pairs. The value of a is less
than unity, and the variation of A is almost entirely controlled by the
dependence of the free-ion concentration on the solution concentra-
tion.
Writing the equilibrium constant for the reaction of Eq. [2], using
brackets to denote concentrations,

Kl = [A~-J/[A+J[B-J
en - a) 1
« 1 [3J
a2e2 "'" a 2e' a

a =&c
we see that the fraction in ionic form depends inversely on the square
root of the solute concentration. The simplifying assumption of a « 1
breaks down at concentrations near K-l; this usually occurs at dilu-
tions outside experimental range.
A second set of reactions involving the ion pairs and free ions gives
conducting triplets and still further clustering at higher concentra-
tions, resulting in a nonlinear increase in the equivalent conductance;
the presence of these reactions results in a minimum in the conduc-
tance curve (see Ref. [2], Chap. 8). Taking the formation of ion trip-
lets into account, the equivalent conductance may be written

[4J

where AQ is the limiting equivalent conductance of the ion triplets,


and K 2 is the equilibrium constant for the reactions AB+A + = A2B+
and AB+B- = AB 2-, assumed for simplicity to have identical equi-
librium properties. At concentrations below the minimum, and for e
~ 11K 1 , the functional variation of A is
322 CHAPTER 17

A = AoVl/ K1c. [6J

Substituting Eq. [6] into the relation between (J and A, we find the
current; it is proportional to the square root of the solute concentra-
tion,

[7]

where V is the voltage, A the area, and d the cell thickness. Fig. 1
shows the equivalent conductance as a function of concentration for

IO.-------~~--------------------------~
.............. 'b...,

''<l,.
X ..... <l.. .... .'Io.n
.__
~ E
~.,
= 58
6ep
-0...,",,- ~.,.J>
-~-=>-_J:Y

EOUIVALENT CONDUCTANCE (Al


I OF TETRA "0 PENTYL AMMONIUM
NITRATE IN p-AZOXYANISOLE x
E ,. 56
7'/ ,. 2 cp

16 4 16'
C mIL

Fig. 1-Equlvalent conductance of tetra-Iso-pentyl ammonium nHrate In Isotropic p-


azoxyanlsole at 152°C, (solid line). The equilibrium constant for Ion-pair for-
mation Is 2 X 10-8 mil. The data Is bracketed between calculated values of
Eq. [4] for two values of the dielectric constant. The variation of the equiva-
lent conductance with dielectric constant Is found In Eq. [8]. (Ref. [3»

the solute tetra-iso- pentyl ammonium nitrate in the nematic p -azox-


yanisole. 6 The problem of obtaining values of Ao in this case by ex-
trapolation to infinite dilution are experimentally complicated be-
cause of the very low concentrations needed to insure that a = 1. Use
can therefore be made of the semi-empirical Walden's rule:
The product of the limiting equivalent conductance and the viscosi-
ty is a constant for each solute, almost independent of temperature.
It is particularly accurate for large ions, but must be corrected slight-
ly for dielectric constant. 4 In an anisotropic solvent, the constant may
be calculated by using the appropriate viscosity. Fig. 2 shows 5 , for p-
azoxyanisole, the reciprocal viscosity (fluidity) parallel to the orien-
ELECTROCHEMISTRY IN NEMATIC SOLVENTS 323

tation, perpendicular to the orientation, and for a nonoriented sam-


ple plotted against the reciprocal temperature.
In the nematic range, the large variations in fluidity on the nematic
side of the nematic-isotropic transition may be related to pretransi-
tional phenomena; when the parallel orientation becomes more disor-
dered, the flow becomes more difficult, while for the perpendicular

TIPAN IN p-AZOXYANISOLE

50

E=0.24eV
20
A REF6 PARALLEL
E=012eV
/1.0 B REF 6 UNORIENTED
10 E =016eV
C C = 3 7 l10-2 mIl
o C = 37l10- 3 m/l
5
E C= 3 7l10- 4 mIl
F C = 3 7 l10-5 mIt
G REF 6 PERPENDICULAR
2 E=057eV

2.2 2.4 2.6


1000/ToK

Fig. 2-Fluldlty of p-azoxyanlsole versus the reciprocal temperature for flow parallel
to the orlentallon, perpendicular to the orientation, and for a nonorlented
sample. The dashed line shows the current data of Fig. 3 normalized to the
IsotropiC range. It Is suggesllve to consider that the varlallon of the acllva-
lion energy of conducllon with concentration depends on the solute Influence
on the orlentallon. (Ref. [5])

orientation, disordering makes the flow easier. This change in orien-


tation can be noticed in the current-reciprocal-temperature graph of
Fig. 3, for the tetra-iso -pentyl ammonium nitrate/p -azoxyanisole
system. Some effects on the fluidity may be related to the concentra-
tion of solute. During the measurements the sample birefringence in-
dicated that the solvent was ordered perpendicular to the electric
field, requiring that the appropriate mobility be for flow perpendicu-
lar to the orientation. With increasing solute concentration, the fluid-
ity properties begin to resemble these for flow parallel to the orienta-
tion. This may be either a direct interaction between the ions and the
solvent or may be due to the influence of the ions on the way the sur-
face affects the bulk orientation.
Ionic equilibrium was also observed in 1-cm-thick cells in the sys-
tem tetra-iso -pentyl ammonium tetra-phenyl borate/methoxy-ben-
324 CHAPTER 17

zylidene p-n -butylaniline (MBBA).6 Walden's rule gave good agree-


ment in the isotropic region, i.e., the product of the conductance and
the viscosity was constant. The samples were in the nonaligned state.
In the nematic range, the experimental values of the mobility near
the transition temperature were five times too low, increasing to ten
times too low at the lowest temperature, compared to the calculated
values. Although no attempt to order the sample was made, ordering

ACTiVATiONI ENERG~.CONOLcTIVITY
OF I
TETRA ISO PENTYL AMMONIUM NITRATE IN
p- AZOXYANISOLE

+--~
...... _+....... 3.7xI0- 2 m/l

~
++"""+~'014ev_

3 7x IO'mll
E'02geV

J(A)
3 7xlO'm/l
E'042eV

-,
10- -

-5
3.7x 10 mIl
E'045eV _

18 I I I I
20 21 2.2 23 24 2.5 26 27
1000/ToK

Fig. 3-Current as a function of the reciprocal temperature In the system tetra-iso-


pentyl ammonium nltratelp-azoxyanisole. Note how the activation energy in
the nematic range decreases with increasing solute concentration. Compare
this variation wHh the Influence of the orientation (Fig. 2). (Ref. [3])

of such samples has been observed under similar experimental condi-


tions. 7 Nonapplicability of Walden's rule has been noted under con-
ditions of unipolar charge injection, where measured values were up
to ten times lower than the calculated ones. 8
The conductance of a solution and, therefore, the current will de-
pend on the square root of the solute concentration over the range for
which ion-pair eq~ilibrium operates. In those devices for which a fi-
nite conductance is required, such as dynamic scattering and the
storage-effect devices,lO departures from Ohm's law may be observed
because of the reorientation of the fluid by field and hydrodynamic
ELECTROCHEMISTRY IN NEMATIC SOLVENTS 325

effects. If the voltage is slowly raised in a dynamic scattering cell that


was originally in the perpendicular homeotropic orientation, the cur-
rent will be controlled.by the fluidity parallel to the field, ~II, until the
threshold voltage for the reorientation to a birefringent condition is
exceeded. The current than becomes dependent on the value of the
fluidity perpehdicular to the field, ~ 1.. Between the threshold for re-
orientation (a dielectric effect) and the onset of domain formation (a
hydrodynamic effect) at another threshold voltage, the application of
a voltage pulse will result in a current transient, as shown by the ar-
rows in Fig. 4. When the voltage is raised still further, the domains

10
..
APPEARANCE OF
BIREFRINGENCE xf"

! "i//
,'I,.x
x~
x:x/x' I
x/DOMAINS
/,/
10' ~--L----'--;IO~--L----'---:IOO~---'------J
V(80Hz rmsl

Fig. 4-Current-voHage characteristic for a dynamic scaHerlng device, original orien-


tation perpendicular horneotroplc. The Ohm's law line Is drawn for mobll"y
baaed on fluidHy perpendicular to the orientation, /L.L.

become unstable and the turbulent dynamic scattering regime is en-


tered. An average mobility,

now controls the current. At sufficiently high fields, the ion-pair


equilibrium is disturbed in favor of dissociation. This, the second
Wein effect, is most easily observed in solvents of low dielectric con-
stant (see Ref. [1], Chap. 4, Sec. 7).
326 CHAPTER 17

Measurements of conductance are usually made using alternating


voltage of a frequency that falls between the dielectric relaxation fre-
quency (see Ref. 1, Chap. 4, Sec. 1), above which ions cannot contrib-
ute to the current, and the inverse transit time,l1 below which, as will
be seen later, there are complications due to polarization of the elec-
trodes.
Variation of the conductance with temperature, both in the tetra-
iso -pentyl ammonium nitrate/p- azoxyanisole and tetra-iso -pentyl
ammonium phenyl borate/MBBA systems, depends almost entirely
on the fluidity, 3 since ionic equilibrium contributes only a small fac-
tor over the ranges of the studies. Because this may not always be the
case, the variation of the equilibrium constant with temperature will
be considered.
For the association-dissociation reaction Eq. [1], the equilibrium
constant is given approximately as 4

[8J

The exponential term is the ratio of the electrostatic to thermal ener-


gy, with a the distance at which the ions can be considered paired, k
is Boltzmann's constant, and f is the average dielectric constant of
the solvent f ave • The ion size parameter a is a constant with tempera-
ture showing little variation with solvent. The variation of K 1 - 1
does depend on the variation of the dielectric constant(s) with tem-
perature. Once the values for a, f(T), IJ,(T), and IJ, (orientation) are
known for a given solute/solvent system, the current as a function of
temperature and concentration may be considered determined. Ex-
perimental and calculated results for the system tetra-iso- pentyl am-
monium bromide in a mixed solvent containing compounds of alkoxy
benzylidine-p- amine phenyl esters (APAPA family)12 are shown in
Fig. 5.
There is another equilibrium system that can lead to ionic conduc-
tion. When an electron-accepting compound is introduced into a sol-
vent that is an electron donor, the reaction to form a donor-acceptor
pair ensues: D + A =' (DA). This donor-acceptor pair may subse-
quently dissociate to give a pair of ions: (DA) =' D+ + A-. This reac-
tion is distinct from ion-pair reactions, since charge transfer does
occur formally, but the role of the low-dielectric solvent is similar in
affecting the equilibrium. A typical electron acceptor, chloranil
(tetrachloro-1,4-benzoquinone), in reacting with MBBA in the dual
role of solvent and electron donor, was able to alter the conduc-
tance. 13 In another case, equal parts of the donor hydroquinone and
ELECTROCHEMISTRY IN NEMATIC SOLVENTS 327

the acceptor p- benzoquinone were dissolved in MBBA;14 this re-


sulted in the formation of ionizable charge transfer complex: HQ +
p -BQ =' HQ . p -BQ =' HQ+ + p- BQ-. The conductance of the solu-
tion became sufficient to produce dynamic scattering. In a 1:1 mix-
ture of MBBA and p- methoxy p-n -hexylaniline, carefully purified,
the current remained constant at 0.15 J.l,A/cm 2 for 4000 hours in her-
metically sealed cells of thickness 25 J.l,m, probably because the con-
ducting species can undergo oxidation and reduction reactions by
electron transfer alone at the appropriate electrodes, thereby revers-

.,
IO ~

.
,i:" T IPA&,.
9 - l i tO'" mil
6 - '. !If 6. CAlC

D _;'1.10· 1-

.0' o - 3. 10' ·

z, za ~o H 34 l.5
IOOO/T

Fig. 5-Current as a function of the reciprocal temperature in the system tetra-iso-


pentyl ammonium bromide/APAPA RT mixture, with concentration as a pa-
rameter. The calculated current shows good agreement with the experimen-
tal data.

ing the formation reactions. The accumulation of neutral products at


the respective electrodes results in a concentration gradient, and
eventually a steady state is set up, allowing reformation of the charge
transfer complex in the bulk. No irreversible reactions can be expect-
ed except as side reactions. With unpurified MBBA, increased cur-
rents were obtained with continued operation.
328 CHAPTER 17

3. Electrochemical Reactions

Kinetic studies of some electrochemical reactions in nematic solvents


have been made, as well as studies of those solvents dissolved in
more polar solvents. It was noticed early that under dc excitation of
dynamic-scattering cells,15 the currents remained relatively constant
and the cells continued to operate for orders of magnitude longer
than would be expected for a simple faradic process, i.e., where an
electrochemical reaction was controlled by the net charge passed
through the cell.
In p- methoxybenzylidene-p -aminophenyl acetate, the dynamic
scattering life was limited by a faradic reaction of low efficiency
which produced polymeric anode films, probably by a free radical
mechanism.1 6 The probable source of the ionic conduction was ion-
ization of intrinsic impurities. Electron injection by a Schottky mech-
anism was proposed,9 but could not be correlated with the activation
energy of the conductance. A possible paradox exists, since in order
to have a high enough field at the electrodes to cause charge injection,
a concentrated electrolytic solution is required, and any injection
would be masked by ionic currents. A dilute solution, with no over-
shadowing ionic currents, would naturally have a lower field across
the double layer, so in order to achieve high enough fields for injec-
tion, large voltages must be applied to the entire cell.
The thickness 17 of the double layer, K-l = [47rac f(fkT)]-1/2, for a
solvent of dielectric constant 5.5 appears in Fig. 6 as a function of
concentration, assuming an equilibrium constant for ion-pair disso-
ciation of 10-7, which is appropriate for MBBA6 or p- azoxyanisole. 3
Estimates of the double-layer field can be made assuming that the
double-layer potential saturates at a few volts. Note that the double-
layer thickness, which depends on (ac) -1/2 = C - 1/4 is almost con-
stant and that its thickness is not insignificant when compared to
typical liquid-crystal-device cell thicknesses (10- 3 cm).
An alternative theory to Schottky emission is the creation of nega-
tive and positive ions from the solvent by oxidation and reduction
reactions at the appropriate electrodes, followed by recombination
after diffusion into the bulk. This mechanism was suggested to ex-
plain the long life of dynamic scattering in compounds of the AP AP A
series. 18 The role of added ionic compounds was unstated.
In undoped purified samples of MBBA, low-field conductance was
attributed to thermal dissociation of trace impurities,19 but at fields
greater than 1500 Vfcm, electrode processes begin to interfere.
Through the use of ion exchange membranes as an electrode coating,
injection effects were supressed. Then one observes at low fields an
ELECTROCHEMISTRY IN NEMATIC SOLVENTS 329

ohmic current due to the natural impurities, if the dissociation rate of


the impurities is fast enough to overcome the rate at which the ions
are deposited on the electrodes. Eventually, the current saturates as
the deposition rate (related to the inverse transit time) exceeds the
generation rate, and at the high fields, the current again begins to in-
crease because of protons injected from the membrane, giving an
imine which is believed to be indentical to that occuring in the first
step of MBBA hydrolysis. That species may be one of the conducting
impurities in un dried MBBA
The effects of direct current on MBBA containing 300 parts per
million of water and, therefore, traces of the hydrolysis products p-
n- butylaniline and p- anisaldehyde were that, first, the p-n butyl-
aniline disappeared by anodic oxidation, producing blackening and

(/)
(/)
!oJ
Z

"X
U

~
C<K, ,k-I .. C- 1I2
''""
~
-'
10
~:J
o
o

CONCENTRATION, mIt

Fig. 8-Double layer thickness K- 1 as a function of concentration for dielectric con-


stant 5.5 at room temperature. The effect of Ion pairing has been Included.
The thickness of the double layer over the range of Interest varies only as
the fourth root of the solute concentration.

then a rapid rise in current.20 The reaction rate depended on the cur-
rent density. Oxidation-reduction reactions of MBBA itself and of
the hydrolysis products were studied in the solvent acetonitrile. It
was found that the oxidation of MBBA and p-n- butylaniline is irre-
versible, while the anisaldehyde is not oxidized; the reduction of
MBBA and the aldehyde produce anion radicals by acceptance of sin-
gle electrons in a reversible reaction, while the p- butylaniline is not
reduced. In the solvent dimethyl formamide,24 reduction of the
MBBA resulted in a radical ion with a measured half-life of 4 sec-
onds.
330 CHAPTER 17

Observations of phenomena related to space-charge accumulation


at the electrodes during dc operation of dynamic scattering was re-
ported and attributed to nonspecific electrode processes. 22 The elec-
trode charging and discharging transients, i.e., the ionic charge accu-
mulated in the double layers, was found to be approximately 10-5
coulomb per cm 2• The kinetics of the discharge were shown to depend
on diffusion from the double layers into the bulk; the charge did not
leave the cell via the electrodes. Fig. 7 shows the charging transients

i;;:
~0r---L-~--~~-------- ____ L-~______~
Ii
.,...

.
...~~0r---L-~~--~----------__~________~
a- Ie) ' couA Ie,.., Z

~~r---~--------------------~--------~
0-
>

T IlliE

Fig. 7-Current and dynamic scallerlng transients upon application of voHage steps
and voHage reversal (typical resuHs for a low conductlvHy sample). The ex-
cess current above steady state corresponds to electrode charge of about
10-5 coulomb/cm 2 •

most easily observed when the conductance of the fluid is less than
10- 10 (ohm-cm)-I. Then the dynamic scattering disappears, also as
the charge accumulates in the double layers. If the voltage is in-
creased or if the polarity is reversed, another transient is produced,
the latter causing a reversal of the potential of the double layer and
releasing the charge, which is transported to the other double layer.
At high enough voltage, dynamic E.cattering resumes, either because
of injection from the electrodes or field-assisted dissociation in the
bulk.
ELECTROCHEMISTRY IN NEMATIC SOLVENTS 331

References

1 H. S. Harned and B. B. Owen, The Physical Chemistry of Electrolytic Solutions, 3rd edition, Rein-
hold Publishing Co., New York (1958). (A general reference for bulk properties of electrolytic solu-
tions.)
2 R. M. Fuoss and F. Assasclna, Electrolytic Conductance, Chap. XVI And XVII, Interscience Pub.,
Inc., New York (1959).
3 A. Sussman, "Ionic Equilibrium and Ionic Conduction in the System Tetra-iso-pentyl Ammonium Ni-
trate p-Azoxyanisole," Mol. Cryst. and Liq. Cryst., 14, p. 182 (1971).
4 J. Barthel, "Conductance of Electrolyte Solutions," Agew. Chern. Internat. Edit., 7, p. 260 (1968)
(special reference to nonaqueous solvents).
5 R. S. Porter and J. F. Johnson, "Orientation of Nematic Mesophases," J. Phys. Chern., 66, p.
1826 (1962).
6 A. Denat, B. Gosse and J. P. Gosse, "Etude du Cristal Liquide p-Methoxybenzilidime p-Butylani-
line," J. Chim. Phys; 2, p. 319 (1973).
7 F. Gaspard, R. Herino and F. Mondon, "Electrohydrodynamic Instabilities in DC fields of a Nematic
Liquid Crystal with Negative Dielectric Anisotropy," Chern. Phys. Lett., 25, p. 449 (1974).
6 J. C. Lacroix and R. Teoazeon, "Sur la Mesure de Mobilities loniques dans un Cristal Liquide Nema-
tique," Comptes Rendus, 278, p. 623 (1974).
9 G. H. Heilmeier, l. A. Zanoni, and l. A. Barton, "Dynamic Scattering-A New Electro-optic Effect
in Certain Classes of Nematic Liquid Crystals," Proc. IEEE,. 56, p. 1162 (1968).
10 G. H. Heilmeier and J. Goldmacher, "A New Electric Field Controlled Reflective Optical Storage
Effect in Mixed Liquid Crystal Systems," Proc. IEEE, 57, p. 34 (1969).
11 A. Sussman, "The Electro-optic Transfer Function in Nematic Liquids," Chapter 16.

12 A. Sussman, unpublished results.


13 A. I. Baise, I. Teucher, and M. M. Labes, "Effect of Charge-Transfer Acceptors on Dynamic Scat-
tering in a Nematic Liquid Crystal," Appl. Phys. Lett., 21, p. 142 (1972). See however: F. Gaspard
and R. Herino, "Comments on 'Effect of Charge-transfer Acceptors on Dynamic Scattering in a
Nematic Liquid Crystal'," Appl. Phys. Lett., 24, p. 252 (1974).
14 Y. Ohnishi and M. Ozutsumi, "Properties of Nematic Liquid Crystals Doped with Hydroquinine and
p-Benzoquinone: Long-term Dynamic Scattering Under DC Excitation," Appl. Phys. Lett., 24, p. 213
(1974).
15 G. H. Heilmeier, l. A. Zanoni, and l. A. Barton, "Further Studies of the Dynamic Scattering Mode
in liquid Crystals and Related Topics," IEEE Trans. E/ec. Dev. ED 17, p. 22 (1970).
16 A. Sussman, "Dynamic Scattering Life in the Nematic Compound p-Methoxy-benzylidene-p-
Amino Phenyl Acetate as Influenced by the Current Density," Appl. Phys. Lett., 21, p. 126 (1972).
17 P. Delahay, Double Layer and Electrode Kinetics, Chap. 3, Interscience-John Wiley & Sons, Inc.
New York (1965).
18 M. Voinov and J. S. Dunnett, "Electrochemistry of Nematic Liquid CrystalS," J. Electrochem. Soc.,
120, p. 922 (1973).
19 G. Briere, R. Herlno, and F. Mondon, "Correlation Between Chemical and Electrochemical Reac-
tivity in the Isotropic Phase of a Liquid Crystalline p-Methoxybenzilidene p-n-Butylaniline," Mol.
Cryst., 19, p. 157 (1972).
20 A. Denat, B. Gosse, and J. Gosse, "Chemical and Electrochemical Stability of p-Methoxybenzili-
dene-p-n-Butylaniline," Chern. Phys. Lett., 18, p. 235 (1973).
21 A. Lomax, R. Hirasawa, and A. J. Bard, "The Electrochemistry of the Liquid Crystal N-(p-Methoxy-
benzilldene)-p-n-Butylanaline (MBBA)," J. E/ectrochem. Soc., 119, p. 1679 (1972).
22 A. Derzhanski and A. G. Petrov, "Inverse Currents and Contact Behavior of Some Nematic liquid
Crystals," Phys. Lett." 3BA, p. 307 (1971).
Lyotropic Liquid Crystals and Biological
Membranes: The Crucial Role of Water

Peter J. Wojtowicz

RCA Laboratories, Princeton, N. J. 08540

1. Introduction

No collection of papers on liquid crystals would be complete without


some discussion of lyotropic liquid crystals and biological mem-
branes. At the present time these two subjects constitute areas of in-
tensive research effort providing a literature of rapidly increasing
size. The high current interest mandates a discussion of these topics,
but at the same time makes it very difficult to select the material to
be presented in the limited space available. The discussion in this
chapter will therefore be confined to two main topics, (1) the composi-
tion and structure of lyotropic liquid crystals and biological mem-
branes and (2) the nature of the principal interactions that give rise
to their existence and stability.
Lyotropic liquid crystals and biological membranes are similar to
the thermotropic liquid crystals described in previous chapters in that

333
334 CHAPTER 18

they are fluid phases that possess considerable molecular order. They
are quite different, however, in that they are necessarily systems of
two or more components being composed of large organic molecules
dissolved in a highly polar solvent, most often water. They are also
different because it is not the intermolecular interaction between the
molecules partaking in the order that is responsible for the formation
of the basic structures of lyotropic liquid crystals and biological
membranes. Rather it is the interaction of the organic molecules with
the aqueous solvent that is most crucial in providing the stability of
these ordered phases. This interaction with the water, the so-called
hydrophobic interaction, was first recognized in the study of the sol-
ubility of simple hydrocarbons in water.! The principles involved
were then successfully applied to the elucidation of the native confor-
mation of the complex proteins. 2 Very recently these principles have
also aided in the understanding of the structure and function of bio-
logical membranes. 3 The importance of the hydrophobic interaction,
however, has not yet been fully appreciated in the case of lyotropic
liquid crystals. 4 One of the intentions of this chapter, therefore, is to
help call attention to the decisive role of the hydrophobic interaction
in providing the stability of lyotropic liquid crystals.
The following sections of this chapter will briefly review the compo-
sition and structure of lyotropic liquid crystals and biological mem-
branes. We will then consider the different interactions that are im-
portant in determining the structure of these phases. The hydropho-
bic interaction will be examined in some detail; the properties of the
water itself will be shown to be the predominant driving force leading
to the existence of lyotropic liquid crystals and biological membranes.
Finally, we will discuss an unusual characteristic of these ordered
phases. Unlike most of the other ordered phases encountered in phys-
ics and chemistry, the lyotropics and the membranes derive their sta-
bility not from a competition between the energy and entropy, but
from a competition between two different kinds of entropy.

2. Lyotropic Liquid Crystals

The literature on the composition and structure of lyotropic liquid


crystals has been extensively reviewed by Winsor.5 A somewhat
shorter review has been given by Brown, Doane and Neff.6 Our dis-
cussions in this section will be based principally on these two papers;
both are recommended to the reader seeking additional information
or further detail.
LYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 335

2.1 Constituents of Lyotropics

Lyotropic liquid crystals are chemical systems composed of two or


more components. Specifically, they are mixtures of amphiphilic
compounds and a polar solvent, most frequently water. Amphiphilic
compounds are characterized by having in the same molecule two
groups that differ greatly in their solubility properties. One part of
the molecule will be hydrophilic, highly soluble in water or other
polar solvents, while the other portion will be lipophilic, highly solu-
ble in hydrocarbon or nonpolar solvents. In the context of our fol-
lowing discussions of the interactions responsible for the stability of
lyotropics it is perhaps more proper to call these latter groups hydro-
phobic, emphasizing their insolubility in water rather than their hy-
drocarbon solubility. Typical hydrophilic groups are -OH, -C0 2 H,
-C0 2Na, -S03K, -O(CH 2-CH 2-O)n H, -N(CH3bBr, and -P0 4-
CH 2 CH 2 -NH 2 • Typical lipophilic or hydrophobic groups are
-Cn H 2n + 1, -C 6 H4 -C n H 2n + 1, and any other radicals containing
long hydrocarbon chains, with or without aromatic rings included. An
example of an amphiphilic compound that has been studied exten-
sively is sodium laurate, whose molecular structure is displayed in
Fig. 1a. The hydrophilic portion is the carboxylic acid group (shown
ionized as it is in aqueous solution), while the hydrophobic part is the
long straight chain hydrocarbon group. In describing the structures of
aggregates of such molecules it is convenient (as shown in Fig. 1a) to
represent the hydrophilic "head" by a black dot and the hydrophobic
"tail" by the zig-zag line.
Depending on the relative strengths of the hydrophilic and lipophi-
lic tendencies of the two different parts of the molecule, amphiphilic
compounds can vary widely in their solubility behavior. They can be
predominantly hydrophilic, water soluble and hydrocarbon insoluble
(such as Cn H2n +1-C0 2 K with n == 1,2,3) or predominantly lipophilic,
hydrocarbon soluble and water insoluble (such as Cn H 2n +1-0H with
n > 12). The most striking amphiphilic properties (solubilization,
formation of micelles, formation of liquid crystals) occurs when the
hydrophilic and hydrophobic tendencies are both strong but evenly
balanced (such as in CnH2n+1-C02Na with n == 8 to 20).
Because of their dual characteristics, amphiphilic compounds are
capable of displaying remarkable solubility properties. They are solu-
ble in both water and hydrocarbons, and show strong co-solvent or
solubilization effects. Moderately concentrated soap solutions, for ex-
ample, can dissolve many different kinds or organic compounds that
will not dissolve in water alone. Similarly, some amphiphilic com-
pounds (such as Aerosol OT) dissolved in hydrocarbon can solubilize
336 CHAPTER 18

water or other polar compounds that do not ordinarily dissolve in hy-


drocarbons. The proper way to describe both situations is to say that
the amphiphilic compound acts as co-solvent for both the water and
the hydrocarbon.

2.2 Micelles
At extreme dilution, amphiphilic molecules are distributed randomly
and uniformly throughout the solution. As the concentration of the

CHS
I<±>
HSC-~-CHS
No <±>
H:zC,
CH2
o 08 /
~ / 0
C
"-C H 2 e 0-"'0
I
I

-
/ /
0
H"C

1
H"c
"-CH" "-
---
CH
/ 0
H"C H2 C
/
"-
CH"
"-0 o'C

H"C
/ /
o·c "-
CH"

CH2
"- CH2"- /
H"C (b)
(0)
/ / "-

n
H2 C H"C CH"
"- "- /
CH" CH" H"C
/ H:zC
/ "-
CH"
H"C /
"- "-
CH" H"C
CHs /
H2 C do"
CH2 "- /
HC
H2C

CH"
/

"- H"c
'"
CH
/

H"C
/
CH" "-
/
"- H"C
C)'2
CH"
"-
H"C
/
CH"
"- H"C
/ CH"
"-
H"C /
"-
CHs HSC

Fig. 1-(a) Molecular structure of a typical amphlphilic molecule, sodium laurate. (b)
Molecular structure of a typical lipid found in membranes, phosphatldylcho-
line (part b after Ref. [10]).

amphiphilic molecules is increased, however, aggregates of molecules


begin to form. Groupings of molecules called micelles arise in which
like is associated with like. Fig. 2 displays the structure of spherical
LYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 337

and cylindrical micelles. In both forms, the hydrophilic heads are as-
sociated with each other on the outer periphery of the aggregates,
while the hydrophobic tails are grouped together in the fluid-like in-
terior of the micelles. Of even greater significance, however, is the ob-
servation that the hydrophilic heads are placed in close association
with the aqueous solvent, while the hydrophobic tails are sequestered
in the interior of the micelles completely out of contact with the
water.

~~;,::\~----------

~~\\\:
.. ..
••••
~
!\....""
....,;8~ .........
"J \•••••
~ :.:.~ ----- -- ---
Fig. 2-Schematlc representation of the structures· of spherical (upper drawings) and
cylindrical (lower drawing) micelles.

Micelles are not entities composed of fixed numbers of molecules


having a fixed geometrical shape. They must be regarded as statisti-
cal in nature, in equilibrium with the surrounding amphiphilic mole-
cules, and fluctuating constantly in size and shape in response to
temperature. On dilution of the mixture, micelles dissociate rapidly,
while on concentrating the solution, more extended micellar struc-
tures appear, eventually forming the many different lyotropic liquid-
crystal phases.
The observed structure of micelles permits the rationalization of
the solubilization property of aqueous solutions of amphiphilic com-
pounds. The interior of the micelles can be thought of as small pock-
ets of essentially pure liquid hydrocarbon. The interiors of the mi-
celles, the pockets of hydrocarbon, are then capable of dissolving
other lipophilic or hydrophobic molecules added to the solution. The
action of soap in cleansing materials of oily or greasy dirt or soil is
precisely of this nature.
338 CHAPTER 18

2.3. Structures of Lyotropics

As the proportion of amphiphilic compound to water increases, an


impressive variety of different lyotropic liquid crystal phases are ob-
served. A systematic classification of the different types and their
structure is presented in References [5] and [6] to which the reader is
referred for the extensive details. In this chapter we shall only attempt
to give the flavor of the situation.
One type of lyotropic structure is the so-called "isotropic" phase.
In this phase spherical micelles form the basic unit of the liquid crys-
tal structure. These are then deployed in either a face-centered cubic
or body-centered cubic arrangement within the fluid aqueous medi-
um. Although the micelles are essentially arranged on a lattice, they
do not touch and the structure is not rigid as in a true crystal. The
presence of the intervening aqueous solvent provides sufficient fluidi-
ty so that this structure (as well as the others to be described below),
though highly ordered, is still properly classified as a liquid. Another
broad class of structures is the so-called "middle" phase. In this
phase the basic structural unit is the rod-like cylindrical micelle of es-
sentially infinite length. These are then disposed in a hexagonal ar-
rangement within the aqueous medium. Related to this category are
the phases composed of rod-like micelles of rectangular cross section
arranged in square or rectangular packings. In all of these structures
the common feature is, of course, the sequestering of the hydrophobic
tails away from contact with the water while allowing the hydrophilic
heads to reside within the aqueous solvent.
When the proportion of water to amphiphilic compound becomes
low, structures of the "reversed" type occur. The basic units here are
spherical or cylindrical reversed micelles in which the hydrophobic
tails are on the outside, while the hydrophilic heads line the interior
which contains the water. Liquid crystalline arrangements of the re-
versed micelles then occur in the several ways outlined above. In all
cases the hydrophobic tails on the exterior of the reversed micelles
are in contact only with each other; the water is sequestered in the
polar interiors.
Among the most interesting of the lyotropic phases are the "lamel-
lar" structures. The most important of these is the so-called "neat"
phase, whose structure is depicted in Fig. 3. In this structure we find
the amphiphilic molecules arranged in double layers of essentially in-
finite extent in two dimensions. The internal structure of the double
layer is such that the hydrophobic tails occupy the interior out of
contact with water, while the hydrophilic heads line the exterior in
LYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 339

contact with the fluid aqueous medium (regions denoted by A in Fig.


3). The double layers then stack periodically along the third dimen-
sion alternating with layers of the aqueous solvent. The thickness of
the double layer is somewhat less than twice the length of the amphi-
philic molecules (layer thickness is thus about 30 to 40 A). The thick-
ness of the intervening aqueous layers is about 20 A. The hydrocar-
bon region within the double layers is essentially fluid. Some experi-
ments, however, show a gradual transition from a fluid-like property
of the hydrocarbon chains to a more rigid type of behavior at lower
temperatures. 7 The overall structure of the neat phase is rather anal-
ogous to that of the smectic liquid crystals.

4 A

1flrrIrrlf1II1Il11rrJrrrrrrfl11
1111111111111111111111111i11111
4 A

l1IfIl!11IIIIIftlllflflllrrIIII
l111il1111Il1iil1111111il1i!li!
OIl A

Fig. 3-Schematlc representation of the structure of a portion of the neat phase. The
regions containing the aqueous solvent are denoted by the A's.

Transitions from one kind of phase to another occur with changes


in both the temperature and the concentration of the amphiphilic
compound in water. A typical phase diagram is that for sodium lau-
rate shown in Fig. 4. The Tc line gives the temperatures of transition
from solid to liquid or liquid crystalline phases. It can be thought of
as representing the depression of the melting point of sodium laurate
by the water. The Ti line gives the temperatures of transition from
liquid crystalline to normal liquid phases. The regions of stability of
the "neat" and "middle" phases are shown. The cross-hatched areas
are the regions of stability of the "isotropic" structures. The interme-
diate regions labelled I are most probably conjugate mixtures of the
adjacent stable phases.
340 CHAPTER 18

3. Biological Membranes

Studies of the structure and functions of biological membranes cur-


rently constitute an area of very active and intensive research. The
literature on this subject is extensive and continually growing. The
material selected for this section is primarily taken from articles by
Rothfield 8 and by Singer. 3 The reader interested in further informa-
tion is encouraged to read these articles as well as others appearing in
the same volume. Also recommended are the more popular articles by
Singer 9 and by Capaldi,1O plus the recent review by Singer.ll

300

NORMAL
250 LIQUID
SOLUTION
oo
W 20
a: NEAT
PHASE
:l
f-
<t
a: 150
w
11.
l
w
f-
100

SOLID
PHASE

50

100 90 80 70 SO 50 40 30 20 10 0
WT % SODIUM LAURATE

Fig. 4-Temperature-composition phase diagram of the sodium-Iaurate-water system


(after Refs. [5] and [6]).

In spite of the extensive knowledge that has been obtained, it is


still difficult to define a biological membrane. Rothfield,8 however,
has provided a particularly succinct description. Reproduced verba-
tim it reads: "Biological membranes are continuous structures sepa-
rating two aqueous phases. They are relatively impermeable to water-
LYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 341

soluble compounds, show a characteristic trilaminar appearance


when fixed sections are examined by electron microscopy, and con-
tain significant amounts of lipids and proteins." We shall enlarge on
this description in the following two sections.

3.1 Constituents of Membranes

The principal constituents of all biological membranes are lipids, pro-


teins, and oligosaccharides. The aqueous environment surrounding
the membrane should properly also be considered one of the major
components of membranes. The constituents that give the mem-
branes their primary structure (and account for approximately half
their mass) are the lipids. A bewildering diversity of lipids is found in
membranes, and anyone membrane will contain several different lip-
ids. The lipids observed to be present belong to a variety of classes in-
cluding phosphatidylcholine, phosphatidylethanolamine, sphin-
gomyelin, glycolipids, cholesterol, etc.
The molecular structure of one example of lipid, phosphatidylcho-
line, is shown in Fig. lb. The molecule is obviously recognized to be
amphiphilic. At the top of the drawing we find the hydrophilic
"head." In the aqueous environment of biological systems, the head is
ionized as shown; the zwitterion (hybrid ion charged both positively
and negatively) in this case is composed of phosphate and trimethyl-
amine groups. In some lipids the hydrophilic head may, however, be
an un-ionized or neutral group. Below the head is the glycerol group
(sometimes called the "backbone"). Attached to the backbone are the
twin hydrophobic "tails." The tails consist of long hydrocarbon
chains. In anyone class of lipids the chains will appear in many dif-
ferent lengths and several degrees of saturation (number of carbon-
carbon double bonds). Their means of attachment can consist of a va-
riety of covalent linkages to the glyceryl phosphate moiety. In dis-
cussing the structure of biological membranes it is convenient (as
shown in Fig. 1b) to represent the hydrophilic head as a large dark
dot and the twin hydrophobic tails as a pair of zig-zag lines.
The second major constituent of biological membranes (accounting
for approximately half their mass) is the proteins. As with the lipids,
a perplexing variety of different proteins are found. Membranes from
different sources are observed to contain large numbers of proteins of
different molecular weights, ranging from less than 15,000 to over
100,000. No single type of protein dominates. While the complete role
of the proteins in the total function of the membranes is still not
completely understood. it does seem clear that they are of lesser con-
sequence in determining the primary structure to be described below.
342 CHAPTER 18

We will, therefore, not give the proteins any further consideration


here; the same holds true for the oligosaccharides.

3.2 Structure of Membranes

In spite of the significant heterogeneity in the molecular structure of


lipids occurring in biological membranes, one important property is
common: all membrane lipids are strongly amphiphilic. Because of
this, one should expect that in the aqueous environment of biological
systems these molecules will aggregate into structures similar to the
lyotropic liquid crystals described above. This is indeed the case.
Aqueous ·solutions of phospholipids and synthetic mixtures of natu-
rally occurring lipids dissolved in water display various forms of liq-
uid-crystal behavior.
The most common structure observed is the so-called "bimolecular
leaflet" or phospholipid bilayer. The structure is analogous to the
neat phase of the lyotropic liquid crystals and is shown schematically
in Fig. 5. The most prominent feature of this structure is the arrange-

.. A

.. A

Fig. 5-Schematlc representation of the structure of a portion of the lipid bilayer com-
ponent of biological membranes. The regions containing the aqueous envi-
ronment are denoted by the A's.

ment of the lipids into two contiguous layers of essentially infinite ex-
tent, such that the hydrophobic tails occupy the interior of the bi-
layers out of contact with the aqueous environment (regions denoted
by A in Fig. 5), while the hydrophilic heads line the exterior in con-
tact with the water. The thickness of the bilayer is approximately
twice the length ofthe individual lipid molecules (40-50 A).
The bimolecular leaflet is not a rigid structural entity. In keeping
with its liquid-crystalline nature, the hydrocarbon interior of the bi-
lYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 343

layer is quite fluid. The fluidity or mobility can be viewed simply as


random movements of the individual hydrocarbon chains of the mol-
ecules in the bilayer. In general, mobility is favored at higher temper-
atures, by greater degrees of unsaturation of the hydrocarbon chains,
and by shorter chain lengths. Gradual transitions from more-fluid
states to less-fluid states are seen in several experiments when the
temperature is lowered. The implications of this fluidity (and its
changes) to the biological function of membranes are many but can-
not be examined here.
The structure of actual biological membranes cannot be as simple
as the phospholipid bilayer. Membranes do, after all, contain large
numbers of proteins of varying weights, shapes, and sizes. One of the
major problems of membrane biology is the question of the detailed
incorporation of both the lipids and the proteins into the overall
membrane structure. The experimental situation is very difficult and
no clearly observed membrane structure is available. Many models,
however, have been proposed. The basic philosophy is that a general
pattern of organization exists and that the heterogeneity and distinc-
tiveness of different membranes can be understood as variations on a
common structural theme. A particularly appealing model of this
kind has been presented by Singer and others. 3 ,9,10
The model is variously called the Lipid-Globular Protein Mosaic
model or the Fluid Mosaic model. The basic structure of the mem-
brane is assumed to be the bimolecular leaflet of lipid molecules. The
leaflet is not considered to be continuous, however. The globular inte-
gral proteins and patches of the lipid bilayer ar.e assumed to be ar-
ranged in an alternating mosaic pattern throughout the membrane.
The hydrophobic portion of the lipids and a large fraction of the non-
polar amino acid residues of the proteins are sequestered from con-
tact with water, mainly in the interior of the membrane. The hydro-
philic groups of the lipids and the ionic residues of the proteins are in
direct contact with the aqueous environment on the exterior of the
membrane. Some of the proteins lie predominantly near the surface
of the membrane, others penetrate the interior. Some of the latter
may penetrate a short distance into the inside; others may extend
clear through the entire thickness of the membrane. The saccharide
components, being hydrophilic, presumably reside on the surface in
direct contact with the aqueous environment.
The Lipid-Globular Protein Mosaic model is based on experimen-
tal data on the conformation of proteins in intact membranes and on
general thermodynamic considerations (maximization of hydrophilic
interactions and minimization of hydrophobic interactions of the lip-
344 CHAPTER 18

ids and proteins with the water). At the present time, there do not
seem to be any data or experimental results that are clearly inconsist-
ent with this model. While there is not yet any direct experimental
evidence for this structural pattern, its consistency and thermody-
namic feasibility recommend this model as a working hypothesis for
further investigation.

4. Interaction of Amphlphillc Compounds with Water

Lyotropic liquid crystals and biological membranes are ordered


assemblies of amphiphilic molecules situated in an aqueous environ-
ment. By analogy with the other ordered or condensed phases en-
countered in physics and chemistry, one might suspect that the sta-
bility of ly6tropics and membranes derive from favorable attractive
interactions between constituent amphiphilic molecules, and that the
water only serves to provide the medium in which the ordered aggre-
gates can reside. This notion is, however, incorrect. The participation
of the water is far from passive. The role of the water is, in fact, cru-
cial to the formation and stability of lyotropics and membranes. The
chief mechanism by which the water acts to promote the various or-
dered structures is the hydrophobic effect.
Simplistically stated, the hydrophobic effect may be defined as the
tendency of water to reject any contact with substances of a nonpolar
or hydrocarbon nature. The existence of this effect was first recog-
nized in the study of the extremely low solubility of hydrocarboQ.s in
water.! The principles involved were later successfully applied to the
elucidation of the native conformation of protein molecules by Kauz-
mann.2 The application of these ideas to the study of membrane struc-
tures has been advanced by Singer. 3 Recently, Tanford 4 published an
entire book on the hydrophobic effect, including the influence of this
interaction on the formation of micelles, lipid bilayers, membranes
and other ordered structures. Aside from Singer's3 and Tanford's4
statements on the decisive role of the hydrophobic effect on lyotro-
pies, the lyotropic liquid-crystal literature seems peculiarly unaware
of this phenomenon. Winsor's5 extensive review with its systematic
analysis (R-theory) of the many lyotropic phases does not take the
hydrophobic effect into account. More recent reviews 6 ,7 of lyotropic
liquid crystals do not mention the phenomenon. We hope that the
present discussion will help to advance the realization of the impor-
tance of the hydrophobic effect to lyotropics. The material of the fol-
lowing sections is taken chiefly from Ref. [3] with some assistance
from Refs. [2] and [4].
LYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 345

4.1 Solubility of Hydrocarbons in Water

Most of the thermodynamic data concerning the hydrophobic inter-


action comes from studies of the solubility of hydrocarbons in water.
An examination of these data provides an understanding of the na-
ture and magnitude of the hydrophobic effect. The knowledge gained
here can then be qualitatively applied to the more complex systems
of interest.
The thermodynamics of solute-solvent interactions is most conve-
niently described in terms of unitary quantities. 12 The unitary free
energy and unitary entropy changes accompanying some process
(such as the transfer of hydrocarbon from nonpolar solvent to water
or the transfer of hydrocarbon from pure hydrocarbon to water) are
the standard free-energy and entropy changes corrected for any
translational entropy terms (the cratic entropy) that are not intrinsic
to the interaction under consideration. The cratic entropy is simply
the entropy of mixing the solute and solvent into an ideal solution.
With the cratic contribution removed, the unitary free energy and en-
tropy contain only contributions to the thermodynamics of the pro-
cess that come from the interaction of the individual solute molecules
with the solvent.

Table l-Thermodynamic Changes -in the Transfer of Hydrocarbons from


Nonpolar Solvents to Water at 25°C (After Ref. [2])
t.Su
t.Fu t.H (caljmole
Process (caljmole) (cal/mol e) OK)

CH. in benzene -- CH. in H 2O +2600 -2800 -18


CH. in ether -- CH. in H 2O +3300 -2400 -19
CH. in CCI. -- CH. in H 2O +2900 -2500 -18
C2Hs in benzene -- C2HS in H 2O +3800 -2200 -20
C 2Hs in CCI. -- C2Hs in H 2O +3700 -1700 -18
C 2H. in benzene -- C2H. in H 2O +2920 -1610 -15
C 2H 2 in benzene -- C2H 2 in H 2O +1870 -190 -7
Liq. propane -- CaHs in H 2O +5050 -1800 -23
Liq. n-butane -- C.H 10 in H 2O +5850 -1000 -23

Table 1 contains a compilation of thermodynamic data for the


transfer of simple hydrocarbons from nonpolar solvents to water. The
unitary free energy, enthalpy, and unitary entropy changes are denot-
ed by !:lFu , !!J{ and ASu , respectively; the temperature is 25°C in all
cases. !:lFu is unfavorably positive in all examples. This is in keeping
with the empirical observation that hydrocarbons do not dissolve in
346 CHAPTER 18

water to any appreciable extent. The t:Jl are, however, exothermic,


demonstrating that the intermolecular interaction energies favor the
solution of hydrocarbon molecules in water. The unfavorable positive
b.Fu can therefore only arise because of a strongly unfavorable uni-
tary entropy (D.Fu = b.H - T b.Su ). Table 1 shows that this is indeed
the case. For all examples, b.Su is large and negative.
The low solubility of hydrocarbons in water is therefore a conse-
quence of the large decrease in unitary entropy accompanying the in-
troduction of such molecules into an aqueous environment. In their
classic investigation, Frank and Evans! concluded that this large en-
tropy decrease must be due to some kind of ordering of the water
molecules around the hydrocarbon molecules dissolved in the water.
The water molecules at the surface of the cavity created by the intro-
duction of the hydrocarbon molecule must be capable of rearranging
themselves in order to regenerate broken hydrogen bonds. In doing
so, however, they create a higher degree of order than existed in the
undisturbed water; the result of this ordering is the decrease in the
unitary entropy. It is important to realize that it is the property of
the water alone (its highly structured nature resulting from the con-
siderable degree of hydrogen bonding) that is responsible for the hy-
drophobic interaction. The hydrophobic effect is relatively insensi-
tive to the precise nature of the nonpolar solutes involved, and, fur-
thermore, is not nearly so pronounced in the case of other polar sol-
vents. Expressing the description of the hydrophobic effect in simple
terms, water rejects contact with hydrocarbon and other nonpolar
groups because not to do so would require the water to increase the
local order of its structure, thereby reducing the entropy of the sys-
tem.

4.2 Solubility of Ionized Species in Water

The interaction of the hydrophilic portions of amphiphilic molecules


with the aqueous solvent is another important factor in the stabiliza-
tion of the structures of lyotropics and membranes. Information on
the hydrophilic interaction may be obtained from thermodynamic
studies of the solubility of simple ionized molecules in water and
other polar solvents.
Simple electrostatic arguments suggest that the free energy of ion-
ized species is inversely proportional to the dielectric constant of the
medium in which the ions reside. This should be so because ions in-
teract more strongly with the molecules of polar solvents than with
those of nonpolar solvents. This trend is indeed observed in thermo-
dynamic data. Table 2 contains a compilation of the solubilities of a
LYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 347

typical charged molecule, the zwitterion of glycine, +H;:lN-CH 2 -C0 2 -,


in various solvents of decreasing polar character. Also included is the
unitary free energy of transfer of this species from water to the other
solvents. LlFu is seen to be strongly positive for all nonaqueous sol-
vents. It is clear that this ionized molecule is at a much lower free en-
ergy in contact with water than with any other solvents.

Table 2~Solubilityand Free Energy of Transfer of Glycine in Various


Solvents at 25°C (After Ref. [3])
Solvent Solubility (mole/liter) t.Fu (cal/mole)
Water 2.886
Formamide 0.0838 1680
Methanol 0.00426 3430
Ethanol 0.00039 4630
Butanol 0.0000959 5190
Acetone 0.0000305 6000

The data in Table 2 reveal another of the unique properties of the


solvent water. The interaction of the water with the ions is not simply
attributable to the high dielectric constant; formamide has almost
the same dielectric constant as water, yet it requires 1680 cal/mole of
free energy to transfer glycine from water to formamide. This
suggests that, as with the nonpolar solutes, ionic species induce sig-
nificant changes in the local order of the water structure.
Another estimate of the importance of the hydrophilic interaction
may be obtained from a consideration of the free-energy difference
between the 'ionized species in water versus the uncharged species in
nonpolar solvent. The thermodynamic data show that it is much
more favorable for the molecules to exist as ionized species in water.
For example, in the case of the carboxyl group, -C0 2 -, it takes a uni-
tary free energy of 3300 cal/mole at 25°C just to protonate it (change
it into the neutral group, -C0 2H) in water at pH 7. These consider-
ations demonstrate that it is a thermodynamic necessity for the hy-
drophilic groups of amphiphilic molecules to be ionized and in direct
contact with the aqueous environment.

4.3 Aggregation of Amphiphilic Compounds

The two fundamental thermodynamic principles described in the


previous sections may now be applied to the question of the stability
of lyotropic liquid crystals and biological membranes.
A very large cratic entropy is associated with the uniform dispersal
348 CHAPTER 18

of amphiphilic molecules throughout the aqueous medium; at ex-


treme dilution this contribution to the free energy will stabilize iso-
lated molecules. But, as the concentration of amphiphilic species is
increased, large amounts of negative unitary entropy are produced by
the hydrophobic effect. The free energy of the system can then be
more effectively minimized by having the amphiphilic molecules ag-
gregate. The cratic entropy is lost, but far larger amounts of unitary
entropy are gained by forming micelles, bilayers, or other lyotropic
structures in which the hydrophobic groups are completely seques-
tered from contact with the water. We noted in previous sections that
all the structures observed did indeed sequester the hydrophobic
groups. These structures, moreover, also arranged all the hydrophilic
groups onto the exterior of the aggregates in direct contact with the
water and away from association with the nonpolar hydrophobic
parts. Thus, the observed structures simultaneously satisfy both of
the important thermodynamic requirements: (1) minimization of the
hydrophobic interaction and (2) maximization of the hydrophilic in-
teraction with the aqueous solvent.
This interpretation is consistent with experiments involving the
addition of nonaqueous solvents to aqueous solutions of amphiphilic
compounds. The addition of 20 to 30 mole % of ethanol to solutions of
amphiphilic compounds significantly reduces the stability of micelles;
isolated molecules become much more favored. The primary effect of
the ethanol is to reduce the hydrophobic effect, thereby diminishing
the stability of aggregates. In addition, conductance measurements
show that the net charge on the amphiphilic molecules is decreased in
30% ethanol because of formation of ion pairs (with the small cations)
in the medium of lower dielectric constant. Thus, the hydrophilic in-
teraction is also reduced on addition of the less polar solvent, and this
also contributes to the destabilization of micelles.
Several other interactions play an important but secondary role in
determining the stability of aggregates of amphiphilic molecules.
These include van der Waals attraction among the hydrocarbon tails,
electrostatic repulsion between similarly charged hydrophilic heads,
electrostatic attraction between zwitterionic heads, and possible hy-
drogen bonding among the polar portions of the molecules. At any
given concentration and temperature, the interplay of these forces
plus the major hydrophobic and hydrophilic interactions determines
the precise form of aggregation of the amphiphilic molecules. A quan-
titative theory of the composition and temperature dependence of
the various structures of the lyotropic phases that includes all of the
above considerations is not yet available. A reconstruction of the R-
LYOTROPIC LIQUID CRYSTALS AND BIOMEMBRANES 349

theoryS to explicitly include the hydrophobic effect along the lines


given by Kauzmann,2 Singer,3 and Tanford 4 would be a highly desir-
able beginning.

5. Conclusion

The composition and structure of lyotropic liquid crystals and bio-


logical membranes have been examined. An understanding of the
varied structures was obtained in terms of the hydrophobic and hy-
drophilic interactions. That is, aggregates of amphiphilic molecules in
aqueous solution always form in such a manner as to minimize the
hydrophobic interaction between the hydrocarbon tails and the
water, while simultaneously maximizing the hydrophilic interaction
of the polar heads with the aqueous solvent. In this way, we saw that
the water was not merely the medium in which these phenomena take
place. Instead, it became clear that it is the unique properties of the
water itself that give rise to the crucial interactions that stabilize lyo-
tropics and membranes.
The ordered structures discussed here are also unusual from anoth-
er point of view. While most of the ordered phases encountered in
physics and chemistry owe their existence to a successful competition
between the energy and the entropy, the ordered structures discussed
here derive their stability from a competition between two different
kinds of entropy. In magnetism, superconductivity, ferroelectricity,
thermotropic liquid crystals, solid-liquid-vapor equilibrium, multi-
component phase separation, etc., the ordered phase is always one of
low energy and low entropy, while the disordered phase has high en-
ergy and high entropy. At low temperatures the free energy (given by
F = E - TS) will be minimized by the low energy ordered state. At
higher temperatures the TS term gains in importance and eventually
the free energy will be minimized by the high entropy disordered
state.
In the case of micelles, lyotropics and membranes, however, the or-
dered state is characterized by having a low cratic entropy and a high
unitary entropy. The disordered state with the molecules dispersed
throughout the solvent, on the other hand, is a state of high cratic en-
tropy and low unitary entropy (because of the hydrophobic effect).
Since the cratic and unitary entropies have different dependencies on
composition and temperature, the minimization of the free energy
will sometimes be accomplished by maximizing the cratic entropy,
and sometimes by maximizing the unitary entropy, depending on the
exact conditions.
350 CHAPTER 18

The outcome of the competition between cratic and unitary en-


tropies, as the temperature and concentration are varied, thus gives
rise to the broad features of the phase diagram shown in Fig. 4. This
situation is somewhat analogous to existence of an ordered nematic
state in the hard rod model. 13 In this case, the two kinds of entropy in
competition are the cratic and orientational entropies.

References

1 H. S. Frank and M. W. Evans, "Free Volume and Entropy in Condensed Systems III," J. Chern.
Phys., 13, p. 507 (1945).
2 W. Kauzmann, "Some Factors in the Interpretation of Protein Denaturation," Adv. Protein Chern.,
14, p. 1 (1959).
3 S. J. Singer, "The Molecular Organization of Biological Membranes," in Structure and Function of
Biological Membranes, ed. by L. I. Rothfield, Academic Press, N. Y. (1971).
4 C. Tanford, The Hydrophobic Effect, John Wiley and Sons, N. Y. (1973).

5 P. A. Winsor, "Binary and Multicomponent Solutions of Amphiphilic Compounds," Chern. Reviews,


68, p. 1 (1968).
6 G. H. Brown, J. W. Doane and V. D. Neff, A Review of the Structure and Physical Properties of Liq-
uid Crystals, CRC Press, Cleveland, Ohio (1971).
7 A. Saupe, "Liquid Crystals," in Annual Reviews of Phys. Chern., ed. by H. Eyring, Annual Reviews,
Inc., Palo Alto, Vol. 24 (1973).
8 L. I. Rothfield, "Biological Membranes: An Overview at the Molecular Level," in Structure and
Function of Biological Membranes, ed. by L. I. Rothfield, Academic Press, N. Y. (1971).
9 S. J. Singer, "The Fluid Mosaic Model of the Structure of Cell Membranes," Science, 175, p. 720
(1972).
10 R. A. Capaldi, "A Dyanmic Model of Cell Membranes," Scientific American, p. 27, March 1974.
11 S. J. Singer, "The Molecular Organization of Membranes," in Annual Review of Biochemistry, ed.
by E. E. Snell, Annual Reviews, Inc., Palo Alto, Vol. 43, 1974.
12 R. W. Gurney, Ionic Processes in Solution, Chapter 5, McGraw·HiII Book Co., Inc., N. Y. (1953).

1~ P. Sheng, "Hard Rod Model of the Nematic·lsotropic Phase Transition," Chapter 5.


Appendix

Table 1-Thermodynamic Properties of Representative Liquid Crystal Materials from Several Different Classes
Class Chemical name Acronym Transition temperatures and enthalpies*
Schiff's bases 4-Methoxybenzylidene-4' -n-butylaniline MBBA X(21,3.2)N(4S,O.1) I
or anils a 4-Ethoxybenzylidene-4' -n-butylaniline EBBA X(36,5.S)N(79,O.2) I
4-Butoxybenzylidene-4' -ethylaniline X(51,1.9)N(65.3,0.1) I
4-0ctyloxybenzylidene-4' -n-hepty laniline X(36,16)S(76,1.7)S(SS,1.5) I
4-Cyanobenzylidene-4' -n-octyloxyaniline CBOOA X(72,9.1)S(Sl,0.012b)N(106,0.2) I
Anisylidene-p-aminophenylacetate AP AP A X(S2,6.0)N(110,O.3) I
Anisylidene-p-aminophenyl-3-methylvalerate X(36,4.6)N(Sl,0.1-0.2) I
w
U1
BiphenylsC 4-n-Pentyl-4' -cyanobiphenyl PCB X(22.5,4.1)N(35,O.1-0.2) I
- 4-n-Nonyl-4' -cyanobiphenyl X( 40.5,8.0 )S( 44.5,0.1-0.3 )N( 4 7 .5,0.1-0.3) I
4-n-Pentoxy-4 -cyanobiphenyl X(4S,6.9)N(67.5,0.1) I
4-n-Octyloxy-4' -cyanobiphenyl X(54.5,5.9)S(67 ,-)N(SO,0.6) I
Azoxybenzenes d 4-4' -Dimethoxyazoxybenzene PAA X(l1S.2,7.1)N(135.5,0.1) I
4-4' -Diethoxyazoxybenzene PAP X(136.6,6A)N(167.5,0.3) I
4-4' +Di-n-heptyloxyazoxybenzene X(74A,9.S )8(95A,OA)N(124.2,0.2-0.3) I
4-4' -Di-n-dodecyloxyazoxybenzene X(81.7,10.1)S(122.0,2.9) I
*The first number in parentheses is the transition temperature in aC, the second number is the enthalpy of transition in kcal/
mol; X = crystal, S = smectic, N = nematic, and I = isotropic.
aMeasurements made at RCA Laboratories by A. W. Levine and M. T. McCaffery.
bSee D. Djurek, J. Baturic-Rubcic, artd K. Franulovic Phys. Rev. Letters 33, 1126 (1974).
cMelting point data taken from D. S. Hulme, E. P. Raynes, and K. J. Harrison, J. Chern. Soc. Chern. Commun. 1974, p. 9S.
Data for 4-n-nonyl-4'-cyanabiphenyl obtained from G. W. Gray, private communication.
dMeasurements made at RCA Laboratories by M. T. McCaffery.
352 APPENDIX

Table 2-Physieal Properties of 4-Methoxybenzylidine-4' -n-butylanaline and


4-n-Pentyl-4' -eyanobiphenyl at Room Temperature
Parameter MBBA PCB
4_7 (a) 19.7(b)
"II
"1
5.4 (a) 6.4 (b)
~X [egs units] 0.97 X 10- 7 (c)
no (5145 A) 1.5616 1.5442

~ ~
no (6328 A) 1.5443 (d) 1.5309 (d)
ne (5145 A) 1.8062 1.7360
ne (6328 A) 1.7582 1.7063
KlI [dynes] 6 X 10- 7 } 6.4 X 10- 7 (b)
K' 2 [dynes] 4 X 10-7 (e)
K33 [dynes] 7.5x 10- 7 lOX 10- 7 (b)
p [g em- 3 ] 1.088 (f)
S (Tc) 0.34 (h) 0.49 (g)
S (R:T.) 0.64 (h) 0.65 (g)
aD. Diguet, F. Rondalez, and G. Durand, C. R. Acad. Sci. B271, 924 (1970).
bMeasurements made at RCA Laboratories by D. Meyerhofer.
cH. Gasparoux and J. Proust, J. Phys. (Paris) 32,953 (1971).
dMeasurements made at Harvard University by Shen Jen and A. E. Bell.
eI. Haller, J. Chern. Phys. 57, 1400 (1972).
fl. Haller, H. A. Huggins, and M. J. Freiser, Mol. Cryst. Liq. Cryst. 16, 53
(1972).
gMeasurements made at RCA Laboratories by A. E. Bell and E. B. Priestley.
hShen Jen, Noel A. Clark, P. S. Pershan, and E. B. Priestley, Phys. Rev. Let-
ters 31, 1552 (1973).
Index

Additives, 224 Coatings,


Addressing, conductive, 220
beam scanning, 261, 270 indium-tin oxide, 220
dual frequency, 266, 267 tin oxide, 226
matrix, 261, 304 Coherence length, 104, 112
television, 268 electric, 115
thermal,273 magnetic, 112
transistor, 268, 269 magnitude, 114
Alignment, Conduction regime, 140, 253
effect of grooves, 225, 226 Contact angle, 229
homeotropic, 222, 291 Correlation length, 172, 178
homogeneous, 222 Cotton-Mouton coefficient, 195
perpendicular, 222, 223, 243, 249, 255
random parallel, 222 Density discontinuity, 70
rubbing, 224, 226, 227, 244 Devices,
sloped evaporation, 229 electrochemical reactions in, 328
uniform parallel, 222, 255 photoconductor control of, 271, 272,
314
Bend, 106 Diamagnetic anisotropy, 74
Birefringence, 242 Dielectric anisotropy, 24, 242, 298, 306,
voltage controlled, 299, 303, 306 309
Bragg scattering, cholesterics, 6, 12, 204, effects of substitu tion, 24
205 related to devices, 25
Bubbles in cells, 239 Dielectric constants, 108, 109
Bundling effect, 70 Dielectric force, 131
Dielectric regime, 253
Cheverons, 253, 303, 305, 310, 312 Dielectric relaxation, 312, 313, 326
Chiral nematic materials, 7, 203 Diffraction by domains, 308
Cholesteric order,S Director,S, 16, 32,47,84,104,204,207,
Cholesterics, 222,312
pitch,6,204 Director field, 104
propagation of light in, 205 Disinclination lines, 112, 120
waveguide regime, 212 Dissolved gases in cells, 239
Classification of liquid crystals, 4 Disordered crystals, 2
Clearing temperature, 38, 167 Dispersion relations, 284

353
354 INDEX

Display devices, 25 Free energy (cont'd)


effects of pressure, 235 cratic, 345
materials for, 24 dielectric, 136
matrix addressed, 261, 304 distortion, 104, 105, 106
Display life, 259 elastic, 136
Distortion parameters, 105 electric, 162
Distribution function, external field, 108
orientational, 35, 51, 71, 72, 73, 76, 84 Landau, 146
smectic-A, 86 magnetic, 162
translational, 85 hematic phase, 38, 51
Domains, 129,300,301,303,310,312 smectic-A phase, 93, 97
threshold for formation, 131, 139 surface, 105
Dynamic scattering, 129,251,256,293, unitary, 345
312,328,330 Freedericksz transition, 115
critical field, 118
Elastic constants, 106, 158, 159, 160 Frequency,
magnitude, 107 critical cut-off, 253, 258, 259, 261
Electric field effect, 104, 242, 292, 303, transit time, 255
385
Electro-optic transfer function, 297, 304, Gas bubbles in cells, 239
305 Gases dissolved in cells, 239
Electrochemical reactions in devices, 328 Grooves, 225, 226
Electrochemistry of nematic solvents, 319 Guest-host effect, 248, 299, 306
Electrohydrodynamic flow, 129,251
Embryo, 183 Hydrodynamic effects, 308
critical radius, 184 Hydrodynamic force, 132
stability, 184, 185 Hydrodynamic instability, 129
Entropy,
change on transition, 40, 94, 98, 167 Index of refraction, anisotropy, 244
cratic, 345, 348, 349 Interactions,
nematic phase, 38, 40, 51 amphiphilic, 335
orientational, 59, 60 electrolyte-solvent, 320
smectic-A phase, 93 hydrophilic, 335, 346
translational, 59, 60 hydrophobic, 335, 344
unitary, 345, 348, 349 intermolecular, 34,46,52,71,88
Equivalent conductance, 320, 322 lipophilic, 335
Euler-Lagrange equation, 65, 66, 111 steric, 59, 63, 70
Evaporation, sloped, 229 Ionic mobility, 325
Excluded volume, 60, 63, 64, 65, 68
Landau expansion, 153, 157, 164
Field effects, 104, 242, 292, 303, 305 Landau free energy, 146
cholesteric pitch variation, 122 Landau parameters, values, 161, 168
cholesteric-nematic transition, 120 Lateral attractions, 20
Field-induced birefringence, 242 Light scattering, 189,287,291
Fluctuations, 41, 168 Lo;al anisotropy, 107
amplitudes, 174, 180, 181, 182 Lyotropic materials, 3, 334, 338
heterophase, 182
homophase, 169 Magnetic birefringence, 193
Fluidity, 3,7,8,319,323,328 Matrix array, 262
Frank elastic constants, 106 Mean field approximation, 35, 47, 89
Free energy, Mesogens, 10, 18
classical gas of hard rods, 66, 68 broadening substituents, 21
INDEX 355

Mesogens (cont'd) Orientational distribution function, 35,51,


central linkages, 10, 11, 18 71,72,73,76,84
terminal groups, 10, 11, 19
Mesophases, 1, 12 Packaging, 219
molecular geometry, 10, 16 Paranematic susceptibility, 194
ordered structures, 16 Partition function, 36,51,61,62,63,64,
structural requiremen ts, 17 65,67,72,92,97,165
thermal stability, 19 Phase diagram, 69
Micelles, 336 lyotropics, 340
Misalignment, alkali-induced, 231 smectic-A, 87, 95
Mixtures, Phase transitions, 1, 38,59,93,98, 99,120,
calculated nematic range, 27 124,249
equilibrium properties, 320 Photoconductor-liquid crystal device, 271,
equivalent conductance, 320 272,314
for display devices, 24 Pitch of cholesterics, 6, 204
temperature ranges, 26 Plastic crystals, 2
Mobility, ionic, 325 Pleochroic dye, 248, 304, 306
Molecular elongation, 10 Polar effects, 19
Molecular order, 4 Polymorphous materials, 9
Molecular rigidity, 10, 11 Potential, mean field, 35,47,71,89,91
Multiplexing, 261 Pressure effects in cells, 235
Pretransitional behavior, 41, 168, 169,189,
Neat phase, 339 195
Nematic order, 4, 5
experimental determination, 77, 78, 79 Rayleigh ratio, 191
macroscopic, 74 Rayleigh scattering, 288, 292
microscopic, 72, 73 Reverse tilt, 247
Nematic phase, 4, 32 Reverse twist, 247
effect of broadening, 21 Rubbing, unidirectional, 224, 226, 227, 244
effect of chirality, 6,7, 16
electrochemistry, 319 Short-range order, 57, 70
homologous series, 22 Silane coupling agents, 224
symmetry, 4, 5, 32, 104 Smectic order, 7
terminal attractions, 20 Smectic phase,
effect of broadening, 21
Odd-even effect, 24 homologous series, 20, 22
Onsager equations, 65, 66 lateral attractions, 20
Optical measurements, 299 symmetry, 7, 8, 84
Optical properties, cholesterics, 203 Solutions,
Optical rotatory power, cholesterics, 205 eqUilibrium properties, 319, 320
Optical waveguides, 212, 281 equivalent conductance, 320, 322
Order parameter, 71, 75, 154, 166 Splay, 106
discontinuity, 39, 70 Storage mode, 255, 299, 314
fluctuations, 168 Structure,
macroscopic, 75 biological membranes, 342
nematic, 32, 106 cholesteric phase. 5, 6, 204, 205
numerical values, 39 lyotropics, 338
smectic-A, 86, 87 mesogens,18
temperature dependence, 37, 39 mesophases, 16
tensor, 75, 154 nematic phase, 4, 32, 74
Ordered fluid mesophases, 12 smectic phases, 7, 84
Ordered fluids, 2 Supercooling, 156, 184, 185, 186, 189
356 INDEX

Surface energy, 188, 189,227 Thermal stability (cont'd)


Surface tension, critical, 228 off-axis substituents, 22
Susceptibility, 106 size of substituents, 20
Symmetry, Thermotropic materials, 3
cholesteric phase,S Thermotropic mesogens, 10
nematic phase, 4,32, 104 Transient response, 258
smectic-A phase, 7, 84 Transition temperature, 19, 20, 21, 22,38,
smectic-B phase, 8 167
smectic-C phase, 7 Twist, 106
director field, 109
Temperature, effect on display devices, distribu tion, 111
236,237 right- and left-handed, 112, 122
Terminal attractions, 20 Twisted nematic cell, 104, 110, 212, 299
Texture, Twisted nematic effect, 245, 306
Grandjean, 222, 249, 256
scattering, 249, 256 Viscous forces, 133, 136
Thermal expansion effects, 236
Thermal stability, 17, 19 Walden's rule, 322, 324
effect of polarity, 19 Wein effect, 325
effect of polarizability, 20 Williams' domains, 129, 252

You might also like